• Photonics Research
  • Vol. 12, Issue 4, 854 (2024)
Yan-Hui Deng1, Yu-Wei Lu1、2, Hou-Jiao Zhang1, Zhong-Hong Shi1, Zhang-Kai Zhou1、3、*, and Xue-Hua Wang1、4、*
Author Affiliations
  • 1State Key Laboratory of Optoelectronic Materials and Technologies, School of Physics, Sun Yat-sen University, Guangzhou 510275, China
  • 2Quantum Science Center of Guangdong–Hong Kong–Macao Greater Bay Area (Guangdong), Shenzhen 518045, China
  • 3e-mail: zhouzhk@mail.sysu.edu.cn
  • 4e-mail: wangxueh@mail.sysu.edu.cn
  • show less
    DOI: 10.1364/PRJ.514576 Cite this Article Set citation alerts
    Yan-Hui Deng, Yu-Wei Lu, Hou-Jiao Zhang, Zhong-Hong Shi, Zhang-Kai Zhou, Xue-Hua Wang. Strong light–matter interactions based on excitons and the abnormal all-dielectric anapole mode with both large field enhancement and low loss[J]. Photonics Research, 2024, 12(4): 854 Copy Citation Text show less

    Abstract

    The room temperature strong coupling between the photonic modes of micro/nanocavities and quantum emitters (QEs) can bring about promising advantages for fundamental and applied physics. Improving the electric fields (EFs) by using plasmonic modes and reducing their losses by applying dielectric nanocavities are widely employed approaches to achieve room temperature strong coupling. However, ideal photonic modes with both large EFs and low loss have been lacking. Herein, we propose the abnormal anapole mode (AAM), showing both a strong EF enhancement of 70-fold (comparable to plasmonic modes) and a low loss of 34 meV, which is much smaller than previous records of isolated all-dielectric nanocavities. Besides realizing strong coupling, we further show that by replacing the normal anapole mode with the AAM, the lasing threshold of the AAM-coupled QEs can be reduced by one order of magnitude, implying a vital step toward on-chip integration of nanophotonic devices.

    1. INTRODUCTION

    The strong coupling system between the photonic modes of micro/nanocavities and quantum emitters (QEs; i.e., molecular excitons, quantum dots, and color centers) is a resultful way to construct room temperature quantum states that can be used to form controllable quantum bits with valuable applications in extensible solid quantum devices and chips [13]. Therefore, in the past decades, the studies of strong coupling have demonstrated advances varying from fundamental physics to integrated optoelectronics, quantum computing, and even quantum chemistry [4], giving rise to applications such as near-unity absorption [5], superconducting qubits [6], and controllable chemical reactions [7].

    Generally, improving the electric field (EF) enhancements [810] and reducing the loss of photonic modes [11] are two vital ways to achieve strong coupling. So, plasmonic nanocavities have made great progress in the study of the strong coupling, due to their natural advantages of intensively strong EF enhancements. For example, not only has room temperature strong coupling based on surface lattice resonances or propagating surface plasmon been observed in many nanoarrays [12], but visible Rabi splitting has also been achieved in metal nanoparticle dimers [13]. Even the strong coupling at single exciton level by increasing the coupling strength has been demonstrated in various plasmonic systems [14], including the single nanorod [8,9], nanoparticle-on-mirror [15], and plasmonic bowtie [13].

    On the other hand, all-dielectric nanostructures with a high-refractive index have been widely used in the study of cavity quantum electrodynamics [16] in recent years due to their abundant electromagnetic modes and low optical loss [17], such as nanospheres [18] and nanodisks [10,19]. In particular, the anapole mode (AM) in dielectric nanocavities has received widespread attention due to the suppressed far field and enhanced near field caused by nonradiative resonance behavior [2022]. The all-dielectric nanocavities have negligible ohmic losses, strong electromagnetic resonances, and higher damage thresholds, enabling many novel functionalities such as nonlinear and quantum regimes, which are regarded as being able to pave the way for nonlinear devices, quantum-entangled light sources, and optical sensors [17]. There are good results on the strong coupling achieved by quasi-bound states in the continuum based on periodic photonic structures [2325], but it is still challenging to obtain individual all-dielectric nanocavities with small loss, as well as large EF enhancements comparable to plasmonic cavities.

    To this end, we propose the silicon (Si) slotted disk-ring (SDR) nanocavity with the abnormal anapole mode (AAM) to construct a strong coupling system. To the best of our knowledge, the AAM is a novel optical mode originating from the hybridization of AM and magnetic quadrupole (MQ) mode. Therefore, the AAM has a typical near-field distribution similar to AM, which exhibits two closed-loop currents with opposite directions along the torus, and a highly curved magnetic loop leading to energy confinement inside the nanodisk [2628]. On the other side, due to the coherent coupling between the MQ mode and AM, the AAM exhibits a radiative peak with a narrower linewidth in the scattering spectrum, which is different from the nonradiative dip of AM. Compared to previous studies, it is found the AAM not only demonstrates an almost minimal loss in individual all-dielectric systems, but also supports EF enhancements comparable to most plasmonic nanocavities. Therefore, based on the AAM, the SDR-QE system can enter the strong coupling region with fewer excitons within a 10–1000 times smaller interaction region than those of previous works with all-dielectric systems. Furthermore, the lasing actions of the coupled systems of QEs and AAM (or AM) have also been theoretically investigated. Due to the advantages of AAM, the lasing threshold and output power of the SDR-QE system are, respectively, one order of magnitude lower and higher than those of the slotted disk-QE (SD-QE) system. Our findings not only contribute to the understanding of the enhanced light–matter interaction, but also help provide a promising platform for future nanophotonic applications such as nanolasers [29,30] and information processing [31].

    2. PRINCIPLES AND METHODS

    A. Numerical Simulation

    Numerical calculations are conducted using Lumerical FDTD solutions. A total-field scattered-field (TFSF) plane wave serves as the excitation source. The excitation source is a normally incident plane wave polarized along the y axis (the short side of the slot), and the origin of the coordinate system coincides with the center of the nanodisk. To simulate the nanostructure placed in an infinite space, perfectly matched layer (PML) boundary conditions are used. In our simulation, the Si nanodisk has an air slot with a width (Wair) of 10 nm and a length of 260 nm. The inner and outer radii of the nanoring are Rin=390 nm and Rout=490  nm, respectively. The height of nanodisk and nanoring is 50 nm. The dielectric constants of the Si are taken from Palik’s book. The surrounding index is n=1 for simulations. The mesh size is 4  nm×4  nm×4  nm for the region around the structures, while the mesh size is 1  nm×1  nm×1  nm and 1  nm×1  nm×0.1  nm around the slot region and the WSe2 monolayer.

    B. Two-Oscillator Fano-Like Model

    The extinction spectrum of the abnormal anapole mode (AAM) can be fitted by employing a Fano-like line shape [i.e., E(ω)=|e(ω)|2] [32] to get e(ω)=ar+j=1,2bjγjeiϕjωωj+iγj,where ar is the background amplitude, bj, γj, ϕj, and ωj are the amplitude, damping, phase, and resonant energy of the oscillator j in the resonant state. A two-oscillator (j=1, 2) model is employed to fit the extinction spectrum, and a consistency was obtained for the calculated and fitted spectra, as shown in Fig. 1(a). The Fano-like line shape of AAM is caused by the coupling of the ED and MQ modes.

    The fitted results. (a) Two-oscillator Fano fitting of the AAM of the SDR (green dashed line). The two individual oscillators are determined by curve fitting (blue solid line), and the red solid line represents the result of the numerical calculation. The SCS of (b) the SDR-WSe2 and (c) the SDR-J-agg fitted by a phenomenological coupled oscillator model. The black dashed lines represent the fitted results based on Eq. (5a), while the red solid lines represent the results of the numerical calculation.

    Figure 1.The fitted results. (a) Two-oscillator Fano fitting of the AAM of the SDR (green dashed line). The two individual oscillators are determined by curve fitting (blue solid line), and the red solid line represents the result of the numerical calculation. The SCS of (b) the SDR-WSe2 and (c) the SDR-J-agg fitted by a phenomenological coupled oscillator model. The black dashed lines represent the fitted results based on Eq. (5a), while the red solid lines represent the results of the numerical calculation.

    C. One-Oscillator Lorentzian Model

    The optical response of the exciton is described, using a classical one-oscillator Lorentzian model, as εex(ω)=ε+fωex2ωex2ω2iωγex,where ε denotes the high-frequency component of the dielectric function of the exciton, ωex and γex are transition frequency and linewidth of the exciton, and f is the oscillator strength, which is proportional to the thickness of the two-dimensional material or the concentration of molecules. We use excitons from two kinds of materials: WSe2 and J-aggregates. According to the experimental reports of WSe2 [33,34] and J-aggregates [16,35], the parameters of these two excitons were artificially unified and set to ε=2.9, ωex=1.666  eV, γex=50  meV, and f=0.4.

    D. Fitting with a Classical Coupled-Oscillator Model

    The coupled oscillator model is a phenomenological model that describes the coupling between resonators and excitons [36,37]. For the SDR-exciton system, a harmonic oscillator is used to describe the resonance of the excitons and the dielectric cavity. Here, the dielectric cavity is described as two harmonic oscillators (cavity 1 and cavity 2) whose resonant frequencies are far apart. Furthermore, the interaction between the dielectric cavity and excitons is actually the near-field interaction of three harmonic oscillators, which can be characterized by the coupling strength of g1 and g2. Due to the resonant frequency of cavity 2 being far from the transition frequency of excitons, which leads to g1g2, we also ignore the interaction between cavity 1 (ca1) and cavity 2 (ca2). Then, the motion equations of three oscillators can be written as x¨ca1(t)+γca1x˙ca1(t)+ωca12xca1+2g1x˙ex(t)=Fca(t),x¨ca2(t)+γca2x˙ca2(t)+ωca22xca2+2g2x˙ex(t)=Fca(t),x¨ex(t)+γexx˙ex(t)+ωex2xex2g1x˙ca1(t)=0,where xca1, xca2, and xex represent the coordinates of the dielectric cavity and the exciton oscillations, and Fca represents the external force. The γca1, γca2, and γex represent the damping rates of the dielectric cavity and the excitons. The ωca1, ωca2, and ωex represent the resonance frequency of the excitons. g1 and g2 represent the coupling strength between the two harmonic oscillators describing the dielectric cavity and the excitons, respectively.

    When the frequency of incident light is ω, the driving force is Fca(t)=Re(Fcaeiωt). At a steady state, xca1(ω), xca2(ω), and xex(ω) can be obtained based on frequency ω as xca1(ω)=(ωex2ω2iγexω)Fca(ω)(ω2ωca12+iγca1ω)(ω2ωex2+iγexω)4ω2g12,xca2(ω)=(ωex2ω2iγexω)Fca(ω)(ω2ωca22+iγca2ω)(ω2ωex2+iγexω)4ω2g22,xex(ω)=ig1ωFca(ω)(ω2ωca12+iγca1ω)(ω2ωex2+iγexω)4ω2g12.

    In a quasi-static approximation, the polarizability α of the nanostructures is introduced and given as αFcaxca. Here, the contribution of the excitons to the total scattering field is negligible. Thus, the scattering cross-section (SCS) of the coupled system can be obtained by solving [37] Csca(ω)=8π3k4|α|2Aω4|xca|2,where the wavevector k=ωnc, and |xca|2=|xca1+xca2|2=|xca1|2+|xca2|2+2|xca1xca2|. Substituting the expression to Eq. (4d), we can obtain the SCS as Csca(ω)Aω4{|ωex2ω2iγexω(ω2ωca12+iγca1ω)(ω2ωex2+iγexω)4ω2g12|2+|ωex2ω2iγexω(ω2ωca22+iγca2ω)(ω2ωex2+iγexω)4ω2g22|2+2|(ωex2ω2iγexω)2[(ω2ωca12+iγca1ω)(ω2ωex2+iγexω)4ω2g12][(ω2ωca22+iγca2ω)(ω2ωex2+iγexω)4ω2g22]|},with the scattering amplitude A [38]. The fitted results of the SCS based on Eq. (5a) are shown in Figs. 1(b) and 1(c). For SDR-WSe2 and SDR-J-aggregates (SDR-J-agg) heterostructures, the coupling strengths are fitted as g1=25  meV (SDR-WSe2), 29 meV (SDR-J-agg), and g2=0.0001  meV. One can find that the fitted results based on the coupled-oscillator model match better with the results of the numerical calculation. Here, due to the influence of the background refractive index of the excitons, the resonant wavelength of the nanocavity will slightly redshift. Notably, if the contribution of excitons is large enough, the scattering cross section should be accurately written as Csca(ω)=8π3k4|α|2Aω4(|xca|2+|xex|2+2|xcaxex|)=Aω4[i,j(|xi|2+|xj|2+2|xixj|)].

    3. RESULTS AND DISCUSSION

    The condition for strong coupling is that the coupling strength g between the photonic mode and resonant excitons can overcome their respective damping decays and form an energy exchange between them [33,36]. Since the coupling strength g is proportional to the localized EF enhancements, to get the g to overpass the system losses of γ [including γca and γex, indicating the linewidths of the cavity and excitons, as shown in Fig. 2(a)], there are generally two ways available: one is to improve the EF and another is to reduce the loss γ [11]. Following these two ideas, we present the AAM and start our investigation by demonstrating its larger EF enhancements and smaller system loss, when compared to the normal AM or other photonic modes in all-dielectric systems.

    From a weak coupling system to a strong coupling system based on all-dielectric nanocavities. (a) Schematic diagrams of theoretical model for strong coupling. (b-i) and (b-ii) show the corresponding absorption and extinction cross sections (ACS, black line; ECS, red line) of the SD in (b-i) and the SDR in (b-ii). The schematic of different configurations is shown in the insets. The Fano fitting of the AAM of the SDR is represented in green dashed line. (c) The EF enhancements |E|/|E0| at the top surface (dashed lines) and center (solid lines) of the SD (black lines) and the SDR (red lines), respectively. The radius of the nanodisk is Rin=220 nm. The width and length of the slot are 10 and 260 nm. The inner and outer radii of the nanoring are Rin=390 nm and Rout=490 nm. Both nanodisk and nanoring have a height that is H=50 nm. The inset shows the minimum Nc to achieve strong coupling based on the SD and the SDR, and the right axis plots the ratio of gcSD/gcSDR. The calculation details of Nc are in Appendix B. (d) The scattering cross section (SCS) of the SD (blue line), the SD-WSe2 (green line), and the SD-J-agg (red line). (e) The SCS of the SDR, the SDR-WSe2, and the SDR-J-agg. The rigorous calculation results from FDTD were well fitted by a phenomenological coupled oscillator model (black dashed line). WSe2 and J-aggregates are placed on the top surface and center of the nanodisk, respectively. The linewidth of excitons is γex=50 meV for both the WSe2 monolayer and J-aggregates.

    Figure 2.From a weak coupling system to a strong coupling system based on all-dielectric nanocavities. (a) Schematic diagrams of theoretical model for strong coupling. (b-i) and (b-ii) show the corresponding absorption and extinction cross sections (ACS, black line; ECS, red line) of the SD in (b-i) and the SDR in (b-ii). The schematic of different configurations is shown in the insets. The Fano fitting of the AAM of the SDR is represented in green dashed line. (c) The EF enhancements |E|/|E0| at the top surface (dashed lines) and center (solid lines) of the SD (black lines) and the SDR (red lines), respectively. The radius of the nanodisk is Rin=220  nm. The width and length of the slot are 10 and 260 nm. The inner and outer radii of the nanoring are Rin=390  nm and Rout=490  nm. Both nanodisk and nanoring have a height that is H=50  nm. The inset shows the minimum Nc to achieve strong coupling based on the SD and the SDR, and the right axis plots the ratio of gcSD/gcSDR. The calculation details of Nc are in Appendix B. (d) The scattering cross section (SCS) of the SD (blue line), the SD-WSe2 (green line), and the SD-J-agg (red line). (e) The SCS of the SDR, the SDR-WSe2, and the SDR-J-agg. The rigorous calculation results from FDTD were well fitted by a phenomenological coupled oscillator model (black dashed line). WSe2 and J-aggregates are placed on the top surface and center of the nanodisk, respectively. The linewidth of excitons is γex=50  meV for both the WSe2 monolayer and J-aggregates.

    As shown in Figs. 2(b-i) and 2(c), for a normal AM (i.e., the SD nanocavity), the loss is 120 meV, while the maximal EF enhancements Emax (black lines) at different locations can range from 11 (surface) to 15 (center). On the other hand, for our proposed AAM generated in the SDR nanocavity [Fig. 2(b-ii) and Fig. 2(c)], due to the constructive coherence between the AM and the MQ mode, its loss can be reduced to 34 meV (see Fig. 1 for the Fano fitting), together with its Emax at different locations ranging from 45 (surface) to 70 (center, red solid line). It is noteworthy that the EF enhancements of 45–70 are comparable to plasmonic systems, such as the nanoarray with a lattice mode [39], the dimer with a localized plasmon mode [40], and the open cavity with a surface plasmon mode [8].

    The low loss and large EF enhancements of AAM are greatly beneficial for its strong coupling with excitons, especially to meet the harsh strong coupling criterion in the emission spectrum (i.e., g>gc=γca2+γex28). Note that it can be named the emission strong coupling condition (i.e., the ESC condition in Appendix A) [11,41]. As shown in the inset of Fig. 2(c), the critical coupling strength gc for multi-exciton strong coupling reduces from 60  meV to 20  meV for an exciton with a 20 Debye dipole moment and a 50  meV linewidth at room temperature [11,42], which means the minimum Nc (i.e., the number of excitons involved in a strong coupling interaction) that meets the requirement of g>gc can be reduced from 32 to 12 (see Appendix B).

    The advantage becomes more obvious with a reduced exciton linewidth, which creates favorable conditions for lowering the required Nc to achieve strong coupling. As a result, when we compare it to the normal AM generated by the SD, we find that when the involved Nc is the same, it is impossible, using the normal AM, to achieve strong coupling effects [Fig. 2(d)]; however, applying AAM will be successful. According to the results shown in Fig. 2(e), the interaction between the SDR and excitons has reached the strong coupling region based on the criterion of both spectral splitting and energy level splitting (see Appendix A) [41,43]. In Figs. 2(d) and 2(e), the WSe2 monolayer (1 nm) is on the top surface of the SDR or SD, and the J-aggregates are filled in the slot with a three-dimensional size of Lx=160, Ly=10, and Lz=2  nm. To fit the asymmetry spectral profiles in our systems, as well as to precisely obtain the system loss and coupling strength, a two-oscillator Fano-like model and the coupled oscillators model [32,3638] are introduced to simulate and analyze the strong coupling system.

    In fact, it is difficult to generate large EF enhancements and small loss simultaneously in optical nanocavities. The reported AM has only achieved great EF enhancements, and it still suffers from high loss [44]. To explain the advantages of the AAM (i.e., a large EF and small loss), we conducted a more in-depth numerical simulation and analysis.

    As shown in Fig. 3, the optical responses of the SD and SDR, including multipole decompositions of the SCS, charge distributions, and near-field enhancements at the resonant wavelength of AM or AAM (λAM or λAAM), are investigated. Through comparative analysis, it can be intuitively found that no matter the far-field response or near-field effect, the SDR nanocavity has overwhelming advantages such as large EF and low loss. The subtle deviation between the sum of multipole contributions (Sum, black) and simulation (FDTD, red) is attributed to other neglected multipole modes and approximate analytical calculations of the SCS. There is a conspicuous dip at 744  nm (namely, the excitation of the AM).

    Optical responses of individual Si SD and Si SDR nanocavities. (a) The contributions from different expansion multipole modes to the scattering spectrum of the SD. They are ED, MD, EQ, MQ, and Sum (blue, pink, olive, cyan, and black solid lines) as well as the simulation calculation (FDTD, red dashed line). The MD contribution approaches 0, and it is covered by the olive line. The inset shows the contribution of MD and EQ. (b) Charge distribution in the SD excited by normal incidence plane wave with polarization along the y axis. The red arrow in the diagram shows the direction where the charge is moving. (c) The resonant near-field profiles on the z=0 nm (center of the slot), z=25 nm (top surface of the SD), x=0 nm, and y=0 nm planes of the SD. (d)–(f) The same contents of the SDR as that in (a)–(c), respectively. The black circles and white lines show the outlines and boundaries of the nanostructures. The origin of the coordinate system is placed at the center of the disk.

    Figure 3.Optical responses of individual Si SD and Si SDR nanocavities. (a) The contributions from different expansion multipole modes to the scattering spectrum of the SD. They are ED, MD, EQ, MQ, and Sum (blue, pink, olive, cyan, and black solid lines) as well as the simulation calculation (FDTD, red dashed line). The MD contribution approaches 0, and it is covered by the olive line. The inset shows the contribution of MD and EQ. (b) Charge distribution in the SD excited by normal incidence plane wave with polarization along the y axis. The red arrow in the diagram shows the direction where the charge is moving. (c) The resonant near-field profiles on the z=0  nm (center of the slot), z=25  nm (top surface of the SD), x=0  nm, and y=0  nm planes of the SD. (d)–(f) The same contents of the SDR as that in (a)–(c), respectively. The black circles and white lines show the outlines and boundaries of the nanostructures. The origin of the coordinate system is placed at the center of the disk.

    The AM can be understood by further investigating the Cartesian multipole contributions into the scattering [4547]. It is observed that the resonance characteristics of the SD is dominated by the electric dipole (ED, blue) mode, accompanied by weak and broad MQ (cyan) mode [Fig. 3(a)]. The contributions from the magnetic dipole (MD, pink) and electric quadrupole (EQ, olive) are even negligible, as shown in the inset. Herein, AM is equivalent to the ED, because the expression for the ED consists of the original ED (P) and the toroidal dipole (T) modes. So, one can see the AM originates from the destructive interference between P and T (see Fig. 4 and Appendix C). The formation of AM presents two opposite circular displacement currents on the left and right half of the SD [Fig. 3(b)], which is generated by bounded electron oscillations caused by the field penetration and phase retardation effects in dielectric nanoparticles [17,44]. We plot EF distributions on the z=0 (top left), y=25  nm (top right), x=0 (bottom left), and y=0 (bottom right) planes of the SD at λAM [Fig. 3(c)]. Considerable EF is localized in the slot and even on the surface of the SD. Nevertheless, the strong coupling phenomenon is still difficult to realize in a single SD system due to the large system loss.

    Cartesian multipole decomposition results of the SCS for the SDR nanocavity including the original electric dipole (P, black solid line) mode, toroidal dipole (T, red solid line) mode, and ED (blue dashed line) mode. When P and T have the same amplitude and opposite phase (i.e., P=−ikT) [47], there is the destructive interference between P and T, resulting in the anapole mode (AM).

    Figure 4.Cartesian multipole decomposition results of the SCS for the SDR nanocavity including the original electric dipole (P, black solid line) mode, toroidal dipole (T, red solid line) mode, and ED (blue dashed line) mode. When P and T have the same amplitude and opposite phase (i.e., P=ikT) [47], there is the destructive interference between P and T, resulting in the anapole mode (AM).

    The same contents of the SDR as those in Figs. 3(a)–3(c) are shown in Figs. 3(d)–3(f), respectively. A sharp peak appears at λAAM=744  nm on the SCS [Fig. 3(d)], which is consistent with the λAM in the SD. The generation of the AAM can be attributed to the coupling of the AM and the MQ mode. On the one side, the far-field is dominated by a radiative MQ mode while the nonradiative AM is still noticeable [Fig. 3(d)]. On the other side, the near-field characteristics of AM are more obvious in the SDR system, since one can find that two pairs of circular displacement currents occur simultaneously in the nanodisk and the nanoring due to the near-field coupling [Fig. 3(e)]. Such a mode with both the near-field characteristics of AM and the radiative far field is called the AAM, which brings two excellent features of low loss and large EF enhancements.

    The low loss is mainly based on the narrow linewidth (<30  meV) of the MQ mode, showing a large suppression of the radiative loss. Two factors contribute to the large EF enhancements of the AAM. First, the existence of two interacting AMs with the phase difference of π/2 brings about the retardation effect in the SDR [21], which results in a stronger near-field effect of the AAM than the AM. Second, the strong far-field responses are usually accompanied by large EF enhancements [48,49]. Therefore, compared to the results in the SD [Fig. 3(c)], more dramatic EF enhancements at the same location appear in the SDR [Fig. 3(f)]. The striking EF enhancements (45 and 70) and the lower system loss will boost the strong coupling between the SDR and the excitons. More importantly, such an advantage is robust, whether it is a material replacement or a dimensional compression (see Fig. 5 and Appendix E).

    SCS of the SDR by varying (a) the refractive index n from 3.4 to 6.0 and (c) the height H from 30 to 100 nm, respectively. (b) and (d) The corresponding EF enhancements at the center of the slot. The other parameters are the same as those in Fig. 2.

    Figure 5.SCS of the SDR by varying (a) the refractive index n from 3.4 to 6.0 and (c) the height H from 30 to 100 nm, respectively. (b) and (d) The corresponding EF enhancements at the center of the slot. The other parameters are the same as those in Fig. 2.

    After identifying the optical properties and origins of the AAM, two types of strong coupling systems have been studied based on the AAM of the SDR system. For the first type, we cover the WSe2 monolayer on the top and bottom surfaces of the SDR [Type 1 in Fig. 6(a)]. For another type, we place the excitonic J-aggregates in the slot center (Type 2). According to Figs. 6(b) and 6(c), one can find that the λAAM in the SDR can be manipulated by varying the radii of the nanodisk (RD) and the nanoring (Rs), and λAAM varies more dramatically when RD is changed compared to changing Rs.

    Strong coupling behaviors of the Si SDR-exciton hybrid system. (a) Simulation model. For Type 1, the excitons are considered as a WSe2 monolayer of 1 nm covered on the top and bottom surfaces of the SDR. For Type 2, the J-aggregates (its sizes are Lx=120 nm, Ly =10 nm, and Lz) are placed in center of the slot. (b), (c) Optical energy diagrams of the SDR by manipulating the radii (b) RD and (c) Rs of the nanodisk and nanoring, where the black dashed lines denote the λAAM. The radius of the nanoring (Rs=Rout−Rin2) is tuned, while the Rout−Rin is fixed at 100 nm for all the rings. (d) Energy dispersions (Type 1, top; Type 2, bottom) of the hybrid states extracted from the simulations [blue dot, upper polariton branch (UP); red dot, lower polariton branch (LP)] and fitted using Eq. (7a) in Appendix B (solid curves). (e) The scattering spectra of Type 2 by varying the thickness Lz of the J-aggregates.

    Figure 6.Strong coupling behaviors of the Si SDR-exciton hybrid system. (a) Simulation model. For Type 1, the excitons are considered as a WSe2 monolayer of 1 nm covered on the top and bottom surfaces of the SDR. For Type 2, the J-aggregates (its sizes are Lx=120  nm, Ly=10  nm, and Lz) are placed in center of the slot. (b), (c) Optical energy diagrams of the SDR by manipulating the radii (b) RD and (c) Rs of the nanodisk and nanoring, where the black dashed lines denote the λAAM. The radius of the nanoring (Rs=RoutRin2) is tuned, while the RoutRin is fixed at 100 nm for all the rings. (d) Energy dispersions (Type 1, top; Type 2, bottom) of the hybrid states extracted from the simulations [blue dot, upper polariton branch (UP); red dot, lower polariton branch (LP)] and fitted using Eq. (7a) in Appendix B (solid curves). (e) The scattering spectra of Type 2 by varying the thickness Lz of the J-aggregates.

    The exciton is modeled as a dispersive medium with the absorption resonance set to 744 nm and a linewidth γex of 50 meV [Fig. 6(a)]. When shifting the original λAAM by varying RD, two peaks of the Rabi splitting shift toward the same directions. By extracting the wavelengths of these peaks, the anticrossing behavior of two SDR-exciton systems can be obtained. From the fitted results, the corresponding gA and gB of Type 1 and Type 2 are 45 and 53 meV, respectively [Fig. 6(d)]. As γca decreases, the ESC condition is also significantly lowered; that is, g>gc becomes easier [8,11,41], so both Type 1 and Type 2 enter the strong coupling regimes. Because of the asymmetric Fano-like line shape of the cavity mode, the line shape of the interaction between the cavity and excitons is not a symmetrical Lorentzian type, resulting in the asymmetry of the detuned dispersion curve.

    Except the two manipulation methods above, changing the thickness of the J-aggregates (i.e., Lz) in Type 2 is also an important way to affect the width of Rabi splitting (i.e., affecting the coupling strength g) because Lz is associated with the number of excitons involved in strong coupling interaction (i.e., Nc). Due to the large EF and the small loss of the AAM, even for the case of Lz=2  nm, the g (g=53.4  meV) between the SDR and J-aggregates can satisfy the ESC condition [Fig. 6(e)]. On the other hand, for the case of Lz=1  nm, the g of the coupled system still meets the strong coupling criterion in a scattering spectrum that is called the SSC condition [11,41] [see Eq. (A5) in Appendix A].

    In previous reports, achieving a Rabi splitting of around 100 meV based on all-dielectric nanocavity-QE strong coupling systems often involves large amounts or even multiple layers of excitons [16,20]. However, for our coupling systems of SDR and J-aggregates, the Nc can be greatly reduced. We use the interaction region Vex (Vex=Lx×Ly×Lz) as a rough criterion to estimate the Nc involved in different strong coupling systems. For the coupling system with Lz=8  nm, the Vex (103  nm3) is one to three orders of magnitude less than that of similar work. The advantages of large EF enhancements and low loss in the all-dielectric SDR nanocavity are expected to supplement the drawback caused by the inherent loss in plasmon-QE strong coupling systems, thereby making it an ideal choice for integrated optoelectronic and quantum devices.

    For the design of integrated quantum devices, we also consider some realistic situations, where the Si SDR-exciton hybrid system is placed on a glass substrate with a refractive index of n=1.46. Figure 7(a) shows the SCS of system I (SDR on substrate), system II, and system III [SDR-WSe2 on substrate, the excitons are placed on the top (II) or bottom (III) surface of the SDR]. Compared to system I (blue line), obvious Rabi splitting can be observed in the SCS of systems II (green line) and III (red line), and the coupling strength (36 meV) is easy to satisfy g>gc. Certainly, the anticrossing phenomenon in the SCS of system III is also demonstrated in Fig. 7(b). These results will bring the diversity of the structure configuration in the experiment, which provides more convenience for on-chip integration.

    Strong coupling behavior and lasing action of the Si SDR-exciton on substrate. (a) The SCS of three hybrid systems: system I (SDR on substrate), system II, and system III (SDR-WSe2 on substrate), respectively. The width of the slot is changed to 30 nm when other parameters are the same as those in Fig. 2(b-ii). (b) Optical energy diagram of system III by manipulating the RD. (c) The SCS of the SDR-J-agg on substrate by varying the Lz of J-aggregates (γex=50 meV). (d) Optical energy diagram of the SDR-J-agg on substrate by manipulating the RD, where the Lz=8 nm. The insets schematically show the geometries of the different hybrid systems. (e) Lasing threshold of SD-QE (navy line) and SDR-QE (red line) versus the QE number. The solid and dashed lines represent structures on SDsub, SDRsub and without a substrate, respectively. The light green area indicates the parameter range where the SDRsub can lase while the SDsub cannot. (f) The corresponding lasing action of SDsub (top panel) and SDRsub (bottom panel) with 2000 QEs and a pump rate of P=20γ0, where γ0 is the spontaneous emission rate of QE. The dipole moment of QEs is 20 Debye, and the linewidth is γex=35 meV [11].

    Figure 7.Strong coupling behavior and lasing action of the Si SDR-exciton on substrate. (a) The SCS of three hybrid systems: system I (SDR on substrate), system II, and system III (SDR-WSe2 on substrate), respectively. The width of the slot is changed to 30 nm when other parameters are the same as those in Fig. 2(b-ii). (b) Optical energy diagram of system III by manipulating the RD. (c) The SCS of the SDR-J-agg on substrate by varying the Lz of J-aggregates (γex=50  meV). (d) Optical energy diagram of the SDR-J-agg on substrate by manipulating the RD, where the Lz=8  nm. The insets schematically show the geometries of the different hybrid systems. (e) Lasing threshold of SD-QE (navy line) and SDR-QE (red line) versus the QE number. The solid and dashed lines represent structures on SDsub, SDRsub and without a substrate, respectively. The light green area indicates the parameter range where the SDRsub can lase while the SDsub cannot. (f) The corresponding lasing action of SDsub (top panel) and SDRsub (bottom panel) with 2000 QEs and a pump rate of P=20γ0, where γ0 is the spontaneous emission rate of QE. The dipole moment of QEs is 20 Debye, and the linewidth is γex=35  meV [11].

    The numerical and fitted results of the Purcell factor for the SD (navy solid line) and the SDR (red solid line) at z=0 nm. The fitted results of the Purcell factor (green dashed lines). (a) and (b) represent structures without SD, SDR and on substrate SDsub, SDRsub, respectively. This comparison shows good agreement.

    Figure 8.The numerical and fitted results of the Purcell factor for the SD (navy solid line) and the SDR (red solid line) at z=0  nm. The fitted results of the Purcell factor (green dashed lines). (a) and (b) represent structures without SD, SDR and on substrate SDsub, SDRsub, respectively. This comparison shows good agreement.

    Similar to Fig. 6(e), the J-aggregates can be deposited into the slot to construct a strong coupling system of SDR and J-aggregates. As the thickness of the deposited J-aggregates increases, the Rabi splitting in the SCS becomes more noticeable [Fig. 7(c)]. The Rabi splitting can reach a staggering value of 104 or even 153 meV, which not only surpasses the strong coupling results reported in all-dielectric systems [20,50], but also can be comparable to the Rabi splitting of the plasmon-exciton strong coupling systems [51]. The anticrossing phenomenon in the SCS still exists when Lz=8  nm [Fig. 7(d)]. It is noticed that larger EF enhancements exist in the center of air slot rather than in its bottom. Therefore, if one chooses the atom clusters as exciton components, allowing them to be suspended in the center of the air slot, more greater values of Rabi splitting can be anticipated.

    Besides the room temperature strong coupling, the features of low loss and large EF enhancements of AAM in the SDR can benefit other applications, such as lasing. Compared to the SD-QE system, three advantages of the SDR-QE in generating lasing can be found. First, the minimal QE numbers to generate lasing are, respectively, 560 and 1600 for the SDR-QE and SD-QE systems. Second, for setting a QE number, the lasing threshold of an SDR-QE system can be at least one order of magnitude lower than that of the SD-QE system [Fig. 7(e)]. Third, in the aspect of lasing output, the SDR-QE system has a faster lasing response and higher output power [Fig. 7(f)]. The response time of the SDR-QE system is 7 ps, which is much faster than that of the SD-QE system (45 ps). Meanwhile, the stable output power of the SDR-QE system is 0.3 mW, which is one order of magnitude higher than that of the SD-QE. These results exhibit the great potential of SDR with AAM in building nanolasers with a low threshold, a high output power, and a fast response. The technical details for the calculation of lasing generation are given in Fig. 8 and Appendix D.

    4. CONCLUSION

    In summary, the resonance coupling between the excitons and the all-dielectric SDR nanocavity with AAM is theoretically investigated. The multipole decompositions and near-field distributions reveal that the AAM is caused by the constructive interference between the common anapole and MQ modes. The calculation results show that the AAM exhibits strong EF enhancements (70-folds) comparable to plasmonic nanocavities, and a low loss (34  meV) that is smaller than a half of the previous best record of isolated all-dielectric nanocavities in the visible region. Therefore, the SDR-QE system can reach a strong coupling with a small number (12) of excitons within an interaction region that is one to three orders of magnitude smaller than that of previous works with all-dielectric systems. Furthermore, compared to the SD-QE system, a lasing threshold at least 10 times lower and an output power one order of magnitude higher can be obtained in the SDR-QE system, showing significantly advanced performance in building a nanolaser. Our findings can further the basic understanding of light–matter interactions in nanosystems and provide avenues to develop nanodevices such as quantum manipulation, low-threshold lasers, and optoelectronic devices.

    APPENDIX A: THE CALCULATION OF RABI SPLITTING

    The coupled-oscillator model is used to imitate the polariton dispersion [10,43]. The eigen-energies of coupled modes can be calculated by (ωcaiγca2ggωexiγex2)(αβ)=ω(αβ).

    Here, ωca and ωex represent the original resonance frequencies of the dielectric nanocavity and excitons, respectively; γca and γex represent the linewidth of the dielectric nanocavity and the excitons extracted from the FDTD simulation (fitted results); g is the coupling strength between the nanocavity and the excitons. The eigenvector components α and β are from the upper and lower hybrid states where the eigenvector components should satisfy |α|2+|β|2=1. Then the eigenenergies can be written as ω±=ωca+ωex2i(γca+γex)4±4g2+(ωcaωexiγca2+iγex2)22.The Rabi splitting energy at ωca=ωex can be calculated as Ω=4g2(γcaγex)24, and the strong coupling condition of the level splitting is expressed as g2>(γcaγex)216.

    The emission strong coupling condition, at which the Rabi splitting occurs in the emission spectrum, is given by [11,41] g>gc=γca2+γex28,where gc is defined as the critical coupling strength.

    The scattering strong coupling conditions in the two channels, at which the Rabi splitting appears in the scattering spectrum, are, respectively, given by [41] Ωca=2g(1+γexγca)(g2+γexγca4)12(g2+γexγca4)γexγca,if  g2>γex28(1+γca/2γex),Ωex=2g(1+γcaγex)(g2+γexγca4)12(g2+γexγca4)γcaγex,if  g2>γca28(1+γex/2γca).

    The larger the dissipation rate of the subsystem is, the more its contribution is to the hybridization spectrum. If γca>γex, Eq. (A4) can be used to confirm that the system has achieved strong coupling. If γca<γex, then Eq. (A5) is used.

    APPENDIX B: THE METHOD TO ESTIMATE THE MINIMUM NUMBER OF EXCITONS INVOLVED IN THE STRONG COUPLING

    For this purpose, the ratio (gc/g0) of the critical coupling strength gc to the coupling rate g0 of a quantum emitter (QE) is defined as the minimum Nc (i.e., the number of excitons involved in strong coupling interaction). The gc can be obtained by using Eq. (A3), and the g0 is calculated as explained below.

    With the classical Green’s function G, which is the solution of the wave equation with a point source at position r, we get ××G(r,r,ω)k02ε(r,ω)G(r,r,ω)=k02Iδ(rr),where I is the unit tensor, k0=ω/c is the wave number, ω is the angular frequency, and c is the speed of light. The Purcell factor (PF) is expressed as [52] PF=Im[n·G(rA,rA,ω)·n]G0,where rA is the location of a QE, n is the unit vector, and G0=Im[n·Gvac(rA,rA,ω)·n]=k03/6π, with Gvac being the Green’s function in vacuum. Then the local coupling strength is given by γ(ω)=1ε0dA·Im[G(rA,rA,ω)]·dA=PFd2ε0G0, where dA=dn with d being the dipole moment of QE. γ(ω) is expected to have a Lorentzian line shape for a single-mode cavity, i.e., γ(ω)=g02γca2(ωωca)2+(γca2)2. Therefore, we can determine the coupling rate g0, the resonant frequency ωca, and the linewidth of cavity γca by curve fitting. For a QE with a dipole moment of d=20 Debye, which is a typical value of quantum dots [42] and J-aggregates [11], we evaluate the parameters based on Fig. 8, as shown in Table 1.

    Curve Fitting Parameters

     SDRSDSDRsubSDsub
    Resonant frequency ωca1.642 eV1.69 eV1.637 eV1.653 eV
    Cavity linewidth γca34.5 meV165 meV87.46 meV233.4 meV
    Coupling rate g01.642 meV1.942 meV1.565 meV1.791 meV

    APPENDIX C: MODE ANALYSIS

    We calculate the contribution of various multipole modes to the SCS of nanoparticles based on the multipole decomposition method. Cartesian multipole moments can be expressed by adopting the induced currents J^ω(r^) as [46] electric dipole moment ED to get pα=1iω{d3r^Jαωj0(kr)+k22d3r^[3(r^·J^ω)rαr2Jαω]j2(kr)(kr)2}.The magnetic dipole moment MD is expressed as mα=32d3r^(r^×J^ω)αj1(kr)kr.The electric quadrupole moment EQ is Qαβe=3iω{d3r^[3(rβJαω+rαJβω2(r^·J^ω)δαβ]j1(kr)kr+2k2d3r^[5rαrβ(r^·J^ω)(rαJβ+rβJα)r2+r2(r^·J^ω)δαβ]j3(kr)(kr)3},and the magnetic quadrupole moment MQ is Qαβm=15d3r^[rα(r^×J^ω)β+rβ(r^×J^ω)α]j2(kr)(kr)2,where α,β=x,y,z, while ω, k, c, r^ correspond to angular frequency, wavenumber, speed of light, and location, respectively. The induced electric current density is obtained by adopting J^ω(r^)=iωε0(εr1)Eω(r^), where Eω(r^) is the electric field distribution, ε0 is vacuum permittivity, and εr is the relative permittivity. Here, a harmonic time dependence exp(iωt) of the induced current is omitted. The electric field distributions Eω(r^)/Einc can be obtained by the FDTD simulation. Einc is the electric field of the incident wave. j1(kr), j2(kr), and j3(kr) are the spherical Bessel functions of the first, second, and third kinds, respectively. The SCS generated by multipole moments can be written as Cscatotal=Cscap+Cscam+CscaQe+CscaQm+=k46πε02|Einc|2[α(|pα|2+|mα|2c)+1120α(|kQαβe|2+|kQαβec|2)+],where pα, mα, Qαβe, and Qαβm are the electric and magnetic dipole moments as well as the electric and magnetic quadrupole moments, respectively.

    The original electric dipole moment P [28] is pα=1iωd3r^Jαω,and the toroidal dipole moment T is Tα=110cd3r^[(r^·J^ω)rα2r2Jαω].

    The power radiated from each multipole can be written as Ip=μ0ω412πc|P|2 and IT=μ0ω612πc3|T|2, where μ0 is vacuum permeability. The SCS for each multipole is Csca=II0, where I0 is the intensity of the excitation plane wave, which can be represented by the electric field amplitude E0 as I0=ε0cE022.

    APPENDIX D: THE MAXWELL–BLOCH EQUATIONS FOR LASING ACTION

    Here, we provide the technical details to study the lasing dynamics. The SD and SDR can be considered as a single-mode cavity, as shown in Appendix B. The Hamiltonian of the system under the rotating wave approximation (RWA) is written as H=ωccc+ω0kNσ+kσk+kNg0k(cσk+σ+kc),where σk and c are the lowering operator of the kth QE and the annihilation operator of cavity mode, respectively. N is the number of QEs. The expectation value of operator O is given by O˙=Tr[iO[ρ,H]+OLLi(ρ)], where ρ is the density matrix, and the Lindblad super-operator LLi(ρ) takes the form LLi(ρ)=12(2LiρLi{LiLi,ρ}),with L1=γ0ηPσ+ for the QE pump, L2=γ0σ for the QE spontaneous emission, L3=γpσz for the QE dephasing, and L5=κcc for the microcavity decay. We thus can obtain the Maxwell–Bloch equations c˙=i(ωciκc2)cikNg0kσk,σ˙k=i(ω0iγ2)σk+ig0kσzkc,σ˙zk=2igck(cσkσ+kc)γ0(σzk+1)+γ0ηP(1σzk),where γ0ηP stands for the incoherent pump rate and γ=2γp+γ0(ηP+1) is the polarization decay of QEs, with η being the EF enhancement factor. We have neglected the QE field correlation by factorizing the second-order expectation values σzkc, cσk, and σ+kc into the products of the first-order expectation values σzkc, cσk, and σ+kc, respectively. We also omit the angle brackets · that indicate the expectation values.

    Considering the case of identical QEs and the resonant QE-cavity coupling, we can obtain the lasing threshold from the Maxwell–Bloch equation Pth=γ0ηn¯+vpn¯vp,where vp=γp/γ0 and n¯=NC/γ0, with C=2g02/κc being the Purcell rate of transfer of population from the QE to the lasing mode. With the parameters in Appendix B, we can obtain the results of Fig. 7(e) and the lasing dynamics shown in Fig. 7(f) by numerically solving the Maxwell–Bloch equations. The output power in Fig. 7(f) is given by Pout=ωcκc|c|2, which is the power of laser field exciting the cavity [53].

    APPENDIX E: THE ROBUST ADVANTAGES OF THE SDR

    The all-dielectric SDR nanocavity with the AAM has robust advantages in reducing radiative loss and improving the EF enhancement. It can be seen from Fig. 5 that, on the one hand, whether the refractive index n of the material (diversity of materials) or the height H of the SDR (ultrathin mate devices) is changed, the radiation linewidth of the AAM is almost unchanged and the EF enhancement is still high. On the other hand, changing the resonant frequency of the SDR will also create the condition for strong coupling between the SDR nanocavity and various excitons.

    References

    [1] A. Reiserer, N. Kalb, G. Rempe. A quantum gate between a flying optical photon and a single trapped atom. Nature, 508, 237-240(2014).

    [2] Z.-K. Zhou, J. Liu, Y. Bao. Quantum plasmonics get applied. Prog. Quantum Electron., 65, 1-20(2019).

    [3] W. Zhang, J.-B. You, J. Liu. Steering room-temperature plexcitonic strong coupling: a diexcitonic perspective. Nano Lett., 21, 8979-8986(2021).

    [4] Q. Zhao, W.-J. Zhou, Y.-H. Deng. Plexcitonic strong coupling: unique features, applications, and challenges. J. Phys. D, 55, 203002(2022).

    [5] D. Jariwala, A. R. Dayoyan, G. Tagliabue. Near-unity absorption in van der Waals semiconductors for ultrathin optoelectronics. Nano Lett., 16, 5482-5487(2016).

    [6] X. Liu, M. C. Hersam. 2D materials for quantum information science. Nat. Rev. Mater., 4, 669-684(2019).

    [7] J. A. Hutchison, T. Schwartz, C. Genet. Modifying chemical landscapes by coupling to vacuum fields. Angew. Chem. Int. Ed. Engl., 51, 1592-1596(2012).

    [8] R. Liu, Z.-K. Zhou, Y.-C. Yu. Strong light-matter interactions in single open plasmonic nanocavities at the quantum optics limit. Phys. Rev. Lett., 118, 237401(2017).

    [9] J.-Y. Li, W. Li, J. Liu. Room-temperature strong coupling between a single quantum dot and a single plasmonic nanoparticle. Nano Lett., 22, 4686-4693(2022).

    [10] K. As’ham, I. Al-Ani, L. Huang. Boosting strong coupling in a hybrid WSe2 monolayer–anapole–plasmon system. ACS Photon., 8, 489-496(2021).

    [11] W. Li, R. Liu, J. Li. Highly efficient single-exciton strong coupling with plasmons by lowering critical interaction strength at an exceptional point. Phys. Rev. Lett., 130, 143601(2023).

    [12] H. Memmi, O. Benson, S. Sadofev. Strong coupling between surface plasmon polaritons and molecular vibrations. Phys. Rev. Lett., 118, 126802(2017).

    [13] K. Santhosh, O. Bitton, L. Chuntonov. Vacuum Rabi splitting in a plasmonic cavity at the single quantum emitter limit. Nat. Commun., 7, 11823(2016).

    [14] G. Zengin, M. Wersall, S. Nilsson. Realizing strong light-matter interactions between single-nanoparticle plasmons and molecular excitons at ambient conditions. Phys. Rev. Lett., 114, 157401(2015).

    [15] R. Chikkaraddy, B. de Nijs, F. Benz. Single-molecule strong coupling at room temperature in plasmonic nanocavities. Nature, 535, 127-130(2016).

    [16] H. Wang, Y. Ke, N. Xu. Resonance coupling in silicon nanosphere–J-aggregate heterostructures. Nano Lett., 16, 6886-6895(2016).

    [17] A. I. Kuznetsov, A. E. Miroshnichenko, M. L. Brongersma. Optically resonant dielectric nanostructures. Science, 354, aag2472(2016).

    [18] S. Lepeshov, M. Wang, A. Krasnok. Tunable resonance coupling in single Si nanoparticle-monolayer WS2 structures. ACS Appl. Mater. Interfaces, 10, 16690-16697(2018).

    [19] R. Verre, D. G. Baranov, B. Munkhbat. Transition metal dichalcogenide nanodisks as high-index dielectric Mie nanoresonators. Nat. Nanotechnol., 14, 679-683(2019).

    [20] S.-D. Liu, J.-L. Fan, W.-J. Wang. Resonance coupling between molecular excitons and nonradiating anapole modes in silicon nanodisk-J-aggregate heterostructures. ACS Photon., 5, 1628-1639(2018).

    [21] Y.-H. Deng, Z.-J. Yang, M.-L. Hu. Boosting an anapole mode response through electromagnetic interactions beyond near-field limit in individual all-dielectric disk-ring nanostructures. New J. Phys., 23, 023004(2021).

    [22] A. E. Miroshnichenko, A. B. Evlyukhin, Y. F. Yu. Nonradiating anapole modes in dielectric nanoparticles. Nat. Commun., 6, 8069(2015).

    [23] M. Qin, J. Duan, S. Xiao. Manipulating strong coupling between exciton and quasibound states in the continuum resonance. Phys. Rev. B, 105, 195425(2022).

    [24] M. Qin, S. Xiao, W. Liu. Strong coupling between excitons and magnetic dipole quasi-bound states in the continuum in WS2-TiO2 hybrid metasurfaces. Opt. Express, 29, 18026-18036(2021).

    [25] K. L. Koshelev, S. K. Sychev, Z. F. Sadrieva. Strong coupling between excitons in transition metal dichalcogenides and optical bound states in the continuum. Phys. Rev. B, 98, 161113(2018).

    [26] C. Gao, S. You, Y. Zhang. Strong coupling of excitons and electric/magnetic toroidal dipole modes in perovskite metasurfaces. Opt. Express, 31, 34143-34153(2023).

    [27] S. You, Y. Zhang, M. Fan. Strong light-matter interactions of exciton in bulk WS2 and a toroidal dipole resonance. Opt. Lett., 48, 1530-1533(2023).

    [28] J. Wang, W. Yang, G. Sun. Boosting anapole-exciton strong coupling in all-dielectric heterostructures. Photon. Res., 10, 1744-1753(2022).

    [29] X. Li, W. Liu, Y. Song. Two-photon-pumped high-quality, single-mode vertical cavity lasing based on perovskite monocrystalline films. Nano Energy, 68, 104334(2020).

    [30] K. Wang, S. Sun, C. Zhang. Whispering-gallery-mode based CH3NH3PbBr3 perovskite microrod lasers with high quality factors. Mater. Chem. Front., 1, 477-481(2017).

    [31] J. S. T. Gongora, A. E. Miroshnichenko, Y. S. Kivshar. Anapole nanolasers for mode-locking and ultrafast pulse generation. Nat. Commun., 8, 15535(2017).

    [32] K. Du, P. Li, K. Gao. Strong coupling between dark plasmon and anapole modes. J. Phys. Chem. Lett., 10, 4699-4705(2019).

    [33] D. Zheng, S. Zhang, Q. Deng. Manipulating coherent plasmon-exciton interaction in a single silver nanorod on monolayer WSe2. Nano Lett., 17, 3809-3814(2017).

    [34] M.-E. Kleemann, R. Chikkaraddy, E. M. Alexeev. Strong-coupling of WSe2 in ultra-compact plasmonic nanocavities at room temperature. Nat. Commun., 8, 1296(2017).

    [35] A. E. Schlather, N. Large, A. S. Urban. Near-field mediated plexcitonic coupling and giant Rabi splitting in individual metallic dimers. Nano Lett., 13, 3281-3286(2013).

    [36] R. Du, H. Hu, T. Fu. How to obtain the correct Rabi splitting in a subwavelength interacting system. Nano Lett., 23, 444-450(2023).

    [37] X. Wu, S. K. Gray, M. Pelton. Quantum-dot-induced transparency in a nanoscale plasmonic resonator. Opt. Express, 18, 23633-23645(2010).

    [38] J. Sun, H. Hu, D. Zheng. Light-emitting plexciton: exploiting plasmon-exciton interaction in the intermediate coupling regime. ACS Nano, 12, 10393-10402(2018).

    [39] V. G. Kravets, A. V. Kabashin, W. L. Barnes. Plasmonic surface lattice resonances: a review of properties and applications. Chem. Rev., 118, 5912-5951(2018).

    [40] T. J. Seok, A. Jamshidi, M. Kim. Radiation engineering of optical antennas for maximum field enhancement. Nano Lett., 11, 2606-2610(2011).

    [41] R. Liu, Z. Liao, Y.-C. Yu. Relativity and diversity of strong coupling in coupled plasmon-exciton systems. Phys. Rev. B, 103, 235430(2021).

    [42] M. D. Leistikow, J. Johansen, A. J. Kettelarij. Size-dependent oscillator strength and quantum efficiency of CdSe quantum dots controlled via the local density of states. Phys. Rev. B, 79, 045301(2009).

    [43] L. Lin, J. Xue, H. Xu. Integrating lattice and gap plasmonic modes to construct dual-mode metasurfaces for enhancing light-matter interaction. Sci. China Mater., 64, 3007-3016(2021).

    [44] Y. Yang, V. A. Zenin, S. I. Bozhevolnyi. Anapole-assisted strong field enhancement in individual all-dielectric nanostructures. ACS Photon., 5, 1960-1966(2018).

    [45] E. A. Gurvitz, K. S. Ladutenko, P. A. Dergachev. The high-order toroidal moments and anapole states in all-dielectric photonics. Laser Photon. Rev., 13, 1800266(2019).

    [46] R. Alaee, C. Rockstuhl, I. Fernandez-Corbaton. An electromagnetic multipole expansion beyond the long-wavelength approximation. Opt. Commun., 407, 17-21(2018).

    [47] C. Zhou, S. Li, M. Fan. Optical radiation manipulation of Si-Ge2Sb2Te5 hybrid metasurfaces. Opt. Express, 28, 9690-9701(2020).

    [48] Y.-H. Deng, Z.-J. Yang, J. He. Plasmonic nanoantenna-dielectric nanocavity hybrids for ultrahigh local electric field enhancement. Opt. Express, 26, 31116-31128(2018).

    [49] H. Xu, Z. Zhu, J. Xue. Giant enhancements of high-order upconversion luminescence enabled by multiresonant hyperbolic metamaterials. Photon. Res., 9, 395-404(2021).

    [50] B. Gerislioglu, A. Ahmadivand. Theoretical study of photoluminescence spectroscopy of strong exciton-polariton coupling in dielectric nanodisks with anapole states. Mater. Today Chem., 16, 100254(2020).

    [51] T. Mahinroosta, S. M. Hamidi. The bull’s-eye structure as a new plexcitonic circular grating. Appl. Phys. A, 128, 1043(2022).

    [52] C. Van Vlack, P. T. Kristensen, S. Hughes. Spontaneous emission spectra and quantum light-matter interactions from a strongly coupled quantum dot metal-nanoparticle system. Phys. Rev. B, 85, 075303(2012).

    [53] E. C. Andre, I. E. Protsenko, A. V. Uskov. On collective Rabi splitting in nanolasers and nano-LEDs. Opt. Lett., 44, 1415-1418(2019).

    Yan-Hui Deng, Yu-Wei Lu, Hou-Jiao Zhang, Zhong-Hong Shi, Zhang-Kai Zhou, Xue-Hua Wang. Strong light–matter interactions based on excitons and the abnormal all-dielectric anapole mode with both large field enhancement and low loss[J]. Photonics Research, 2024, 12(4): 854
    Download Citation