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

Resonance fluorescence engineering in hybrid systems consist of biexciton quantum dots and anisotropic metasurfaces

Open Access Open Access

Abstract

The resonance fluorescence properties in the steady-state regime are investigated for a driven cascaded exciton-biexciton quantum dot coupled to the two-dimensional black phosphorus metasurfaces. It is shown that for the material parameters under consideration, both the elliptic and hyperbolic dispersion patterns of the surface plasmon modes are achievable according to the variation of the carrier concentration. Further study on the Purcell factor indicates unequal enhancements in the spontaneous decay of the orthogonal in-plane dipoles. Motivated by this intriguing phenomenon, we then investigate the steady-state properties of the driven quantum dot, where the populations of the dressed levels are highly tunable by engineering the anisotropy of the surfaces. As a result, the manipulation of the carrier concentration will lead to strong modifications in the resonance fluorescence. Under certain conditions, one can observe the squeezing of two-mode noise spectra with different resonances and polarizations. Although at the expense of declines in the photon-sideband detunings, it is feasible to enhance the two-mode squeezing by gate doping. Our proposal can be easily extended to other hybrid systems containing anisotropic metasurfaces, which are important for the development of quantum information science.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Optical nanostructures are efficient architectures in tailoring the electromagnetic vacuum and controlling light-matter interactions. Among them, a novel kind of deep subwavelength polariton modes-surface plasmon modes (SPM), has been reported and observed in various structures [13]. It is known that the excitation of the SPM can lead to dramatically large density of the states near the supported boundaries, which often accompanied by prominent enhancements in the Purcell factor [4,5]. In this aspect, systems consist of quantum emitters coupled to plasmonic platforms are of both fundamental and practical interest due to their ability in modifying dipole radiations. They also give rise to potential applications in quantum domains such as superradiance [6], single photon sources [7], quantum entanglement [8], resonance fluorescence [9] and quantum interference [10]. The two-dimensional metasurfaces are uniaxial structures with either opposite or the same sign in the effective surface conductivities. So far, surface polariton modes with both the elliptical and hyperbolic dispersion topologies have been reported in metasurfaces composed of silver [11], graphene [12], black phosphorus (BP) [13,14], van der Waals materials [15] and hexagonal boron nitride [16]. Among them, two-dimensional anisotropic metasurfaces like BP and graphene heterostructures possess actively tunable electronic and optical properties. Based on these novel features, they have found numerous applications in plasmonics and optoelectronics [1720].

As a building block of understanding the mechanism lies behind the light scattering in the steady-state regime, the resonance fluorescence of driven systems has attracted attentions since it was proposed [2123]. Many intriguing phenomena demonstrating the quantum features of dipole radiations via coherent pumpings have been reported, such as squeezing in resonance fluorescence [24], photon antibunching [25] and ultra-coherent photons generation [26]. Squeezed light-a nonclassical state of radiation, now becomes an important resource for applications in quantum information areas [27,28]. The squeezed states of light can be generated by reducing the variance of the electric field below the shot-noise limit, which are useful in the quantum metrology [29,30], gravitational wave detections [31], quantum key distributions [32,33], continuous variable quantum computings [34,35] and quantum sensing [36]. There are various theoretical methods in producing the squeezed states of light, and a wide range of experiments have been accomplished to explore their practical applications. These include the Gaussian squeezed states as well as non-Gaussian types, where the latter often behaves more ’non-classically’ and can be generated via non-Gaussian operations on the Gaussian coherent states [3739]. Moreover, squeezed state generations via multiphoton processes containing four-wave mixing in atomic and solid-state systems [4042], optical parametric amplifiers [43,44], cavity optomechanical or cavity quantum electrodynamical systems [4548], and Bose-Einstein condensations [49] are also intensely investigated by scientists.

It was first predicted by Walls and Zoller in 1981 that non-Gaussian quadrature squeezing can be generated in the resonance fluorescent field of a driven two-level quantum emitter, which is originated from the buildup and survival of the steady-state coherence in the weak excitation regime [50]. Recently, this nonclassical phenomenon was also observed in the semiconductor quantum dot platform [51], which possesses higher photon collection efficiency compared to the atomic approach [52]. In the theoretical work, Grünwald and Vogel demonstrated the possibility in increasing the squeezing via atomic-state purifications in atom-cavity coupled systems. It was found that the atomic coherence is important in optimizing the squeezing in resonance fluorescence [53]. The results also indicate the enhancement of squeezing in the radiation field from a coherent driven quantum emitter by coupling dipoles to an optimal environment, which is no longer limited to weakly driven systems. Besides, they predicted the large enhancement of squeezing by coupling quantum emitters to the surface plasmon modes supported by the gold nanospheres [54]. The antenna effect [55] boosts the squeezing in the far field, which is robust against the dephasing of the solid-state system owing to the strong light-matter interactions at the plasmon resonances. However, few works focus on the generation of squeezed light with high tunability through exciton-environment couplings in driven QD systems [56], where the surface plasmon modes [57,58] serve as the tunable photonic reservoir.

In this work we investigate the anisotropic Purcell effect as well as the resonance fluorescence properties of a driven cascaded exciton-biexciton quantum dot coupled to anisotropic metasurfaces. The central frequency of the drive is tuned to half of the biexciton energy such that the two-photon resonance condition is satisfied. Starting from the Green’s functions, the influence of the carrier concentration on the Purcell factor [59] is investigated. We show that the plasmon resonances are highly tunable, and the Purcell factors of the orthogonal in-plane dipoles exhibit large differences in the frequency domain. Based on these intriguing properties, we then study the steady-state populations of the driven QD. It is shown that different energy levels can be selectively populated by manipulating the carrier concentration of the metasurfaces [60]. As the result, the resonance fluorescent field exhibits asymmetric Mollow peaks. Finally we find that two-mode squeezing of the photon pairs with different resonances and polarizations can be generated and further enhanced according to the variations in the carrier concentration.

Our paper is organized as follows. In Sec. 2, the model is introduced and the Green functions of the electromagnetic field are derived by applying the boundary conditions. Then the influence of the carrier concentration on the Purcell factor is discussed. In Sec. 3. the Bloch equations of the dressed QD are obtained according to the master equation, where the steady-state populations, the resonance fluorescence and the two-mode noise spectra at different carrier concentrations are studied. Finally we present our conclusions in Sec. 4. and an appendix is given to gain insights into the main results of this work.

2. Anisotropic Purcell effect mediated by anisotropic metasurfaces

The system under consideration can be well described by a biexciton QD coupled to the BP metasurfaces, where the QD is driven by a laser and the conductivity of the metasurfaces takes the form $\overset {\leftrightarrow }{\boldsymbol {\kappa }}=diag\left (\kappa _{xx},\kappa _{yy}\right )$. As illustrated by Fig. 1, the QD can be effectively treated as a four-level transition system, with bare states defined as the biexciton state $\left |B\right \rangle$, the $X-$exciton state $\left |X\right\rangle$, the $Y-$exciton state $\left |Y\right \rangle$ and the ground state $\left |G\right \rangle$. Here the small energy splitting between two excitonic states is neglected [61] and the resonant frequency is taken as $\omega _{e}=400\,\textrm {THz}$, the energy of the biexciton state is equal to $2\hbar \omega _{e}-\chi$ owing to the existence of the binding energy $\chi$ [62]. For linearly polarized excitations, it is possible to induce the transitions from the ground state to the biexciton state through two orthogonal channels, with photonic polarizations along the $x-$ or $y-$direction. The $y-$polarized transition dipoles $\boldsymbol {p}_{y1}$ and $\boldsymbol {p}_{y2}$ couple to the drive with Rabi frequencies $\Omega _{1}$ and $\Omega _{2}$, where the undressed transition dipoles are denoted by $\boldsymbol {p}_{x1}$ and $\boldsymbol {p}_{x2}$. All the dipoles are modeled in the point dipole approximation and we take $\boldsymbol {p}_{x1}=\boldsymbol {p}_{y1}=60\,\textrm {D}$, $\boldsymbol {p}_{x2}=\boldsymbol {p}_{y2}$ for further calculations. In this configuration, the spontaneous emission of the QD takes place via both the $x$- and $y$-polarized transitions, where the relevant strength is modified by the metasurfaces owing to the strong coupling with the SPM. Then within the rotating wave approximation, after rotating with the laser’s frequency the unperturbed and the interaction Hamiltonians appear to be

$$\hat{H}_{0}' = \hbar\Delta_{L}\left(\hat{\sigma}_{XX}+\hat{\sigma}_{YY}\right)+\hbar\left[\Omega_{1}\hat{\sigma}_{YG}+\Omega_{2}\hat{\sigma}_{BY}+\textrm{H.c.}\right]$$
and
$$\hat{H}_{I}' ={-}\bigg[\left(\hat{\sigma}_{YG}\boldsymbol{p}_{y1}+\hat{\sigma}_{BY}\boldsymbol{p}_{y2}+\hat{\sigma}_{XG}\boldsymbol{p}_{x1}+\hat{\sigma}_{BX}\boldsymbol{p}_{x2}\right)\cdot\left.\int d\omega\vec{p}_{1}\hat{\mathbf{E}}^{(+)}\left(\boldsymbol{r}_{Q},\omega\right)e^{i\left(\omega_{L}-\omega\right)t}+\textrm{H.c.}\right].$$
In Eqs. (1) and (2), $\hat {\sigma }_{jk}=\left |j\right \rangle \left \langle k\right |$ ($j,k=G,X,Y,B$) are Pauli operators of the bare QD, $\boldsymbol {r}_{Q}$ is the spatial position of the QD, $\Delta _{L}=\omega _{e}-\omega _{L}$ is the frequency detuning between the driving field and the single exciton state, the relation $\hbar \Delta _{L}=\chi /2$ should be satisfied in the two-photon resonance case. The positive frequency part of the field operator can be written as [63]
$$\hat{\mathbf{E}}^{(+)}\left(\boldsymbol{r}_{Q},\omega\right) = i\omega\mu_{0} \int d\boldsymbol{r}\overset{\leftrightarrow}{\mathbf{G}}(\boldsymbol{r}_{Q},\boldsymbol{r},\omega)\cdot\hat{\mathbf{j}}(\boldsymbol{r},\omega),$$
where $\hat {\mathbf {j}}(\boldsymbol {r},\omega )=\omega \sqrt {\hbar \varepsilon _{0}\textrm {Im}\left [\varepsilon (\boldsymbol {r},\omega )\right ]/\pi }\,\hat {\mathbf {f}}(\boldsymbol {r},\omega )$ denotes the noise currents evaluated at the spatial position $\boldsymbol {r}$ and the frequency $\omega$, it fulfills the following wave equation
$$\left[\nabla\times\frac{1}{\mu(\mathbf{r},\omega)}\nabla\times{-}\frac{\omega^{2}}{c^{2}}\varepsilon(\mathbf{r},\omega)\right]\hat{\mathbf{E}}(\mathbf{r},\omega) = i\omega\mu_{0}\,\hat{\mathbf{j}}(\mathbf{r},\omega).$$
In our notation, $\textrm {Im}\left [\varepsilon (\boldsymbol {r},\omega )\right ]$ is the imaginary part of the permittivity, the bosonic operators $\hat {\mathbf {f}}^{\dagger }(\boldsymbol {r},\omega )$ and $\hat {\mathbf {f}}(\boldsymbol {r},\omega )$ represent the creation and annihilation of an exciton in the photonic reservoir [63]. The Green’s dyadic $\overset {\leftrightarrow }{\mathbf {G}}(\boldsymbol {r}_{Q},\boldsymbol {r},\omega )$, which is responsible for the observable field components at the QD’s position $\boldsymbol {r}_{Q}$ due to the existence of an exciton at $\boldsymbol {r}$, can be derived by imposing the boundary conditions in the presence of the metasurfaces according to the methods mentioned in the work [64], with the forms
$$\begin{aligned} \overset{\leftrightarrow}{\mathbf{G}}_{yy(xx)}\left(\mathbf{r},\mathbf{r}'\right) = & \frac{i}{4\pi^{2}}\int_{-\infty}^{\infty}\int_{-\infty}^{\infty}g_{yy(xx)}^{R}\left(\mathbf{k}\right)(2p)^{{-}1}\left[k_{0}^{2}-k_{y(x)}^{2}\right] e^{ip\left(z+z'\right)}e^{{-}i\mathbf{k}\left(\boldsymbol{\rho}-\boldsymbol{\rho}'\right)}dk_{x}dk_{y}\\ &+\overset{\leftrightarrow}{\mathbf{G}}_{yy(xx)}^{P}\left(\mathbf{r},\mathbf{r}'\right), \end{aligned}$$
$$g_{yy}^{R}\left(\mathbf{k}\right) = \Big[2k_{y}^{2}\kappa_{xx}-2\left(k_{y}^{2}-k_{0}^{2}\right)\kappa_{yy}-k_{0}p\kappa_{xx}\kappa_{yy}\eta\Big]/D(\mathbf{k}),$$
$$g_{xx}^{R}\left(\mathbf{k}\right) = \Big[2k_{0}^{2}\kappa_{xx}-2k_{x}^{2}\left(\kappa_{xx}-\kappa_{yy}\right)-k_{0}p\kappa_{xx}\kappa_{yy}\eta\Big]/D(\mathbf{k}),$$
$$\overset{\leftrightarrow}{\mathbf{G}}_{yy(xx)}^{P}\left(\mathbf{r},\mathbf{r}'\right) = \frac{i}{2\pi}\int_{-\infty}^{\infty}k^{3}(2p)^{{-}1}e^{ip\left|z-z'\right|}J_{0}\left(k\rho\right)dk, .$$

 figure: Fig. 1.

Fig. 1. Interaction between a driven QD and the anisotropic metasurfaces (here we consider BP structures): the QD couples to the anisotropic SPM supported by the surfaces. The pump frequency is reasonably chosen such that two-photon resonance condition is satisfied in the transition channel $\left |G\right \rangle \leftrightarrow \left |Y\right \rangle \leftrightarrow \left |B\right \rangle$. In the energy level diagram, the spontaneous emission of the exciton and biexciton states are modified by the anisotropy of the surfaces, the corresponding decay rates are denoted by $\gamma _{jj}^{y}$ ($y-$polarized transitions) and $\gamma _{jj}^{x}$ ($x-$polarized transitions).

Download Full Size | PDF

The Green’s functions $\overset {\leftrightarrow }{\mathbf {G}}_{yy}\left (\mathbf {r},\mathbf {r}'\right )$ and $\overset {\leftrightarrow }{\mathbf {G}}_{xx}\left (\mathbf {r},\mathbf {r}'\right )$, which are diagonal components of the Green’s dyadic, plays an important role in determining the spontaneous decay property of the driven QD. Both of them can be separated into the principle and the scattered parts (denoted by the superscript $P$ and $R$, respectively), where the former part originates from the dipole radiation in the free space and the later part reflects the influence of the metasurfaces on light-matter couplings owing to the excitation of the SPM. Notice that the principle part is determined by the isotropic radial wave number $k$ and the scattered part depends unequally on the wave numbers along two optical axes ($\mathbf {k}=k_{x}\mathbf {e}_{x}+k_{y}\mathbf {e}_{y}$, $\mathbf {e}_{x(y)}$ is unit vectors in the $x(y)$ direction), which can be clearly explained by exploring the dispersion of the SPM. Other parameters are: $\eta =\sqrt {\mu _{0}/\varepsilon _{0}}$ ($\varepsilon _{0}$ and $\mu _{0}$ are vacuum permittivity and permeability), $k_{0}$ is the vacuum wave number and $p=\sqrt {k_{0}^{2}-k_{x}^{2}-k_{y}^{2}}$ is the $z-$component of the wave vector. The denominator of the reflection coefficient $g_{yy(xx)}^{R}\left (\mathbf {k}\right )$ is denoted by $D(\mathbf {k})=2\kappa _{xx}\left (k_{x}^{2}-k_{0}^{2}\right )+2\kappa _{yy}\left (k_{y}^{2}-k_{0}^{2}\right )+4pk_{0}\eta ^{-1}\left (1+\kappa _{xx}\kappa _{yy}\eta ^{2}/4\right )$, which determines the isofrequency contour of the SPM. It is known that in heterostructures containing BP or graphene [1214], the intraband excitations can be well described by the Drude model [65] and the interband transitions at high frequencies are defined by the step absorption function. In the following we consider the dispersive BP layers with the conductivity components [66]

$$\kappa_{ll} = \frac{in_{c}e^{2}}{m_{l}\left(\omega+i\omega_{r}\right)}+\mathcal{I_{\mathit{l}}}\left[\Theta\left(\omega-\varpi_{l}\right)+\frac{i}{\pi}\ln\frac{\left|\omega-\varpi_{l}\right|}{\left|\omega+\varpi_{l}\right|}\right](l=x,y),$$
where the first and second term on the right-hand side of the above equation originates from the intraband electron motions and the interband transitions, respectively. Here $m_{l}$ is the effective mass of the electron along the $l$ direction, $\omega _{r}$ is the intraband relaxation rate and $\varpi _{l}$ is the frequency of the interband transitions related to the $ll$ component of the conductivity, $\mathcal {I_{\mathit {l}}}$ represents the strength of the interband transition and $n_{c}$ accounts for the carrier concentration that is experimentally tunable via electric or chemical dopings [60,67]. It should be pointed out that different crystallographic directions have distinct symmetries for the anisotropic metasurfaces. Thus the dipole moment for the electron interband transition becomes anisotropic, which leads to the discrepancy between the interband transition frequencies $\varpi _{x}$ and $\varpi _{y}$. Generally, for real anisotropic metasurfaces, Eq. (9) can be used to evaluate the conductivity of the multilayer BP [66,68].

It has been demonstrated that two-dimensional structures are good candidates in supporting anisotropic excitations of the SPM [6971]. In most cases, the increase in the carrier concentration will lead to declines in the real part of the conductivity due to Pauli blocking, whereas the imaginary part suffers prominent growths at the same time. The imaginary part of the conductivity often ranges from $\mu \textrm {S}$ to $m\textrm {S}$ for different doping levels, where the real part is negligibly small [14,72] for the strong excitation cases of the SPM. Owing to this reason, metasurfaces with high carrier concentrations can exhibit strong electromagnetic responses and low dissipations. In the proposed systems, it is instructive to inspect the Purcell factor [59] in order to understand the behavior of light-matter coupling in the presence of the BP metasurfaces. In the proper scheme of the field quantization, the spontaneous decay of dipoles with different polarizations is often related to the imaginary part of the diagonal Green’s functions [73]. Based on this property, the Purcell factors can be defined as $\Gamma _{y(x)}/\Gamma _{0}=6\pi \left (c\,\omega _{e}^{-1}\right )^{3}\textrm {Im}[\overset {\leftrightarrow }{\mathbf {G}}_{yy(xx)}\left (\mathbf {r}_{Q},\mathbf {r}_{Q},\omega _{e}\right )]$ ($\Gamma _{y(x)}$ denotes the spontaneous decay rate of the $y(x)$-polarized dipole), where the vacuum decay rate takes the form $\Gamma _{0}=\omega _{e}^{3}p_{y(x)}^{2}/3\varepsilon _{0}\pi \hbar c^{3}$. In the previous work, we found that the total coupling between quantum emitters and the SPM supported by anisotropic metasurfaces can be realized under certain conditions [78], where the radiations into both the free vacuum modes and the absorption of materials are negligible.

In Figs. 2(a) and 2(b) we inspect the Purcell factor of dipoles with different polarizations as a function of the excitation frequency and the carrier concentration, where the distance between the QD and the metasurface is assumed to be $20\,\textrm {nm}$ such that the SPM has a dominant contribution on the dipole radiations. The results demonstrate strong enhancements in the Purcell factor of the $y-$oriented dipole, which is mainly due to the Drude-like behavior of the conductivity components for excitation frequencies below the threshold of the interband transitions (see Fig. S1 for more details). Also we notice that although the spontaneous emission of the $x-$polarized dipoles is enhanced, the corresponding Purcell factor is much smaller than that of the $y-$polarized dipoles. However, the situation changes quite a lot at higher frequencies. In this regime, the Purcell factor of the $x-$polarized dipoles becomes much larger for small carrier concentrations, which indicates a predominant contribution of the $x-$polarized dipoles on the spontaneous decay. The unequal enhancements in the Purcell factors can be attributed to the deformation of the isofrequency contours toward one of the optical axes (see Fig. S1 for more details). For example, at low excitation frequencies the isofrequency contours of the surface plasmon modes become quasi-elliptical, which are elongated along the $y$ axis. As the result, the SPM are mostly guided in the directions with small angles respect to the $x$ axis, which is the main reason for the prominent enhancement in the Purcell factor of the $y-$polarized dipole. On the other hand, it is shown that the variations in the carrier concentration can effectively modify the Purcell factors. A novel feature is that the the plasmon resonance suffers blueshifts with the increase in the carrier concentration, where declines in the strength of the Purcell factors take place at the same time. That is to say, the in-plane coupling between the dipoles and the BP metasurfaces turns out to be highly anisotropic, where the coupling strength is highly tunable via the gate doping.

 figure: Fig. 2.

Fig. 2. Purcell factors of dipoles with different polarizations coupled to the BP metasurfaces, different carrier concentrations $n_{c}=\left (0.2,0.5,1,1.5,2\right )\times 10^{18}/\textrm {m}^{2}$ are considered to stress the modification of the metasurfaces on the dipole radiation. The spatial coordination of the QD is $\mathbf {r}_{Q}=\left (0,0,20\right )\textrm {nm}$, the material parameters are $m_{x}=0.2m_{0}$, $m_{y}=0.7m_{0}$, $\omega _{r}=2.5\textrm {THz}$, $\varpi _{x}/2\pi =240\textrm {THz}$, $\varpi _{y}/2\pi =80\textrm {THz}$, $\mathcal {I_{\mathit {x}}}=1.5\kappa _{0}$ and $\mathcal {I_{\mathit {y}}}=3.5\kappa _{0}$, here $\kappa _{0}=e^{2}/4\hbar$ is the conductance quantum and $m_{0}$ is the free electron mass. (a) The Purcell factor of $y-$polarized dipoles. (b) The Purcell factor of $x-$polarized dipoles.

Download Full Size | PDF

3. Tunable resonance fluorescence and noise spectra in the steady-state regime

Based on the above discussions, the application of the anisotropic metasurfaces indeedly opens a new avenue for controlling light-matter interactions. In this section, we will mainly focus on its influence on the radiation properties of the driven QD. It is known that the dynamical and spectral features of the driven systems can be well understood by working in the dressed state picture. After diagonalizing the unperturbed Hamiltonian defined in Eq. (1), the dressed state basis [74] of the system under consideration can be expressed in forms of the combinations of the bare QD operators as

$$\left|+\right\rangle = \sqrt{-E_{-}/E_{t}}\left(E_{1}\left|G\right\rangle +E_{+}\left|Y\right\rangle +E_{2}\left|B\right\rangle \right)/E_{e},$$
$$\left|-\right\rangle = \sqrt{E_{+}/E_{t}}\left(E_{1}\left|G\right\rangle +E_{-}\left|Y\right\rangle +E_{2}\left|B\right\rangle \right)/E_{e},$$
$$\left|M\right\rangle = \left({-}E_{2}\left|G\right\rangle +E_{1}\left|B\right\rangle \right)/E_{e}.$$
In the above expressions, the effective Rabi energies are denoted by $E_{e}=\hbar \Omega _{e}=\hbar \sqrt {\Omega _{1}^{2}+\Omega _{2}^{2}}=\sqrt {E_{1}^{2}+E_{2}^{2}}$ and $E_{t}=\hbar \Omega _{t}=\sqrt {4E_{e}^{2}+\chi ^2/4}$, where $E_{+}=\hbar \Omega _{+}=(\chi +2E_{t})/4$, $E_{-}=\hbar \Omega _{-}=(\chi -2E_{t})/4$ and $E_{M}=0$ are energies of the dressed states $\left |+\right \rangle$, $\left |-\right \rangle$ and $\left |M\right \rangle$. Then after applying canonical transformation on the interaction Hamiltonian defined in Eq. (2) (the picture transformation takes the form $\hat {V}_{I}(t)=\hat {U}^{\dagger }\hat {H}_{I}'\hat {U}$, with the unitary operator $\hat {U}(t)=\exp \left \{ -i\left [\int d\mathbf {r}\int \hbar \omega \hat {\mathbf {f}}^{\dagger }(\boldsymbol {r},\omega )\hat {\mathbf {f}}(\boldsymbol {r},\omega )d\omega +\hbar \omega _{L}\left (\hat {\sigma }_{XX}+\hat {\sigma }_{YY}+2\hat {\sigma }_{BB}\right )\right ]t\right \}$), and performing the second-order perturbation theory by tracing out the reservoir degrees through the relation $\dot {\hat {\rho }}(t)=-\hbar ^{-2}\int _{0}^{t}\textrm {Tr}_{R}\left \{ \left [\hat {V}_{I}(t),\left [\hat {V}_{I}(t-\tau ),\rho (t-\tau )\right ]\right ]\right \} d\tau$ [75], under the Born-Markovian and secular [76] approximations the brief master equation for the reduced density operator of the driven QD is
$$\dot{\hat{\rho}} ={-}\frac{i}{\hbar}[\hat{H}_{0}',\hat{\rho}]-\Gamma_{+{+}}\left(\hat{\sigma}_{+{+}}\hat{\rho}\hat{\sigma}_{-{-}}+\hat{\sigma}_{-{-}}\hat{\rho}\hat{\sigma}_{+{+}}\right. \left.-\frac{1}{2}\boldsymbol{\mathcal{L}}_{+{+}}-\frac{1}{2}\boldsymbol{\mathcal{L}}_{-{-}}\right)+\frac{1}{2}\sum_{m\neq n={+},-,M,X}\Gamma_{mn}\boldsymbol{\mathcal{L}}_{mn}.$$
Here the Lamb shifts of the energy levels have been neglected since they do not affect the main physics, the bath is assumed to be at zero temperature thus the traces of the field can be described by $\textrm {Tr}_{R}[\hat {\mathbf {f}}^{\dagger }(\boldsymbol {r},\omega )\hat {\mathbf {f}}(\boldsymbol {r}',\omega ')]=0$ and $\textrm {Tr}_{R}[\hat {\mathbf {f}}^{\dagger }(\boldsymbol {r},\omega )\hat {\mathbf {f}}(\boldsymbol {r}',\omega ')]=\delta (\boldsymbol {r}-\boldsymbol {r}')\delta (\omega -\omega ')$. The operator $\boldsymbol {\mathcal {L}}_{mn}=2\hat {\sigma }_{nm}\hat {\rho }\hat {\sigma }_{mn}-\hat {\sigma }_{mn}\hat {\sigma }_{nm}\hat {\rho }-\hat {\rho }\hat {\sigma }_{mn}\hat {\sigma }_{nm}$, where the spontaneous decay rates $\Gamma _{mn}$ related to the transition $\left |m\right \rangle \rightarrow \left |n\right \rangle$ in the dressed state basis appear to be
$${$\begin{aligned} \Gamma_{+{+}} & = \Gamma_{-{-}} = \frac{2}{\Omega_{t}^{2}}\left[\Omega_{1}^{2}\gamma_{11}^{y}\left(0\right)+2\Omega_{1}\Omega_{2}\gamma_{12}^{y}\left(0\right)+\Omega_{2}^{2}\gamma_{22}^{y}\left(0\right)\right],\\ \Gamma_{+{-}} & = \frac{2M}{\left(\Omega_{t}\Omega_{e}\right)^{2}}\left[\frac{\Omega_{1}\Omega_{+}}{\Omega_{2}\Omega_{-}}\gamma_{11}^{y}\left(\Omega_{t}\right)-2\gamma_{12}^{y}\left(\Omega_{t}\right)\right.+\left.\frac{\Omega_{2}\Omega_{-}}{\Omega_{1}\Omega_{+}}\gamma_{22}^{y}\left(\Omega_{t}\right)\right],\\ \Gamma_{-{+}} & = \frac{2M}{\left(\Omega_{t}\Omega_{e}\right)^{2}}\left[\frac{\Omega_{1}\Omega_{-}}{\Omega_{2}\Omega_{+}}\gamma_{11}^{y}\left(-\Omega_{t}\right)-2\gamma_{12}^{y}\left(-\Omega_{t}\right)\right.\left.+\frac{\Omega_{2}\Omega_{+}}{\Omega_{1}\Omega_{-}}\gamma_{22}^{y}\left(-\Omega_{t}\right)\right],\\ \Gamma_{+M} & = \frac{2\Omega_{2}^{2}\Omega_{+}}{\Omega_{t}\Omega_{e}^{2}}\gamma_{11}^{y}\left(\Omega_{+}\right),\Gamma_{M+}=\frac{2\Omega_{1}^{2}\Omega_{+}}{\Omega_{t}\Omega_{e}^{2}}\gamma_{22}^{y}\left(-\Omega_{+}\right),\\ \Gamma_{X-} & = \frac{2\Omega_{1}^{2}\Omega_{+}}{\Omega_{t}\Omega_{e}^{2}}\gamma_{11}^{x}\left(\Omega_{+}\right),\Gamma_{-X}=\frac{2\Omega_{2}^{2}\Omega_{+}}{\Omega_{t}\Omega_{e}^{2}}\gamma_{22}^{x}\left(-\Omega_{+}\right),\\ \Gamma_{M-} & = \frac{2\Omega_{1}^{2}\Omega_{-}}{\Omega_{t}\Omega_{e}^{2}}\gamma_{22}^{y}\left(\Omega_{-}\right),\Gamma_{-M}=\frac{2\Omega_{2}^{2}\Omega_{-}}{\Omega_{t}\Omega_{e}^{2}}\gamma_{11}^{y}\left(-\Omega_{-}\right),\\ \Gamma_{+X} & = \frac{2\Omega_{2}^{2}\Omega_{-}}{\Omega_{t}\Omega_{e}^{2}}\gamma_{22}^{x}\left(\Omega_{-}\right),\Gamma_{X+}=\frac{2\Omega_{1}^{2}\Omega_{-}}{\Omega_{t}\Omega_{e}^{2}}\gamma_{11}^{x}\left(-\Omega_{-}\right),\\ \Gamma_{XM} & = \frac{2\Omega_{2}^{2}}{\Omega_{t}^{2}}\gamma_{11}^{x}\left(\Delta_{L}\right),\quad\,\Gamma_{MX}=\frac{2\Omega_{1}^{2}}{\Omega_{t}^{2}}\gamma_{22}^{x}\left(-\Delta_{L}\right), \end{aligned}$}$$
in the above equations the coefficient $M=\prod _{m=1,2,+,-}\Omega _{m}$, the reservoir-assisted decay rates of dipoles with $s$ polarization are denoted by $\gamma _{jl}^{s}\left (\omega \right )=\boldsymbol {p}_{sj}\cdot \textrm {Im}[\overset {\leftrightarrow }{\mathbf {G}}\left (\boldsymbol {r}_{Q},\boldsymbol {r}_{Q},\omega _{L}+\omega \right )]\cdot \boldsymbol {p}^{\ast }_{sl}(\omega _{L}+\omega )^{2}/\hbar \pi \varepsilon _{0}c^{2}$. Obviously, the transitions between different energy states of the driven QD can be clearly distinguished via their detunings from the drive. Starting from Eq. (13), it is easy to acquire the following Bloch equations
$$\dot{\hat{\sigma}}_{nn} ={-}\sum_{m\neq n={+},-,M,X}\Gamma_{nm}\hat{\sigma}_{nn}+\sum_{l\neq n={+},-,M,X}\Gamma_{ln}\hat{\sigma}_{ll},\;(n={+},-,M,X)$$
$$\dot{\hat{\sigma}}_{jl} = \,f_{jl}\hat{\sigma}_{jl},\;(j\neq l={+},-,M,X)$$
where the diagonal components $\sigma _{nn}$ in Eq. (15) represent the populations of different energy states, and the off-diagonal components $\sigma _{jl}$ in Eq. (16) are polarization operators accounting for the spectral features of the resonance fluorescent field. It should be pointed out that the coefficients $f_{jl}$ often contain an imaginary part related to the resonant frequency of the sideband transitions $\left |j\right \rangle \rightarrow \left |l\right \rangle$ and a real part corresponding to the half-widths of the spectra.

In order to investigate the influence of the anisotropic Purcell effect on the population distributions of the QD, in Fig. 3 we plot the steady-state populations versus the Rabi detuning $\Delta _{R}$ by considering different carrier concentrations, where other parameters are the same as in Fig. 2. It can be seen from Fig. 3(a) that for low carrier concentrations, there are prominent population differences between the state $\left |+\right \rangle$ and other energy states when $\Delta _{R}<0$. When the carrier concentration increases to $n_c=2\times 10^{17}/m^{2}$, the steady-state population of the dressed QD is rapidly accumulated to the state $\left |X\right\rangle$ with the increase of the Rabi amplitude $\Omega _{2}$, which is shown in Fig. 3(b).

 figure: Fig. 3.

Fig. 3. Steady-state populations of the energy states $\left |+\right \rangle$, $\left |-\right \rangle$, $\left |M\right \rangle$ and $\left |X\right \rangle$ (denoted by $\sigma _{++}^{ss}$, $\sigma _{--}^{ss}$, $\sigma _{MM}^{ss}$ and $\sigma _{XX}^{ss}$), as a function of the Rabi detuning ($\Delta _{R}=\Omega _{2}-\Omega _{1}$, the Rabi energy $\Omega _{1}=20\;\textrm {meV}$). The influence of the carrier concentration on the population distributions are studied: (a) $n_{c}=10^{17}/\textrm {m}^{2}$, (b) $n_{c}=2\times 10^{17}/\textrm {m}^{2}$, (c) $n_{c}=1.1\times 10^{18}/\textrm {m}^{2}$ and (d) $n_{c}=2.2\times 10^{18}/\textrm {m}^{2}$, other parameters are the same as in Fig. 2.

Download Full Size | PDF

Figures 3(c) and 3(d) demonstrate the behavior of the steady-state populations at large carrier concentrations. Obviously, the increases in the carrier concentration can effectively redistribute the populations of the energy states. This can be well understood by combining the results shown in Fig. 2, where the growth of the carrier concentration leads to blueshifts in the resonant peak of the Purcell factor. As the result, different transition channels can be selectively enhanced without changing the driving conditions. In other words, the anisotropic metasurfaces act as a tunable reservoir coupled to the dressed QD, which provides unique capabilities to control the steady-state properties of the driven system. For example, as indicated by Fig. 3(c), the dressed state $\left |+\right \rangle$ is largely populated for the Rabi detuning $\Delta _{R}=-8.3\;\textrm {meV}$. However, under the same driving conditions a prominent population inversion between the state $\left |-\right \rangle$ and other states can be achieved by only increasing the carrier concentration, with steady-state population $\rho _{--}^{ss}\approx 0.97$ (see Fig. 3(d)). Thus it is clear that the steady-state populations of the dressed levels can be well controlled by manipulating the material parameters or driving conditions.

For the system under consideration, the steady-state spectrum of the resonance fluorescence is proportional to the Fourier transformation of the correlation function $\lim _{t\rightarrow \infty }\langle \hat {\mathbf {E}}_{-}\left (\boldsymbol {r},t+\tau \right )\cdot \hat {\mathbf {E}}_{+}(\boldsymbol {r},t) \rangle$, where $\hat {\mathbf {E}}_{+}(\boldsymbol {r},t)$ and $\hat {\mathbf {E}}_{-}(\boldsymbol {r},t)$ are positive and negative parts of the dipole field in the far zone [75], respectively. In our notation, the resonance fluorescence originates from the sideband transitions can be generally written as

$$\textrm{S}_{sid}(\delta_{L})=\textrm{S}_{M-}(\delta_{L})+\textrm{S}_{M+}(\delta_{L})+\textrm{S}_{MX}(\delta_{L})+\textrm{S}_{X+}(\delta_{L})+\textrm{S}_{X-}(\delta_{L})+\textrm{S}_{+{-}}(\delta_{L}),$$
where the spectral functions corresponding to the transitions between different energy states (denoted by $\textrm {S}_{ mn}(\delta _{L}),m\neq n=+,X,M,-$) are given in Eq. (5), with the transformation of the two-time correlation functions defined by
$$\mathit{L}_{jl,mn}(\delta_{L})=\textrm{Re}\int_{0}^{\infty}d\tau\left\langle \hat{\sigma}_{jl}\left(\tau\right),\hat{\sigma}_{mn}\left(0\right)\right\rangle e^{{-}i\delta_{L}\tau}.$$

We now focus on how the BP metasurfaces modify the resonance fluorescence of the driven QD. The analysis will be carried out in the dressed state picture, where the function $\textrm {S}_{jl}(\delta _{L})$ corresponds to the spectrum of the two-photon emission via the transition $\left |j\right \rangle \rightarrow \left |l\right \rangle \rightarrow \left |j\right \rangle$). It is known that the two-time correlation functions can be evaluated by applying the quantum regression theory [77] on the Bloch equations given in Eqs. (14) and (15). Within the secular approximation, the couplings between the populations and coherences are neglected, thus the resonance fluorescence spectra can be decomposed as a sum of Lorentzian peaks. Numerical results are displayed in Fig. 4, where the influence of the carrier concentrations is taken into account. The resonance fluorescence spectra of different transitions are plotted separately to clearly illustrate the modification the metasurfaces. Notice that the peak locations of the spectra are solely determined by the driving condition and thus are insensitive to the carrier concentration. However, the peak values are strongly modified as the carrier concentration changes. As the result, the peak signals can exhibit tunable asymmetries. For example, as depicted in Fig. 4(f), distinct Mollow peaks of the spectrum $\textrm {S}_{+-}(\delta _{L})$ can be observed when $n_{c}$ varies from $10^{17}/\textrm {m}^2$ to $2.2\times 10^{18}/\textrm {m}^2$. This fact is exactly a signature of the selective enhancement in the dipole transitions. More detailed, from Fig. 3(a) it is clear that the state $\left |+\right \rangle$ is largely populated under the conditions $\hbar \Delta _{R}=-3\;\textrm {meV}$ and $n_{c}=10^{17}/\textrm {m}^2$, which is the reason for the appearance of the strong fluorescence peak at the right sideband. On the contrary, when $n_{c}$ increases to $2.2\times 10^{18}/\textrm {m}^2$ the population of the state $\left |-\right \rangle$ becomes prominent, thus a strong peak at the left sideband occurs. Similar modifications on the resonance fluorescence spectra corresponds to other sideband transitions can also be found under certain conditions, which are illustrated by Figs. (4(a))–(4(e)).

 figure: Fig. 4.

Fig. 4. Steady-state resonance fluorescence spectra versus the photon-laser detuning ($\delta _{L}=\omega -\omega _{L}$) and the carrier concentration of the metasurfaces, the material parameters are the same as used in Fig. 2. The resonance fluorescence spectra denoting distinct two-photon emissions are plotted: (a) $\left |M\right \rangle \rightarrow \left |-\right \rangle \rightarrow \left |M\right \rangle$, $\hbar \Delta _{R}=-1\;\textrm {meV}$, (b) $\left |M\right \rangle \rightarrow \left |+\right \rangle \rightarrow \left |M\right \rangle$, $\hbar \Delta _{R}=-3\;\textrm {meV}$, (c) $\left |M\right \rangle \rightarrow \left |X\right \rangle \rightarrow \left |M\right \rangle$, $\hbar \Delta _{R}=3\;\textrm {meV}$, (d) $\left |X\right \rangle \rightarrow \left |+\right \rangle \rightarrow \left |X\right \rangle$, $\hbar \Delta _{R}=3\;\textrm {meV}$, (e) $\left |X\right \rangle \rightarrow \left |-\right \rangle \rightarrow \left |X\right \rangle$, $\hbar \Delta _{R}=-1\;\textrm {meV}$, and (f) $\left |+\right \rangle \rightarrow \left |-\right \rangle \rightarrow \left |+\right \rangle$, $\hbar \Delta _{R}=-3\;\textrm {meV}$. The carrier concentrations are taken as $n_{c}=(0.2,1.1,2.2)\times 10^{18}/\textrm {m}^2$ for subfigures (a), (c), (d), (e) and $n_{c}=(0.1,1.1,2.2)\times 10^{18}/\textrm {m}^2$ for subfigures (b), (f).

Download Full Size | PDF

In the driven systems, quadrature squeezing occurs when the noise spectra exhibit negative peaks. Notice that the photon pairs generated via distinct two-photon transitions are correlated in the frequency, thus by transforming the spectral functions into the frequency domain, the two-mode noise spectra can be defined through

$$\begin{aligned} \textrm{S}_{mn}^{in(out)}\left(\delta_{mn}\right) = & \textrm{Re}\int_{0}^{\infty}\left[C_{mn}^{i(o)}\left\langle \hat{\sigma}_{mn}\left(\tau\right)e^{{-}i\omega_{mn}\tau},\hat{\sigma}_{nm}\left(0\right)\right\rangle e^{i\omega_{1}\tau}\right.\\ &+\left.C_{nm}^{i(o)}\left\langle \hat{\sigma}_{nm}\left(\tau\right)e^{{-}i\omega_{nm}\tau},\hat{\sigma}_{mn}\left(0\right)\right\rangle e^{i\omega_{2}\tau}\right]d\tau\\ = & C_{mn}^{i(o)}\,\mathit{L}_{mn,nm}(-\delta_{mn})+C_{nm}^{i(o)}\,\mathit{L}_{nm,mn}(\delta_{mn}). \end{aligned}$$
In the above equation, $\textrm {S}_{mn}^{in(out)}\left (\delta _{mn}\right )$ ($m\neq n=+,-,M,X$) denotes the in-phase (out-of-phase) two-mode noise spectra corresponding to the transitions $\left |m\right \rangle \rightarrow \left |n\right \rangle \rightarrow \left |m\right \rangle$ or $\left |n\right \rangle \rightarrow \left |m\right \rangle \rightarrow \left |n\right \rangle$. Here $C_{mn}^{i(o)}$ and $C_{nm}^{i(o)}$ are transforming coefficients of different quadratures from the bare basis to the dressed basis (see Eq. (8) in the Appendix for their explicit forms), $\omega _{mn}$ is the resonant frequency of the transition $\left |m\right \rangle \rightarrow \left |n\right \rangle$, where $\omega _{1}$ and $\omega _{2}$ denote the frequencies of the emitted photon pairs satisfying the relation $\omega _{1}+\omega _{2}=2\omega _{L}$. Obviously, $\delta _{mn}=\omega _{1}-\omega _{mn}=-\left (\omega _{2}-\omega _{nm}\right )$ indicates the frequency detuning of the sideband photons from their resonances. After straightforward deducion, the explicit forms of the two-mode noise spectra are proved to be
$$\textrm{S}_{M+(-)}^{in}\left[\delta_{M+(-)}\right] = \frac{D_{M+(-)}\left[\delta_{M+(-)}\right]\Omega_{+(-)}\Delta_{R}}{\Omega_{t}\Omega_{e}^{2}}\left[\Omega_{1}\sigma_{MM}^{ss}-\Omega_{2}\sigma_{+{+}(-{-})}^{ss}\right],$$
$$\textrm{S}_{MX}^{in}\left(\delta_{MX}\right) = \frac{D_{MX}\left(\delta_{MX}\right)\Delta_{R}}{\Omega_{t}^{2}}\left(\Omega_{1}\sigma_{MM}^{ss}-\Omega_{2}\sigma_{XX}^{ss}\right),$$
$$\textrm{S}_{X+(-)}^{out}\left[\delta_{X+(-)}\right] = \frac{D_{X+(-)}\left[\delta_{X+(-)}\right]\Delta_{R}}{2\Omega_{e}^{2}}\left[\Omega_{1}\sigma_{XX}^{ss}-\Omega_{2}\sigma_{+{+}(-{-})}^{ss}\right],$$
$$\begin{aligned} \textrm{S}_{+{-}}^{out}\left(\delta_{+{-}}\right) = & -\frac{D_{+{-}}\left(\delta_{+{-}}\right)}{\Omega_{t}^{2}}\left[\frac{\Delta_{R}}{\Omega_{e}^{2}}\left(\Omega_{1}\Omega_{+}^{2}-\Omega_{2}\Omega_{-}^{2}\right)+\Omega_{1}^{2}\right]\sigma_{+{+}}^{ss}\\ &-\frac{D_{-{+}}\left(\delta_{-{+}}\right)}{\Omega_{t}^{2}}\left[\frac{\Delta_{R}}{\Omega_{e}^{2}}\left(\Omega_{1}\Omega_{-}^{2}-\Omega_{2}\Omega_{+}^{2}\right)+\Omega_{2}^{2}\right]\sigma_{-{-}}^{ss}, \end{aligned}$$
where the function $D_{mn}\left (\delta _{mn}\right )=\textrm {Re}[f_{mn}]/(\textrm {Re}[f_{mn}]^{2}+\delta _{mn}^{2})$ determines the robustness of the two-mode noise spectra on the photon-sideband detunings. Obviously, the generation of squeezing in the resonance fluorescence strongly relies on the steady-state populations under certain conditions.

To provide a general scheme for the generation of two-mode squeezing, we plot the noise spectra as the function of the photon-sideband detunings under different carrier concentrations, where the leapfrog arrows denote different two-photon transitions and the transition frequencies are also given (the level energies are $E_{+}$, $\chi /2$, $0$ and $E_{-}$ for the states $\left |+\right \rangle$, $\left |X\right \rangle$, $\left |M\right \rangle$ and $\left |-\right \rangle$). As indicated by the results, the carrier concentration of the BP metasurfaces plays a key role in the generation and the enhancement of squeezing. In Fig. 5(a), the noise spectrum $\textrm {S}_{M-}^{in}$ exhibits remarkable squeezing for $n_{c}=2.2\times 10^{18}/\textrm {m}^2$ at the Rabi detuning $\Delta _{R}=-1\;\textrm {meV}$. Recalling the result of Fig. 3(d), it is clear that in this case strong population inversion can be observed, where the steady-state population of the state $\left |-\right \rangle$ is much larger than that of the state $\left |M\right \rangle$. Thus the term $\Omega _{1}\sigma _{MM}^{ss}-\Omega _{2}\sigma _{--}^{ss}$ is effectively modified and leads to the enhancement of squeezing in the noise spectrum $\textrm {S}_{M-}^{in}$. In addition, the result also indicates a photon-sideband detuning of $0.25\;\textrm {meV}$ in order to obtain half of the maximum squeezing.

 figure: Fig. 5.

Fig. 5. Two-mode noise spectra of the in-phase and out-of-phase quadratures versus the photon-sideband detunings under different carrier concentrations, where the diagram of the related two-photon transitions are also drawn (including the transition channels and the resonant frequencies, the emissions with orthogonal polarizations can be distinguished from the arrows marked in different colors). The material parameters are the same as in Fig. 3, with the Rabi detunings taken as: (a) $\Delta _{R}=-1\;\textrm {meV}$, (b) $\Delta _{R}=-2\;\textrm {meV}$, (c) $\Delta _{R}=1\;\textrm {meV}$, (d) $\Delta _{R}=10\;\textrm {meV}$, (e) $\Delta _{R}=6\;\textrm {meV}$ and (f) $\Delta _{R}=-12\;\textrm {meV}$.

Download Full Size | PDF

The production of squeezing via other two-photon transitions are demonstrated in Figs. 5(b)–5(f). From previous discussions one may notice that the coupling between the QD and the metasurfaces is important in controlling the steady-state populations. As the result, two-mode squeezing can be produced and largely enhanced by manipulating the carrier concentration, where the photon-sideband detuning ranges from $0.014\;\textrm {meV}$ (Fig. 5(f)) to $0.57\;\textrm {meV}$ (Fig. 5(e)) in order to achieve half degree of the maximum squeezing. Moreover, Figs. 5(b) and 5(d) clearly illustrate the enhancements in the squeezing via manipulations on the carrier concentration. This behavior can be attributed to the prominent population accumulations in one of the energy states corresponds to the two-photon emissions. For example, in Fig. 5(d) the photon-sideband detuning of the noise spectrum strongly depends on the real part of the polarization coefficient $f_{X+}=i\Omega _{-}-\left (\Gamma _{++}+\sum _{m=+,M,-}\Gamma _{{X}\scriptstyle m}+\sum _{n=X,M,-}\Gamma _{{+}\scriptstyle n}\right )$. The decrease in the carrier concentration then results in the accumulation of the population into the state $\left |+\right \rangle$ (see Fig. 3), which indicates a decline in the decay rate from the state $\left |+\right \rangle$ to other energy states, i.e., the decay term $\sum _{n=X,M,-}\Gamma _{{+}\scriptstyle n}$ turns out to be smaller according to the steady-state solutions given in Eq. (15). Thus in one aspect, gains in the steady-state population $\sigma _{++}^{ss}$ prominently enhances the squeezing. On the other hand, the declines in the level decay rates reduces the tolerance of the photon-sideband detunings in the generation of squeezing. As demonstrated by the green and blue solid lines in Fig. 5, squeezing can be largely enhanced at the cost of declines in the tolerance of the photon-sideband detunings.

4. Conclusions

In conclusion, we have investigated the Purcell effect and the properties of the resonance fluorescent field in hybrid systems consist of a driven four-level biexciton QD coupled to the BP metasurfaces. The dispersive model of the material is introduced [66] to describe the optical response of the metasurfaces at different frequencies. Within the framework of the quantization scheme in the presence of inhomogeneous dissipative media [63], the dipole decay rates are expressed in terms of the Green’s functions [73]. We have shown that the Purcell factors of the orthogonal in-plane dipoles display unequal enhancements in different frequency regions for certain material parameters. In the meantime, the plasmon resonance exhibits prominent shifts according to the variation of the carrier concentration,which is experimentally achievable by the methods of biasing [79] or chemical doping [80]. Further studies indicate that the energy levels of the driven QD can be selectively populated by engineering the anisotropy of the metasurfaces. Based on this novel feature, we have explored the steady-state properties of the resonance fluorescence and the noise spectra under the two-photon resonance condition. As demonstrated, the resonance fluorescence of the driven QD are strongly modified by the metasurfaces owing to the couplings between transition dipoles and the SPM. Moreover, the generation of two-mode squeezing in the resonance fluorescence relies on the control of the steady-state populations of the driven system, which is feasible by manipulating the carrier concentration of the materials. We have also found that, although at the cost of declines in the tolerance of photon-sideband detunings, the degrees of squeezing can be largely enhanced according to the variation in the carrier concentration.

Beyond the BP metasurfaces considered in this work, it has been reported that anisotropic polariton modes can also be supported by two-dimensional or nanostructured metasurfaces like arrays of graphene nanoribbons [12], carbon phosphide [81] and van der Waals layers [15,82]. From this point of view, our proposal is promising in developing other tunable hybrid systems containing the metasurfaces mentioned above. After considering all the merits, we believe that our proposal is useful for the realization of integrated photonic platforms with high tunability, which may have potential applications in the atomic state manipulation, squeezed light generation and other quantum technologies.

Appendix A: Derivations of the spectral functions

The rigorous master equation for the reduced density operator of the driven QD, which can be derived by straightforward deductions according to the second-order perturbation theory, appears to be

$$\begin{aligned} \dot{\hat{\rho}} = & -\frac{i}{\hbar}\left[\hat{H}_{0}',\hat{\rho}\right]+\int_{0}^{\infty}d\omega\int_{0}^{t}d\tau e^{{-}i\left(\omega-\omega_{_{L}}\right)\tau} \Big[\gamma_{11}^{x}\left(\omega\right)\mathcal{O}_{_{GX,XG}}(\tau) +\gamma_{22}^{x}\left(\omega\right)\mathcal{O}_{_{XB,BX}}(\tau)\Big]\\ &\int_{0}^{\infty}d\omega\int_{0}^{t}d\tau e^{{-}i\left(\omega-\omega_{_{L}}\right)\tau}\Big\{ \gamma_{11}^{y}\left(\omega\right)\mathcal{O}_{_{GY,YG}}(\tau)+\gamma_{22}^{y}\left(\omega\right)\mathcal{O}_{_{YB,BY}}(\tau)+\gamma_{12}^{y}\left(\omega\right)\\ &\times\Big[\mathcal{O}_{_{GY,BY}}(\tau)+\mathcal{O}_{_{YB,YG}}(\tau)\Big]\Big\}. \end{aligned}$$
Here the operator $\mathcal {O}_{jl,mn}(\tau )=\hat {\sigma }_{jl}(-\tau )\hat {\rho }\hat {\sigma }_{mn}-\hat {\sigma }_{mn}\hat {\sigma }_{jl}(-\tau )\hat {\rho }$, where the time-dependent Pauli operators are easily obtained via the transformation $\hat {\sigma }_{jl}(\tau )=\exp (i\hat {H}_{0}'\tau /\hbar )\hat {\sigma }_{jl}\exp (-i\hat {H}_{0}'\tau /\hbar )$ and can be expressed in the combinations of the dressed operators. Starting from Eq. (24), after applying the secular approximation, the brief master equation and Bloch equations of the QD are shown in Eqs. (13), (15) and (16). The coefficients related to the polarization operators in Eq. (16) are
$$\begin{aligned} &f_{X-({+}M)}= i\Omega_{+}-\frac{1}{2}\Gamma_{+{+}}-\Gamma_{X-({+}M)}^{S},\\ &f_{+X(M-)}= i\Omega_{-}-\frac{1}{2}\Gamma_{+{+}}-\Gamma_{+X(M-)}^{S},\\ &f_{+{-}}= i\Omega-2\Gamma_{+{+}}-\Gamma_{+{-}}^{S},\\ &f_{XM}= i\Delta_{L}-\Gamma_{XM}^{S}, \end{aligned}$$
where the coefficients $\Gamma _{mn}^{S}\;(m\neq n$) are defined as $\Gamma _{mn}^{S}=(\sum _{j\neq m}\Gamma _{mj}+\sum _{l\neq n}\Gamma _{nl})/2$. The incoherent part of the resonance fluorescence spectrum for the driven system under consideration can be written as
$$\begin{aligned} \textrm{S}_{tot}(\omega)=&\textrm{Re}\sum_{P=X,Y}\lim_{t\rightarrow\infty}\int_{0}^{\infty}d\tau\Big[\left\langle \hat{\sigma}_{PG}\left(t+\tau\right),\hat{\sigma}_{GP}\left(t\right)\right\rangle +\left\langle \hat{\sigma}_{BP}\left(t+\tau\right),\hat{\sigma}_{PB}\left(t\right)\right\rangle\\ &+\left\langle \hat{\sigma}_{PG}\left(t+\tau\right),\hat{\sigma}_{PB}\left(t\right)\right\rangle +\left\langle \hat{\sigma}_{BP}\left(t+\tau\right),\hat{\sigma}_{GP}\left(t\right)\right\rangle \Big]e^{{-}i\left(\omega-\omega_{L}\right)\tau}. \end{aligned}$$
The numerical evaluation of $\textrm {S}_{tot}(\omega )$ requires the explicit forms of the two-time correlation functions, which can be derived with the application of the quantum regression theorem [77]. After arrangements, the resonance fluorescence spectrum can be expressed in the following form
$$\textrm{S}_{tot}(\delta_{L})=\textrm{S}_{cen}(\delta_{L})+\textrm{S}_{M-}(\delta_{L})+\textrm{S}_{M+}(\delta_{L})+\textrm{S}_{MX}(\delta_{L})+\textrm{S}_{X+}(\delta_{L})+\textrm{S}_{X-}(\delta_{L})+\textrm{S}_{+{-}}(\delta_{L}),$$
where the central peak signal denoted by ($\textrm {S}_{cen}(\delta _{L})$) is responsible for the elastic scattering of the drive and other terms originate from distinct two-photon transitions can be identified as different sidebands in the spectrum. They can be evaluated through
$$\begin{aligned} \textrm{S}_{cen}(\delta_{L})=&\frac{\Omega_{e}^{2}}{\Omega_{t}^2}\left[\mathit{L}_{+{+},+{+}}(\delta_{L})+\mathit{L}_{-{-},-{-}}(\delta_{L})-\mathit{L}_{+{+},-{-}}(\delta_{L})-\mathit{L}_{-{-},+{+}}(\delta_{L})\right],\\ \textrm{S}_{M-}(\delta_{L})=&\frac{\Omega_{-}^2}{\Omega_{e}^2\Omega_{t}^2}\Big[\Omega_{1}^2\mathit{L}_{M-,-M}(\delta_{L})+\Omega_{2}^2\mathit{L}_{-M,M-}(\delta_{L})\Big],\\ \textrm{S}_{M+}(\delta_{L})=&\frac{\Omega_{+}^2}{\Omega_{e}^2\Omega_{t}^2}\Big[\Omega_{1}^2\mathit{L}_{M+,+M}(\delta_{L})+\Omega_{2}^2\mathit{L}_{+M,M+}(\delta_{L})\Big],\\ \textrm{S}_{MX}(\delta_{L})=&\frac{1}{\Omega_{t}^2}\Big[\Omega_{1}^2\mathit{L}_{MX,XM}(\delta_{L})+\Omega_{2}^2\mathit{L}_{XM,MX}(\delta_{L})\Big],\\ \textrm{S}_{X+}(\delta_{L})=&\frac{\Omega_{-}^2}{\Omega_{e}^2\Omega_{t}^2}\Big[\Omega_{1}^2\mathit{L}_{X+,X+}(\delta_{L})+\Omega_{2}^2\mathit{L}_{+X,X+}(\delta_{L})\Big],\\ \textrm{S}_{X-}(\delta_{L})=&\frac{\Omega_{+}^2}{\Omega_{e}^2\Omega_{t}^2}\Big[\Omega_{1}^2\mathit{L}_{X-,-X}(\delta_{L})+\Omega_{2}^2\mathit{L}_{-X,X-}(\delta_{L})\Big],\\ \textrm{S}_{+{-}}(\delta_{L})=&\frac{1}{\Omega_{e}^2\Omega_{t}^2}\Big[\left(\Omega_{1}^2\Omega_{-}^2+\Omega_{2}^2\Omega_{+}^2\right)\mathit{L}_{-{+},+{-}}(\delta_{L})+\left(\Omega_{1}^2\Omega_{+}^2+\Omega_{2}^2\Omega_{-}^2\right)\mathit{L}_{+{-},-{+}}(\delta_{L})\Big]. \end{aligned}$$
Here the two-time correlation functions $\mathit {L}_{mn,jl}(\delta _{L})$ are defined in Eq. (18).

According to the definitions, the in-phase ($\textrm {S}^{in}(\omega )$) and out-of-phase ($\textrm {S}^{out}(\omega )$) components of the normally-order noise spectra for the system under consideration can be expanded as

$$\begin{aligned} \textrm{S}^{in}(\omega)=&\textrm{Re}\int_{0}^{\infty}d\tau\cos(\omega\tau)\Big[C_{d}^{i}\left\langle \hat{\sigma}_{+{+}}\left(\tau\right)-\hat{\sigma}_{-{-}}\left(\tau\right),\right.\left.\hat{\sigma}_{+{+}}\left(0\right)-\hat{\sigma}_{-{-}}\left(0\right)\right\rangle\\ &+\sum_{m\neq n={+},M,-}C_{mn}^{i}\left\langle \hat{\sigma}_{mn}\left(\tau\right),\hat{\sigma}_{nm}\left(0\right)\right\rangle+\sum_{j={+},M,-}\Big[C_{\scriptscriptstyle{X}\scriptstyle j}^{i}\left\langle \hat{\sigma}_{\scriptscriptstyle{X}\scriptstyle j}\left(\tau\right),\hat{\sigma}_{jX}\left(0\right)\right\rangle\\ &+C_{jX}^{i}\left\langle \hat{\sigma}_{jX}\left(\tau\right),\hat{\sigma}_{\scriptscriptstyle{X}\scriptstyle j}\left(0\right)\right\rangle\Big] \end{aligned}$$
and
$$\begin{aligned} \textrm{S}^{out}(\omega)=&\textrm{Re}\int_{0}^{\infty}d\tau\cos(\omega\tau)\Big[\sum_{m\neq n={+},M,-}C_{mn}^{o}\left\langle \hat{\sigma}_{mn}\left(\tau\right),\hat{\sigma}_{nm}\left(0\right)\right\rangle\\ &+\sum_{j={+},M,-}\Big[C_{\scriptscriptstyle{X}\scriptstyle j}^{o}\left\langle \hat{\sigma}_{\scriptscriptstyle{X}\scriptstyle j}\left(\tau\right),\hat{\sigma}_{jX}\left(0\right)\right\rangle+C_{jX}^{o}\left\langle \hat{\sigma}_{jX}\left(\tau\right),\hat{\sigma}_{\scriptscriptstyle{X}\scriptstyle j}\left(0\right)\right\rangle\Big]. \end{aligned}$$
The coefficients related to the weights of the spectral functions are of the forms
$$\begin{aligned} &C_{d}^{i}=\frac{2\left(\Omega_{1}+\Omega_{2}\right)^2}{\Omega_{t}^2},\\ &C_{MX}^{i(o)}=\frac{\Omega_{1}\Omega_{2}}{\Omega_{t}^2}\left(\frac{\Omega_{1}}{\Omega_{2}}\pm1\right),\\ &C_{XM}^{i(o)}=\frac{\Omega_{1}\Omega_{2}}{\Omega_{t}^2}\left(\frac{\Omega_{2}}{\Omega_{1}}\pm1\right),\\ &C_{+(-)M}^{i(o)}=\frac{\Omega_{1}\Omega_{2}\Omega_{+(-)}}{\Omega_{e}^2\Omega_{t}}\left(\frac{\Omega_{2}}{\Omega_{1}}\mp1\right),\\ &C_{M+(-)}^{i(o)}=\frac{\Omega_{1}\Omega_{2}\Omega_{+(-)}}{\Omega_{e}^2\Omega_{t}}\left(\frac{\Omega_{1}}{\Omega_{2}}\mp1\right),\\ &C_{+(-)X}^{i(o)}=\frac{\Omega_{1}\Omega_{2}\Omega_{-(+)}}{\Omega_{e}^2\Omega_{t}}\left(\frac{\Omega_{2}}{\Omega_{1}}\pm1\right),\\ &C_{X+(-)}^{i(o)}=\frac{\Omega_{1}\Omega_{2}\Omega_{-(+)}}{\Omega_{e}^2\Omega_{t}}\left(\frac{\Omega_{1}}{\Omega_{2}}\pm1\right),\\ &C_{+{-}}^{i(o)}=\frac{M}{\Omega_{e}^2\Omega_{t}^2}\left(\frac{\Omega_{1}\Omega_{+}}{\Omega_{2}\Omega_{-}}+\frac{\Omega_{2}\Omega_{-}}{\Omega_{1}\Omega_{+}}\mp\frac{\Omega_{1}}{\Omega_{2}}\pm\frac{\Omega_{-}}{\Omega_{+}}\pm\frac{\Omega_{\scriptscriptstyle+}}{\Omega_{-}}\right),\\ &C_{-{+}}^{i(o)}=\frac{M}{\Omega_{e}^2\Omega_{t}^2}\left(\frac{\Omega_{1}\Omega_{-}}{\Omega_{2}\Omega_{+}}+\frac{\Omega_{2}\Omega_{+}}{\Omega_{1}\Omega_{-}}\mp\frac{\Omega_{2}}{\Omega_{1}}\pm\frac{\Omega_{-}}{\Omega_{+}}\pm\frac{\Omega_{+}}{\Omega_{-}}\right), \end{aligned}$$
where the explicit form of the coefficient $M$ has been given in Eq. (14). In the frame rotating with the laser’s frequency, with the assumption of slowly varied transition operators of the dressed QD (adopting the transformation $\hat {\sigma }_{mn}(t)=\hat {\sigma }_{mn}'(t)\exp (-i\omega _{mn}t)$, $\hat {\sigma }_{mn}'(t)$ is the slowly varied operator), one can obtain the noise spectra describing different two-photon transitions shown in Eq. (19).

Funding

National Natural Science Foundation of China (11874287, 11947044); Huaqiao University (605-50Y19046); National Key Research and Development Program of China (2021YFA1400602).

Disclosures

The authors declare no conflicts of interest.

Data availability

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

Supplemental document

See Supplement 1 for supporting content.

References

1. D. K. Gramotnev and S. I. Bozhevolnyi, “Plasmonics beyond the diffraction limit,” Nat. Photonics 4(2), 83–91 (2010). [CrossRef]  

2. R. F. Oulton, V. J. Sorger, D. A. Genov, D. F. P. Pile, and X. Zhang, “A hybrid plasmonic waveguide for subwavelength confinement and long-range propagation,” Nat. Photonics 2(8), 496–500 (2008). [CrossRef]  

3. S. A. Maier and H. A. Atwater, “Plasmonics: Localization and guiding of electromagnetic energy in metal/dielectric structures,” J. Appl. Phys. 98(1), 011101 (2005). [CrossRef]  

4. H. Iwase, D. Englund, and J. Vuckovic, “Analysis of the Purcell effect in photonic and plasmonic crystals with losses,” Opt. Express 18(16), 16546–16560 (2010). [CrossRef]  

5. A. P. Slobozhanyuk, P. Ginzburg, D. A. Powell, I. Iorsh, A. S. Shalin, P. Segovia, A. V. Krasavin, G. A. Wurtz, V. A. Podolskiy, P. A. Belov, and A. V. Zayats, “Purcell effect in hyperbolic metamaterial resonators,” Phys. Rev. B 92(19), 195127 (2015). [CrossRef]  

6. D. Dzsotjan, A. S. Sørensen, and M. Fleischhauer, “Quantum emitters coupled to surface plasmons of a nanowire: A Green’s function approach,” Phys. Rev. B 82(7), 075427 (2010). [CrossRef]  

7. H. Lian, Y. Gu, J. J. Ren, F. Zhang, L. J. Wang, and Q. H. Gong, “Efficient single photon emission and collection based on excitation of gap surface plasmons,” Phys. Rev. Lett. 114(19), 193002 (2015). [CrossRef]  

8. A. Gonzalez-Tudela, D. Martin-Cano, E. Moreno, L. Martin-Moreno, C. Tejedor, and F. J. Garcia-Vidal, “Entanglement of two qubits mediated by one-dimensional plasmonic waveguides,” Phys. Rev. Lett. 106(2), 020501 (2011). [CrossRef]  

9. F. Carre no, M. A. Antón, V. Yannopapas, and E. Paspalakis, “Control of resonance fluorescence of a four-level quantum emitter near a plasmonic nanostructure,” Phys. Rev. A 95(4), 043825 (2017). [CrossRef]  

10. V. Yannopapas, E. Paspalakis, and N. V. Vitanov, “Plasmon-induced enhancement of quantum interference near metallic nanostructures,” Phys. Rev. Lett. 103(6), 063602 (2009). [CrossRef]  

11. A. A. High, R. C. Devlin, A. Dibos, M. Polking, D. S. Wild, J. Perczel, N. P. de Leon, M. D. Lukin, and H. Park, “Visible-frequency hyperbolic metasurface,” Nature 522(7555), 192–196 (2015). [CrossRef]  

12. J. S. Gomez-Diaz and A. Alú, “Flatland optics with hyperbolic metasurfaces,” ACS Photonics 3(12), 2211–2224 (2016). [CrossRef]  

13. T. Low, A. S. Rodin, A. Carvalho, Y. J. Jiang, H. Wang, F. N. Xia, and A. H. C. Neto, “Tunable optical properties of multilayer black phosphorus thin films,” Phys. Rev. B 90(7), 075434 (2014). [CrossRef]  

14. E. van Veen, A. Nemilentsau, A. Kumar, R. Roldan, M. I. Katsnelson, T. Low, and S. J. Yuan, “Tuning Two-Dimensional Hyperbolic Plasmons in Black Phosphorus,” Phys. Rev. Appl. 12(1), 014011 (2019). [CrossRef]  

15. P. N. Li, I. Dolado, F. J. Alfaro-Mozaz, F. Casanova, L. E. Hueso, S. Liu, J. H. Edgar, A. Y. Nikitin, S. Vélez, and R. Hillenbrand, “Infrared hyperbolic metasurface based on nanostructured van der Waals materials,” Science 359(6378), 892–896 (2018). [CrossRef]  

16. A. Woessner, M. B. Lundeberg, Y. Gao, A. Principi, P. Alonso-González, M. Carrega, K. Watanabe, T. Taniguchi, G. Vignale, M. Polini, J. Hone, R. Hillenbrand, and F. H. L. Koppens, “Highly confined low-loss plasmons in graphene-boron nitride heterostructures,” Nat. Mater. 14(4), 421–425 (2015). [CrossRef]  

17. Q. L. Hong, F. Xiong, W. Xu, Z. H. Zhu, K. Liu, X. D. Yuan, J. F. Zhang, and S. Q. Qin, “Towards high performance hybrid two-dimensional material plasmonic devices: strong and highly anisotropic plasmonic resonances in nanostructured graphene-black phosphorus bilayer,” Opt. Express 26(17), 22528–22535 (2018). [CrossRef]  

18. P. K. Venuthurumilli, P. D. Ye, and X. F. Xu, “Plasmonic resonance enhanced polarization-sensitive photodetection by black phosphorus in near infrared,” ACS Nano 12(5), 4861–4867 (2018). [CrossRef]  

19. A. N. Abbas, B. L. Liu, L. Chen, Y. Q. Ma, S. Cong, N. Aroonyadet, M. Koepf, T. Nilges, and C. W. Zhou, “Black phosphorus gas sensors,” ACS Nano 9(5), 5618–5624 (2015). [CrossRef]  

20. Y. J. Cai, K. D. Xu, N. X. Feng, R. R. Guo, H. J. Lin, and J. F. Zhu, “Anisotropic infrared plasmonic broadband absorber based on graphene-black phosphorus multilayers,” Opt. Express 27(3), 3101–3112 (2019). [CrossRef]  

21. B. R. Mollow, “Power spectrum of light scattered by two-level systems,” Phys. Rev. 188(5), 1969–1975 (1969). [CrossRef]  

22. G. Wrigge, I. Gerhardt, J. Hwang, G. Zumofen, and V. Sandoghdar, “Efficient coupling of photons to a single molecule and the observation of its resonance fluorescence,” Nat. Phys. 4(1), 60–66 (2008). [CrossRef]  

23. S. Ates, S. M. Ulrich, S. Reitzenstein, A. Loffler, A. Forchel, and P. Michler, “Post-selected indistinguishable photons from the resonance fluorescence of a single quantum dot in a microcavity,” Phys. Rev. Lett. 103(16), 167402 (2009). [CrossRef]  

24. Z. Ficek and S. Swain, “Squeezing in anomalous resonance fluorescence,” J. Opt. Soc. Am. B 14(2), 258–262 (1997). [CrossRef]  

25. L. Hanschke, L. Schweickert, J. C. L. Carreno, E. Schoell, K. D. Zeuner, T. Lettner, E. Z. Casalengua, M. Reindl, S. F. C. da Silva, R. Trotta, J. J. Finley, A. Rastelli, E. del Valle, F. P. Laussy, V. Zwiller, K. Muller, and K. D. Jons, “Origin of antibunching in resonance fluorescence,” Phys. Rev. Lett. 125(17), 170402 (2020). [CrossRef]  

26. R. Proux, M. Maragkou, E. Baudin, C. Voisin, P. Roussignol, and C. Diederichs, “Measuring the photon coalescence time window in the continuous-wave regime for resonantly driven semiconductor quantum dots,” Phys. Rev. Lett. 114(6), 067401 (2015). [CrossRef]  

27. S. L. Braunstein and P. van Loock, “Quantum information with continuous variables,” Rev. Mod. Phys. 77(2), 513–577 (2005). [CrossRef]  

28. C. Weedbrook, S. Pirandola, R. Garcia-Patron, N. J. Cerf, T. C. Ralph, J. H. Shapiro, and S. Lloyd, “Gaussian quantum information,” Rev. Mod. Phys. 84(2), 621–669 (2012). [CrossRef]  

29. W. Wang, Y. Wu, Y. Ma, W. Cai, L. Hu, X. Mu, Y. Xu, Z. J. Chen, H. Wang, Y. P. Song, H. Yuan, C. L. Zou, L. M. Duan, and L. Sun, “Heisenberg-limited single-mode quantum metrology in a superconducting circuit,” Nat. Commun. 10(1), 4382 (2019). [CrossRef]  

30. J. D. Zhang, Z. J. Zhang, L. Z. Cen, M. Yu, S. Li, F. Wang, and Y. Zhao, “Effects of imperfect elements on resolution and sensitivity of quantum metrology using two-mode squeezed vacuum state,” Opt. Express 25(21), 24907–24916 (2017). [CrossRef]  

31. H. Grote, K. Danzmann, K. L. Dooley, R. Schnabel, J. Slutsky, and H. Vahlbruch, “First long-term application of squeezed states of light in a gravitational-wave observatory,” Phys. Rev. Lett. 110(18), 181101 (2013). [CrossRef]  

32. F. H. Xu, X. F. Ma, Q. Zhang, H. K. Lo, and J. W. Pan, “Secure quantum key distribution with realistic devices,” Rev. Mod. Phys. 92(2), 025002 (2020). [CrossRef]  

33. I. Derkach, V. C. Usenko, and R. Filip, “Squeezing-enhanced quantum key distribution over atmospheric channels,” New J. Phys. 22(5), 053006 (2020). [CrossRef]  

34. M. L. Gu, C. Weedbrook, N. C. Menicucci, T. C. Ralph, and P. van Loock, “Quantum computing with continuous-variable clusters,” Phys. Rev. A 79(6), 062318 (2009). [CrossRef]  

35. U. L. Andersen, J. S. Neergaard-Nielsen, P. van Loock, and A. Furusawa, “Hybrid discrete- and continuous-variable quantum information,” Nat. Phys. 11(9), 713–719 (2015). [CrossRef]  

36. C. Xu, L. D. Zhang, S. T. Huang, T. X. Ma, F. Liu, H. Yonezawa, Y. Zhang, and M. Xiao, “Sensing and tracking enhanced by quantum squeezing,” Photonics Res. 7(6), A14–A26 (2019). [CrossRef]  

37. P. Marek, H. Jeong, and M. S. Kim, “Generating “squeezed” superpositions of coherent states using photon addition and subtraction,” Phys. Rev. A 78(6), 063811 (2008). [CrossRef]  

38. G. F. Huang, W. W. Deng, H. T. Tan, and G. L. Cheng, “Generation of squeezed states and single-phonon states via homodyne detection and photon subtraction on the filtered output of an optomechanical cavity,” Phys. Rev. A 99(4), 043819 (2019). [CrossRef]  

39. V. Svarc, J. Hlousek, M. Novakova, J. Fiurasek, and M. Jezek, “Feedforward-enhanced Fock state conversion with linear optics,” Opt. Express 28(8), 11634–11644 (2020). [CrossRef]  

40. G. X. Li, H. T. Tan, and M. Macovei, “Enhancement of entanglement for two-mode fields generated from four-wave mixing with the help of the auxiliary atomic transition,” Phys. Rev. A 76(5), 053827 (2007). [CrossRef]  

41. C. F. McCormick, V. Boyer, E. Arimondo, and P. D. Lett, “Strong relative intensity squeezing by four-wave mixing in rubidium vapor,” Opt. Lett. 32(2), 178–180 (2007). [CrossRef]  

42. Z. Q. Yang, P. Saurabh, F. Schlawin, S. Mukamel, and K. E. Dorfman, “Multidimensional four-wave-mixing spectroscopy with squeezed light,” Appl. Phys. Lett. 116(24), 244001 (2020). [CrossRef]  

43. H. Vahlbruch, M. Mehmet, K. Danzmann, and R. Schnabel, “Detection of 15 dB squeezed states of light and their application for the absolute calibration of photoelectric quantum efficiency,” Phys. Rev. Lett. 117(11), 110801 (2016). [CrossRef]  

44. Z. J. Wan, J. X. Feng, Y. J. Li, and K. S. Zhang, “Comparison of phase quadrature squeezed states generated from degenerate optical parametric amplifiers using PPKTP and PPLN,” Opt. Express 26(5), 5531–5540 (2018). [CrossRef]  

45. S. P. Liu, W. X. Yang, Z. H. Zhu, T. Shui, and L. Li, “Quadrature squeezing of a higher-order sideband spectrum in cavity optomechanics,” Opt. Lett. 43(1), 9–12 (2018). [CrossRef]  

46. K. N. Qu and G. S. Agarwal, “Generating quadrature squeezed light with dissipative optomechanical coupling,” Phys. Rev. A 91(6), 063815 (2015). [CrossRef]  

47. P. B. Li and F. L. Li, “Engineering squeezed states of microwave radiation with circuit quantum electrodynamics,” Phys. Rev. A 83(3), 035807 (2011). [CrossRef]  

48. R. Stassi, S. Savasta, L. Garziano, B. Spagnolo, and F. Nori, “Output field-quadrature measurements and squeezing in ultrastrong cavity-QED,” New J. Phys. 18(12), 123005 (2016). [CrossRef]  

49. J. Esteve, C. Gross, A. Weller, and S. Giovanazzi, “Squeezing and entanglement in a Bose-Einstein condensate,” Nature 455(7217), 1216–1219 (2008). [CrossRef]  

50. D. F. Walls and P. Zoller, “Reduced quantum fluctuations in resonance fluorescence,” Phys. Rev. Lett. 47(10), 709–711 (1981). [CrossRef]  

51. C. H. H. Schulte, J. Hansom, A. E. Jones, C. Matthiesen, C. Le Gall, and M. Atature, “Quadrature squeezed photons from a two-level system,” Nature 525(7568), 222–225 (2015). [CrossRef]  

52. Z. H. Lu, S. Bali, and J. E. Thomas, “Observation of squeezing in the phase-dependent fluorescence spectra of two-level atoms,” Phys. Rev. Lett. 81(17), 3635–3638 (1998). [CrossRef]  

53. P. Grünwald and W. Vogel, “Optimal squeezing in resonance fluorescence via atomic-state purification,” Phys. Rev. Lett. 109(1), 013601 (2012). [CrossRef]  

54. D. Martin-Cano, H. R. Haakh, K. Murr, and M. Agio, “Large suppression of quantum fluctuations of light from a single emitter by an optical nanostructure,” Phys. Rev. Lett. 113(26), 263605 (2014). [CrossRef]  

55. S. Kuhn, U. Hakanson, L. Rogobete, and V. Sandoghdar, “Enhancement of single-molecule fluorescence using a gold nanoparticle as an optical nanoantenna,” Phys. Rev. Lett. 97(1), 017402 (2006). [CrossRef]  

56. F. Hargart, M. Muller, K. Roy-Choudhury, S. L. Portalupi, C. Schneider, S. Hofling, M. Kamp, S. Hughes, and P. Michler, “Cavity-enhanced simultaneous dressing of quantum dot exciton and biexciton states,” Phys. Rev. B 93(11), 115308 (2016). [CrossRef]  

57. L. Ferrari, C. H. Wu, D. Lepage, X. Zhang, and Z. W. Liu, “Hyperbolic metamaterials and their applications,” Prog. Quantum Electron. 40, 1–40 (2015). [CrossRef]  

58. S. Dai, Q. Ma, M. K. Liu, T. Andersen, Z. Fei, M. D. Goldflam, M. Wagner, K. Watanabe, T. Taniguchi, M. Thiemens, F. Keilmann, G. C. A. M. Janssen, S. E. Zhu, P. Jarillo-Herrero, M. M. Fogler, and D. N. Basov, “Graphene on hexagonal boron nitride as a tunable hyperbolic metamaterial,” Nat. Nanotechnol. 10(8), 682–686 (2015). [CrossRef]  

59. E. M. Purcell, H. C. Torrey, and R. V. Pound, “Resonance absorption by nuclear magnetic moments in a solid,” Phys. Rev. 69(1-2), 37–38 (1946). [CrossRef]  

60. G. Long, S. G. Xu, X. B. Cai, Z. F. Wu, T. Y. Han, J. X. Z. Lin, C. Cheng, Y. Cai, X. R. Wang, and N. Wang, “Gate-tunable strong-weak localization transition in few-layer black phosphorus,” Nanotechnology 29(3), 035204 (2018). [CrossRef]  

61. W. Langbein, P. Borri, U. Woggon, V. Stavarache, D. Reuter, and A. D. Wieck, “Control of fine-structure splitting and biexciton binding in InxGa1-x As quantum dots by annealing,” Phys. Rev. B 69(16), 161301 (2004). [CrossRef]  

62. M. Glässl, A. M. Barth, K. Gawarecki, P. Machnikowski, M. D. Croitoru, S. Lüker, D. E. Reiter, T. Kuhn, and V. M. Axt, “Biexciton state preparation in a quantum dot via adiabatic rapid passage: comparison between two control protocols and impact of phonon-induced dephasing,” Phys. Rev. B 87(8), 085303 (2013). [CrossRef]  

63. A. Drezet, “Quantizing polaritons in inhomogeneous dissipative systems,” Phys. Rev. A 95(2), 023831 (2017). [CrossRef]  

64. A. Y. Nikitin, F. J. Garcia-Vidal, and L. Martin-Moreno, “Analytical expressions for the electromagnetic dyadic Green’s function in graphene and thin layers,” IEEE J. Sel. Top. Quant. 19(3), 4600611 (2013). [CrossRef]  

65. J. P. Nong, W. Wei, W. Wang, G. L. Lan, Z. G. Shang, J. M. Yi, and L. L. Tang, “Strong coherent coupling between graphene surface plasmons and anisotropic black phosphorus localized surface plasmons,” Opt. Express 26(2), 1633–1644 (2018). [CrossRef]  

66. T. Low, R. Roldán, H. Wang, F. Xia, P. Avouris, L. M. Moreno, and F. Guinea, “Plasmons and screening in monolayer and multilayer black phosphorus,” Phys. Rev. Lett. 113(10), 106802 (2014). [CrossRef]  

67. B. C. Yang, B. S. Wan, Q. H. Zhou, Y. Wang, W. T. Hu, W. M. Lv, Q. Chen, Z. M. Zeng, F. S. Wen, J. Y. Xiang, S. J. Yuan, J. L. Wang, B. S. Zhang, W. H. Wang, J. Y. Zhang, B. Xu, Z. S. Zhao, Y. J. Tian, and Z. Y. Liu, “Te-doped black phosphorus field-effect transistors,” Adv. Mater. 28(42), 9408–9415 (2016). [CrossRef]  

68. J. S. Qiao, X. H. Kong, Z. X. Hu, F. Yang, and W. Ji, “High-mobility transport anisotropy and linear dichroism in few-layer black phosphorus,” Nat. Commun. 5(1), 4475 (2014). [CrossRef]  

69. X. Huang, Y. Q. Cai, X. W. Feng, W. C. Tan, D. M. N. Hasan, L. Chen, N. Chen, L. Wang, L. Huang, T. J. Duffin, C. A. Nijhuis, Y. W. Zhang, C. Lee, and K. W. Ang, “Black phosphorus carbide as a tunable anisotropic plasmonic metasurface,” ACS Photonics 5(8), 3116–3123 (2018). [CrossRef]  

70. R. Hayn, T. Wei, V. M. Silkin, and J. van den Brink, “Plasmons in anisotropic Dirac systems,” Phys. Rev. Mater. 5(2), 024201 (2021). [CrossRef]  

71. W. L. Ma, P. Alonso-Gonzalez, S. J. Li, A. Y. Nikitin, J. Yuan, J. Martin-Sanchez, J. Taboada-Gutierrez, I. Amenabar, P. N. Li, and S. Velez, “In-plane anisotropic and ultra-low-loss polaritons in a natural van der Waals crystal,” Nature 562(7728), 557–562 (2018). [CrossRef]  

72. P. A. Huidobro, M. Kraft, S. A. Maier, and J. B. Pendry, “Graphene as a tunable anisotropic or isotropic plasmonic metasurface,” ACS Nano 10(5), 5499–5506 (2016). [CrossRef]  

73. A. Drezet, “Description of spontaneous photon emission and local density of states in the presence of a lossy polaritonic inhomogeneous medium,” Phys. Rev. A 95(4), 043844 (2017). [CrossRef]  

74. Z. Ficek and S. Swain, “Quantum interference in optical fields and atomic radiation,” J. Mod. Opt. 49(1-2), 3–42 (2002). [CrossRef]  

75. M. O. Scully and M. S. Zubairy, Quantum Optics (Cambridge University, 1997).

76. S. Smart and S. Swain, “3-level atom in a squeezed vacuum.II. resonance fluorescence,” J. Mod. Opt. 41(6), 1055–1077 (1994). [CrossRef]  

77. P. R. Rice and H. J. Carmichael, “Nonclassical effects in optical-spectra,” J. Opt. Soc. Am. B 5(8), 1661–1668 (1988). [CrossRef]  

78. W. Fang and Y. P. Yang, “Directional dipole radiations and long-range quantum entanglement mediated by hyperbolic metasurfaces,” Opt. Express 28(22), 32955–32977 (2020). [CrossRef]  

79. L. K. Li, G. J. Ye, V. Tran, R. X. Fei, G. R. Chen, H. C. Wang, J. Wang, K. Watanabe, T. Taniguchi, L. Yang, X. H. Chen, and Y. B. Zhang, “Quantum oscillations in a two-dimensional electron gas in black phosphorus thin films,” Nat. Nanotechnol. 10(7), 608–613 (2015). [CrossRef]  

80. L. Huang and K. W. Ang, “Black phosphorus photonics toward on-chip applications,” Appl. Phys. Rev. 7(3), 031302 (2020). [CrossRef]  

81. M. Dehdast, Z. Valiollahi, M. Neek-Amal, B. Van Duppen, F. M. Peeters, and M. Pourfath, “Tunable natural terahertz and mid-infrared hyperbolic plasmons in carbon phosphide,” Carbon 178, 625–631 (2021). [CrossRef]  

82. Z. B. Zheng, F. S. Sun, W. C. Huang, J. Y. Jiang, R. Z. Zhan, Y. L. Ke, H. J. Chen, and S. Z. Deng, “Phonon Polaritons in Twisted Double-Layers of Hyperbolic van der Waals Crystals,” Nano Lett. 20(7), 5301–5308 (2020). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       The optical response of the black phosphorus metasurfaces and the isofrequency contours of the surface plasmon modes. The Purcell factor in the isotropic topological dispersion case is also included.

Data availability

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

Cited By

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

Alert me when this article is cited.


Figures (5)

Fig. 1.
Fig. 1. Interaction between a driven QD and the anisotropic metasurfaces (here we consider BP structures): the QD couples to the anisotropic SPM supported by the surfaces. The pump frequency is reasonably chosen such that two-photon resonance condition is satisfied in the transition channel $\left |G\right \rangle \leftrightarrow \left |Y\right \rangle \leftrightarrow \left |B\right \rangle$. In the energy level diagram, the spontaneous emission of the exciton and biexciton states are modified by the anisotropy of the surfaces, the corresponding decay rates are denoted by $\gamma _{jj}^{y}$ ($y-$polarized transitions) and $\gamma _{jj}^{x}$ ($x-$polarized transitions).
Fig. 2.
Fig. 2. Purcell factors of dipoles with different polarizations coupled to the BP metasurfaces, different carrier concentrations $n_{c}=\left (0.2,0.5,1,1.5,2\right )\times 10^{18}/\textrm {m}^{2}$ are considered to stress the modification of the metasurfaces on the dipole radiation. The spatial coordination of the QD is $\mathbf {r}_{Q}=\left (0,0,20\right )\textrm {nm}$, the material parameters are $m_{x}=0.2m_{0}$, $m_{y}=0.7m_{0}$, $\omega _{r}=2.5\textrm {THz}$, $\varpi _{x}/2\pi =240\textrm {THz}$, $\varpi _{y}/2\pi =80\textrm {THz}$, $\mathcal {I_{\mathit {x}}}=1.5\kappa _{0}$ and $\mathcal {I_{\mathit {y}}}=3.5\kappa _{0}$, here $\kappa _{0}=e^{2}/4\hbar$ is the conductance quantum and $m_{0}$ is the free electron mass. (a) The Purcell factor of $y-$polarized dipoles. (b) The Purcell factor of $x-$polarized dipoles.
Fig. 3.
Fig. 3. Steady-state populations of the energy states $\left |+\right \rangle$, $\left |-\right \rangle$, $\left |M\right \rangle$ and $\left |X\right \rangle$ (denoted by $\sigma _{++}^{ss}$, $\sigma _{--}^{ss}$, $\sigma _{MM}^{ss}$ and $\sigma _{XX}^{ss}$), as a function of the Rabi detuning ($\Delta _{R}=\Omega _{2}-\Omega _{1}$, the Rabi energy $\Omega _{1}=20\;\textrm {meV}$). The influence of the carrier concentration on the population distributions are studied: (a) $n_{c}=10^{17}/\textrm {m}^{2}$, (b) $n_{c}=2\times 10^{17}/\textrm {m}^{2}$, (c) $n_{c}=1.1\times 10^{18}/\textrm {m}^{2}$ and (d) $n_{c}=2.2\times 10^{18}/\textrm {m}^{2}$, other parameters are the same as in Fig. 2.
Fig. 4.
Fig. 4. Steady-state resonance fluorescence spectra versus the photon-laser detuning ($\delta _{L}=\omega -\omega _{L}$) and the carrier concentration of the metasurfaces, the material parameters are the same as used in Fig. 2. The resonance fluorescence spectra denoting distinct two-photon emissions are plotted: (a) $\left |M\right \rangle \rightarrow \left |-\right \rangle \rightarrow \left |M\right \rangle$, $\hbar \Delta _{R}=-1\;\textrm {meV}$, (b) $\left |M\right \rangle \rightarrow \left |+\right \rangle \rightarrow \left |M\right \rangle$, $\hbar \Delta _{R}=-3\;\textrm {meV}$, (c) $\left |M\right \rangle \rightarrow \left |X\right \rangle \rightarrow \left |M\right \rangle$, $\hbar \Delta _{R}=3\;\textrm {meV}$, (d) $\left |X\right \rangle \rightarrow \left |+\right \rangle \rightarrow \left |X\right \rangle$, $\hbar \Delta _{R}=3\;\textrm {meV}$, (e) $\left |X\right \rangle \rightarrow \left |-\right \rangle \rightarrow \left |X\right \rangle$, $\hbar \Delta _{R}=-1\;\textrm {meV}$, and (f) $\left |+\right \rangle \rightarrow \left |-\right \rangle \rightarrow \left |+\right \rangle$, $\hbar \Delta _{R}=-3\;\textrm {meV}$. The carrier concentrations are taken as $n_{c}=(0.2,1.1,2.2)\times 10^{18}/\textrm {m}^2$ for subfigures (a), (c), (d), (e) and $n_{c}=(0.1,1.1,2.2)\times 10^{18}/\textrm {m}^2$ for subfigures (b), (f).
Fig. 5.
Fig. 5. Two-mode noise spectra of the in-phase and out-of-phase quadratures versus the photon-sideband detunings under different carrier concentrations, where the diagram of the related two-photon transitions are also drawn (including the transition channels and the resonant frequencies, the emissions with orthogonal polarizations can be distinguished from the arrows marked in different colors). The material parameters are the same as in Fig. 3, with the Rabi detunings taken as: (a) $\Delta _{R}=-1\;\textrm {meV}$, (b) $\Delta _{R}=-2\;\textrm {meV}$, (c) $\Delta _{R}=1\;\textrm {meV}$, (d) $\Delta _{R}=10\;\textrm {meV}$, (e) $\Delta _{R}=6\;\textrm {meV}$ and (f) $\Delta _{R}=-12\;\textrm {meV}$.

Equations (31)

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

H ^ 0 = Δ L ( σ ^ X X + σ ^ Y Y ) + [ Ω 1 σ ^ Y G + Ω 2 σ ^ B Y + H.c. ]
H ^ I = [ ( σ ^ Y G p y 1 + σ ^ B Y p y 2 + σ ^ X G p x 1 + σ ^ B X p x 2 ) d ω p 1 E ^ ( + ) ( r Q , ω ) e i ( ω L ω ) t + H.c. ] .
E ^ ( + ) ( r Q , ω ) = i ω μ 0 d r G ( r Q , r , ω ) j ^ ( r , ω ) ,
[ × 1 μ ( r , ω ) × ω 2 c 2 ε ( r , ω ) ] E ^ ( r , ω ) = i ω μ 0 j ^ ( r , ω ) .
G y y ( x x ) ( r , r ) = i 4 π 2 g y y ( x x ) R ( k ) ( 2 p ) 1 [ k 0 2 k y ( x ) 2 ] e i p ( z + z ) e i k ( ρ ρ ) d k x d k y + G y y ( x x ) P ( r , r ) ,
g y y R ( k ) = [ 2 k y 2 κ x x 2 ( k y 2 k 0 2 ) κ y y k 0 p κ x x κ y y η ] / D ( k ) ,
g x x R ( k ) = [ 2 k 0 2 κ x x 2 k x 2 ( κ x x κ y y ) k 0 p κ x x κ y y η ] / D ( k ) ,
G y y ( x x ) P ( r , r ) = i 2 π k 3 ( 2 p ) 1 e i p | z z | J 0 ( k ρ ) d k , .
κ l l = i n c e 2 m l ( ω + i ω r ) + I l [ Θ ( ω ϖ l ) + i π ln | ω ϖ l | | ω + ϖ l | ] ( l = x , y ) ,
| + = E / E t ( E 1 | G + E + | Y + E 2 | B ) / E e ,
| = E + / E t ( E 1 | G + E | Y + E 2 | B ) / E e ,
| M = ( E 2 | G + E 1 | B ) / E e .
ρ ^ ˙ = i [ H ^ 0 , ρ ^ ] Γ + + ( σ ^ + + ρ ^ σ ^ + σ ^ ρ ^ σ ^ + + 1 2 L + + 1 2 L ) + 1 2 m n = + , , M , X Γ m n L m n .
$ Γ + + = Γ = 2 Ω t 2 [ Ω 1 2 γ 11 y ( 0 ) + 2 Ω 1 Ω 2 γ 12 y ( 0 ) + Ω 2 2 γ 22 y ( 0 ) ] , Γ + = 2 M ( Ω t Ω e ) 2 [ Ω 1 Ω + Ω 2 Ω γ 11 y ( Ω t ) 2 γ 12 y ( Ω t ) + Ω 2 Ω Ω 1 Ω + γ 22 y ( Ω t ) ] , Γ + = 2 M ( Ω t Ω e ) 2 [ Ω 1 Ω Ω 2 Ω + γ 11 y ( Ω t ) 2 γ 12 y ( Ω t ) + Ω 2 Ω + Ω 1 Ω γ 22 y ( Ω t ) ] , Γ + M = 2 Ω 2 2 Ω + Ω t Ω e 2 γ 11 y ( Ω + ) , Γ M + = 2 Ω 1 2 Ω + Ω t Ω e 2 γ 22 y ( Ω + ) , Γ X = 2 Ω 1 2 Ω + Ω t Ω e 2 γ 11 x ( Ω + ) , Γ X = 2 Ω 2 2 Ω + Ω t Ω e 2 γ 22 x ( Ω + ) , Γ M = 2 Ω 1 2 Ω Ω t Ω e 2 γ 22 y ( Ω ) , Γ M = 2 Ω 2 2 Ω Ω t Ω e 2 γ 11 y ( Ω ) , Γ + X = 2 Ω 2 2 Ω Ω t Ω e 2 γ 22 x ( Ω ) , Γ X + = 2 Ω 1 2 Ω Ω t Ω e 2 γ 11 x ( Ω ) , Γ X M = 2 Ω 2 2 Ω t 2 γ 11 x ( Δ L ) , Γ M X = 2 Ω 1 2 Ω t 2 γ 22 x ( Δ L ) , $
σ ^ ˙ n n = m n = + , , M , X Γ n m σ ^ n n + l n = + , , M , X Γ l n σ ^ l l , ( n = + , , M , X )
σ ^ ˙ j l = f j l σ ^ j l , ( j l = + , , M , X )
S s i d ( δ L ) = S M ( δ L ) + S M + ( δ L ) + S M X ( δ L ) + S X + ( δ L ) + S X ( δ L ) + S + ( δ L ) ,
L j l , m n ( δ L ) = Re 0 d τ σ ^ j l ( τ ) , σ ^ m n ( 0 ) e i δ L τ .
S m n i n ( o u t ) ( δ m n ) = Re 0 [ C m n i ( o ) σ ^ m n ( τ ) e i ω m n τ , σ ^ n m ( 0 ) e i ω 1 τ + C n m i ( o ) σ ^ n m ( τ ) e i ω n m τ , σ ^ m n ( 0 ) e i ω 2 τ ] d τ = C m n i ( o ) L m n , n m ( δ m n ) + C n m i ( o ) L n m , m n ( δ m n ) .
S M + ( ) i n [ δ M + ( ) ] = D M + ( ) [ δ M + ( ) ] Ω + ( ) Δ R Ω t Ω e 2 [ Ω 1 σ M M s s Ω 2 σ + + ( ) s s ] ,
S M X i n ( δ M X ) = D M X ( δ M X ) Δ R Ω t 2 ( Ω 1 σ M M s s Ω 2 σ X X s s ) ,
S X + ( ) o u t [ δ X + ( ) ] = D X + ( ) [ δ X + ( ) ] Δ R 2 Ω e 2 [ Ω 1 σ X X s s Ω 2 σ + + ( ) s s ] ,
S + o u t ( δ + ) = D + ( δ + ) Ω t 2 [ Δ R Ω e 2 ( Ω 1 Ω + 2 Ω 2 Ω 2 ) + Ω 1 2 ] σ + + s s D + ( δ + ) Ω t 2 [ Δ R Ω e 2 ( Ω 1 Ω 2 Ω 2 Ω + 2 ) + Ω 2 2 ] σ s s ,
ρ ^ ˙ = i [ H ^ 0 , ρ ^ ] + 0 d ω 0 t d τ e i ( ω ω L ) τ [ γ 11 x ( ω ) O G X , X G ( τ ) + γ 22 x ( ω ) O X B , B X ( τ ) ] 0 d ω 0 t d τ e i ( ω ω L ) τ { γ 11 y ( ω ) O G Y , Y G ( τ ) + γ 22 y ( ω ) O Y B , B Y ( τ ) + γ 12 y ( ω ) × [ O G Y , B Y ( τ ) + O Y B , Y G ( τ ) ] } .
f X ( + M ) = i Ω + 1 2 Γ + + Γ X ( + M ) S , f + X ( M ) = i Ω 1 2 Γ + + Γ + X ( M ) S , f + = i Ω 2 Γ + + Γ + S , f X M = i Δ L Γ X M S ,
S t o t ( ω ) = Re P = X , Y lim t 0 d τ [ σ ^ P G ( t + τ ) , σ ^ G P ( t ) + σ ^ B P ( t + τ ) , σ ^ P B ( t ) + σ ^ P G ( t + τ ) , σ ^ P B ( t ) + σ ^ B P ( t + τ ) , σ ^ G P ( t ) ] e i ( ω ω L ) τ .
S t o t ( δ L ) = S c e n ( δ L ) + S M ( δ L ) + S M + ( δ L ) + S M X ( δ L ) + S X + ( δ L ) + S X ( δ L ) + S + ( δ L ) ,
S c e n ( δ L ) = Ω e 2 Ω t 2 [ L + + , + + ( δ L ) + L , ( δ L ) L + + , ( δ L ) L , + + ( δ L ) ] , S M ( δ L ) = Ω 2 Ω e 2 Ω t 2 [ Ω 1 2 L M , M ( δ L ) + Ω 2 2 L M , M ( δ L ) ] , S M + ( δ L ) = Ω + 2 Ω e 2 Ω t 2 [ Ω 1 2 L M + , + M ( δ L ) + Ω 2 2 L + M , M + ( δ L ) ] , S M X ( δ L ) = 1 Ω t 2 [ Ω 1 2 L M X , X M ( δ L ) + Ω 2 2 L X M , M X ( δ L ) ] , S X + ( δ L ) = Ω 2 Ω e 2 Ω t 2 [ Ω 1 2 L X + , X + ( δ L ) + Ω 2 2 L + X , X + ( δ L ) ] , S X ( δ L ) = Ω + 2 Ω e 2 Ω t 2 [ Ω 1 2 L X , X ( δ L ) + Ω 2 2 L X , X ( δ L ) ] , S + ( δ L ) = 1 Ω e 2 Ω t 2 [ ( Ω 1 2 Ω 2 + Ω 2 2 Ω + 2 ) L + , + ( δ L ) + ( Ω 1 2 Ω + 2 + Ω 2 2 Ω 2 ) L + , + ( δ L ) ] .
S i n ( ω ) = Re 0 d τ cos ( ω τ ) [ C d i σ ^ + + ( τ ) σ ^ ( τ ) , σ ^ + + ( 0 ) σ ^ ( 0 ) + m n = + , M , C m n i σ ^ m n ( τ ) , σ ^ n m ( 0 ) + j = + , M , [ C X j i σ ^ X j ( τ ) , σ ^ j X ( 0 ) + C j X i σ ^ j X ( τ ) , σ ^ X j ( 0 ) ]
S o u t ( ω ) = Re 0 d τ cos ( ω τ ) [ m n = + , M , C m n o σ ^ m n ( τ ) , σ ^ n m ( 0 ) + j = + , M , [ C X j o σ ^ X j ( τ ) , σ ^ j X ( 0 ) + C j X o σ ^ j X ( τ ) , σ ^ X j ( 0 ) ] .
C d i = 2 ( Ω 1 + Ω 2 ) 2 Ω t 2 , C M X i ( o ) = Ω 1 Ω 2 Ω t 2 ( Ω 1 Ω 2 ± 1 ) , C X M i ( o ) = Ω 1 Ω 2 Ω t 2 ( Ω 2 Ω 1 ± 1 ) , C + ( ) M i ( o ) = Ω 1 Ω 2 Ω + ( ) Ω e 2 Ω t ( Ω 2 Ω 1 1 ) , C M + ( ) i ( o ) = Ω 1 Ω 2 Ω + ( ) Ω e 2 Ω t ( Ω 1 Ω 2 1 ) , C + ( ) X i ( o ) = Ω 1 Ω 2 Ω ( + ) Ω e 2 Ω t ( Ω 2 Ω 1 ± 1 ) , C X + ( ) i ( o ) = Ω 1 Ω 2 Ω ( + ) Ω e 2 Ω t ( Ω 1 Ω 2 ± 1 ) , C + i ( o ) = M Ω e 2 Ω t 2 ( Ω 1 Ω + Ω 2 Ω + Ω 2 Ω Ω 1 Ω + Ω 1 Ω 2 ± Ω Ω + ± Ω + Ω ) , C + i ( o ) = M Ω e 2 Ω t 2 ( Ω 1 Ω Ω 2 Ω + + Ω 2 Ω + Ω 1 Ω Ω 2 Ω 1 ± Ω Ω + ± Ω + Ω ) ,
Select as filters


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