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

Optical bistability in a heterodimer composed of a quantum dot and a metallic nanoshell

Open Access Open Access

Abstract

We theoretically explore the conditions for generating optical bistability (OB) in a heterodimer comprised of a semiconductor quantum dot (SQD) and a metallic nanoshell (MNS). The MNS is made of a metallic nanosphere as a core and a dielectric material as a shell. For the specific hybrid system considered, the bistable effect appears only if the frequency of the pump field is equal to (or slightly less than) the exciton frequency for a proper shell thickness. Bistability phase diagrams, when plotted, show that the dipole-induced bistable region can be greatly broadened by changing the shell thickness of the MNS in a strong exciton-plasmon coupling regime. In particular, we demonstrate that the multipole polarization not only narrows the bistable zone but also enlarges the corresponding thresholds for a given intermediate scaled pumping intensity. On the other hand, when the SQD couples strongly with the MNS, the multipole polarization can also significantly broaden the bistable region and induce a great suppression of the FWM (four-wave mixing) signal for a fixed shell thickness. These interesting findings offer a fresh understanding of the bistability conditions in an SQD/MNS heterodimer, and may be useful in the fabrication of high-performance and low-threshold optical bistable nanodevices.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Exploiting exciton-plasmon interaction to manipulate optical properties of metal-semiconductor nanostructures is of fundamental importance [15]. One of these nanostructures of interest is a nanodimer composed of a semiconductor quantum dot (SQD) and a metal nanoparticle (MNP), which has been subjected to increasing investigation [69]. Since the pioneering work of Zhang and coworkers who demonstrated that the exciton-plasmon interaction can lead to some interesting phenomena such as Förster energy transfer, exciton energy shift, and nonlinear Fano resonance [10], a lot of other intriguing optical features arisen from strong exciton-plasmon coupling have also been observed [1121]. Examples include irreversible ultrafast properties [11], radiative and nonradiative energy transfer [12], tunable emission polarization [13], gain without inversion [14,15], magnetically induced optical transparency [16], controllable Kerr nonlinearity [17], enhanced coherent plasmonic field [11], modified four-wave mixing (FWM) effect [18,19], improved resonance fluorescence properties [20,21], and so on.

Optical bistability (OB) in the SQD/MNP hybrid nanosystem has also been extensively investigated both theoretically and experimentally [2232]. In a seminal work, Artuso and Bryant found that the bistable effect can be controlled by only changing the sizes of two NPs in a strongly coupled SQD/MNP nanodimer [22]. Malyshev et al. demonstrated that OB can be observed by measuring the optical hysteresis of Rayleigh scattering intensity [23]. Li et al. observed the bistable feature by detecting third-order nonlinear absorption and refraction [24,25]. Nugroho et al. proposed a fresh way to reveal the bistability mechanism and calculated the switching time between two stable branches in the SQD/MNP heterodimer [26]. Also, Asadpour et al. contributed some significant results in this area. They not only studied the role of a two-dimensional array of metal-coated dielectric nanospheres in modulating OB and optical multistability (OM) of a four-level quantum system embedded in a unidirectional ring cavity [27,28], but also demonstrated that the switching between OB and OM is feasible in a unidirectional cavity consisting of a hybrid SQD-MNP molecule [29]. Paspalakis et al. calculated the optical rectification coefficient and found that it is bivalued when the bistable effect arises [30]. In an interesting work of Mari et al., they reported that the fluorescence spectrum is strongly correlated to the initial state of the system when the strong transition and bistable regions appear [31]. In a core-shell NP-QD nanosystem, Miri et al. plotted bistability phase diagrams and found that the borders between Fano, double peaks, weak transition, strong transition, and bistability regions are strongly correlated to the thickness and chemical composition of both core and shell layers [32].

As is well known, surface plasmons of the naked MNP are present at its surface, however, surface plasmons of the metallic nanoshell (MNS) made by a metallic core and a dielectric shell appear at the interface between the core and the shell. Additionally, the dielectric environment of the MNS is obviously different from that of the naked MNP. It is conceivable that these features of the MNS will bring forth some novel optical phenomena in a SQD/MNS heterodimer. Considering this, in this paper we will explore the conditions for producing OB in a heterodimer composed of a SQD and a MNS. We investigate the influence of the shell thickness of the MNS on OB when the pump field is resonant or near-resonant with the exciton transition. We map out bistability phase diagrams in the system’s parameter subspace. We also discuss the role of the multipole polarization in controlling OB.

2. Theoretical model and method

As depicted schematically in Fig. 1, a heterodimer under consideration is composed of a SQD and a MNS. The heterodimer is simultaneously subjected to two applied fields consisting of a strong pump field (Epu, ωpu) and a weak probe field (Epr, ωpr). The MNS is made by a metallic core and a dielectric shell. The radius of the metallic core is denoted as R1, the total radius of the MNS is denoted as R2, and the thickness of the dielectric shell is R2 − R1. The SQD can be modeled as a two level system with a ground state |g > and a single exciton state |e > . The dielectric constants for the background, SQD, core and shell of the MNS refer, respectively, to εb, ε1, εm and εs. The center-to-center distance between the SQD and the MNS is denoted as d.

 figure: Fig. 1.

Fig. 1. Schematic of a heterodimer consisting of a SQD and a MNS.

Download Full Size | PDF

The total Hamiltonian of the SQD-MNS heterodimer in the rotating frame takes the form [24,33]:

$$H = \hbar {\Delta _{pu}}{\sigma _z} - \mu ({E_{SQD}}{\sigma _{10}} + E_{SQD}^\ast {\sigma _{01}}), $$
wherein Δpu = ωexωpu refers to the detuning between the exciton and the pump field, σij = |i><j| (i, j = 0, 1) denotes the transition operator from |i > to |j>, w = 2σz = σ11 σ00 represents the population inversion of a SQD exciton. ESQD corresponds to the total field experienced by the SQD, which is given by [3436]
$$\scalebox{0.9}{$\begin{aligned} {{\tilde{E}}_{SQD}} &= \frac{1}{{{\varepsilon _{eff}}}}\left[ {1 + \frac{{2R_2^3{\gamma_1}(\omega )}}{{{d^3}}}} \right]({{E_{pu}} + {E_{pr}}{e^{ - i\delta t}}} ) + \frac{1}{{8\pi {\varepsilon _0}{\varepsilon _b}{\varepsilon _{eff}}}} \times \sum\limits_{n = 1}^\infty {n({n + 1} ){{({n + 1} )}^2}\frac{{R_2^{2n + 1}{\gamma _n}(\omega )}}{{{d^{2n + 4}}}}{P_{SQD}}} \\ &= {\Pi _1}({{E_{pu}} + {E_{pr}}{e^{ - i\delta t}}} )+ {\Pi _2}{P_{SQD}} \end{aligned},$}$$
where Π1 = [1 + 2γ1(ω)R23/d3]/εeff, Π2 = [Σnn(n + 1)(n + 1)2γn(ω)PSQDR22n+ 1]/[8πε0εbεeffd2n+ 4]. εeff = (ε1 + 2εb)/(3εb) denotes the effective dielectric constant of the SQD, PSQD refers to the dipole moment of the SQD, and δ = ωprωpu represents the probe-pump detuning. The nth order (n = 1, 2, 3…) polarization factor γn(ω) can be written as [32,36,37]
$${\gamma _n}(\omega )= \frac{{n[{{\varepsilon_{mcn}}(\omega )- {\varepsilon_b}} ]}}{{n{\varepsilon _{mcn}}(\omega )+ ({n + 1} ){\varepsilon _b}}}, $$
$${\varepsilon _{mcn}} = {\varepsilon _m} \times \frac{{[{n{\varepsilon_m} + ({n + 1} ){\varepsilon_s}} ]R_2^{2n + 1} + n({{\varepsilon_m} - {\varepsilon_s}} )R_1^{2n + 1}}}{{[{n{\varepsilon_m} + ({n + 1} ){\varepsilon_s}} ]R_2^{2n + 1} - ({n + 1} )({{\varepsilon_m} - {\varepsilon_s}} )R_1^{2n + 1}}}, $$
By using the Heisenberg equation of motion and the commutation relations of different operators, the corresponding semiclassical equations read as follows
$$\begin{aligned} \frac{{d\left\langle {{\sigma_z}} \right\rangle }}{{dt}} &={-} {\Gamma _1}\left( {\left\langle {{\sigma_z}} \right\rangle + 1/2} \right) + i{\Omega _{pu}}\left[ {{\Pi _1}\left\langle {{\sigma_{10}}} \right\rangle - \Pi _1^\ast \left\langle {{\sigma_{01}}} \right\rangle } \right]\\ &+ i{\Omega _{pu}}\left( {{\Pi _1}\left\langle {{\sigma_{10}}} \right\rangle {e^{ - i\delta t}} - \Pi _1^\ast \left\langle {{\sigma_{01}}} \right\rangle {e^{i\delta t}}} \right) - 2{\mathop{\rm Im}\nolimits} G\left\langle {{\sigma_{01}}} \right\rangle \left\langle {{\sigma_{10}}} \right\rangle \end{aligned}, $$
$$\frac{{d\left\langle {{\sigma_{01}}} \right\rangle }}{{dt}} ={-} ({\Gamma _2} + i{\Delta _{pu}})\left\langle {{\sigma_{01}}} \right\rangle - 2i{\Pi _1}{\Omega _{pu}}\left\langle {{\sigma_z}} \right\rangle - 2i{\Pi _1}{\Omega _{pr}}{e^{ - i\delta t}}\left\langle {{\sigma_z}} \right\rangle - 2iG\left\langle {{\sigma_z}} \right\rangle \left\langle {{\sigma_{01}}} \right\rangle, $$
Where <σz > and <σ01> are the expectation values of operators σz and σ01, respectively. $\Omega_{p u}=\mu E_{p u} / \hbar, \; \text{and} \; \Omega_{p r}=\mu E_{p r} / \hbar,$ are the Rabi frequencies of the pump field and the probe field, respectively. $G=\mu^2 \Pi_2 / \hbar$ denotes the feedback parameter. Γ1 and Γ2 correspond, respectively, to the exciton relaxation rate and the exciton dephasing rate. To solve Eqs. (5) and (6), we make the following ansatz: $<\sigma_z>$ = σz(0) + σz(−)eiδt +σz(+)eiδt and <σ01> = σ01(0) + σ01(−)eiδt +σ01(+)eiδt. The formula of the FWM signal can be derived after substituting these expressions into Eqs. (5) and (6). By doing that, one can obtain
$$|{FWM} |= \left|{\frac{{\sigma_{01}^{( - )}}}{{E_{pr}^\ast {\hbar^{ - 1}}\Gamma _2^{ - 1}}}} \right|= \frac{{2{t_3}({{t_4}\sigma_z^{(0)} - {t_1}\sigma_{01}^{(0)}} )}}{{\mu {\hbar ^{ - 1}}\Gamma _2^{ - 1}({{t_1}{t_5}{t_7} - {t_2}{t_4}{t_7} + {t_1}{t_6}} )}}. $$
where w0 = 2σz(0), σ01(0) = {Im[Π1] − iRe[Π1]}Ωpuw0/{Γ2 − Im[G]w0 + iΔpu + iRe[G]w0}, t1 = {Γ2 − Im[G]w0 + i(δ − Δpu) − iRe[G]w0}, t2 = {Im[Π1] + iRe[Π1]}μΩpu + {Im[G] +iRe[G]}σz(0)*, $t_3=\mu^2\left\{\operatorname{Im}\left[\Pi_1\right]+i \operatorname{Re}\left[\Pi_1\right]\right\} / \hbar$, t4 = −{4σ01(0)Im[G] + 2μΩpuRe[Π1]} +2ΩpuRe[Π1], t5 = ( + Γ1)μ2, t6 = 2μΩpuIm[Π1] + 4σ01(0)*Im[G] + 2ΩpuRe[Π1], t7 = [Γ2 +i(δ + Δpu + Gw0)]/{μΩpuIm[Π1] + σ01(0)Im[G]) − ΩpuRe[Π1] − 01(0)Re[G]}.

The population inversion (w0) of the exciton in the SQD can be derived by [24]

$${\Gamma _1}({{w_0} + 1} )\{{{{[{{\Gamma _2} - ({{\mathop{\rm Im}\nolimits} G} ){w_0}} ]}^2} + {{[{{\Delta _{pu}} + ({Re G} ){w_0}} ]}^2}} \}+ 4{|{{\Pi _1}} |^2}\Omega _{pu}^2{\Gamma _2}{w_0} = 0 .$$

3. Results and discussions

We consider a hybrid heterodimer consisting of a CdSe QD and a MNS. The MNS is made of an Au core and the SiO2 shell. We take εm as the bulk dielectric constant of Au [38]. The exciton resonant frequency is chosen to 2.5 eV, which is close to the broad plasmon frequency of gold. The parameters are taken as εb = 1, εs = 2.16 [39], ε1 = 6, μ = 40 D, Γ1 = 1.25 ns−1, Γ2 = 3.33 ns−1 [10,22], R1 = 10 nm.

To begin, we study the influence of the shell thickness of the MNS on the FWM signal in the dipole approximation (N = 1). The variation of the exciton-population inversion w0 versus the exciton-pump field detuning Δpu for different shell thicknesses Δr is plotted in Fig. 2(a). As Δr = 0 nm, the spectrum of w0 exhibits a single-peaked structure. As Δr increases to 3 nm, the spectrum varies from a single-peaked structure to an S-shaped bistable one. However, a fresh bistable curve appears with Δr increasing to 5 nm or more. The physics behind Fig. 2(a) can be partly understood from the hysteresis loops. Take Δr = 3 nm as an example, when the exciton-pump field detuning Δpu increases, the system firstly follows the lower stable branch and jumps to the upper branch at a critical of Δpu. Upon sweeping Δpu back, the system remains on the upper branch and then transitions to the lower one at the other critical value. A hysteresis loop has been completed. The corresponding FWM spectra are plotted in Fig. 2(b). Before the appearance of OB, the FWM spectrum exhibits two asymmetric peaks, which are labeled by P1 and P2 peaks. As Δr = 3 nm, an asymmetric “U-shaped” bistable curve is observed. Such a curve traces out a crossed path ①→②→③→④→⑤. To gain a deeper insight into the physics behind in Fig. 2(b), the variations of the maximum value and position of the P1 peak versus Δr are presented in Fig. 2(c). As Δr increases, the P1 peak value firstly increases, reaching a maximum value ∼ 12.6 × 10−30 at Δr = 0.75 nm, then it decreases and reaches a minimum value when Δr = 0.92 nm. With a further increase in Δr, the P1 peak value increases again and then declines during 0.92 nm < Δr < 1.39 nm. Specially, the P1 peak disappears completely as Δr = 1.39 nm. This indicates that a proper Δr can help to modulate the FWM signal. In addition, the P1 peak position keeps almost unchanged. As we expected, the bistable effect occurs as Δr ≥ 1.39 nm. Figure 2(d) shows a charming bistability phase diagram in the system’s parameter subspace [Δr, Δpuc, d = 20 nm, Ipu = 200 GHz2]. As Δr is enlarged, the width of the bistable region firstly increases, and then decreases, finally increases continuously. As Δr increases from 1.39 nm to 6 nm, the bistable region varies from [-0.002 meV, −0.002 meV] to [-0.001 meV, 0.029 meV], suggesting that a large Δr may make the conditions for generating OB easier to achieve.

 figure: Fig. 2.

Fig. 2. Population inversion w0 (a) and the FWM signal (b), plotted as a function of the exciton-pump field detuning Δpu for different Δr. Δr denotes the SiO2 shell thickness. The inset in Fig. 2(b) shows the corresponding optical hysteresis loop. (c) Variations of the maximum value and position of the P1 peak versus Δr. (d) Bistability phase diagram in the system’s parameter space [Δr, Δpuc, d = 20 nm, Ipu = 200 GHz2]. Ipu = Ωpu2 denotes the scaled pumping intensity, and Δpuc refers to the critical value where the bistable effect just happens. The parameters used in Figs. 2(a)-(c) are N = 1, d = 20 nm, and Ipu = 200 GHz2.

Download Full Size | PDF

Inspired by Fig. 2(b), we find that the pumping frequency (i.e. Δpuc) and the SiO2 shell thickness (i.e. Δr) have a great impact on the bistable effect. Considering this, next we will mainly explore the Δr-dependent bistable behaviors in three different examples: (i) Δpuc = 0.05 meV, (ii) Δpuc = −0.05 meV, and (iii) Δpuc = 0 meV. Also, the influence of the multipole effect on OB is also considered. In Fig. 3(a) we study the variation of w0 as a function of the SQD-MNS distance d in the first example (i.e. Δpuc = 0.05 meV). Under the conditions of N = 1 and Δr = 0 nm, we observe that w0 exhibits a wide peak upon increasing d. With a further increase of Δr, this peak moves to a direction with a larger d. Unexpectedly, a standard S-shaped bistable curve appears as Δr = 7 nm. However, the situation for N = 20 is rather different. Herein, the bistable effect only arises as Δr = 0 nm. The corresponding FWM spectra are plotted in Fig. 3(b). We note that the lineshapes of the FWM spectra in the dipole approximation (N = 1) are similar to those of w0, however, the FWM spectrum evolves from a single-peaked curve into a fish-hook shaped bistable curve at Δr = 7 nm. Herein, the single peak shown in Fig. 3(b) is labeled as the P3 peak. To further clarify the contribution of the shell thickness to the P3 peak, in Fig. 3(c) we explore the variations of the maximum value and position of the P3 peak as a function of Δr. For N = 1, as Δr increases from 0 to 5 nm, the maximum value of the P3 peak increases by 5.6 times, and then it undergoes a sharp increase and reaches to a maximum value at Δr = 6 nm. The position of the P3 peak moves to a direction with a larger d as Δr increases. After eliminate the influence of the dipole polarization, we can draw a conclusion that the multipole polarization (i.e. N > 1) plays an insignificant role in controlling the P3 peak. Figure 3(d) presents the corresponding bistability phase diagram in the system’s parameter space [Δr, dc, Δpu = 0.05 meV, Ipu = 200 GHz2]. In the dipole approximation (N = 1), the bistable effect just appears as ΔrD = 6.14 nm and dcD = 21.07 nm. As Δr increases to 7.35 nm, the width of the bistable region reaches a maximum value of Δdc = 2.23 nm. Due to the multipole effect (N = 20), the bistable effect can only be observed in a region of Δr ∈ [0, 0.79] nm. This indicates that the shrinkage of the bistable region can be mainly ascribed to the multipole polarization.

 figure: Fig. 3.

Fig. 3. Population inversion w0 (a) and the FWM signal (b), plotted as a function of the SQD-MNS distance d for N = 1, 20 and different Δr. (c) Variations of the maximum value and position of the P3 peak versus Δr. (d) Bistability phase diagram in the system’s parameter space [Δr, dc, Δpu = 0.05 meV, Ipu = 200 GHz2]. dc denotes the bistable threshold. The parameters used in Figs. 3(a)-(c) are Δpu = 0.05 meV and Ipu = 200 GHz2.

Download Full Size | PDF

We study how the SQD-MNS distance and the SiO2 shell thickness affect the FWM signal in the second example (i.e. Δpuc = −0.05 meV). The curve of w0 versus d exhibits a wide peak, and w0 varies slightly (w0 ≈ −1) for N = 1 and Δr = 0 nm, indicating that there is a lack of a powerful mechanism to promote the population inversion of excitons in the current regime. However, due to the multipole effect (N = 20) and an increase of Δr, the peak of w0 moves to a direction with a larger d. In Fig. 4(b), the corresponding FWM signal is plotted against d for different Δr. We observe that, in the dipole approximation (N = 1), the contribution of Δr to the FWM signal differs significantly in both sides of the critical value d0 (i.e. d0 = 26.8 nm). As d < d0, the FWM signal is weakened with Δr increasing. However, as d > d0, the FWM signal is enhanced as Δr increases. Moreover, the magnitude of the FWM signal increases slightly and the FWM peak shifts to the direction with a larger d as Δr increases. This critical value d0 will be modified by the multipole polarization. Herein, d0 is equal to 40.5 nm as N = 20. Obviously, the results presented in Figs. 4(a)–4(b) are in sharp contrast to that in Figs. 3(a)–3(b). In the current regime, the bistable effect disappears completely, providing a fresh understanding for the bistability conditions.

 figure: Fig. 4.

Fig. 4. Population inversion w0 (a) and the FWM signal (b), plotted as a function of the SQD-MNS distance d for N = 1, 20 and different Δr. The parameters used are Δpu = -0.05 meV and Ipu = 200 GHz2.

Download Full Size | PDF

Finally, we focus on exploring the FWM behavior in the third example (i.e. Δpu = 0 meV). In Fig. 5(a) we observe that, in the dipole approximation (N = 1), the bistable effect occurs and the bistable region is broadened as Δr increases. Compared to the case of N = 1, the multipole polarization (N = 20) induces these bistable regions to move to a direction with a larger d. In Fig. 5(b), we plot the corresponding FWM signal as a function of d by changing the value of Δr. As expected, a serial of “U-shaped” bistable curves appear with increasing Δr (N = 1). The bistable effect arises in the region of d ∈ [12.56 nm, 17.94 nm] when the SiO2 shell of the MNS is absent (i.e. Δr = 0 nm). When Δr is enlarged to 3 nm, the bistable region becomes d ∈ [12.24 nm, 21.49 nm]. However, the trend for N = 20 is quite different. The FWM spectra assisted by the multipole effect are reshaped and switched from “U-shaped” bistable curves to “fish-hook shaped” ones. Additionally, the FWM signal is weakened at the positions of dc0 and dc1 as Δr increases (dc0 and dc1 refer to the lower and upper bistable thresholds, respectively). To further reveal the controllability of OB, we map out a bistability phase diagram in the system’s parameter subspace [Δr, dc, Δpu = 0 meV, Ipu = 200 GHz2] for N = 1 and N = 10, as shown in Fig. 5(c). We can see that, as N = 1 and Δr is increased, dc0 and dc1 both increase. It is significant to highlight one point that the value of d should take into account the sizes of the SQD and the MNS (i.e. dR1 + rSQD). Considering this, the width of the bistable region keeps almost unchanged in the dipole approximation (N = 1). Additionally, we note that, the FWM spectrum is completely different from that in Fig. 4(b), indicating that the bistable feature in the SQD-MNS heterodimer is strongly influenced by the presence of the resonance (i.e. Δpu = 0 meV). In the current regime, the multipole effect becomes more significant. The bistable region for N = 20 becomes narrower and is distributed in an area where dc1 and dc0 are both larger in comparison with that for N = 1. Taking the case of N = 1 and Δr = 6 nm for example, a wide bistable region of d ∈ [18.0 nm, 24.47 nm] appears. Unexpectedly, the width of the bistable region for N = 20 is shrunk to nearly 20% of that for N = 1. This indicates that the multipole polarization leads to a significant shrinkage of the bistable region.

 figure: Fig. 5.

Fig. 5. Population inversion w0 (a) and the FWM signal (b), plotted as a function of the SQD-MNS distance d for N = 1, 20 and different Δr. (c) Bistability phase diagram in the system’s parameter space [Δr, dc, Δpu = 0 meV, Ipu = 200 GHz2] for N = 1, 20. The parameters used in Figs. 5(a)-(b) are Δpu = 0 meV and Ipu = 200 GHz2.

Download Full Size | PDF

We proceed to study the variations of w0 and the FWM signal versus Ipu for different Δr in a strong exciton-plasmon coupling regime. Figure 6(a) shows that the width of the bistable region strongly replies on the value of Δr for N = 1. The larger Δr, the wider the bistable region correlated to the scaled pumping intensity Ipu. Due to the multipole effect (N = 20), the bistable region broadens greatly and moves to a direction with a larger Ipu for a given Δr. It is not difficult to find that the conditions for generating OB in our scheme can be satisfied easily if the coupling between two NPs is sufficiently strong and the pump field is resonant with the exciton transition. The corresponding FWM spectra are plotted in Fig. 6(b). An interesting result seen here is that at the locations of thresholds the FWM signals are weakened greatly with Δr increasing. For a given Δr, the multipole polarization leads to a significant reduction of the FWM signal in comparison to that in the dipole approximation. In addition, the FWM signal will be weakened as Δr increases, indicating that a thicker shell is not conducive to the enhancement of the FWM signal. Analogically, we plot a bistability phase diagram in the parameter subspace [Δr, Ipuc, Δpu = 0 meV, d = 20 nm] for N = 1 and N = 20 (Fig. 6(c)). As N = 1 and Δr is increased, the lower bistable threshold Ipuc0 changes slightly, however, the upper bistable threshold Ipuc1 undergoes a sharp rise accompanied by a strong broadening of the bistable region. For Δr = 0 nm, the bistable region is located at Ipu ∈ [19.8 GHz2, 49.8 GHz2], while this region is widened more than 30 times as Δr increases to 5 nm. As a comparison, the lower bistable threshold in the SQD-MNS nanodimer is only 20% of that reported in the SQD-MNP molecule [24]. To further uncover the influence of the multipole effect on OB in the strong exciton-plasmon coupling regime, we make a comparison of the bistable region in two cases of N = 20 and N = 1. The results show that, for Δr = 5 nm, the width of the bistable region for N = 20 is broadened by more than five orders of magnitude compared to that for N = 1. From these results, we can draw a conclusion that the multipole polarization can broaden greatly the bistable region as the SQD couples strongly with the MNS.

 figure: Fig. 6.

Fig. 6. Population inversion w0 (a) and the FWM signal (b), plotted as a function of Log[Ipu] for N = 1, 20 and different Δr (Ipu denotes the scaled pumping intensity). (c) Bistability phase diagram in the system’s parameter space [Δr, Ipuc, Δpu = 0 meV, d = 20 nm]. The parameters used in Figs. 6(a)-(b) are Δpu = 0 and d = 20 nm.

Download Full Size | PDF

4. Conclusions

We conducted a theoretical study of the conditions for achieving OB in a SQD/MNS heterodimer. Three different examples have been discussed under the dipole and multipole approximations. We find that the frequency of the pump field has an important impact on the generation of OB. We have also calculated bistability phase diagrams in the system’s parameter subspace and found that the dipole-induced bistable region is broadened effectively as the shell thickness Δr increases. We also demonstrated that the multipole polarization not only can narrow the bistable zone but also enlarge d-correlated bistability thresholds for a given intermediate scaled pumping intensity Ipu. For a given shell thickness Δr, the Ipu-related bistable region can be broadened greatly due to the multipole effect in a strong exciton-plasmon coupling regime. This work may open new horizons for fabricating high-performance and low-threshold optical bistable nanodevices.

Funding

Research Foundation of Education Bureau of Hunan Province (20B602, 21B0253); Natural Science Foundation of Hunan Province (2020JJ4935); National Natural Science Foundation of China (11404410).

Disclosures

The authors declare no conflicts of interest.

Data availability

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

References

1. S. Nie and S. R. Emory, “Probing single molecules and single nanoparticles by surface-enhanced Raman scattering,” Science 275(5303), 1102–1106 (1997). [CrossRef]  

2. J. Lee, P. Hernandez, J. Lee, A. O. Govorov, and N. A. Kotov, “Exciton-plasmon interactions in molecular spring assemblies of nanowires and wavelength-based protein detection,” Nat. Mater. 6(4), 291–295 (2007). [CrossRef]  

3. J. T. Zhang, Y. Tang, K. Lee, and M. Ouyang, “Tailoring light-matter-spin interactions in colloidal hetero-nanostructures,” Nature 466(7302), 91–95 (2010). [CrossRef]  

4. H. F. Xu, Z. M. Zhu, J. C. Xue, Q. Q. Zhan, Z. K. Zhou, and X. H. Wang, “Giant enhancements of high-order upconversion luminescence enabled by multiresonant hyperbolic metamaterials,” Photonics Res. 9(3), 395–404 (2021). [CrossRef]  

5. Q. Zhao, W. J. Zhou, Y. H. Deng, Y. Q. Zheng, Z. H. Shi, L. K. Ang, Z. K. Zhou, and L. Wu, “Plexcitonic strong coupling: unique features, applications, and challenges,” J. Phys. D: Appl. Phys. 55(20), 203002 (2022). [CrossRef]  

6. J. W. Tang, J. Xia, M. D. Fang, F. L. Bao, G. J. Cao, J. Q. Shen, J. L. Evans, and S. L. He, “Selective far-field addressing of coupled quantum dots in a plasmonic nanocavity,” Nat. Commun. 9(1), 1705 (2018). [CrossRef]  

7. W. X. Yang, A. X. Chen, Z. W. Huang, and R. K. Lee, “Ultrafast optical switching in quantum dot-metallic nanoparticle hybrid systems,” Opt. Express 23(10), 13032–13040 (2015). [CrossRef]  

8. D. Stefanatos, A. Smponias, I. Thanopulos, and E. Paspalakis, “Efficient exciton generation in a semiconductor quantum-dot–metal-nanoparticle composite structure using conventional chirped pulses,” Phys. Rev. A 105(5), 052604 (2022). [CrossRef]  

9. H. R. Hamedi, V. Yannopapas, G. Juzeliūnas, and E. Paspalakis, “Coherent optical effects in a three-level quantum emitter near a periodic plasmonic nanostructure,” Phys. Rev. B 106(3), 035419 (2022). [CrossRef]  

10. W. Zhang, A. O. Govorov, and G. W. Bryant, “Semiconductor-metal nanoparticle molecules: Hybrid excitons and the nonlinear Fano effect,” Phys. Rev. Lett. 97(14), 146804 (2006). [CrossRef]  

11. S. M. Sadeghi, W. J. Wing, and R. R. Gutha, “Control of plasmon fields via irreversible ultrafast dynamics caused by interaction of infrared laser pulses with quantum-dot-metallic-nanoparticle molecules,” Phys. Rev. A 92(2), 023808 (2015). [CrossRef]  

12. M. T. Cheng, S. D. Liu, H. J. Zhou, Z. H. Hao, and Q. Q. Wang, “Coherent exciton-plasmon interaction in the hybrid semiconductor quantum dot and metal nanoparticle complex,” Opt. Lett. 32(15), 2125–2127 (2007). [CrossRef]  

13. M. T. Cheng, S. D. Liu, and Q. Q. Wang, “Modulating emission polarization of semiconductor quantum dots through surface plasmon of metal nanorod,” Appl. Phys. Lett. 92(16), 162107 (2008). [CrossRef]  

14. S. M. Sadeghi, “Gain without inversion in hybrid quantum dot-metallic nanoparticle systems,” Nanotechnology 21(45), 455401 (2010). [CrossRef]  

15. D. X. Zhao, Y. Gu, J. R. Wu, J. X. Zhang, T. C. Zhang, B. D. Gerardot, and Q. H. Gong, “Quantum-dot gain without inversion: Effects of dark plasmon-exciton hybridization,” Phys. Rev. B 89(24), 245433 (2014). [CrossRef]  

16. J. H. Li, S. T. Shen, C. L. Ding, and Y. Wu, “Magnetically induced optical transparency in a plasmon-exciton system,” Phys. Rev. A 103(5), 053706 (2021). [CrossRef]  

17. S. G. Kosionis and E. Paspalakis, “Control of self-Kerr nonlinearity in a driven coupled semiconductor quantum dot-metal nanoparticle structure,” J. Phys. Chem. C 123(12), 7308–7317 (2019). [CrossRef]  

18. E. Paspalakis, S. Evangelou, S. G. Kosionis, and A. F. Terzis, “Strongly modified four-wave mixing in a coupled semiconductor quantum dot-metal nanoparticle system,” J. Appl. Phys. 115(8), 083106 (2014). [CrossRef]  

19. J. B. Li, M. D. He, and L. Q. Chen, “Four-wave parametric amplification in semiconductor quantum dot-metallic nanoparticle hybrid molecules,” Opt. Express 22(20), 24734–24741 (2014). [CrossRef]  

20. F. Carreño, M. A. Yannopapas, V. Antón, and E. Paspalakis, “Resonance fluorescence spectrum of a $\Lambda$-type quantum emitter close to a metallic nanoparticle,” Phys. Rev. A 94(1), 013834 (2016). [CrossRef]  

21. S. T. Shen, Z. M. Wu, J. H. Li, and Y. Wu, “Insights into Fano-type resonance fluorescence from quantum-dot-metal-nanoparticle molecules with a squeezed vacuum,” Phys. Rev. A 104(1), 013717 (2021). [CrossRef]  

22. R. D. Artuso and G. W. Bryant, “Optical response of strongly coupled quantum dot-metal nanoparticle systems: Double peaked Fano structure and bistability,” Nano Lett. 8(7), 2106–2111 (2008). [CrossRef]  

23. A. V. Malyshev and V. A. Malyshev, “Optical bistability and hysteresis of a hybrid metal-semiconductor nanodimer,” Phys. Rev. B 84(3), 035314 (2011). [CrossRef]  

24. J. B. Li, N. C. Kim, M. T. Cheng, L. Zhou, Z. H. Hao, and Q. Q. Wang, “Optical bistability and nonlinearity of coherently coupled exciton-plasmon systems,” Opt. Express 20(2), 1856–1861 (2012). [CrossRef]  

25. J. B. Li, S. Liang, M. D. He, L. Q. Chen, X. J. Wang, and X. F. Peng, “A tunable bistable device based on a coupled quantum dot-metallic nanoparticle nanosystem,” Appl. Phys. B 120(1), 161–166 (2015). [CrossRef]  

26. B. S. Nugroho, A. A. Iskandar, V. A. Malyshev, and J. Knoeste, “Bistable optical response of a nanoparticle heterodimer: Mechanism, phase diagram, and switching time,” J. Chem. Phys. 139(1), 014303 (2013). [CrossRef]  

27. S. H. Asadpour and H. Rahimpour Soleimani, “Phase dependence of optical bistability and multistability in a four-level quantum system near a plasmonic nanostructure,” J. Appl. Phys. 119(2), 023102 (2016). [CrossRef]  

28. S. H. Asadpour and H. R. Soleimani, “Optical bistability and multistability in a four-level quantum system in the presence of plasmonic nanostructure,” Phys. E 75, 112–117 (2016). [CrossRef]  

29. G. Solookinejad, M. Jabbari, M. Nafar, E. Ahmadi, and S. H. Asadpour, “Incoherent control of optical bistability and multistability in a hybrid system: Metallic nanoparticle-quantum dot nanostructure,” J. Appl. Phys. 124(6), 063102 (2018). [CrossRef]  

30. F. Carreño, M. A. Antón, and E. Paspalakis, “Nonlinear optical rectification and optical bistability in a coupled asymmetric quantum dot-metal nanoparticle hybrid,” J. Appl. Phys. 124(11), 113107 (2018). [CrossRef]  

31. A. Mohammadzadeh and M. F. Miri, “Resonance fluorescence of a hybrid semiconductor-quantum-dot-metal-nanoparticle system driven by a bichromatic field,” Phys. Rev. B 99(11), 115440 (2019). [CrossRef]  

32. Z. Naeimi, A. Mohammadzadeh, and M. F. Miri, “Optical response of a hybrid system composed of a quantum dot and a core-shell nanoparticle,” J. Opt. Soc. Am. B 36(8), 2317–2324 (2019). [CrossRef]  

33. R. W. Boyd, Nonlinear Optics, (Academic, San Diego, CA, 2006).

34. J. Y. Yan, W. Zhang, S. Q. Duan, X. G. Zhao, and A. O. Govorov, “Optical properties of coupled metal-semiconductor and metal-molecule nanocrystal complexes: Role of multipole effects,” Phys. Rev. B 77(16), 165301 (2008). [CrossRef]  

35. X. N. Liu, N. Kongsuwan, X. G. Li, D. X. Zhao, Z. M. Wu, O. Hess, and X. H. Zhang, “Tailoring the third-order nonlinear optical property of a hybrid semiconductor quantum dot-metal nanoparticle: From saturable to Fano-enhanced absorption,” J. Phys. Chem. Lett. 10(24), 7594–7602 (2019). [CrossRef]  

36. B. S. Nugroho, A. A. Iskandar, V. A. Malyshev, and J. Knoester, “Plasmon-assisted two-photon absorption in a semiconductor quantum dot-metallic nanoshell composite,” Phys. Rev. B 102(4), 045405 (2020). [CrossRef]  

37. M. R. Singh, P. D. Persaud, and S. Yastrebov, “A study of two-photon florescence in metallic nanoshells,” Nanotechnology 31(26), 265203 (2020). [CrossRef]  

38. E. D. Palik, Handbook of Optical Constants of Solids, (Academic, New York, 1985).

39. M. R. Singh and P. D. Persaud, “Dipole-dipole interaction in two-photon spectroscopy of metallic nanohybrids,” J. Phys. Chem. C 124(11), 6311–6320 (2020). [CrossRef]  

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 (6)

Fig. 1.
Fig. 1. Schematic of a heterodimer consisting of a SQD and a MNS.
Fig. 2.
Fig. 2. Population inversion w0 (a) and the FWM signal (b), plotted as a function of the exciton-pump field detuning Δpu for different Δr. Δr denotes the SiO2 shell thickness. The inset in Fig. 2(b) shows the corresponding optical hysteresis loop. (c) Variations of the maximum value and position of the P1 peak versus Δr. (d) Bistability phase diagram in the system’s parameter space [Δr, Δpuc, d = 20 nm, Ipu = 200 GHz2]. Ipu = Ωpu2 denotes the scaled pumping intensity, and Δpuc refers to the critical value where the bistable effect just happens. The parameters used in Figs. 2(a)-(c) are N = 1, d = 20 nm, and Ipu = 200 GHz2.
Fig. 3.
Fig. 3. Population inversion w0 (a) and the FWM signal (b), plotted as a function of the SQD-MNS distance d for N = 1, 20 and different Δr. (c) Variations of the maximum value and position of the P3 peak versus Δr. (d) Bistability phase diagram in the system’s parameter space [Δr, dc, Δpu = 0.05 meV, Ipu = 200 GHz2]. dc denotes the bistable threshold. The parameters used in Figs. 3(a)-(c) are Δpu = 0.05 meV and Ipu = 200 GHz2.
Fig. 4.
Fig. 4. Population inversion w0 (a) and the FWM signal (b), plotted as a function of the SQD-MNS distance d for N = 1, 20 and different Δr. The parameters used are Δpu = -0.05 meV and Ipu = 200 GHz2.
Fig. 5.
Fig. 5. Population inversion w0 (a) and the FWM signal (b), plotted as a function of the SQD-MNS distance d for N = 1, 20 and different Δr. (c) Bistability phase diagram in the system’s parameter space [Δr, dc, Δpu = 0 meV, Ipu = 200 GHz2] for N = 1, 20. The parameters used in Figs. 5(a)-(b) are Δpu = 0 meV and Ipu = 200 GHz2.
Fig. 6.
Fig. 6. Population inversion w0 (a) and the FWM signal (b), plotted as a function of Log[Ipu] for N = 1, 20 and different Δr (Ipu denotes the scaled pumping intensity). (c) Bistability phase diagram in the system’s parameter space [Δr, Ipuc, Δpu = 0 meV, d = 20 nm]. The parameters used in Figs. 6(a)-(b) are Δpu = 0 and d = 20 nm.

Equations (8)

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

$$H = \hbar {\Delta _{pu}}{\sigma _z} - \mu ({E_{SQD}}{\sigma _{10}} + E_{SQD}^\ast {\sigma _{01}}), $$
$$\scalebox{0.9}{$\begin{aligned} {{\tilde{E}}_{SQD}} &= \frac{1}{{{\varepsilon _{eff}}}}\left[ {1 + \frac{{2R_2^3{\gamma_1}(\omega )}}{{{d^3}}}} \right]({{E_{pu}} + {E_{pr}}{e^{ - i\delta t}}} ) + \frac{1}{{8\pi {\varepsilon _0}{\varepsilon _b}{\varepsilon _{eff}}}} \times \sum\limits_{n = 1}^\infty {n({n + 1} ){{({n + 1} )}^2}\frac{{R_2^{2n + 1}{\gamma _n}(\omega )}}{{{d^{2n + 4}}}}{P_{SQD}}} \\ &= {\Pi _1}({{E_{pu}} + {E_{pr}}{e^{ - i\delta t}}} )+ {\Pi _2}{P_{SQD}} \end{aligned},$}$$
$${\gamma _n}(\omega )= \frac{{n[{{\varepsilon_{mcn}}(\omega )- {\varepsilon_b}} ]}}{{n{\varepsilon _{mcn}}(\omega )+ ({n + 1} ){\varepsilon _b}}}, $$
$${\varepsilon _{mcn}} = {\varepsilon _m} \times \frac{{[{n{\varepsilon_m} + ({n + 1} ){\varepsilon_s}} ]R_2^{2n + 1} + n({{\varepsilon_m} - {\varepsilon_s}} )R_1^{2n + 1}}}{{[{n{\varepsilon_m} + ({n + 1} ){\varepsilon_s}} ]R_2^{2n + 1} - ({n + 1} )({{\varepsilon_m} - {\varepsilon_s}} )R_1^{2n + 1}}}, $$
$$\begin{aligned} \frac{{d\left\langle {{\sigma_z}} \right\rangle }}{{dt}} &={-} {\Gamma _1}\left( {\left\langle {{\sigma_z}} \right\rangle + 1/2} \right) + i{\Omega _{pu}}\left[ {{\Pi _1}\left\langle {{\sigma_{10}}} \right\rangle - \Pi _1^\ast \left\langle {{\sigma_{01}}} \right\rangle } \right]\\ &+ i{\Omega _{pu}}\left( {{\Pi _1}\left\langle {{\sigma_{10}}} \right\rangle {e^{ - i\delta t}} - \Pi _1^\ast \left\langle {{\sigma_{01}}} \right\rangle {e^{i\delta t}}} \right) - 2{\mathop{\rm Im}\nolimits} G\left\langle {{\sigma_{01}}} \right\rangle \left\langle {{\sigma_{10}}} \right\rangle \end{aligned}, $$
$$\frac{{d\left\langle {{\sigma_{01}}} \right\rangle }}{{dt}} ={-} ({\Gamma _2} + i{\Delta _{pu}})\left\langle {{\sigma_{01}}} \right\rangle - 2i{\Pi _1}{\Omega _{pu}}\left\langle {{\sigma_z}} \right\rangle - 2i{\Pi _1}{\Omega _{pr}}{e^{ - i\delta t}}\left\langle {{\sigma_z}} \right\rangle - 2iG\left\langle {{\sigma_z}} \right\rangle \left\langle {{\sigma_{01}}} \right\rangle, $$
$$|{FWM} |= \left|{\frac{{\sigma_{01}^{( - )}}}{{E_{pr}^\ast {\hbar^{ - 1}}\Gamma _2^{ - 1}}}} \right|= \frac{{2{t_3}({{t_4}\sigma_z^{(0)} - {t_1}\sigma_{01}^{(0)}} )}}{{\mu {\hbar ^{ - 1}}\Gamma _2^{ - 1}({{t_1}{t_5}{t_7} - {t_2}{t_4}{t_7} + {t_1}{t_6}} )}}. $$
$${\Gamma _1}({{w_0} + 1} )\{{{{[{{\Gamma _2} - ({{\mathop{\rm Im}\nolimits} G} ){w_0}} ]}^2} + {{[{{\Delta _{pu}} + ({Re G} ){w_0}} ]}^2}} \}+ 4{|{{\Pi _1}} |^2}\Omega _{pu}^2{\Gamma _2}{w_0} = 0 .$$
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.