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

Tunable plasmonically induced transparency with giant group delay in gain-assisted graphene metamaterials

Open Access Open Access

Abstract

We propose a graphene metamaterial consisting of several layers of longitudinally separated graphene nanoribbon array embedded into gain-assisted medium, demonstrating electromagnetically induced transparency-like spectra. Combined with finite-difference time-domain simulations, the transfer matrix method and temporal coupled-mode theory are adopted to quantitatively describe its transmission characteristics. These transmission characteristics can be tuned by altering the gain level in medium layer and the Fermi energy level in graphene. Additionally, it is the incorporation between gain medium and graphene nanoribbons with optimized geometrical parameters and Fermi energy level that the destructive interference between high order graphene plasmonic modes can be obtained, suggesting drastic phase transition with giant group delay and ultra-high group index up to 180 ps and 104, respectively. Our results can achieve efficient slow light effects for better optical buffers and other nonlinear applications.

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

1. Introduction

Electromagnetically induced transparency (EIT) in atomic medium is a fascinating quantum interference effect, giving rise to an ultra-narrow and steep transparency window within a broad frequency band [1,2]. This intriguing phenomenon promises a serious of applications in slow light [312], optical nonlinearity [13,14], and switches [15,16]. Unfortunately, the EIT effect in application is limited by harsh conditions which are usually desired to maintain quantum states coherence, such as low temperature, strong gas lasers and macroscopic devices [1,17]. These requirements are unsuitable for the implementation of chip-scale. To overcome this issue, the analogues of EIT realized in classical optical systems have attracted wide attention, such as photonic crystals [18,19], metamaterials [20,21], and waveguide-cavity configurations [22]. All-optical analogue of EIT photonic crystals (PhCs) with ultra-high quality factor demonstrates group time delay up to 20 ps in experiment due to its low intrinsic loss in the silicon-based cavity [18]. Although the electromagnetic modes can be confined within the wavelength scale, PhCs are still far distant to the goal for controlling and manipulating light in nanoscale.

Surface plasmon polaritons (SPPs), collective electronic excitation, are strongly confined in the metal-dielectric interface. These evanescent waves are usually adopted in subwavelength optical devices due to their wavelengths being far below classical diffraction limit [23,24]. Plasmonically induced transparency (PIT) can be realized in metallic structures via destructive interference between bright and dark plasmonic modes or phase-coupling between two space-separated resonators [25]. Although plasmonic nanostructures possess the unique characteristics of giant near-field enhancement and ultra-compact configuration, the moderate slow light is obtained due to the intrinsic losses of the metal [26]. Usually, the group refractive index is less than 102 [17]. Consequently, the realizations of enhanced nonlinearities, enlarged spectral sensitivity, and increased phase shifts can be achieved by reducing the bandwidth to a small value [2731]. Recently, it is a feasible way to compensate for the intrinsic Ohmic loss and pave the way to metallic metamaterials by resorting to the gain medium [3234]. For example, Ambati et al. demonstrated a experiment of stimulated emission of SPPs with erbium doped glass as a gain-assisted medium in 2008 [35]. After that, Dong et al. demonstrated a loss-compensated resonance enhancement by using the gain-assisted PIT in a metallic metamaterial, and the Q-factor is up to 666.67 [36]. On the other hand, a two-dimensional material, graphene, demonstrates many excellent properties such as strong confinement, low resistive damping, and optical transparency, which gain wide attention from researchers in devise fields [3740]. Up to now, a large number of graphene plasmonic metamaterials have been demonstrated to achieve the PIT effect [41]. Some nanostructured graphene are used to achieve optical devices with high slow light, as the intrinsic loss of graphene is much smaller than that of metallic structures in the mid-infrared regimes. For example, Gao et al. proposed a terahertz graphene metasurface with a group index as high as 935 in the four graphene ribbons [42]. In 2021, Cui et al. realized dual-PIT effect with a higher group index of ∼2553 in a hexagonal graphene coupled metasurface [5]. Then, can we combine gain material with graphene to obtain higher slow light?

In this paper, we demonstrate a graphene metamaterial possessing a sharp EIT-like window with a Q-factor of ∼104. The high-Q transparency window is accomplished by employing the high-order mode to minimize the radiative losses, and gain medium to partially compensate the intrinsic Ohmic losses. In addition, the tunability of EIT-like window can be investigated by varying the gain level in dielectric layer and the Fermi energy level of GNR. The transfer matrix method and temporal coupled-mode theory are adopted to quantitatively describe the observed EIT-like responses, and these responses are in good with the finite-difference time-domain simulations.

2. Structure and method

The designed graphene metamaterials schematically show in Fig. 1, which consist of several layers of vertically separated graphene nanoribbon (GNR) array embedded into gain-assisted dielectric. The surface conductivity of graphene is described by σ = (e2ħ2) iEFj/ (ω +  −1), where τ = μEFj / F2 is the intrinsic inelastic lifetime (μ = 2.5 × 104 cm2 V−1s−1, and νF = 106 ms−1, see Ref. [29,43]). In the FDTD simulation, the single-layer graphene is modeled as an effective dielectric layer with a thickness (H) of 1 nm with a dielectric constant εr = 1 + /(ωε0H). The dielectric constant of single-layer graphene can also be dynamically adjusted by changing its Fermi energy EFj through electrostatic doping [44]. Here, according to a direct experimental evidence of stimulated emission of surface plasmon polaritons (SPPs) with gain medium [35], the Er3+ ions are doped in the dielectric spacer as a gain-assisted medium when the gain is needed, and the complex permittivity of this medium is ε = ε` + ``. The real part of the permittivity is taken to be ε` = 2 and the gain is introduced through the imaginary part ε``. Since the gain as a function of wavelength can be approximately depicted as a Gaussian profile, the maximum gain coefficient α0 is located at 2800 nm wavelength and the gain bandwidth is γ0 = 0.03ω0 [45].

 figure: Fig. 1.

Fig. 1. (a) The schematic illustration of the graphene metamaterials consisting of several layers of longitudinally separated graphene nanoribbon (GNR) array embedded into gain-assisted dielectric. Wj and hj represent the width of GNRj (j = 1,2,…) and the separation between GNRj and GNRj+1, respectively. P is the period. (b) Cross section of the proposed one unit cell with j-layer GNRs in the metamaterials. The dashed lines are reference planes of GNRs.

Download Full Size | PDF

As shown schematically in Fig. 1(a), when a linearly polarized light propagates along the + y-axis direction, the plasmon resonances at a specific wavelength will be excited in GNRs. For a unit cell of graphene metamaterials in Fig. 1(b), the E(j)d,in and E(j)d,out (j = 1, 2, …) are represent the amplitudes of the incident and outgoing lights of the GNRs. At GRAj (“j” represents the jth nanoribbon), the subscripts d = “+” or “−” indicate that the wave propagates in the orientations of the + y-axis and the −y-axis and the incident and outgoing waves are denoted by the subscripts “in” and “out”, respectively. When the incident light with a frequency ω0j is inputted only from the upper (E(j)-,in = 0), the temporal normalized mode amplitude of j-th graphene nanoribbon aj can be written as

$$\frac{{d{a_j}}}{{dt}} = ( - i{\omega _{0j}} - {\kappa _{o,j}} - {\kappa _{e,j}}){a_j} + {e^{i{\theta _j}}}\sqrt {{\kappa _{e,j}}} E_{ + ,\textrm{in}}^{(j )} + {e^{i{\theta _j}}}\sqrt {{\kappa _{e,j}}} E_{\textrm{ - },\textrm{in}}^{(j )},$$
where ω0j stands for the resonant frequency of corresponding GNRj, κo,j represents the decay rate due to the internal Ohmic loss and κe1,2 represents the decay rate due to the escaping energy to the dielectric spacer. θj is the phase of the coupling coefficient. To gain a relatively simple theoretical model, the coupling loss is not considered for a single GNR and adjacent dielectric. The incoming waves into jth GNR should satisfy the relationships as follows
$$E_{\textrm{ - },\textrm{in}}^{(j )} = E_{\textrm{ - },\textrm{out}}^{(j )} + {e^{ - i{\theta _j}}}\sqrt {{\kappa _{e,j}}} {a_j},$$
$$E_{ + ,\textrm{in}}^{(j )} = E_{ + ,\textrm{out}}^{(j )} + {e^{ - i{\theta _j}}}\sqrt {{\kappa _{e,j}}} {a_j}.$$

Combined with the above Eqs. (1), (2), and (3), the complex transmission and reflection coefficients tj and rj of the single-layer GNRs structure can be described as

$${t_j}(\omega )= \frac{{i({\omega _{0j}} - \omega ) + {\kappa _{o,j}}}}{{i({\omega _{0j}} - \omega ) + {\kappa _{o,j}} + {\kappa _{e,j}}}},\quad {r_j}(\omega )= \frac{{{\kappa _{e,j}}}}{{i({\omega _{0j}} - \omega ) + {\kappa _{o,j}} + {\kappa _{e,j}}}}.$$

As shown in Fig. 1(b), the transport characteristics of the proposed F-P microcavity with two spatially separated GNRs (i.e., GRA1 and GRA2) can be described by the transfer matrix model. The transfer equation can be expressed as

$$\left( {\begin{array}{c} {E_{\textrm{ - },\textrm{in}}^{(j )}}\\ {E_{ + ,\textrm{out}}^{(j )}} \end{array}} \right) = {M_j}{S_{j - 1}}{M_{j - 1}}{S_{j - 2}}\ldots {M_2}{S_1}{M_1}\left( {\begin{array}{c} {E_{ + ,\textrm{in}}^{(1 )}}\\ {E_{\textrm{ - },\textrm{out}}^{(1 )}} \end{array}} \right).$$

Matrice Sj represents the propagate wave in the hj-thickness dielectric interlayer and Mj represents in the graphene interface. These matrices are given by

$${S_j} = \left( {\begin{array}{cc} 0&{{e^{i\beta {h_j}}}}\\ {{e^{\textrm{ - }i\beta {h_j}}}}&0 \end{array}} \right), \quad {M_j} = \left( {\begin{array}{cc} {{{ - {r_j}} / {{t_j}}}}&{{1 / {{t_j}}}}\\ {1 + {{{r_j}} / {{t_j}}}}&{{{{r_j}} / {{t_j}}}} \end{array}} \right).$$

Here, β(ω) stands for the propagation constant. When the gain medium is introduced, the imaginary part of β(ω) will be a negative value (i.e., β = β` + iβ`` and β``< 0). Considering the dual-layer GNRs system with the dielectric interlayer, the transmission can be obtained by

$${T_{\textrm{dual}}} = \left|{\frac{{E_{ + ,\textrm{out}}^{(2 )}}}{{E_{ + ,\textrm{in}}^{(1 )}}}} \right|= {\left|{\frac{{{e^{i2\beta {h_1}}}{t_1}{t_2}}}{{1 - {r_1}{r_2}{e^{i2\beta {h_1}}}}}} \right|^2}.$$

Then, Eq. (4) is substituted into Eq. (7), we have,

$${T_{\textrm{dual}}} = {\left|{\frac{{({i({{\omega_1} - \omega } )+ {\kappa_{o1}}} )({i({{\omega_2} - \omega } )+ {\kappa_{o2}}} )\textrm {exp} ({2i\beta^{\prime}{h_1} - 2\beta^{\prime\prime}{h_1}} )}}{{({i({{\omega_1} - \omega } )+ {\kappa_{o1}} + {\kappa_{e1}}} )({i({{\omega_2} - \omega } )+ {\kappa_{o2}} + {\kappa_{e2}}} )- {\kappa_{e1}}{\kappa_{e2}}\textrm {exp} ({2i\beta^{\prime}{h_1} - 2\beta^{\prime\prime}{h_1}} )}}} \right|^2}.$$

In addition, the numerical transmission spectra can be obtained by using the two dimensional (2D) finite-difference time-domain (FDTD) method. And perfectly matched layers (PML) boundaries are used in the y-axis directions, as well as periodic boundary conditions (P = 283.6 nm) are set at x-axis directions. The thickness of single-layer GNR is taken to 1 nm, and the maximum mesh size is set as Δx = 0.4 nm and Δy = 0.2 nm, which ensure the convergence of numerical results.

3. Results and discussions

Figure 2(a) demonstrates the transmission spectra of the dual-layer GNRs system at δEF = 0.3 meV (i.e., EF2 - EF1 = 0.3 meV, EF1 and EF2 represent the Fermi energy levels of two layer ribbons) for different gain coefficients α0. In the numerical calculations, at the expance of efficient pump power, the imaginary part of permittivity ε`` will be a negative value. When the pump power is increased, the value of -ε`` will rise and the medium will eventually show the significant gain effect. Here, the gain medium is embedded into the graphene nanoribbon array, for which the active characteristic can be described by the gain coefficient α0 = −(2π/λ)Im((ε` + iε``)1/2) [36,4648]. It is clearly shows that compare with the lower value of α0, the PIT window is more prominent, and the bandwidth is narrower with the higher value of α0 because increasing the value of α0 decreases the intrinsic Ohmic loss. Noteworthily, a gain level within 11.54-13.34 cm−1 can realize a sufficiently large transmittance in the transparency window. In Fig. 2(b), the values of radiative and intrinsic damping rates related to the graphene material are obtained by theoretical fittings. The radiative damping rates κe1 and κe2 do not vary significantly with increasing α0, whereas the damping rate of the intrinsic loss κo1 and κo2 decrease markedly from 8.98 cm−1. In addition, due to the compensation of the gain medium for the intrinsic losses, the damping rate is gradually dominated by the radiative damping. On the other hand, with the increase of α0, the value of intrinsic loss κo1 and κo2 are getting closer and the value of intrinsic loss κo1 is always smaller than κo2 because the Fermi energy level of GNR1 is smaller than GNR2.

 figure: Fig. 2.

Fig. 2. (a) Transmission spectra of the dual-layer GNRs system based on theoretical fittings (solid lines) and numerical calculations (scattered open dots) for different gain coefficients α0. The other structural parameters are h1 = 1.98 µm, W1 = W2 = 141.8 nm, and the Fermi energy levels (EF1 and EF2) of the dual-layer GNRs are set as 799.85 and 800.15 meV. (b) Damping rates κo1(2) and κe1(2) are the parameters used in the transmission fitting of the graphene metamaterial under different gain coefficients α0.

Download Full Size | PDF

As shown in Fig. 3(a), under the gain level α0 = 12.83 cm−1, the tunability of EIT-like response can be investigated by adjusting the δEF, which can be realized by alkaline surface doping in experiment. A trade-off between peak transmission and the full width at half maximum (FWHM) at an EIT-like window is apparent for the dual-layer GNRs system. The transmission spectrum has a small split due to the absence of δEF in an almost lossless environment. As δEF gradually increases from 0.2 meV to 0.5 meV, the transmission undergoes a large modulation, with an increase in the transmission peak. Ultimately, an EIT-like window with a transmission amplitude of 77% is obtained at 2800 nm when δEF reaches 0.5 meV. Significantly reducing the bandwidth of the transparency window is valuable for slow light applications. Because sufficient gain level is chosen in the system, the bandwidth of EIT-like window remains narrow and the Q-factor originally 5.9×103 at δEF = 0.5 meV, reaches the maximum value of 2×104. To demonstrate the physical picture of the dual-layer GNRs system, we plot the x-component of the electric field distribution (i.e., Ex) with δEF = 0.3 meV in Figs. 3(b)–3(d). When the incoming lights pass through the system, as depicted in Figs. 3(b) and 3(d), strong higher-order resonances indicate the effective excitation of surface plasmon in single GNR at the transmission dips λ1 and λ2. As illustrated in Fig. 3(c), both two-layer GNRs are excited at the central wavelength λ12 of the transmission window, leading to Fabry-Perot (F-P) resonance between the two nanoribbons. According to the F-P model, the transparency peak with the wavelength detuning δλ = |λ2 - λ1| appears at the resonance wavelength λ12 = (λ1+λ2)/2 when the phase difference β`h1 equals nπ (n = 1, 2, 3…..). In our investigation, the wavelengths of transmission dips are λ1 = 2799.71 nm and λ2 = 2800.12 nm, and the resonance wavelength λ12 = 2799.92 nm.

 figure: Fig. 3.

Fig. 3. (a) Transmission spectra of the dual-layer GNRs system by adjusting the difference of Fermi energy level (i.e., δEF). δEF = 0 meV, δEF = 0.2 meV, δEF = 0.3 meV, δEF = 0.4 meV, and δEF = 0.5 meV. The red solid curves are the theoretical fitting results and the blue spheres are the simulation results calculated by FDTD method. The Fermi energy levels (EF1 and EF2) of two-layer GNRs are set as 800 and 800 meV, 799.9 and 800.1 meV, 799.85 and 800.15 meV, 799.8 and 800.2 meV, 799.75 and 800.25 meV. And the other parameters are h1 = 1.98 µm and W1 = W2 = 141.8 nm. (b)-(d) Numerical field distributions (Ex) with the incident wavelengths at λ1 = 2799.71 nm, λ12 = 2799.92 nm, and λ2 = 2800.12 nm. The inset (black dash squares) shows the field distribution around the 2nd-layer GNR.

Download Full Size | PDF

It is well known that the EIT-like effect also supports slow light in plasmonic system. The slow-light effect can be described by group index ng, which is expressed as

$${n_g} = \frac{c}{{{v_g}}} = \frac{c}{L}{\tau _g} = \frac{c}{L}\frac{{\textrm{d}\varphi (\omega )}}{{\textrm{d}\omega }}.$$

Here, vg is the group velocity in the system. τg and φ(ω) stand for the group delay time and transmission phase shift. L is the length of the system. Figure 4 shows group indices are extracted from transmission spectra. As shown in Fig. 4(a), the slow light in the transparency peak is numerically investigated with EF1 = 800.1 meV, EF2 = 799.9 meV, W1 = W2 = 141.8 nm, and h1 = 1.98 µm. It indicates that the decreasing phase is steepest at the position of the transparency peak. In Eq. (9), the value ng is determined by the phase slope. This system exhibits a maximum group index over 104 around the transparency-peak wavelength, as shown in Fig. 4(a). As δEF varies, it can be predicted from Fig. 4(b) that the maximum group index will occur near δEF = 0.15 meV. And the maximum value of ng is about 3.4×104, which corresponding optical delay of the transparency peak resonance reaches ∼230 ps.

 figure: Fig. 4.

Fig. 4. (a) Transmission phase shifts (black dashed line) and group indices (blue solid line) in the system with h1 = 1.98 µm. EF1 = 799.9 meV, EF2 = 800.1 meV and W1 = W2 = 141.8 nm. (b) Group indices are extracted from transmission spectra for different δEF.

Download Full Size | PDF

As presented in Fig. 5(a), a periodic relation between the optical delay of transmission peak and vertical separation is theoretically calculated. The maximum optical delay is about 180 ps can be obtained at vertical separations equal to 0.99 µm and 1.98 µm, which is agree well with the consequence exhibited by the spectrum. In addition, the evolution of the transmission spectrum with separation h is as detailed in Fig. 5(b). The result shows that the transparency peaks obviously display and redshift around h1 = 0.99 µm and 1.98 µm. The periodic evolution of transmission characteristics can be explained by the F-P model. Finally, to verify the F-P model can be extended to multi-layer GNRs system, Fig. 6 reveals the transmission properties in triple-layer GNRs system with different vertical separations hj (j = 1, 2). When j = 3, the transmission efficiency can be obtained by

$${T_{\textrm{triple}}} = {\left|{\frac{{\exp({i2\beta ({{h_1} + {h_2}} )} ){t_1}{t_2}{t_3}}}{{1 - {r_1}{r_2}\exp ({i2\beta {h_1}} )- {r_2}{r_3}\exp ({i2\beta {h_2}} )- {r_1}{r_3}({{r_2} + {t_2}} )\exp({i2\beta ({{h_1} + {h_2}} )} )}}} \right|^2}.$$

 figure: Fig. 5.

Fig. 5. (a) Variation of optical delay of transmission peak with the separation h changed. (b) Evolution of transmission spectra as a function of separation h, and the other parameters are EF1 = 799.9 meV, EF2 = 800.1 meV, and W1 = W2 = 141.8 nm.

Download Full Size | PDF

 figure: Fig. 6.

Fig. 6. (a) Simulation (blue spheres) and theoretical (red solid curve) transmission spectra in the triple-layer-GNRs system with W1 = W2 = W3 = 141.8 nm, h1 = 1.98 µm, h2 = 1.97 µm, and δEF = 0.3 meV. (b) Optical delay (blue line) in this system. (c) and (d) Numerical field distributions (Ex) at the two transparency-resonance wavelengths: λ11 = 2799.5 nm and λ22 = 2800.11 nm.

Download Full Size | PDF

As shown in Fig. 6(a), the theoretical fittings based on Eq. (10) show good agreement with the numerical simulations. In addition, as shown in Fig. 6(b), the maximal optical delay reach to 96 ps and group index up to 1.44×104. In this system, the separation between the 2nd and 3rd-layer GNR can be set as 1979.96 nm and the width of 3rd-layer GNR is W3 = 141.8 nm. The Fig. 6(a) shows that there are two transparency windows at λ11 = 2799.5 nm and λ22 = 2800.11 nm, and the line and the dot are fit well, namely, the theoretical model is in a good agreement with the simulation results. As shown in Figs. 6(c) and 6(d), similar to a dual-layer GNRs system, the electric field distributions show that two F-P cavities at the wavelength of EIT-like peaks. Thus, it can be conjectured that the multi-layer GNRs system can also be implemented to achieve the graphene-based multiple EIT-like responses.

4. Conclusions

In summary, ultra-narrowband EIT-like effect in a gain-assisted graphene plasmonic metamaterial has been theoretically and numerically investigated. The EIT window can be tuned by varying the gain level in medium layer and altering the Fermi energy level in graphene. Moreover, based on the gain medium and high-order graphene plasmonic resonances, radiative and non-radiative loss can be reduced by the proposed system, and the metamaterials can be used to achieve drastic phase transition and ultra-high group index up to 104. It should be possible, therefore, to achieve efficient slow light effects for better optical buffers and other nonlinear applications.

Funding

National Natural Science Foundation of China (11947062); Natural Science Foundation of Hunan Province (2020JJ5551, 2021JJ40523).

Disclosures

The authors declare no conflicts of interest.

Data Availability

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

References

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

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

3. L. V. Hau, S. E. Harris, Z. Dutton, and C. H. Behroozi, “Light speed reduction to 17 metres per second in an ultracold atomic gas,” Nature 397(6720), 594–598 (1999). [CrossRef]  

4. K. Totsuka, N. Kobayashi, and M. Tomita, “Slow light in coupled-resonator-induced transparency,” Phys. Rev. Lett. 98(21), 213904 (2007). [CrossRef]  

5. W. Cui, Y. Wang, Z. He, and H. He, “Strong slow-light effect for a hexagonal graphene coupled metasurface in terahertz,” Results Phys. 26, 104356 (2021). [CrossRef]  

6. C. Sui, B. Han, T. Lang, X. Li, X. Jing, and Z. Hong, “Electromagnetically induced transparency in an all-dielectric metamaterial-waveguide with large group index,” IEEE Photonics J. 9(5), 1–8 (2017). [CrossRef]  

7. C. Marco, G. Matteo, O. Liam, and G. Dario, “Electromagnetically induced transparency from first-order dynamical systems,” Phys. Rev. B 104(20), 205434 (2021). [CrossRef]  

8. N. Ahmer, “Reversible fast to slow-light transition originating in the optical analog of EIA-EIT transformation in optical resonators,” OSA Continuum 4(11), 2771–2783 (2021). [CrossRef]  

9. F. Gao, P. Yuan, S. Gao, J. Deng, Z. Sun, G. Jin, G. Zeng, and B. Yan, “Active Electromagnetically Induced Transparency Effect in Graphene-Dielectric Hybrid Metamaterial and Its High-Performance Sensor Application,” Nanomaterials 11(8), 2032 (2021). [CrossRef]  

10. J. Zhang, Z. Li, L. Shao, F. Xiao, and W. Zhu, “Active modulation of electromagnetically induced transparency analog in graphene-based microwave metamaterial,” Carbon 183, 850–857 (2021). [CrossRef]  

11. J. Zhu, Q. Zhang, and G. Huang, “Quantum squeezing of slow-light solitons,” Phys. Rev. A 103(6), 063512 (2021). [CrossRef]  

12. D. H. Minh, N. L. T. Yen, K. D. Xuan, and N. H. Bang, “Controllable ultraslow optical solitons in a degenerated two-level atomic medium under EIT assisted by a magnetic field,” Sci. Rep. 10(1), 1–9 (2020). [CrossRef]  

13. Y. Niu and S. Gong, “Enhancing kerr nonlinearity via spontaneously generated coherence,” Phys. Rev. A 73(5), 053811 (2006). [CrossRef]  

14. H. R. Hamedi and M. R. Mehmannavaz, “Switching feature of EIT-based slow light giant phase-sensitive Kerr nonlinearity in a semiconductor quantum well,” Phys. E 66, 309–316 (2015). [CrossRef]  

15. F. K. Le and A. Rauschenbeutel, “Nanofiber-based all-optical switches,” Phys. Rev. A 93(1), 013849 (2016). [CrossRef]  

16. J. Chen, P. Wang, C. Chen, Y. Lu, H. Ming, and Q. Zhan, “Plasmonic EIT-like switching in bright-dark-bright plasmon resonators,” Opt. Express 19(7), 5970–5978 (2011). [CrossRef]  

17. S. Zhang, D. A. Genov, Y. Wang, M. Liu, and X. Zhang, “Plasmon-induced transparency in metamaterials,” Phys. Rev. Lett. 101(4), 047401 (2008). [CrossRef]  

18. X. Yang, M. Yu, D. L. Kwong, and C. W. Wong, “All-optical analog to electromagnetically induced transparency in multiple coupled photonic crystal cavities,” Phys. Rev. Lett. 102(17), 173902 (2009). [CrossRef]  

19. Z. Chai, X. Hu, C. Li, H. Yang, and Q. Gong, “On-chip multiple electromagnetically induced transparencies in photon-plasmon composite nanocavities,” ACS Photonics 3(11), 2068–2073 (2016). [CrossRef]  

20. S. Y. Chiam, R. Singh, C. Rockstuhl, F. Lederer, W. Zhang, and A. A. Bettiol, “Analogue of electromagnetically induced transparency in terahertz metamaterial,” Phys. Rev. B 80(15), 153103 (2009). [CrossRef]  

21. R. Singh, C. Rockstuhl, F. Lederer, and W. Zhang, “Coupling between a dark and a bright eigenmode in a terahertz metamaterial,” Phys. Rev. B 79(8), 085111 (2009). [CrossRef]  

22. H. Lu, X. Liu, and D. Mao, “Plasmonic analog of electromagnetically induced transparency in multi-nanoresonator-coupled waveguide systems,” Phys. Rev. A 85(5), 053803 (2012). [CrossRef]  

23. J. M. Pitarke, V. M. Silkin, E. V. Chulkov, and P. M. Echenique, “Theory of surface plasmons and surface-plasmon polaritons,” Rep. Prog. Phys. 70(1), 1–87 (2007). [CrossRef]  

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

25. R. D. Kekatpure, E. S. Barnard, W. Cai, and M. L. Brongersma, “Phase-coupled plasmon induced transparency,” Phys. Rev. Lett. 104(24), 243902 (2010). [CrossRef]  

26. X. Y. Yang, X. Y. Hu, Z. Chai, C. C. Lu, H. Yang, and Q. H. Gong, “Tunable ultracompact chip-integrated multichannel filter based on plasmon-induced transparencies,” Appl. Phys. Lett. 104(22), 221114 (2014). [CrossRef]  

27. T. Baba, “Slow light in photonic crystals,” Nat. Photonics 2(8), 465–473 (2008). [CrossRef]  

28. J. Sancho, J. Bourderionnet, J. Lloret, S. Combrié, I. Gasulla, S. Xavier, S. Sales, P. Colman, G. Lehoucq, D. Dolfi, J. Capmany, and A. De Rossi, “Integrable microwave filter based on a photonic crystal delay line,” Nat. Commun. 3(1), 1075 (2012). [CrossRef]  

29. L. Meng, D. Zhao, Y. Yang, F. J. G. De Abajo, Q. Li, Z. Ruan, and M. Qiu, “Gain-assisted plasmon resonance narrowing and its application in sensing,” Phys. Rev. Appl. 11(4), 044030 (2019). [CrossRef]  

30. Z. Y. Li, S. Butun, and K. Aydin, “Ultranarrow band absorbers based on surface lattice resonances in nanostructured metal surfaces,” ACS Nano 8(8), 8242–8248 (2014). [CrossRef]  

31. Y. L. Liao and Y. Zhao, “Ultra-narrowband dielectric metamaterial absorber with ultra-sparse nanowire grids for sensing applications,” Sci. Rep. 10(1), 1480 (2020). [CrossRef]  

32. Y. Yang, I. I. Kravchenko, D. P. Briggs, and J. Valentine, “All-dielectric metasurface analogue of electromagnetically induced transparency,” Nat. Commun. 5(1), 5753 (2014). [CrossRef]  

33. W. Tang, L. Wang, X. Chen, C. Liu, A. Yu, and W. Lu, “Dynamic metamaterial based on the graphene split ring high-Q Fano-resonnator for sensing applications,” Nanoscale 8(33), 15196–15204 (2016). [CrossRef]  

34. T. T. Kim, H. D. Kim, R. Zhao, S. S. Oh, T. Ha, D. S. Chung, Y. H. Lee, B. Min, and S. Zhang, “Electrically tunable slow light using graphene metamaterials,” ACS Photonics 5(5), 1800–1807 (2018). [CrossRef]  

35. M. Ambati, S. H. Nam, E. Ulin-Avila, D. A. Genov, G. Bartal, and X. Zhang, “Observation of stimulated emission of surface plasmon polaritons,” Nano Lett. 8(11), 3998–4001 (2008). [CrossRef]  

36. Z. G. Dong, H. Liu, J. X. Cao, T. Li, S. M. Wang, S. N. Zhu, and X. Zhang, “Enhanced sensing performance by the plasmonic analog of electromagnetically induced transparency in active metamaterials,” Appl. Phys. Lett. 97(11), 114101 (2010). [CrossRef]  

37. Z. Fei, A. S. Rodin, G. O. Andreev, W. Bao, A. S. McLeod, M. Wagner, L. M. Zhang, Z. Zhao, M. Thiemens, G. Dominguez, M. M. Fogler, A. H. Castro Neto, C. N. Lau, F. Keilmann, and D. N. Basov, “Gate-tuning of graphene plasmons revealed by infrared nano-imaging,” Nature 487(7405), 82–85 (2012). [CrossRef]  

38. B. Zhang, H. Li, H. Xu, M. Zhao, C. Xiong, C. Liu, and K. Wu, “Absorption and slow-light analysis based on tunable plasmon-induced transparency in patterned graphene metamaterial,” Opt. Express 27(3), 3598–3608 (2019). [CrossRef]  

39. E. Gao, Z. Liu, H. Li, H. Xu, Z. Zhang, X. Luo, C. Xiong, C. Liu, B. Zhang, and F. Zhou, “Dynamically tunable dual plasmon-induced transparency and absorption based on a single-layer patterned graphene metamaterial,” Opt. Express 27(10), 13884–13894 (2019). [CrossRef]  

40. E. Gao, Z. Liu, H. Li, H. Xu, Z. Zhang, X. Zhang, X. Luo, and F. Zhou, “Dual plasmonically induced transparency and ultra-slow light effect in m-shaped graphene-based terahertz metasurfaces,” Appl. Phys. Express 12(12), 126001 (2019). [CrossRef]  

41. S. Y. Xiao, T. Wang, T. Liu, X. Yan, Z. Li, and C. Xu, “Active modulation of electromagnetically induced transparency analogue in terahertz hybrid metal-graphene metamaterials,” Carbon 126, 271–278 (2018). [CrossRef]  

42. E. Gao, H. Li, Z. Liu, C. Xiong, C. Liu, B. Ruan, M. Li, and B. Zhang, “Terahertz multifunction switch and optical storage based on triple plasmon-induced transparency on a single-layer patterned graphene metasurface,” Opt. Express 28(26), 40013–40023 (2020). [CrossRef]  

43. H. Yan, X. Li, B. Chandra, G. Tulevski, Y. Wu, M. Freitag, W. Zhu, P. Avouris, and F. Xia, “Tunable infrared plasmonic devices using graphene/insulator stacks,” Nat. Nanotechnol. 7(5), 330–334 (2012). [CrossRef]  

44. F. Javier and G. Abajo, “Graphene plasmonics: challenges and opportunities,” ACS Photonics 1(3), 135–152 (2014). [CrossRef]  

45. W. Jin, C. Khandekar, A. Pick, A. G. Polimeridis, and A. W. Rodriguez, “Amplified and directional spontaneous emission from arbitrary composite bodies: A self-consistent treatment of Purcell effect below threshold,” Phys. Rev. B: Condens. Matter Mater. Phys. 93(12), 125415 (2016). [CrossRef]  

46. M. P. Nezhad, K. Tetz, and Y. Fainman, “Gain assisted propagation of surface plasmon polaritons on planar metallic waveguides,” Opt. Express 12(17), 4072–4079 (2004). [CrossRef]  

47. Z. G. Dong, H. Liu, T. Li, Z. H. Zhu, S. M. Wang, J. X. Cao, S. N. Zhu, and X. Zhang, “Optical loss compensation in a bulk left-handed metamaterial by the gain in quantum dots,” Appl. Phys. Lett. 96(4), 044104 (2010). [CrossRef]  

48. T. Liu, C. Zhou, and S. Xiao, “Gain-assisted critical coupling for enhanced optical absorption in graphene,” Nanotechnology 32(20), 205202 (2021). [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. (a) The schematic illustration of the graphene metamaterials consisting of several layers of longitudinally separated graphene nanoribbon (GNR) array embedded into gain-assisted dielectric. Wj and hj represent the width of GNRj (j = 1,2,…) and the separation between GNRj and GNRj+1, respectively. P is the period. (b) Cross section of the proposed one unit cell with j-layer GNRs in the metamaterials. The dashed lines are reference planes of GNRs.
Fig. 2.
Fig. 2. (a) Transmission spectra of the dual-layer GNRs system based on theoretical fittings (solid lines) and numerical calculations (scattered open dots) for different gain coefficients α0. The other structural parameters are h1 = 1.98 µm, W1 = W2 = 141.8 nm, and the Fermi energy levels (EF1 and EF2) of the dual-layer GNRs are set as 799.85 and 800.15 meV. (b) Damping rates κo1(2) and κe1(2) are the parameters used in the transmission fitting of the graphene metamaterial under different gain coefficients α0.
Fig. 3.
Fig. 3. (a) Transmission spectra of the dual-layer GNRs system by adjusting the difference of Fermi energy level (i.e., δEF). δEF = 0 meV, δEF = 0.2 meV, δEF = 0.3 meV, δEF = 0.4 meV, and δEF = 0.5 meV. The red solid curves are the theoretical fitting results and the blue spheres are the simulation results calculated by FDTD method. The Fermi energy levels (EF1 and EF2) of two-layer GNRs are set as 800 and 800 meV, 799.9 and 800.1 meV, 799.85 and 800.15 meV, 799.8 and 800.2 meV, 799.75 and 800.25 meV. And the other parameters are h1 = 1.98 µm and W1 = W2 = 141.8 nm. (b)-(d) Numerical field distributions (Ex) with the incident wavelengths at λ1 = 2799.71 nm, λ12 = 2799.92 nm, and λ2 = 2800.12 nm. The inset (black dash squares) shows the field distribution around the 2nd-layer GNR.
Fig. 4.
Fig. 4. (a) Transmission phase shifts (black dashed line) and group indices (blue solid line) in the system with h1 = 1.98 µm. EF1 = 799.9 meV, EF2 = 800.1 meV and W1 = W2 = 141.8 nm. (b) Group indices are extracted from transmission spectra for different δEF.
Fig. 5.
Fig. 5. (a) Variation of optical delay of transmission peak with the separation h changed. (b) Evolution of transmission spectra as a function of separation h, and the other parameters are EF1 = 799.9 meV, EF2 = 800.1 meV, and W1 = W2 = 141.8 nm.
Fig. 6.
Fig. 6. (a) Simulation (blue spheres) and theoretical (red solid curve) transmission spectra in the triple-layer-GNRs system with W1 = W2 = W3 = 141.8 nm, h1 = 1.98 µm, h2 = 1.97 µm, and δEF = 0.3 meV. (b) Optical delay (blue line) in this system. (c) and (d) Numerical field distributions (Ex) at the two transparency-resonance wavelengths: λ11 = 2799.5 nm and λ22 = 2800.11 nm.

Equations (10)

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

d a j d t = ( i ω 0 j κ o , j κ e , j ) a j + e i θ j κ e , j E + , in ( j ) + e i θ j κ e , j E  -  , in ( j ) ,
E  -  , in ( j ) = E  -  , out ( j ) + e i θ j κ e , j a j ,
E + , in ( j ) = E + , out ( j ) + e i θ j κ e , j a j .
t j ( ω ) = i ( ω 0 j ω ) + κ o , j i ( ω 0 j ω ) + κ o , j + κ e , j , r j ( ω ) = κ e , j i ( ω 0 j ω ) + κ o , j + κ e , j .
( E  -  , in ( j ) E + , out ( j ) ) = M j S j 1 M j 1 S j 2 M 2 S 1 M 1 ( E + , in ( 1 ) E  -  , out ( 1 ) ) .
S j = ( 0 e i β h j e  -  i β h j 0 ) , M j = ( r j / t j 1 / t j 1 + r j / t j r j / t j ) .
T dual = | E + , out ( 2 ) E + , in ( 1 ) | = | e i 2 β h 1 t 1 t 2 1 r 1 r 2 e i 2 β h 1 | 2 .
T dual = | ( i ( ω 1 ω ) + κ o 1 ) ( i ( ω 2 ω ) + κ o 2 ) exp ( 2 i β h 1 2 β h 1 ) ( i ( ω 1 ω ) + κ o 1 + κ e 1 ) ( i ( ω 2 ω ) + κ o 2 + κ e 2 ) κ e 1 κ e 2 exp ( 2 i β h 1 2 β h 1 ) | 2 .
n g = c v g = c L τ g = c L d φ ( ω ) d ω .
T triple = | exp ( i 2 β ( h 1 + h 2 ) ) t 1 t 2 t 3 1 r 1 r 2 exp ( i 2 β h 1 ) r 2 r 3 exp ( i 2 β h 2 ) r 1 r 3 ( r 2 + t 2 ) exp ( i 2 β ( h 1 + h 2 ) ) | 2 .
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.