Next Article in Journal
Thermal, Mechanical and Electrical Properties of Carbon Fiber Fabric and Graphene Reinforced Segmented Polyurethane Composites
Next Article in Special Issue
Metalens Eyepiece for 3D Holographic Near-Eye Display
Previous Article in Journal
Preparation of Fe3O4-Ag Nanocomposites with Silver Petals for SERS Application
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Symmetric Graphene Dielectric Nanowaveguides as Ultra-Compact Photonic Structures

College of Physics and Electronic Engineering, Zhengzhou Normal University, Zhengzhou 450044, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2021, 11(5), 1281; https://doi.org/10.3390/nano11051281
Submission received: 1 April 2021 / Revised: 7 May 2021 / Accepted: 10 May 2021 / Published: 13 May 2021
(This article belongs to the Special Issue Nanophotonics and Integrated Optics Devices)

Abstract

:
A symmetric graphene plasmon waveguide (SGPWG) is proposed here to achieve excellent subwavelength waveguiding performance of mid-infrared waves. The modal properties of the fundamental graphene plasmon mode are investigated by use of the finite element method. Due to the naturally rounded tips, the plasmon mode in SGPWG could achieve a normalized mode field area of ~10−5 (or less) and a figure of merit over 400 by tuning the key geometric structure parameters and the chemical potential of graphene. In addition, results show that the modal performance of SGPWG seems to improve over its circular counterparts. Besides the modal properties, crosstalk analysis indicates that the proposed waveguide exhibits extremely low crosstalk, even at a separation distance of 64 nm. Due to these excellent characteristics, the proposed waveguide has promising applications in ultra-compact integrated photonic components and other intriguing nanoscale devices.

1. Introduction

Plasmonic waveguides (PWGs) [1], which can confine and guide light at the subwavelength scale, are one of the components necessary to realize ultra-high-density photonic integration [2]. The traditional noble metallic PWGs, such as metal stripe/nanowire waveguides [3,4,5], channel/wedge plasmon waveguides [6], gap plasmon waveguides [7,8,9], dielectric-loaded plasmon waveguides [10,11,12], and hybrid PWGs [1,13,14,15,16,17,18], have been intensively studied in the near-infrared and visible bands. However, in the mid- and far-infrared bands, the plasmonic effects of metals (which were modeled as perfect electric conductors) are very weak and the electromagnetic optical response cannot be dynamically adjusted, which imposes restrictions on their applications at the nanoscale [19,20]. Thus, finding appropriate materials for PWGs is an urgent need.
To address this critical challenge, researchers have suggested some available materials with tunable properties for exciting surface plasmons [21], including graphene [22,23,24], transition metal dichalcogenides [25,26,27], bulk Dirac semimetals [28,29], borophene [30,31], etc. Among these, graphene plasmons (GPs) have attracted widespread attention due to their advantages, including strong light–matter interactions, deep subwavelength field confinement, and tunable optical properties [23,24]. Benefiting from these excellent characteristics, graphene has served as an effective nanoscale waveguiding platform in the infrared region. More importantly, the combination of graphene and silicon-on-insulator (SOI) waveguides make it possible to design graphene-based photonic integration devices [32], such as waveguides [33,34,35,36,37,38,39], sensors [40,41], filters [42], modulators [43,44], etc. Recently, graphene–SiO2–Si coaxial-like waveguides [45], graphene layer–SiO2–Si planar structures [46,47,48,49,50], and graphene-coated nanowires integrated with SiO2 or Si substrates [51,52,53,54,55,56] were presented to demonstrate ultra-compact photonic integrated circuits in the mid- and far-infrared bands.
The combination of graphene and silicon-based waveguides not only provides additional degrees of freedom to tune the modal properties, but also leads to strong coupling between the graphene layer and silicon layer, which massively reduces the modal field area. To realize ultra-compact photonic integration, the modal field area should be small enough to circumvent the interference between neighboring structures. Therefore, highly concentrated modal fields are preferred. Here, we propose a symmetric graphene plasmon waveguide (SGPWG) to achieve nanoscale waveguiding of mid-infrared waves. The results of our simulations show that the elliptical nanowire-based SGPWG performs better than its circular counterparts and is suitable for ultra-compact photonics integration. We first introduce the waveguide structure and methods, then comprehensively evaluate the proposed waveguide in the light of geometric and physical parameters. We also discuss the dependences of modal properties on the ratio of the semi-major axis to the semi-minor axis. Finally, we present the crosstalk analysis and briefly compare the mode characteristics of the elliptical nanowire-based SGPWG with those of the circular nanowire-based SGPWG.

2. Waveguide Structure and Methods

Figure 1 shows the schematic of the proposed SGPWG, which is composed of two graphene-coated elliptical nanowires symmetrically placed on each side of a thin Si slab with a small gap distance, h. The semi-minor axis and semi-major axis of the elliptical nanowire are a and b, respectively. The thickness and width of the Si slab are H and W, respectively. The relative permittivities of the dielectric nanowire, Si slab, and surrounding silica are set as ε1 = 2, ε2 = 12.25 (with an approximated refractive index of 3.5), and ε3 = 2.25, respectively [57,58,59]. The dielectric nanowires are coated by the monolayer graphene. In this work, the monolayer graphene is simulated as an electric field-induced surface current (J = σgE) on the nanowire surfaces [60]. Graphene’s surface conductivity (σg) is obtained from Kubo’s formula, which consists of the intra- and inter-band contributions [56], namely σg = σintra + σinter, where:
σ intra = 2 i e 2 k B T π 2 ( ω + i / τ ) ln [ 2 cosh ( E F 2 k B T ) ]
σ inter = e 2 4 [ 1 2 + 1 π arctan ( ω 2 E F 2 k B T ) i 2 π ln ( ω + 2 E F ) 2 ( ω 2 E F ) 2 + ( 2 k B T ) 2 ]
The electron relaxation time τ is set at 0.5 ps according to [61]. The temperature is T = 300 K; ω =f, with f being the frequency of the incident light; EF is the chemical potential; is the reduced Plank’s constant; kB is the Boltzmann’s constant; and e = 1.6 × 10−19 C.
The GP mode propagates along the z-direction with a complex propagation constant kz, where: kz = k0Neff; k0 = 2π/λ0, where λ0 is the wavelength in air; and Neff is the complex effective mode index. The propagation distance is calculated by LP = λ0/[2πIm(Neff)], where Im(Neff) denotes the imaginary part of Neff. The normalized mode size is defined as AN = Aeff/A0, where A0 = λ02/4, and Aeff is defined as
A eff = W ( r ) d 2 r / max { W ( r ) }
where W(r) represents the energy density of the plasmon mode [1]. Figure of merit (FoM) [62] is defined as LP/(Aeff/π)1/2. The modal properties are investigated by the finite element method (FEM) software COMSOL Multiphysics, which is capable of precisely modeling complex plasmonic structures and thus widely used [1,60,63].

3. Results and Discussion

The time-averaged Poynting vector along the z-direction is given as Sz = 1/2Re(E × M*)|z, where Re(·) is the real part, the superscript of M is the complex conjugate, and E and M are the electric and magnetic field vectors. Figure 2a–c depicts the 2D energy density (Sz) distributions of the fundamental mode in SGPWG for different gap distances (h = 2, 5, 10 nm), where a = 50 nm, b = 100 nm, W = 400 nm, H = 20 nm, EF = 0.5 eV, and f = 30 THz. It could be seen that the optical energy of the plasmon mode is mainly concentrated between the nanowire and the Si layer. For h = 2, 5, and 10 nm, the peak values of the energy density are 0.36, 0.13 and 0.06 W/m2, respectively. To get an intuitive view, Figure 2d,e depicts the normalized energy density distributions along the x (at y = 0 nm) and y (at x = 0 nm) directions (shown in Figure 1b), respectively. Clearly, the energy is mainly restricted between the nanowires. With increasing h, the field confinement weakens, and the energy density becomes more dispersed. Both features are consistent with the coupling behavior between metal nanoparticles [64]. This is because the coupling strength between the GP mode and silicon layer decreases when h increases. In order to increase the degree of field confinement, the gap should be at the deep subwavelength scale (e.g., 2~20 nm), which in turn can help to avoid crosstalk between optical signals in the photonic integrated circuits.
As elliptical nanowires are involved in the proposed SGPWG, we need to consider different b/a values. We chose four representative b/a values throughout the paper, namely b/a = 0.5, 1, 2, and 5. Although the case of b/a = 1 has already been investigated in [52], we show that the b/a > 1 setting presented an increased subwavelength optical energy transmission performance as compared to the b/a = 1 setting.
Figure 3 shows the relationship between the fundamental mode properties and the gap distance h at different b/a values (b/a = 0.5, 1, 2, and 5). The parameters were a = 50 nm, W = 400 nm, H = 20 nm, EF = 0.5 eV, and f = 30 THz. From Figure 3a, it can be seen that the real parts of the effective mode indices (neff) decrease with the increase in h for b/a = 0.5, 1, 2, and 5. For the case of b/a = 0.5, the coupling between the plasmon mode and the Si layer is much stronger, and then gradually weakens with the increase in b/a. For the four cases considered here, LP increases with the increase in h. As a whole, the propagation distance also increases with the increase in b/a, as shown in Figure 3b. Figure 3c,d shows the normalized mode area and figure of merit, respectively. As stated above, the smaller the gap distance, the better the field confinement. Therefore, one can see that AN varies from 3.76 × 10−6 to 8.305 × 10−5 when h increases from 2 to 12 nm. At a fixed h value, the normalized mode area decreases with the increase in b/a. Finally, the plots of figure of merit shown in Figure 3d indicate that the overall performance of the SGPWG can be improved when h is reduced.
Based on above simulations, we show that the cases of b/a > 1 (orange and purple lines) show better performances compared with the cases of b/a = 1 (red lines) and 0.5 (blue lines) in terms of LP, AN, and FoM. In other words, these results show that the modal performance of elliptical nanowire-based SGPWG is better than its circular counterparts. Noticing that the FoM decreases with the increase in the gap distance h, we set h = 5 nm in the following experiments to maintain the good performance of this structure.
Figure 4 shows the relationship between the fundamental mode properties and the thickness of the silicon layer at b/a = 0.5, 1, 2, and 5 when a = 50 nm, W = 400 nm, h = 5 nm, EF = 0.5 eV, and f = 30 THz. The modal properties are similar to those presented in Figure 3, except for Figure 4b, where the propagation length seems to have a maximum value when H is around 30 nm. As show in Figure 4b,c, when H increases, LP first reaches a maximum value and then slightly decreases, while the normalized mode area AN increases monotonically. With increasing H, the corresponding effective mode index and FoM decrease significantly, as shown in Figure 4a,d. Hence, we set H = 20 nm to maintain a large FoM value in the following experiments. Apparently, b/a > 1 settings (orange and purple lines) perform better than b/a = 1 (red lines) and 0.5 (blue lines) settings in terms of LP, AN, and FoM.
The frequency-dependent mode characteristics of the SGPWG are shown in Figure 5 at different b/a values. We set a = 50 nm, W = 400 nm, H = 20 nm, h = 5 nm, and EF = 0.5 eV. As shown in Figure 5a,b, when the frequency varies from 20 to 40 THz, neff increases linearly and LP decreases monotonically. This is because at higher frequencies, the large absorption of graphene leads to the increase in propagation loss. From Figure 5c, we see that the normalized mode field area (AN) increases with increasing frequency, maintaining a level of ~10−5. As for the figure of merit shown in Figure 5d, increasing frequency degenerates the overall performance. In other words, higher frequencies lead to higher neff and AN values, but lower LP and FoM values. Once again, we showed that b/a > 1 settings (orange and purple lines) perform better than b/a = 1 (red lines) and 0.5 (blue lines) settings in terms of LP, AN, and FoM.
To present the tunability of the GP mode in the proposed SGPWG, the mode properties versus EF are presented in Figure 6. In [65], the chemical potential of graphene could reach a value of 1.77 eV; thus, EF varied from 0.4 to 1.6 eV. The other parameters were a = 50 nm, W = 400 nm, H = 20 nm, h = 5 nm, and f = 30 THz. As shown in Figure 6a,b, when EF changes from 0.4 to 1.6 eV, both neff and modal loss decrease monotonically. The latter is due to the fact that when the chemical potential is increased, the interband contribution of σg is drastically reduced, thus leading to the reduction of propagation loss. However, the chemical potential seems to have a limited effect on AN (see Figure 6c), since the modal field area changes a little. In addition, the figure of merit shown in Figure 6d increases rapidly with EF increase, and reaches a value of 850 when EF = 1.6 eV and b/a = 5. Overall, when b/a ranges from 0.5 to 5, LP and FoM increase and AN decreases, indicating that the elliptical nanowire-based SGPWG performs better than its circular counterparts. A recent report showed that the graphene samples fabricated by large scale methods show a relaxation time of only τ = 0.05 ps [66]. Hence, we also studied the modal properties when τ = 0.05 ps (see black lines) and b/a = 5 for comparison. As seen in Figure 6b, LP substantially decreases while neff and AN are nearly unchanged (see Figure 6a,c), and the black lines are overlapped with the purple lines. These results indicate that the shorter relaxation time degenerates the overall performance of the SGPWG (see Figure 6d). This is because the imaginary part of the equivalent relative permittivity of monolayer graphene increases about ten times.
If ultra-compact photonic integration is attempted, highly localized modal fields could help reduce the crosstalk between neighboring structures. Here, we studied the crosstalk between neighboring structures by considering a system consisting of two parallel SGPWGs with an edge-to-edge distance of S (40~100 nm) as shown in Figure 7a. The parameters were a = 50 nm, W = 800 nm, H = 20 nm, h = 5 nm, EF = 0.5 eV, and f = 30 THz. The electric field (Ey) distributions of the symmetric and antisymmetric modes are depicted in Figure 7b,c with S = 40 nm and b/a = 2. Based on the coupled mode theory, the crosstalk is estimated by the coupling length LC, which is the length required for complete power transfer from one waveguide to the other, and is given by LC = λ0/(2|neff,s - neff,as|), where neff,s and neff,as denote the real parts of the effective mode indices of the symmetric and antisymmetric modes, respectively [7]. Figure 7d shows the normalized coupling lengths (LC/LP) with respect to S for different b/a values. Usually, when LC/LP approaches 10 (see the black dashed line), it is assumed that no coupling happens between the neighboring components. When LC/LP reaches 10, the corresponding edge-to-edge distance are about 44.0, 55.5, 61.7, and 64.0 nm for b/a = 0.5, 1, 2, and 5, respectively. It is worth noting that when S is above 64 nm, LC/LP is always larger than 10, which indicates that the proposed SGPWG shows extremely low crosstalk between neighboring components and is suitable for ultra-compact photonics integration.
Next, we briefly compare the mode characteristics among three different kinds of waveguides, including circular nanowire-based SGPWGs with R = a (Waveguide A) and R = b (Waveguide C), and an elliptical nanowire-based SGPWG (Waveguide B). The schematics of the three SGPWGs are shown in Figure 8a, and we set a = 50 nm, b = 150 nm, W = 400 nm, h = 5 nm, H = 20 nm, EF = 0.5 eV, and f = 30 THz. Figure 8b shows the 2D energy density distributions of the fundamental mode in three SGPWGs. The results show that Waveguide B exhibits much stronger energy confinement than Waveguides A and C, which can be intuitively seen from Table 1 (lower AN). As shown in Table 1, the plasmon mode in Waveguide B has the smallest normalized mode area and largest LP compared with the other waveguides.
Finally, we briefly introduce the fabrication process of the SGPWG. Recent reports [67,68] showed that graphene-coated elliptical nanowires can be experimentally made by coating a dielectric nanowire with a monolayer of graphene due to van der Waals forces. Then, based on modern semiconductor fabrication technology, the substrate silica can be deposited on top of the buffer layer [69]. After that, one GCNW is placed on the silica substrate, then the GCNW is covered by silica deposition with a thickness of 2b + h. Next, a silicon layer with a thickness of H (about 10 nm) could be deposited on top of the silica using plasma-enhanced chemical vapor deposition (PECVD) technology [70]. Then, h-thick silica is deposited on the silicon layer before transferring another GCNW. Finally, the second GCNW is covered by silica deposition with a thickness of over 2b.

4. Conclusions

In summary, a symmetric graphene plasmon waveguide is proposed and investigated. The simulation results show that a normalized mode field area of ∼10-5 and a figure of merit of ~400 can be achieved by optimizing the parameters. The mode characteristics could be dynamically tuned by changing the chemical potential of graphene by a DC bias voltage or chemical doping. By varying the ratio of b/a, the simulation results show that elliptical nanowire-based SGPWGs (b/a > 1) show better performance compared with the cases of b/a ≤1 in terms of LP, AN, and FoM. In addition, crosstalk analysis suggests that the proposed SGPWGs have extremely small energy couplings between neighboring components, even at a separation distance of 64 nm. These findings could have potential applications for ultra-compact photonic integration and subwavelength optoelectronic devices in the mid-infrared band.

Author Contributions

Conceptualization, D.T.; writing—original draft preparation, D.T.; Y.W.; writing—review and editing, Y.W.; T.X.; visualization, H.W.; Q.S.; supervision, D.T.; Y.T. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Key Scientific Research Project of Henan College (21A140029), the Young Backbone Teacher Training Program of Zhengzhou Normal University (QNGG-20774), the Open Research Fund of Zhengzhou Normal University, and the Scientific Research Starting Foundation of Zhengzhou Normal University.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Oulton, R.F.; Sorger, V.J.; Genov, D.A.; Pile, D.F.P.; Zhang, X. A hybrid plasmonic waveguide for subwavelength confinement and long-range propagation. Nat. Photon. 2008, 2, 496–500. [Google Scholar] [CrossRef] [Green Version]
  2. Cheben, P.; Halir, R.; Schmid, J.H.; Atwater, H.A.; Smith, D.R. Subwavelength integrated photonics. Nature 2018, 560, 565–572. [Google Scholar] [CrossRef]
  3. Verhagen, E.; Polman, A.; Kuipers, L.K. Nanofocusing in laterally tapered plasmonic waveguides. Opt. Express 2008, 16, 45–57. [Google Scholar] [CrossRef] [PubMed]
  4. Yang, L.; Li, P.; Li, Z. Plasmonic polarization beam splitting based on single silver nanowire. Opt. Express 2019, 27, 3851–3860. [Google Scholar] [CrossRef] [PubMed]
  5. Lu, F.; Zhang, W.; Huang, L.; Liang, S.; Mao, D.; Gao, F.; Mei, T.; Zhao, J. Mode evolution and nanofocusing of grating-coupled surface plasmon polaritons on metallic tip. Opto-Electron. Adv. 2018, 1, 180010. [Google Scholar] [CrossRef]
  6. Moreno, E.; Rodrigo, S.G.; Bozhevolnyi, S.I.; Martín-Moreno, L.; García-Vidal, F.J. Guiding and focusing of electromagnetic fields with wedge plasmon polaritons. Phys. Rev. Lett. 2008, 100, 023901. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Veronis, G.; Fan, S. Crosstalk between three-dimensional plasmonic slot waveguides. Opt. Express 2008, 16, 2129–2140. [Google Scholar] [CrossRef] [Green Version]
  8. Gao, H.; Cao, Q.; Zhu, M.; Teng, D.; Shen, S. Nanofocusing of terahertz wave in a tapered hyperbolic metal waveguide. Opt. Express 2014, 22, 32071–32081. [Google Scholar] [CrossRef]
  9. Zhang, Q.; Hao, H.; Ren, J.; Zhang, F.; Gong, Q.; Gu, Y. A quantum phase gate capable of effectively collecting photons based on a gap plasmon structure. Nanoscale 2020, 12, 10082–10089. [Google Scholar] [CrossRef]
  10. Steinberger, B.; Hohenau, A.; Ditlbacher, H.; Stepanov, A.L.; Drezet, A.; Aussenegg, F.R.; Leitner, A.; Krenn, J.R. Dielectric stripes on gold as surface plasmon waveguides. Appl. Phys. Lett. 2006, 88, 094104. [Google Scholar] [CrossRef]
  11. Krasavin, A.V.; Zayats, A.V. Three-dimensional numerical modeling of photonic integration with dielectric-loaded SPP waveguides. Phys. Rev. B 2008, 78, 045425. [Google Scholar] [CrossRef]
  12. Han, Z.; Radko, I.P.; Mazurski, N.; Desiatov, B.; Beermann, J.; Albrektsen, O.; Levy, U.; Bozhevolnyi, S.I. On-chip detection of radiation guided by dielectric-loaded plasmonic waveguides. Nano Lett. 2015, 15, 476–480. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Dai, D.; He, S. A silicon-based hybrid plasmonic waveguide with a metal cap for a nano-scale light confinement. Opt. Express 2009, 17, 16646–16653. [Google Scholar] [CrossRef]
  14. Chen, L.; Zhang, T.; Li, X.; Huang, W. Novel hybrid plasmonic waveguide consisting of two identical dielectric nanowires symmetrically placed on each side of a thin metal film. Opt. Express 2012, 20, 20535–20544. [Google Scholar] [CrossRef] [PubMed]
  15. Alam, M.Z.; Aitchison, J.S.; Mojahedi, M. A marriage of convenience: Hybridization of surface plasmon and dielectric waveguide modes. Laser Photonics Rev. 2014, 8, 394–408. [Google Scholar] [CrossRef]
  16. Zhang, L.; Pan, C.; Zeng, D.; Yang, Y.; Yang, Y.; Junxian, M. A hybrid-plasmonic-waveguide-based polarization-independent directional coupler. IEEE Access 2020, 8, 134268–134275. [Google Scholar] [CrossRef]
  17. Teng, D.; Cao, Q.; Wang, K. An extension of the generalized nonlocal theory for the mode analysis of plasmonic waveguides at telecommunication frequency. Journal of Optics 2017, 19, 055003. [Google Scholar] [CrossRef] [Green Version]
  18. Tuniz, A.; Bickerton, O.; Diaz, F.J.; Kasebier, T.; Kley, E.B.; Kroker, S.; Palomba, S.; de Sterke, C.M. Modular nonlinear hybrid plasmonic circuit. Nat. Commun. 2020, 11, 2413. [Google Scholar] [CrossRef] [PubMed]
  19. Huang, T.J.; Yin, L.Z.; Zhao, J.; Du, C.H.; Liu, P.K. Amplifying evanescent waves by dispersion-induced plasmons: Defying the materials limitation of superlens. ACS Photonics 2020, 7, 2173–2181. [Google Scholar] [CrossRef]
  20. Gao, Y.; Shadrivov, I.V. Second harmonic generation in graphene-coated nanowires. Opt. Lett. 2016, 41, 3623–3626. [Google Scholar] [CrossRef] [PubMed]
  21. Barnes, W.L.; Dereux, A.; Ebbesen, T.W. Surface plasmon subwavelength optics. Nature 2003, 424, 824–830. [Google Scholar] [CrossRef]
  22. Novoselov, K.S.; Fal, V.I.; Colombo, L.; Gellert, P.R.; Schwab, M.G.; Kim, K. A roadmap for graphene. Nature 2012, 490, 192–200. [Google Scholar] [CrossRef] [PubMed]
  23. Fan, Y.C.; Shen, N.H.; Zhang, F.L.; Zhao, Q.; Wu, H.J.; Fu, Q.H.; Wei, Z.Y.; Li, H.Q.; Soukoulis, C.M. Graphene plasmonics: A platform for 2D Optics. Adv. Opt. Mater. 2019, 7, 1800537. [Google Scholar] [CrossRef] [Green Version]
  24. Teng, D.; Wang, K.; Li, Z. Graphene-coated nanowire waveguides and their applications. Nanomaterials 2020, 10, 229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Zheng, K.; Yuan, Y.F.; He, J.J.; Gu, G.Q.; Zhang, F.; Chen, Y.; Song, J.; Qu, J.L. Ultra-high light confinement and ultra-long propagation distance design for integratable optical chips based on plasmonic technology. Nanoscale 2019, 11, 4601–4613. [Google Scholar] [CrossRef]
  26. Shan, H.; Yu, Y.; Wang, X.; Luo, Y.; Zu, S.; Du, B.; Han, T.Y.; Li, B.W.; Li, Y.; Wu, J.R.; et al. Fang, Z. Direct observation of ultrafast plasmonic hot electron transfer in the strong coupling regime. Light Sci. Appl. 2019, 8, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. He, X.; Liu, F.; Lin, F.; Xiao, G.; Shi, W. Tunable MoS2 modified hybrid surface plasmon waveguides. Nanotechnology 2018, 30, 125201. [Google Scholar] [CrossRef] [PubMed]
  28. Jiang, Y.; Shi, C.; Wang, J. A hybrid plasmonic terahertz waveguide with ridge structure base on Bulk-Dirac-semimetal. Opt. Commun. 2020, 475, 126239. [Google Scholar] [CrossRef]
  29. He, X.; Liu, F.; Lin, F.; Lin, F.; Shi, W. Tunable 3D Dirac-semimetals supported Mid-IR hybrid plasmonic waveguides. Opt. Lett. 2021, 46, 472–475. [Google Scholar] [CrossRef]
  30. Dereshgi, S.A.; Liu, Z.; Aydin, K. Anisotropic localized surface plasmons in borophene. Opt. Express 2020, 28, 16725–16739. [Google Scholar] [CrossRef] [PubMed]
  31. Lian, C.; Hu, S.Q.; Zhang, J.; Cheng, C.; Yuan, Z.; Gao, S.; Meng, S. Integrated plasmonics: Broadband dirac plasmons in borophene. Phys. Rev. Lett. 2020, 125, 116802. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, J.; Xing, Z.; Chen, X.; Cheng, Z.Z.; Li, X.J.; Liu, T. Recent progress in waveguide-integrated graphene photonic devices for sensing and communication applications. Front. Phys. 2020, 8, 37. [Google Scholar] [CrossRef]
  33. Yu, P.; Fesenko, V.I.; Tuz, V.R. Dispersion features of complex waves in a graphene-coated semiconductor nanowire. Nanophotonics 2018, 7, 925–934. [Google Scholar] [CrossRef]
  34. Saeed, M.; Ghaffar, A.; Alkanhal, M.A.S.; Alqahtani, A.H.; Khan, Y.; ur Rehman, S. Plasmon modes supported by metamaterial-filled monolayer graphene cylindrical waveguides. J. Opt. Soc. Am. B 2020, 37, 3515–3525. [Google Scholar] [CrossRef]
  35. Xing, R.; Jian, S. Numerical analysis on the multilayer nanoring waveguide pair. IEEE Photon. Technol. Lett. 2016, 28, 2779–2782. [Google Scholar] [CrossRef]
  36. Teng, D.; Wang, K.; Li, Z.; Zhao, Y. Graphene-coated nanowire dimers for deep subwavelength waveguiding in mid-infrared range. Opt. Express 2019, 27, 12458–12469. [Google Scholar] [CrossRef]
  37. Wu, D.; Tian, J. Study on the plasmonic characteristics of bow-tie type graphene-coated nanowire pair. Optik 2018, 156, 689–695. [Google Scholar] [CrossRef]
  38. Wang, X.; Wang, J.; Ma, T.; Liu, H.; Wang, F. Plasmonic characteristics of suspended graphene-coated wedge porous silicon nanowires with Ag partition. Chin. Phys. B 2021, 30, 014207. [Google Scholar] [CrossRef]
  39. Teng, D.; Wang, K. Theoretical analysis of terahertz dielectric–loaded graphene waveguide. Nanomaterials 2021, 11, 210. [Google Scholar] [CrossRef] [PubMed]
  40. Jabbarzadeh, F.; Habibzadeh-Sharif, A. High performance dielectric loaded graphene plasmonic waveguide for refractive index sensing. Opt. Commun. 2021, 479, 126419. [Google Scholar] [CrossRef]
  41. Zhu, J.; Jiang, F.; Yunbai, Q. Sense of surface plasmon polarization waveguide of graphene. Plasmonics 2019, 14, 1903–1910. [Google Scholar] [CrossRef]
  42. Liao, B.; Guo, X.; Hu, H.; Liu, N.; Chen, K.; Yang, X.; Dai, Q. Ultra-compact graphene plasmonic filter integrated in a waveguide. Chin. Phys. B 2018, 27, 094101. [Google Scholar] [CrossRef]
  43. Xu, Y.; Li, F.; Kang, Z.; Huang, D.; Zhang, X.; Tam, H.-Y.; Wai, P.K.A. Hybrid graphene-silicon based polarization-insensitive electro-absorption modulator with high-modulation efficiency and ultra-broad bandwidth. Nanomaterials 2019, 9, 157. [Google Scholar] [CrossRef] [Green Version]
  44. Hao, R.; Jiao, J.; Peng, X.; Zhen, Z.; Dagarbek, R.; Zou, Y.; Li, E. Experimental demonstration of a graphene-based hybrid plasmonic modulator. Opt. Lett. 2019, 44, 2586–2587. [Google Scholar] [CrossRef] [PubMed]
  45. Liu, J.P.; Zhai, X.; Xie, F.; Wang, L.L.; Xia, S.X.; Li, H.J.; Luo, X.; Shang, X.J. Analytical model of mid-infrared surface plasmon modes in a cylindrical long-range waveguide with double-layer graphene. J. Lightwave Technol. 2017, 35, 1971–1979. [Google Scholar] [CrossRef]
  46. Hajati, M.; Hajati, Y. Investigation of plasmonic properties of graphene multilayer nano-ribbon waveguides. Appl. Opt. 2016, 55, 1878–1884. [Google Scholar] [CrossRef] [PubMed]
  47. Wang, Y.; Liu, H.; Wang, S.; Cai, M.; Ma, L. Optical transport properties of graphene surface plasmon polaritons in mid-infrared band. Crystals 2019, 9, 354. [Google Scholar] [CrossRef] [Green Version]
  48. Ctyroky, J.; Petráek, J.; Kuzmiak, V.; Kwiecien, P.; Richter, I. Silicon waveguides with graphene: Coupling of waveguide mode to surface plasmons. J. Opt. 2020, 22, 095801. [Google Scholar] [CrossRef]
  49. Chen, X.; Wang, Y.; Xiang, Y.; Jiang, G.; Wang, L.; Bao, Q.; Zhang, H.; Liu, Y.; Wen, S.; Fan, D. A broadband optical modulator based on a graphene hybrid plasmonic waveguide. J. Lightwave Technol. 2016, 34, 4948–4953. [Google Scholar] [CrossRef]
  50. Hasan, K.B.M.R.; Islam, M.A.; Alam, M.S. Design of a broadband single mode hybrid plasmonic waveguide incorporating silicon nanowire. Opt. Mater. Express 2020, 10, 2783–2799. [Google Scholar] [CrossRef]
  51. Hajati, M.; Hajati, Y. High-performance and low-loss plasmon waveguiding in graphene-coated nanowire with substrate. J. Opt. Soc. Am. B 2016, 33, 2560–2565. [Google Scholar] [CrossRef]
  52. Hajati, M.; Hajati, Y. Plasmonic characteristics of two vertically coupled graphene-coated nanowires integrated with substrate. Appl. Opt. 2017, 56, 870–875. [Google Scholar] [CrossRef]
  53. Sun, M.; Tian, J.; Lan, X.; He, Z.; Liu, J. Transmission properties of two vertically coupled double-graphene-coated nanowires integrated with substrate. Optik 2019, 185, 242–247. [Google Scholar] [CrossRef]
  54. Teng, D.; Guo, J.; Yang, Y.; Ma, W.; Wang, K. Study of modal properties in graphene-coated nanowires integrated with substrates. Appl. Phys. B 2020, 126, 173. [Google Scholar] [CrossRef]
  55. Teng, D.; Yang, Y.; Guo, J.; Ma, W.; Tang, Y.; Wang, K. Efficient guiding mid-infrared waves with graphene-coated nanowire based plasmon waveguides. Res. Phys. 2020, 17, 103169. [Google Scholar] [CrossRef]
  56. Teng, D.; Wang, K.; Huan, Q.; Chen, W.; Li, Z. High-performance light transmission based on graphene plasmonic waveguides. J. Mater. Chem. C 2020, 8, 6832–6838. [Google Scholar] [CrossRef]
  57. Chandler-Horowitz, D.; Amirtharaj, P.M. High-accuracy, midinfrared (450 cm− 1⩽ ω ⩽ 4000 cm− 1) refractive index values of silicon. J. Appl. Phys. 2005, 97, 123526. [Google Scholar] [CrossRef]
  58. Yang, X.; Liu, X.; Yu, S.; Gan, L.; Zhou, J.; Zeng, Y. Permittivity of undoped silicon in the millimeter wave range. Electronics 2019, 8, 886. [Google Scholar] [CrossRef] [Green Version]
  59. Yu, H.; Wang, H.; Xiong, Q.; Mei, J.; Zhang, Y.; Wang, Y.; Lai, J.; Chen, C. Photothermal switch of sub-microsecond response: A monolithic-integrated ring resonator and a metasurface absorber in silicon photonic crystals. Opt. Lett. 2020, 45, 1806–1809. [Google Scholar]
  60. Ono, M.; Hata, M.; Tsunekawa, M.; Nozaki, K.; Sumikura, H.; Chiba, H.; Notomi, M. Ultrafast and energy-efficient all-optical switching with graphene-loaded deep-subwavelength plasmonic waveguides. Nat. Photon. 2020, 14, 37–43. [Google Scholar] [CrossRef] [Green Version]
  61. Nikitin, A.Y.; Alonso-González, P.; Hillenbrand, R. Efficient coupling of light to graphene plasmons by compressing surface polaritons with tapered bulk materials. Nano Lett. 2014, 14, 2896–2901. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Buckley, R.; Berini, P. Figures of merit for 2D surface plasmon waveguides and application to metal stripes. Opt. Express 2007, 15, 12174–12182. [Google Scholar] [CrossRef] [PubMed]
  63. Said, A.; Atia, K.S.R.; Obayya, S.S.A. On modeling of plasmonic devices: Overview. J. Opt. Soc. Am. B 2020, 37, A163–A174. [Google Scholar] [CrossRef]
  64. Caridad, J.M.; Winters, S.; McCloskey, D.; Duesberg, G.S.; Donegan, J.F.; Krstić, V. Hot-volumes as uniform and reproducible sers-detection enhancers in weakly-coupled metallic nanohelices. Sci. Rep. 2017, 7, 45548. [Google Scholar] [CrossRef] [Green Version]
  65. Kanahashi, K.; Tanaka, N.; Shoji, Y.; Maruyama, M.; Jeon, I.; Kawahara, K.; Ishihara, M.; Hasegawa, M.; Ohta, H.; Ago, H.; et al. Formation of environmentally stable hole-doped graphene films with instantaneous and high-density carrier doping via a boron-based oxidant. npj 2D Mater. Appl. 2019, 3, 44. [Google Scholar] [CrossRef] [Green Version]
  66. Whelan, P.R.; Shen, Q.; Zhou, B.; Serrano, I.G.; Kamalakar, M.V.; Mackenzie, D.M.; Ji, J.; Huang, D.; Shi, H.; Luo, D.; et al. Fermi velocity renormalization in graphene probed by terahertz time-domain spectroscopy. 2D Mater. 2020, 7, 035009. [Google Scholar] [CrossRef]
  67. Cao, T.; Li, Y.; Tian, L.; Liang, H.; Qin, K. Fast switching “On/Off” chiral surface plasmon polaritons in graphene-coated Ge2Sb2Te5 nanowire. ACS Appl. Nano Mater. 2018, 1, 759–767. [Google Scholar] [CrossRef]
  68. Chen, B.; Meng, C.; Yang, Z.; Li, W.; Lin, S.; Gu, T.; Guo, X.; Wang, D.; Yu, S.; Wong, C.W.; et al. Graphene coated ZnO nanowire optical waveguides. Opt. Express 2014, 22, 24276–24285. [Google Scholar] [CrossRef]
  69. Vitanov, P.; Ivanova, T.; Dikov, H. Low-temperature deposition of ultrathin SiO2 films on Si substrates. J. Phys. Conf. Ser. 2014, 514, 012010. [Google Scholar] [CrossRef]
  70. Flöry, N.; Ma, P.; Salamin, Y.; Emboras, A.; Taniguchi, T.; Watanabe, K.; Leuthold, J.; Novotny, L. Waveguide-integrated van der Waals heterostructure photodetector at telecom wavelengths with high speed and high responsivity. Nat. Nanotechnol. 2020, 15, 118–124. [Google Scholar] [CrossRef]
Figure 1. Schematic of the proposed SGPWG. (a) Three-dimensional (3D) view, (b) two-dimensional (2D) cross-section.
Figure 1. Schematic of the proposed SGPWG. (a) Three-dimensional (3D) view, (b) two-dimensional (2D) cross-section.
Nanomaterials 11 01281 g001
Figure 2. Energy density distribution. (ac) Two-dimensional energy density distributions when h = 2, 5, and 10 nm. (de) Normalized energy density distributions along the x (at y = 0 nm) and y (at x = 0 nm) directions (shown in Figure 1b).
Figure 2. Energy density distribution. (ac) Two-dimensional energy density distributions when h = 2, 5, and 10 nm. (de) Normalized energy density distributions along the x (at y = 0 nm) and y (at x = 0 nm) directions (shown in Figure 1b).
Nanomaterials 11 01281 g002
Figure 3. Modal properties of h at different b/a values. (a) neff, (b) LP, (c) AN, (d) FoM.
Figure 3. Modal properties of h at different b/a values. (a) neff, (b) LP, (c) AN, (d) FoM.
Nanomaterials 11 01281 g003
Figure 4. Modal properties of H at different b/a values. (a) neff, (b) LP, (c) AN, (d) FoM.
Figure 4. Modal properties of H at different b/a values. (a) neff, (b) LP, (c) AN, (d) FoM.
Nanomaterials 11 01281 g004
Figure 5. Modal properties of f at different b/a values. (a) neff, (b) LP, (c) AN, (d) FoM.
Figure 5. Modal properties of f at different b/a values. (a) neff, (b) LP, (c) AN, (d) FoM.
Nanomaterials 11 01281 g005
Figure 6. Modal properties of EF at different b/a values. (a) neff, (b) LP, (c) AN, (d) FoM. Note that the black lines overlap the purple lines in (a) and (c).
Figure 6. Modal properties of EF at different b/a values. (a) neff, (b) LP, (c) AN, (d) FoM. Note that the black lines overlap the purple lines in (a) and (c).
Nanomaterials 11 01281 g006
Figure 7. Crosstalk between SGPWGs. (a) Coupling system; (b) and (c) for normalized Ey field distributions of symmetric and antisymmetric modes when b/a = 2 and S = 40 nm; (d) LC/LP with respect to S for different b/a values.
Figure 7. Crosstalk between SGPWGs. (a) Coupling system; (b) and (c) for normalized Ey field distributions of symmetric and antisymmetric modes when b/a = 2 and S = 40 nm; (d) LC/LP with respect to S for different b/a values.
Nanomaterials 11 01281 g007
Figure 8. Comparison of three SGPWGs. (a) Cross-section of Waveguide A (R = a = 50 nm), B (a = 50 nm, b = 150 nm), and C (R = b = 150 nm); (b) 2D energy density distributions of (a).
Figure 8. Comparison of three SGPWGs. (a) Cross-section of Waveguide A (R = a = 50 nm), B (a = 50 nm, b = 150 nm), and C (R = b = 150 nm); (b) 2D energy density distributions of (a).
Nanomaterials 11 01281 g008
Table 1. Waveguiding performance comparison of three SGPWGs.
Table 1. Waveguiding performance comparison of three SGPWGs.
WaveguideNeffLP/μmAN
A58.845 + 0.642i2.4772.512 × 10−5
B52.428 + 0.608i2.6161.699 × 10−5
C65.394 + 0.658i2.4173.501 × 10−5
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Teng, D.; Wang, Y.; Xu, T.; Wang, H.; Shao, Q.; Tang, Y. Symmetric Graphene Dielectric Nanowaveguides as Ultra-Compact Photonic Structures. Nanomaterials 2021, 11, 1281. https://doi.org/10.3390/nano11051281

AMA Style

Teng D, Wang Y, Xu T, Wang H, Shao Q, Tang Y. Symmetric Graphene Dielectric Nanowaveguides as Ultra-Compact Photonic Structures. Nanomaterials. 2021; 11(5):1281. https://doi.org/10.3390/nano11051281

Chicago/Turabian Style

Teng, Da, Yuncheng Wang, Tianzi Xu, Huayu Wang, Qinqin Shao, and Yanan Tang. 2021. "Symmetric Graphene Dielectric Nanowaveguides as Ultra-Compact Photonic Structures" Nanomaterials 11, no. 5: 1281. https://doi.org/10.3390/nano11051281

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop