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

Enhanced second harmonic generation by an atomically thin MoS2 sheet attached to a resonant metasurface

Open Access Open Access

Abstract

Two-dimensional transition metal dichalcogenides have shown large second-order nonlinear responses due to their broken crystal inversion symmetry. However, their nonlinear interaction with light is restricted to an atomically thin layer. Placing a sheet of transition metal dichalcogenides on a resonant metasurface enhances the field interacting with the nonlinear material thus compensating for this shortcoming. But, it remains a challenge to tune resonances such, that they coincide with fundamental and second harmonic frequencies simultaneously. Here we demonstrate two independent methods to achieve that goal and numerically illustrate our findings for a MoS2 layer combined with silicon nitride photonic crystals. We numerically demonstrate 20-fold and 170-fold enhancement of second-harmonic generation compared with a design based on a single resonant structure. Although we focus on that specific configuration our approach can likewise be applied to other dielectrics combined with highly nonlinear 2D materials.

© 2024 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Nonlinear wave mixing processes are employed for sensing [14], sum- and difference frequency generation [57], and the generation of entangled photon pairs [810]. In recent years, monolayers of transition metal dichalcogenides (TMDCs) such as MoS2 have shown great promise as they can be combined with various photonic platforms including silicon compounds [11,12], but also show a strong second-order nonlinear response caused by their intrinsically broken inversion symmetry [1315]. Hence, TMDCs allow to design of the linear and nonlinear response of photonic platforms to a certain extent independently. But, unfortunately, conversion efficiencies of TMDCs are limited by their atomic thickness. Thus, design strategies to enhance light-matter interaction in TMDCs are required. Here we focus on second harmonic generation (SHG) aiming at the up conversion of a fundamental harmonic (FH) into a second harmonic (SH) wave at double the frequency. But, our approach can likewise be applied to other nonlinear wave mixing processes based on light matter interaction in TMDCs.

Numerical simulations have shown that SHG in TMDCs is enhanced by orders of magnitude if the 2D material sheet is combined with resonant photonic structures [16,17] as optical cavities [18,19], waveguides [20,21], resonant plasmonic structures [2224] and dielectric metasurfaces [25,26]. Experiments confirmed these predictions for one-dimensional (1D) and two-dimensional (2D) photonic crystals (PCs) [25,27]. All these structures have in common that they are resonant at the FH frequency. Further improvement is expected for double resonant structures simultaneously enhancing FH and SH fields. However, such an approach requires a kind of phase matching and is usually hindered by dissimilar refractive indices at FH and SH wavelengths. In a recent theoretical work, the authors proposed to circumvent that problem by using arrays of metallic nanostructures the dipolar and quadrupolar modes of which are resonant at FH and SH frequencies [28]. Numerical simulations suggest four orders of magnitude enhancement compared with the bare structure. In another experimental paper, the authors report 16-fold enhancement based on a TMDC embedded in a layered cavity structure and Bragg reflectors with stop bands at FH and SH frequencies [29].

Here we follow a different and scalable approach based on a very simple all-dielectric planar waveguide which is decorated with periodically arranged trenches and a TMDC on top. We show that deliberately breaking the symmetry in such 1D and 2D metasurfaces can be used to match resonances at FH and SH frequencies. In the case of the 2D structure limits of achievable enhancement are not set by the structure itself, but by fabrication tolerances and the linewidth of the exciting laser.

We numerically determine the resulting SH enhancement by a generalized source finite-difference time-domain (GS-FDTD) approach, which is well suited for low conversion efficiencies and allows us to apply the FDTD method to nonlinear 2D materials by introducing the concept of a nonlinear generalized source [30].

2. Results and discussion

Based on the GS-FDTD method, the nonlinear optical interaction is modeled by conducting the nonlinear simulation based on two distinct linear FDTD simulations. Since the 2D TMDC is modeled as a generalized surface current, its thin layer compared to the rest of the structure can be neglected during the simulations. In the initial linear simulation, the excitation via a continuous-wave (CW) source at the FH wavelength 880 nm is considered. It excites a second order polarization in the MoS2 layer via the susceptibility tensor of the D3h point-group $\chi _{x^{\prime} x^{\prime} x^{\prime} }^{(2 )} ={-} \chi _{x^{\prime} y^{\prime} y^{\prime} }^{(2 )} ={-} \chi _{y^{\prime} y^{\prime} x^{\prime} }^{(2 )} ={-} \chi _{y^{\prime} x^{\prime} y^{\prime} }^{(2 )} = 5{\; }{10^{ - 9}}\textrm{m}\; {\textrm{V}^{ - 1}}$ where $x^{\prime}$ corresponds to the arm-chair and $y^{\prime}$ to the zig-zag direction as reported in [2,31]. The time derivative of this second-order nonlinear polarization acts as a nonlinear generalized source current in another FDTD simulation modeling SHG in the structure. The employed generalized source current models the electromagnetic response of the 2D TMDC. Consequently, the need for a bulk layer representation of the 2D material is eliminated in simulation [30].

The dielectric materials (silica, silicon nitride) used in the simulations have negligible dispersion in the investigated frequency domain. Respective refractive indices were taken from [32,33].

2.1 MoS2 on 1D PCs

First, we assume a simple grating structure with a MoS2 monolayer on top (see a principal sketch of the structure in Fig. 1(a)). To induce resonances at FH and SH wavelength we merge, two nonequivalent unit cells (Fig. 1(b) and 1(c)) into a single bigger one (Fig. 1(d)).

 figure: Fig. 1.

Fig. 1. 1D PCs with a MoS2 monolayer placed on top (a) principal sketch and (b-d) construction of the unit cell of a double resonant grating by combining a symmetric 1D PCs (b) with a period of Λ to couple the FH light into the guiding layer and one with a shorter period Λ’ (c) to couple the SH light. Thus, a unit cell of an asymmetric double periodic grating (d) with a period of Λ is formed. (e) Linear reflection spectra of symmetric and asymmetric 1D PCs shown in (b-d) for normal incident and polarization parallel to the grating bars.

Download Full Size | PDF

We start with a grating on top of a high index silicon nitride layer (see Fig. 1(b), period Λ = 540 nm, width w = 500 nm, depth h = 120 nm, thickness of silicon nitride layer on silica d = 180 nm). For an excitation under normal incidence at FH frequency its first diffracted order excites the fundamental guided mode of the high index layer. As a consequence, a leaky wave resonance (LWR) appears at FH frequency causing a sharp peak in the reflection spectrum at 880 nm (see dashed blue line in Fig. 1(e)). A further LWR, which is caused by the excitation of the same waveguide mode via the second order of the grating shows up at 480 nm. However, due to waveguide dispersion, this second LWR does not match the SH wavelength and no resonant enhancement of the SH wave is obtained.

In contrast, a grating with a much shorter period (see the blue area in Fig. 1(c), period Λ’ = 270 nm, width w’ = 130 nm) excites the waveguide mode via its first diffracted order at SH frequency as demonstrated by a peak in reflection at 440 nm (see dashed green line in Fig. 1(e)). But this grating now lacks the resonance at FH frequency.

To induce resonances at FH and SH frequencies we partially combine both structures (yellow and blue areas in Fig. 1(c)). Thus, we break the symmetry in the unit cell by slightly increasing the width of one of the bars (see Fig. 1(d), widths dw = 60 nm). The resulting asymmetric structure has still the right period Λ = 540 nm to excite the FH mode. But, it shows also signatures of the high frequency grating (displayed in Fig. 1(c)). In particular it still reproduces the LWR at 440 nm. Hence, this double periodic grating (see Fig. 1(d)) shows peaks in reflection at both FH and SH wavelengths (see red line in Fig. 1(e)).

In addition, a further resonance shows up in the reflection spectrum at 800 nm. These two LWRs around the FH frequency are caused by coupling forward and backward propagating guided modes by the second order of the grating (see intensity profiles of resulting super modes in Figs. 2(a) and 2(b)). Two similar LWR states, a symmetric and an antisymmetric one, also exist for the symmetric grating as displayed in Fig. 1(b), but only the symmetric one can be excited under normal incidence (see Fig. 1(e), blue dashed line). Breaking the symmetry of the unit cells makes both modes optically accessible (see Fig. 1(d)), although they still differ considerably with respect to their field shapes. Maxima are found either at the bars or the trenches. Consequently, the LWR at 880 nm, which is mainly localized inside the bars, is very sensitive to an increase in their width. In contrast, the LWR at SH frequency is located merely, below the grating trenches (see Fig. 2(c)) and is almost not affected by changes in the bars. Hence, the width of the bars is a convenient tuning parameter to shift the LWR at 880 nm with respect to that at 440 nm thus giving us a tuning parameter to obtain resonances exactly at FH and SH frequencies.

 figure: Fig. 2.

Fig. 2. Mode profiles of the asymmetric 1D PC at selected resonances (red line in Fig. 1(e)) via first-order diffraction (a) at FH wavelength, (b) at 800 nm, and (c) second-order diffraction at SH wavelength.

Download Full Size | PDF

The double resonant 1D PC structure provides not only a strong excitation of TMDCs placed on top by the FH field but also an efficient extraction of generated SH by resonant scattering towards free space. The asymmetric double resonant 1D PC structure allows for 20 times more efficient SHG compared to the symmetric 1D PC (see Fig. 1(b)), which is only resonant for the FH frequency (see Fig. 3) and 4600 times more enhancement compared with SH generation in the structure being resonant at SH wavelength only (see Fig. 1(c)).

 figure: Fig. 3.

Fig. 3. Normalized SH intensity from MoS2 for varying wavelengths of the FH driving field. The electric field is polarized parallel to the bars and the TMDC sheet is assumed to be oriented such that its armchair axis is parallel to the bars.

Download Full Size | PDF

2.2 SHG in MoS2 on a 2D PCs

The grating structure discussed above does not utilize the full flexibility of a 2D metasurface covered by a TMDC sheet. When considering a genuine 2D metasurface the strong polarization dependence of SHG in TMDCs must be considered. It is determined by the second-order susceptibility tensor with D3h point-group symmetry as introduced above. The generated SH is parallel to the driving FH field only if both are directed in the armchair direction. If the FH field deviates from that direction by an angle $\varphi $ the SH polarization turns into the opposite direction without changing its magnitude. Therefore, the part of the SH field pointing into the excitation direction is modulated like $\propto \cos ({3\varphi } )$.

Provided that the orientation of the TMDC sheet is chosen correctly a SH waveguide mode propagating orthogonally to the original FH mode can be excited. Hence, a 2D PC allows disentangling the two resonances at FH and SH wavelength as we will demonstrate below.

Again, we study an asymmetric, but 2D PC. Here, asymmetry refers to two different periods Λ and Λ implemented into orthogonal directions (see the unit cell in Fig. 4(a), etching width: 490nm, etching depth: h = 20nm, silicon nitride layer of thickness d=140nm). The grating period Λ in the x-direction is adjusted to create a resonance at FH wavelength for z-polarized FH light exciting the structure under normal incidence (red line in Fig. 4(b)). In contrast grating periods Λ in the z-direction is chosen such that a grating resonance is created at SH wavelength for x-polarized light (blue line in Fig. 4(b),). The TMDC sheet on top is oriented such that its armchair direction has a 30-degree angle with FH z-polarized excitation. For that orientation of the TMDC, a z-polarized FH driving field will generate an SH with x-polarization, both fields being resonant for their specific propagation directions.

 figure: Fig. 4.

Fig. 4. (a) Schematic of a unit cell of asymmetric 2D PC with periods of Λ= 540 and Λ’ = 240 nm, etching widths of w = 490 nm, and w’ = 200 nm (red and blue arrows indicate FH and SH polarization, respectively). (b) Reflection spectra of asymmetric 2D PCs under normal incident.

Download Full Size | PDF

For comparison, we study a symmetric 2D PC with a square lattice the period of which is adjusted such that it resonates at FH wavelength (period 540nm). It has a resonance at 880nm, but no resonance at SH wavelength.

Again, the asymmetric double resonant structure performs 170 times better than the symmetric one with the square pattering (Fig. 5).

 figure: Fig. 5.

Fig. 5. Normalized SH intensity from MoS2 at the SH wavelength.

Download Full Size | PDF

Please note that SH generation efficiency in this structure is critically determined by the etching depth and width. The shallower the etching, the higher the quality factor and the narrower the line width of respective resonances. In principle, even much higher q-factors can be obtained, but technical limitations of manufacturing, as well as the linewidth of the exciting laser, will set the limits for this kind of structure. But, for the assumed asymmetric 2D configuration SH generation is again about 60 times more efficient than that in the already optimized asymmetric 1D PC discussed first. This underlines the great potential offered by the second available transverse dimension.

3. Conclusion

To conclude, coupling a TMDC monolayer with a metasurface can greatly enhance SH generation, particularly if the linear structure shows resonant behavior at both FH and SH wavelengths. We have demonstrated two approaches to obtain that goal: First, we combined two 1D gratings thus breaking the mirror symmetry within the unit cell. In a second approach, we employed the second dimension of a metasurface and the polarization sensitivity of SHG in TMDCs. By using a rectangular lattice with dissimilar base lengths, we designed an optical structure that is resonant at FH and SH frequencies to enhance SH generation compared with structures that support only FH or SH resonances. The two optimization schemes demonstrated in this paper can likewise be applied to other excitation frequencies and even to other frequency conversion processes, as long as one aims at resonances at all interacting frequencies e.g. in spontaneous parametric down-conversion.

Funding

Deutsche Forschungsgemeinschaft.

Acknowledgments

We acknowledge the financial support via project A2 in the CRC 1375 “NOA-Nonlinear optics down to atomic scales” – project number: 398 816 777 funded by the Deutsche Forschungsgemeinschaft (DFG).

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper may be obtained from the authors upon reasonable request.

References

1. L. Yilei, Y. Rao, K. F. mak, et al., “Probing symmetry properties of few-layer MoS2 and h-BN by optical second-harmonic generation,” Nano Lett. 13(7), 3329–3333 (2013). [CrossRef]  

2. N. Kumar, S. Najmaei, Q. Cui, et al., “Second harmonic microscopy of monolayer MoS2,” Phys. Rev. B 87(16), 161403 (2013). [CrossRef]  

3. X. Yin, Z. Ye, D. A. Chenet, et al., “Edge nonlinear optics on a MoS2 atomic monolayer,” Science 344(6183), 488–490 (2014). [CrossRef]  

4. E. Mishina, N. Sherstyuk, S. Lavrov, et al., “Observation of two polytypes of MoS2 ultrathin layers studied by second harmonic generation microscopy and photoluminescence,” Appl. Phys. Lett. 106(13), 131901 (2015). [CrossRef]  

5. Z. Hao, J. Wang, S. Ma, et al., “Sum-frequency generation in on-chip lithium niobate microdisk resonators,” Photonics Res. 5(6), 623–628 (2017). [CrossRef]  

6. X. Ye, S. Liu, Y. Chen, et al., “Sum-frequency generation in lithium-niobate-on-insulator microdisk via modal phase matching,” Opt. Lett. 45(2), 523–526 (2020). [CrossRef]  

7. J. Chen, Y. M. Sua, Z. Ma, et al., “Efficient parametric frequency conversion in lithium niobate nanophotonic chips,” OSA Continuum 2(10), 2914–2924 (2019). [CrossRef]  

8. A. Vaziri, G. Weihs, and A. Zeilinger, “Experimental two-photon, three-dimensional entanglement for quantum communication,” Phys. Rev. Lett. 89(24), 240401 (2002). [CrossRef]  

9. S. Tanzilli, W. Tittel, M. Halder, et al., “A photonic quantum information interface,” Nature 437(7055), 116–120 (2005). [CrossRef]  

10. S. Zaske, A. Lenhard, C. A. Keßler, et al., “Visible-to-telecom quantum frequency conversion of light from a single quantum emitter,” Phys. Rev. Lett. 109(14), 147404 (2012). [CrossRef]  

11. N. Youngblood and M. Li, “Integration of 2D materials on a silicon photonics platform for optoelectronics applications,” Nanophotonics 6(6), 1205–1218 (2016). [CrossRef]  

12. Z. Sun, A. Martinez, and F. Wang, “Optical modulators with 2D layered materials,” Nat. Photonics 10(4), 227–238 (2016). [CrossRef]  

13. A. Säynätjoki, L. Karvonen, H. Rostami, et al., “Ultra-strong non-linear optical processes and trigonal warping in MoS2 layers,” Nat. Commun. 8(1), 893 (2017). [CrossRef]  

14. L. M. Malard, T. V. Alencar, A. Paula, et al., “Observation of intense second harmonic generation from MoS2 atomic crystals,” Phys. Rev. B 87(20), 201401 (2013). [CrossRef]  

15. M. L. Trolle, Y. Tsao, K. Pedersen, et al., “Observation of excitonic resonances in the second harmonic spectrum of MoS2,” Phys. Rev. B 92(16), 161409 (2015). [CrossRef]  

16. J. A. Armstrong, N. Bloembergen, J. Ducuing, et al., “Interactions between Light Waves in a Nonlinear Dielectric,” Phys. Rev. 127(6), 1918–1939 (1962). [CrossRef]  

17. A. Ashkin, G. Boyd, and J. Dziedzic, “Resonant optical second harmonic generation and mixing,” IEEE J. Quant. Electron. 2(6), 109–124 (1966). [CrossRef]  

18. J. K. Day, M. H. Chung, Y. H. Lee, et al., “Microcavity enhanced second harmonic generation in 2D MoS2,” Opt. Mater. Express 6(7), 2360–2365 (2016). [CrossRef]  

19. F. Yi, M. Ren, J. Reed, et al., “Optomechanical Enhancement of Doubly Resonant 2D Optical Nonlinearity,” Nano Lett. 16(3), 1631–1636 (2016). [CrossRef]  

20. H. Chen, V. Corboliou, A. S. Solntsev, et al., “Enhanced Second-Harmonic Generation from Two-Dimensional MoSe2 on a Silicon Waveguide,” Light: Sci. Appl. 6(10), e17060 (2017). [CrossRef]  

21. J. Chen, J. Tan, G. Wu, et al., “Tunable and Enhanced Light Emission in Hybrid Ws2-Optical-FiberNanowire Structures,” Light: Sci. Appl. 8(1), 8 (2019). [CrossRef]  

22. J. Shi, W. Liang, S. S. Raja, et al., “Plasmonic Enhancement and Manipulation of Optical Nonlinearity in Monolayer Tungsten Disulfide,” Laser Photonics Rev. 12(10), 1800188 (2018). [CrossRef]  

23. J. Chen, K. Wang, H. Long, et al., “Tungsten Disulfide−Gold Nanohole Hybrid Metasurfaces for Nonlinear Metalenses in the Visible Region,” Nano Lett. 18(2), 1344–1350 (2018). [CrossRef]  

24. Z. Wang, Z. Dong, H. Zhu, et al., “Selectively Plasmon Enhanced Second-Harmonic Generation from Monolayer Tungsten Diselenide on Flexible Substrates,” ACS Nano 12(2), 1859–1867 (2018). [CrossRef]  

25. J. Franz, F. Löchner, A. George, et al., “Hybrid Dielectric Metasurfaces for Enhancing Second-Harmonic Generation in Chemical Vapor Deposition Grown MoS2 Monolayers,” ACS Photonics 8(1), 218–227 (2021). [CrossRef]  

26. R. Mupparapu, T. Bucher, and I. Staude, “Integration of two-dimensional transition metal dichalcogenides with Mie-resonant dielectric nanostructures,” Adv. Phys.: X 5(1), 1734083 (2020). [CrossRef]  

27. Z. Zhang, L. Zhang, R. Gogna, et al., “Large enhancement of second-harmonic generation in MoS2 by one dimensional photonic crystals,” Solid State Commun. 322, 114043 (2020). [CrossRef]  

28. Y. Xie, L. Yang, J. Du, et al., “Giant enhancement of second-harmonic generation in hybrid metasurface coupled MoS2 with Fano-resonance effect,” Nanoscale Res. Lett. 17(1), 97 (2022). [CrossRef]  

29. S. Shradha, F. Abtahi, Z. Gan, et al., “Towards Double Resonant Cavity Enhanced Second Harmonic Generation in Monolayer MoS2,” Adv. Opt. Mater. 1, 2300907 (2023). [CrossRef]  

30. J. You, E. Threlfall, D. Gallagher, et al., “Computational analysis of dispersive and nonlinear 2D materials by using a GS-FDTD method,” J. Opt. Soc. Am. B 35(11), 2754–2763 (2018). [CrossRef]  

31. Y. Dai, Y. Wang, S. Das, et al., “Electrical control of interband resonant nonlinear optics in monolayer MoS2,” ACS Nano 14(7), 8442–8448 (2020). [CrossRef]  

32. K. Luke, Y. Okawachi, M. R. E. Lamont, et al., “Broadband mid-infrared frequency comb generation in a Si3N4 microresonator,” Opt. Lett. 40(21), 4823–4826 (2015). [CrossRef]  

33. Y. Arosa and R. de la Fuente, “Refractive index spectroscopy and material dispersion in fused silica glass,” Opt. Lett. 45(15), 4268–4271 (2020). [CrossRef]  

Data availability

Data underlying the results presented in this paper may be obtained from the authors upon reasonable request.

Cited By

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

Alert me when this article is cited.


Figures (5)

Fig. 1.
Fig. 1. 1D PCs with a MoS2 monolayer placed on top (a) principal sketch and (b-d) construction of the unit cell of a double resonant grating by combining a symmetric 1D PCs (b) with a period of Λ to couple the FH light into the guiding layer and one with a shorter period Λ’ (c) to couple the SH light. Thus, a unit cell of an asymmetric double periodic grating (d) with a period of Λ is formed. (e) Linear reflection spectra of symmetric and asymmetric 1D PCs shown in (b-d) for normal incident and polarization parallel to the grating bars.
Fig. 2.
Fig. 2. Mode profiles of the asymmetric 1D PC at selected resonances (red line in Fig. 1(e)) via first-order diffraction (a) at FH wavelength, (b) at 800 nm, and (c) second-order diffraction at SH wavelength.
Fig. 3.
Fig. 3. Normalized SH intensity from MoS2 for varying wavelengths of the FH driving field. The electric field is polarized parallel to the bars and the TMDC sheet is assumed to be oriented such that its armchair axis is parallel to the bars.
Fig. 4.
Fig. 4. (a) Schematic of a unit cell of asymmetric 2D PC with periods of Λ= 540 and Λ’ = 240 nm, etching widths of w = 490 nm, and w’ = 200 nm (red and blue arrows indicate FH and SH polarization, respectively). (b) Reflection spectra of asymmetric 2D PCs under normal incident.
Fig. 5.
Fig. 5. Normalized SH intensity from MoS2 at the SH wavelength.
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.