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

High-speed mid-infrared graphene electro-optical modulator based on suspended germanium slot waveguides

Open Access Open Access

Abstract

The mid-infrared (MIR) region is attracting increasing interest for on-chip synchronous detection and free-space optical (FSO) communications. For such applications, a high-performance electro-optical modulator is a crucial component. In this regard, we propose and investigate a graphene-based electro-absorption modulator (EAM) and microring modulator (MRM) using the suspended germanium waveguide platform. The modulators are designed for the second atmospheric window (8 to 12 µm). The incorporation of double-layer graphene on the suspended slot waveguide structure allows for the significant enhancement of light-graphene interaction, theoretically achieving a 3-dB bandwidth as high as 78 GHz. The EAM shows a calculated modulation depth of 0.022-0.045 dB/µm for the whole operation wavelength range. The MRM exhibits a calculated extinction ratio as high as 68.9 dB and a modulation efficiency of 0.59 V·cm around 9 µm. These modulators hold promise for constructing high-speed FSO communication and on-chip spectroscopic detection systems in the MIR atmospheric window.

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

1. Introduction

The mid-infrared (MIR) region, typically defined as the wavelength range of 2 to 20 µm, covers the molecular fingerprint of many molecules, making it a useful tool for molecular absorption spectroscopy in civilian and military applications. This region also includes two crucial atmospheric transmission windows at 3-5 µm and 8-12 µm, which make it an ideal choice for both terrestrial and space-based applications. Thus, the MIR region has great potential for chemical or biological sensing [1], astronomy [2], industrial processes control [3], thermal imaging [4], and free-space optical (FSO) communications [5]. To develop MIR photonics applications, a suitable waveguide platform is needed. Silicon photonics has gained significant attention for its potential in creating integrated devices at the near-infrared (NIR) telecom band [610], especially for optical interconnects and communication systems. However, the use of silicon-on-insulator (SOI) material platform for longer wavelengths (beyond 3.8 µm) is restricted due to the high absorption of silicon dioxide [1112]. Fortunately, germanium, also a group IV material, offers an intrinsic transparent window (2-14 µm). As silicon starts to absorb light from 8 µm, germanium is the preferred material for integrated platforms particular for longer wavelengths up to 14 µm [13]. It has similar properties to silicon, such as wide transparency ranges, high refractive indexes, and mechanical robustness [14]. Moreover, germanium-based integrated devices can be fabricated using the mature CMOS fabrication technology.

Germanium-based passive integrated devices have been successfully demonstrated experimentally in the MIR region, such as grating couplers [15], resonators [16], arrayed waveguide gratings [17], crystal cavities [18], and polarization rotators [19]. However, active devices, which are crucial for the development of MIR integrated systems, such as high-speed light modulation and detection, are less explored. For example, high-performance modulators could offer significant potential for FSO transmission applications and spectroscopic detection systems. Despite its potential, MIR light modulation remains largely unexploited. In recent years, high-speed optical modulators have been demonstrated using external free-space Stark modulation [20] or directly-modulated quantum cascade lasers [21]. However, these modulators are not easily scalable due to their complicated fabrication processes, which rely on quantum wells. Furthermore, the electro-absorption modulator (EAM) based on the germanium-on-silicon (Ge-on-Si) platform has achieved modulation up to 8 µm [22], but its modulation bandwidth is limited to only 60 MHz due to unoptimized Ohmic contacts between doping areas and metal electrodes. Therefore, there is a pressing need to explore new effects that can facilitate the development of high-speed, integrated MIR modulators.

In the meantime, 2D materials, including graphene, hexagonal boron nitride (hBN), transition metal dichalcogenides (TMDs), and black phosphorus (BP), have emerged as promising materials for optoelectronic components in integrated photonics [2326]. These materials can change the complex refractive index when an electric field is applied, making them useful for integrated light modulation applications [2731]. Graphene, in particular, has been extensively studied for light modulation applications due to its strong interaction with light, gate-variable in-plane optical conductivity [32], high carrier mobility, and ultra-broadband compatibility [33]. It can be easily integrated on top of various photonic platforms, such as silicon, germanium, or other dielectric waveguides, to realize efficient light modulation. Although using graphene modulators based on group IV integrated platforms have been demonstrated experimentally over the last decade [3439], waveguide-integrated graphene modulators at the MIR region have yet to be explored. Therefore, this study aims to exploit a high-speed germanium-based graphene modulator for the MIR atmospheric window.

In this work, we propose and numerically investigate double-layer graphene (DLG) EAM and a microring modulator (MRM) based on suspended germanium slot waveguides. The suspended slot waveguide structure is utilized to minimize cladding material absorption and increase light interaction, which is expected to improve the modulation bandwidth as the overlap length of DLG is reduced. The DLG EAM is designed to have a wide operation window that covers the entire second atmospheric window, with a modulation depth ranging from 0.022-0.045 dB/µm. On the other hand, the MRM is designed to have a very high extinction ratio (ER) of approximately 68.9 dB and a modulation efficiency of 0.59 V·cm at around 9 µm. The calculated 3-dB bandwidth of these modulators using the equivalent resistor-capacitor (RC) circuit model is approximately 78 GHz, which indicates their potential for high-speed modulation applications.

2. Device design

Figure 1(a) schematically illustrates the proposed design for a high-speed DLG EAM based on a suspended germanium slot waveguide. The cross-section of the device is depicted in Fig. 1(b). The modulator is based on the Ge-on-Si platform. A dielectric layer of zinc sulfide (ZnS), which is transparent in the 1-14 µm range [40] is used as the top layer. Subsequent to the deposition and planarization of the ZnS layer, the silicon layer is removed using a tetramethylammonium hydroxide (TMAH) solution through etched holes. This step is crucial for eliminating the significant absorption of silicon in the second atmospheric window. The germanium layer has a thickness of 500 nm, and the entire device is etched to a depth of 400 nm to ensure efficient mode overlap with the DLG overlayer. The thickness of ZnS between the DLG is selected to tZnS = 30 nm. To maximize the evanescent coupling of the optical mode with the DLG, the thickness of the waveguide top cladding is te = 5 nm.

 figure: Fig. 1.

Fig. 1. Schematic of the proposed EAM. (a) The three-dimensional structure. (b) The cross-section.

Download Full Size | PDF

Figure 2(a)-(c) show the simulated mode profiles at graphene Fermi level EF (also chemical potential µ) = 0 eV in the Ge-on-Si rib, suspended germanium rib, and suspended germanium slot waveguide. Figure 2(d) shows the absorption of graphene in a suspended germanium rib waveguide and Ge-on-Si rib waveguide when the germanium layer height (hGe) is 1 µm and the etch depth (hetch) is 0.6 µm at a wavelength of 9 µm, calculated using the graphene Kubo formula [41]. As the energy distribution of the Ge-on-Si rib waveguide is leaked to the silicon substrate due to the relatively low index contrast between the germanium layer and silicon substrate, the mode overlap of the evanescent field with the upper graphene is weaker, resulting in lower light absorption [42]. Thus, suspended germanium waveguides and thinner germanium layers are employed to improve the light-graphene interaction. Figure 2(e) depicts the absorption of graphene in a suspended germanium slot and rib waveguide when the germanium layer height is 0.5 µm and the etch depth is 0.1 µm at the wavelength of 9 µm. The electric field is tightly confined in the slot region, enabling stronger interaction of graphene with the propagating fundamental TE mode in slot waveguides than in rib waveguides. Additionally, rib waveguides require a larger DLG overlap width (the width of the graphene-ZnS-graphene capacitor) than slot waveguides, resulting in limited electro-optical (EO) 3-dB bandwidth with increasing capacitance. Therefore, suspended germanium slot waveguides are used in our device.

 figure: Fig. 2.

Fig. 2. (a)-(c) The electric field magnitude |E| distributions of the fundamental TE mode at graphene Fermi level EF = 0 eV in the Ge-on-Si rib, suspended germanium rib, and suspended germanium slot waveguide, respectively. (d) The graphene absorption with different Fermi levels in (a) and (b). (e) The graphene absorption with different Fermi levels in (b) and (c).

Download Full Size | PDF

To optimize the modulation performance, different rib widths wrib and slot widths wslot are considered. Figure 3(a) shows the calculated absorption of the DLG on the suspended germanium slot waveguide with varying slot widths as a function of rib width at the graphene Fermi level EF of 0 eV and the wavelength of 9 µm. The figure suggests that wrib = 2.8 µm is the optimal width with maximum absorption. The calculated absorption and group index of the EAM as a function of slot width when the rib width is 2.8 µm are displayed in Fig. 3(b), showing that the interaction area reduces with the increasing slot width. While decreasing it will enhance the interaction strength, which in turn reduces the group index, leading to an increase in group velocity and resulting in a decline in interaction time, thus weakening the interaction between the mode field and graphene. Therefore, there is a trade-off between the enhancement of interaction strength and the reduction of interaction area, and we choose wslot as 200 nm to achieve the maximum absorption. Considering that the width of the DLG overlap (wm) is set to be the same as the slot width, we have optimized the graphene width wgraphene = 6.2 µm to maintain a proper distance between the metal contacts and the slot waveguide. This is important because longer wavelengths lead to larger mode fields, which could cause excess optical losses due to the metal contacts. All the parameters of the proposed EAM are presented in Table 1.

 figure: Fig. 3.

Fig. 3. (a) Absorption of the EAM for different slot widths as a function of rib width. (b) Absorption and group index of the EAM as a function of slot width of a rib width ∼ 2.8 µm.

Download Full Size | PDF

Tables Icon

Table 1. Parameters of the proposed EAM

3. Modulation performance

Figure 4(a) shows the calculated modulation depth of the proposed DLG EAM as a function of the Fermi level of graphene for different wavelengths, using the optimized structure. The modulation depth is defined as the difference value at graphene Fermi level EF = 0 eV and 0.15 eV, while the insertion loss (IL) is defined as the value at graphene Fermi level EF = 0.15 eV. The EAM's operation window covers the entire second atmospheric window (8-12 µm), with modulation depth ranging from 0.022 to 0.045 dB/µm and IL ranging from 0.0085 to 0.0095 dB/µm. For example, a 300-µm-long EAM can achieve 6.6-13.5 dB modulation depth in the 8-12 µm range, with a moderate IL of 2.55-2.85 dB. The calculated IL primarily comprises the absorption contributions from both graphene and the metal electrodes. The metal electrode structure consists of a 5 nm thickness of titanium (Ti) and a 35 nm layer of gold (Au), which introduces an intrinsic loss of approximately 0.0015 dB/µm to the total loss. Furthermore, it is worth noting that in an experimental device, the IL is likely to be higher than the calculated value due to additional losses originating from factors such as sidewall scattering, material defects at the Si/Ge interface, and the presence of a native oxide layer or imperfections on the germanium surface. However, previous measurements have reported losses in the range of 2-5 dB/cm for germanium waveguides operating within the wavelength range of 7-12 µm [4244]. Considering the relatively short length of our device, we anticipate that the experimental loss will have a negligible influence.

 figure: Fig. 4.

Fig. 4. Modulation depth of the EAM as a function of graphene Fermi level with (a) different wavelengths at graphene scattering time = 100 fs (b) different graphene scattering times at wavelength λ = 9 µm.

Download Full Size | PDF

The maximum achievable modulation depth per unit length depends on the mobility of graphene and previous research was based on a scattering time of 100 fs. In Fig. 4(b), the EAM modulation depth at the wavelength of 9 µm is illustrated for scattering times of 50 fs, 100 fs, and 1 ps. For a high mobility graphene (scattering time = 1 ps) with changing Fermi level from 0 to 0.15 eV, a value of 0.046 dB/µm is obtained, while for a low mobility graphene (scattering time = 50 fs) within the same Fermi level range, the value is 0.037 dB/µm.

To improve the modulation depth, an MRM is also investigated for wavelength near 9 µm, and its structure is presented in Fig. 5(a). An all-pass type microring resonator is proposed with a radius of 60 µm that incorporates a 50 µm overlap length of graphene on a section of the ring waveguide. The transmission spectra are characterized by:

$$T\left( \lambda \right)\textrm{ = }{\left| {\frac{{\left( {\sqrt {1\textrm{ - }{\kappa ^2}} \textrm{ - }{\textrm{e}^{ - r + i\phi \left( \lambda \right)}}} \right)}}{{\left( {1 - \sqrt {1\textrm{ - }{\kappa ^2}} \cdot {\textrm{e}^{ - r + i\phi \left( \lambda \right)}}} \right)}}} \right|^2}$$
where κ is the coupling coefficient between the straight waveguide and the ring resonator, and e-r and φ(λ) are the loss and phase delay in the cavity per round, respectively. The loss and phase delay is determined by the imaginary part and real part of the effective index, respectively. Based on the aforementioned analysis, Fig. 5(b) displays the calculated transmission spectra for the MRM designed with a power-coupling coefficient κ2 of 0.0885 at various gate voltages. Figure 5(c) shows the modulation depth as a function of wavelength for different bias voltages in this design, and the IL and ER are estimated to be 0.3 dB and 68.9 dB, respectively, at a bias of 1.8 V. It is important to note that the absorption caused by the metal electrodes is only 0.01 dB in our case. Additionally, if the metal electrodes are positioned away from the ring waveguide, this absorption can be eliminated. The modulation efficiency VπLπ is 0.59 V·cm for Q = 2,413.

 figure: Fig. 5.

Fig. 5. (a) Schematic of the proposed MRM. (b) Transmission spectra of the MRM with different applied voltages. (c) Modulation depth of the MRM with different applied voltages.

Download Full Size | PDF

The incident light is modulated by changing the applied voltage V related to the graphene Fermi level EF, according to the following formula:

$${E_F} = {\mathop{\rm sgn}} (V )\hbar {v_F}\sqrt {\frac{{\pi {\varepsilon _0}{\varepsilon _{ZnS}}|V |}}{{{t_{ZnS}}e}}} $$
where ${\hbar}$ is the reduced Planck constant, νF is the Fermi velocity of 106 m/s, e is the elementary charge, ε0 and εZnS are the permittivity in the vacuum and ZnS, respectively. Figure 6(a) shows the correspondence between applied voltage and graphene Fermi level. These modulators exhibit very low drive voltage as low as 1.8 V as they work at the graphene Fermi level of 0-0.15 eV.

 figure: Fig. 6.

Fig. 6. (a) Drive voltage of the proposed modulators versus the graphene Fermi level. (b) The equivalent RC circuit of the proposed modulators.

Download Full Size | PDF

The equivalent RC circuit of the EAM is shown in Fig. 6(b), and the EO 3-dB bandwidth of the modulator can be calculated by [45,46]:

$$BW = \frac{1}{{2\pi }}[{({{R_C} + {R_S}} )\cdot {C_{graphene}}} ]$$
where RC is the contact resistance between graphene and metal, RS is the sheet resistance of graphene, and Cgraphene is the capacitance of the graphene-ZnS-graphene structure. Among them, RC is given by:
$${R_C} = 2{R_{C1}} = \frac{{2{\rho _C}}}{L}$$
where contact resistivity ρc is selected as ρc = 500 Ω·µm and L is the length of the modulator. And RS is given by:
$${R_S} = 2{R_{S1}} = \frac{{2{R_\square } \cdot \left( {{w_{graphene}} - {w_m}} \right)}}{L}$$
where sheet resistance R is selected as R = 500 Ω/□. Besides, the Cgraphene can be calculated by:
$${C_{graphene}} = {\varepsilon _0}{\varepsilon _{ZnS}}\frac{{{w_m}L}}{{{t_{ZnS}}}}$$

Consequently, the calculated 3-dB bandwidths of the EAM and the MRM are as high as 78 GHz by subtitling RC, RS and Cgraphene into Eq. (2), which exhibit a splendid dynamic characteristic. We acknowledge that the experimental bandwidth of the proposed modulators may be slightly lower than the calculated value due to the load impedance [45].

4. Discussion

Figure 7 illustrates the dependence of graphene's complex relative permittivity on the Fermi level for various wavelengths, using the graphene Kubo formula. When the graphene Fermi level EF is lower than half the photon energy of incident light µc (about 0.40 eV, 0.14 eV, 0.08 eV, 0.05 eV at 1.55 µm, 4.5 µm, 8 µm, and 12 µm, respectively), graphene exhibits significant absorption, and the absorption rapidly decreases if the EF is larger than half the µc. At 9 µm, the graphene absorption reduces with increasing graphene Fermi level from 0-0.15 eV. This reduction in absorption could allow effective light modulation. Conversely, the absorption of DLG magnifies when graphene Fermi level exceeds 0.18 eV (as shown in Fig. 2(e)), this is because the surface plasmons polaritons (SPP) of graphene is excited on the waveguides. Therefore, our graphene modulators operate within the Fermi level range of 0-0.15 eV, eliminating the presence of graphene SPP. Consequently, these modulators require an ultra-low drive voltage of 1.8 V.

 figure: Fig. 7.

Fig. 7. Complex relative permittivity (real part and imaginary part) of graphene versus Fermi level with different wavelengths.

Download Full Size | PDF

Table 2 compares the latest integrated MIR modulators designed for the 8-12 µm range. MIR light modulation was achieved based on Ge-on-Si and Graded SiGe waveguide using the free-carrier plasma dispersion effect [22,48]. Until now, very few broadband MIR modulators have been demonstrated, except for the recent report of the Stark EAM with a bandwidth of up to 9 GHz [5]. However, these quantum wells-based Stark modulators require relatively complex fabrication processes that are impractical for large-scale integration. In theory, it is possible to achieve MIR modulators with very large 3-dB bandwidths [47,49,51]. The EAMs relying on the free-carrier plasma dispersion effect necessitate long lengths (typically a few millimeters) to achieve a modulation depth higher than 1 dB, resulting in significant IL [50,52]. Additionally, the manufacture of graphene hybrid plasmonic EAMs, as described in Ref. [47] and [49], requires precise dimensional parameters and extremely high-quality graphene. In contrast, our proposed EAM and MRM could provide high modulation depth, low IL, and large 3-dB bandwidth, and more importantly, they can be manufactured by well-established CMOS fabrication technology and thus are cost-efficient.

Tables Icon

Table 2. Performance comparison of on-chip integrated MIR electro-optic modulators operating at the second atmospheric window (8-12 µm)

5. Conclusion

To summarize, suspended germanium slot waveguide-based MIR modulators (EAM and MRM) using DLG for the second atmospheric window are investigated. Suspended slot waveguides are introduced to reduce propagated loss, strengthen the graphene absorption, and increase the modulation bandwidth. The simulations show that the modulation depth of the EAM is 0.022-0.045 dB/µm for the 8-12 µm wavelength region. And the calculated results of the MRM show a very high extinction ratio of ∼ 68.9 dB, and modulation efficiency of ∼ 0.59 V·cm at around 9 µm. The EO 3-dB bandwidth of the modulators is up to 78 GHz. It is believed that the proposed modulators can promote the development of spectroscopic detection systems and FSO communications in the MIR atmospheric window.

Appendix 1: Schematic illustration of the proposed fabrication processes

The schematic illustration of the fabrication processes for our proposed devices is shown in Fig. 8. First, a germanium slot waveguide is fabricated on the Ge-on-Si wafer via electron beam lithography (EBL) and inductively coupled plasma (ICP) etching processes (Figs. 8 (a) and (b)). Then, the ZnS film is deposited on the chip using electron beam evaporation (EBE) (Fig. 8 (c)) [53]. After the ZnS film deposition, a flat surface of the chip can be achieved using chemical mechanical polishing (CMP) to ensure a good attachment of graphene The Ge and ZnS layers are then etched using ICP to form wet etched holes (Fig. 8 (d)) [54], after which the silicon substrate is removed using TMAH solution through the aforementioned holes (Fig. 8 (e)). Subsequently, the bottom layer graphene is transferred onto the chip's flat surface using a wet transfer method (Fig. 8 (f)), and then patterned via EBL and oxygen ion etching processes [55]. An electrode that establishes contact with the bottom layer graphene is created using a lift-off method (Fig. 8 (g)). Subsequently, a very thin ZnS dielectric layer is deposited using EBE (Fig. 8 (h)), and the ZnS film on the electrode can be removed by wet etching using acid. Finally, the top layer graphene and electrode are fabricated once more using the aforementioned method (Figs. 8 (i) and (j)). Although our design incorporates a suspended structure, the deposition of the ZnS film and CMP process on the germanium slot waveguide enhance the feasibility of fabrication. Additionally, the deposition of a very thin dielectric layer, along with the transferring and patterning of the graphene layer, as well as the lift-off process of the electrodes, have already been successfully demonstrated on suspended waveguides in Ref. [45].

 figure: Fig. 8.

Fig. 8. Fabrication steps for the proposed graphene modulators. (a) Ge-on-Si wafer. (b) germanium slot waveguide. (c) ZnS deposition and planarization. (d) Wet etch holes etching. (e) Si trenches etching. (f) Bottom layer graphene transferring and patterning. (g) Bottom layer graphene contact electrode. (h) ZnS dielectric layer deposition and the ZnS on the electrode surface removing. (i) Top layer graphene transferring and patterning. (j) Top layer graphene contact electrode.

Download Full Size | PDF

Appendix 2: Analysis of the overlay accuracy with the graphene layers

Practically, the overlay accuracy in EBL during the fabrication processes could result in variations to the graphene overlap of the fabricated devices. we have considered six cases of fabrication tolerances related to the overlay accuracy: (a) shifting the bottom layer graphene forward by 50 nm while keeping the top layer graphene unchanged, (b) shifting the bottom layer graphene backward by 50 nm while keeping the top layer graphene unchanged, (c) shifting the top layer graphene forward by 50 nm without altering the bottom layer graphene, (d) shifting the top layer graphene backward by 50 nm without altering the bottom layer graphene, (e) shifting both the top and bottom layer graphene forward by 50 nm, and (f) shifting both the top and bottom layer graphene backward by 50 nm. In Fig. 9, we compare these six cases and observe only slight changes in the modulation depth of the device due to variations in the overlay accuracy. Additionally, in our calculations, the bandwidth of all six cases, dependent on the width of the double-layer graphene overlap (wm), is calculated to > 53 GHz, enabling high-speed modulation in the mid-infrared range. Thus, despite the potential impact of overlay accuracy on device performance, our proposed devices still exhibit good performance.

 figure: Fig. 9.

Fig. 9. Modulation depth of the EAM at wavelength λ = 9 µm with the overlay accuracy consideration. Case a: shifting the bottom layer graphene forward by 50 nm while keeping the top layer graphene unchanged. Case b: shifting the bottom layer graphene backward by 50 nm while keeping the top layer graphene unchanged. Case c: shifting the top layer graphene forward by 50 nm without altering the bottom layer graphene. Case d: shifting the top layer graphene backward by 50 nm without altering the bottom layer graphene. Case e: shifting both the top and bottom layer graphene forward by 50 nm. Case f: shifting both the top and bottom layer graphene backward by 50 nm.

Download Full Size | PDF

Funding

Key Technologies Research and Development Program (2022YFB2803600); National Natural Science Foundation of China (62175080); Key Research and Development Program of Hubei Province (2021BAA005).

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. B. Schwarz, P. Reininger, D. Ristanić, H. Detz, A. M. Andrews, W. Schrenk, and G. Strasser, “Monolithically integrated mid-infrared lab-on-a-chip using plasmonics and quantum cascade structures,” Nat. Commun. 5(1), 4085 (2014). [CrossRef]  

2. L. Labadie and O. Wallner, “Mid-infrared guided optics: a perspective for astronomical instruments,” Opt. Express 17(3), 1947–1962 (2009). [CrossRef]  

3. G. Reich, “Near-infrared spectroscopy and imaging: basic principles and pharmaceutical applications,” Adv. Drug Delivery Rev. 57(8), 1109–1143 (2005). [CrossRef]  

4. D. Zhang, C. Li, C. Zhang, M. N. Slipchenko, G. Eakins, and J.-X. Cheng, “Depth-resolved mid-infrared photothermal imaging of living cells and organisms with submicrometer spatial resolution,” Sci. Adv. 2(9), 1600521 (2016). [CrossRef]  

5. P. Didier, H. Dely, T. Bonazzi, O. Spitz, E. Awwad, E. Rodriguez, A. Vasanelli, C. Sirtori, and F. Grillot, “High-capacity free-space optical link in the midinfrared thermal atmospheric windows using unipolar quantum devices,” Adv,” Photonics 4(5), 056004 (2022). [CrossRef]  

6. G. T. Reed, G. Mashanovich, F. Y. Gardes, and D. J. Thomson, “Silicon optical modulators,” Nat. Photonics 4(8), 518–526 (2010). [CrossRef]  

7. Z. Zhou, R. Chen, X. Li, and T. Li, “Development trends in silicon photonics for data centers,” Opt. Fiber Technol. 44, 13–23 (2018). [CrossRef]  

8. P. Dong, Y. K. Chen, G. H. Duan, and D. T. Neilson, “Silicon photonic devices and integrated circuits,” Nanophotonics 3(4-5), 215–228 (2014). [CrossRef]  

9. T. Li, J. Zhang, H. Yi, W. Tan, Q. Long, Z. Zhou, X. Wang, and H. Wu, “Low-voltage, high speed, compact silicon modulator for BPSK modulation,” Opt. Express 21(20), 23410–23415 (2013). [CrossRef]  

10. T. Li, D. Wang, J. Zhang, Z. Zhou, F. Zhang, X. Wang, and H. Wu, “Demonstration of 6.25 Gbaud advanced modulation formats with subcarrier multiplexed technique on silicon Mach-Zehnder modulator,” Opt. Express 22(16), 19818–19823 (2014). [CrossRef]  

11. G. Z. Mashanovich, F. Y. Gardes, D. J. Thomson, Y. Hu, L. Ke, M. Nedeljkovic, J. Soler Penades, A. Z. Khokhar, C. J. Mitchell, S. Stankovic, R. Topley, S. A. Reynolds, W. Yun, B. Troia, V. M. N. Passaro, C. G. Littlejohns, T. Dominguez Bucio, P. R. Wilson, and G. T. Reed, “Silicon Photonic Waveguides and Devices for Near- and Mid-IR Applications,” IEEE J. Select. Topics Quantum Electron. 21(4), 407–418 (2015). [CrossRef]  

12. M. Nedeljkovic, R. Soref, and G. Z. Mashanovich, “Free-Carrier Electrorefraction and Electroabsorption Modulation Predictions for Silicon Over the 1-14 µm Infrared Wavelength Range,” IEEE Photonics J. 3(6), 1171–1180 (2011). [CrossRef]  

13. M. Nedeljkovic, R. Soref, and G. Z. Mashanovich, “Predictions of Free-Carrier Electroabsorption and Electrorefraction in Germanium,” IEEE Photonics J. 7(3), 1–14 (2015). [CrossRef]  

14. R. Soref, “Mid-infrared photonics in silicon and germanium,” Nat. Photonics 4(8), 495–497 (2010). [CrossRef]  

15. C. Alonso-Ramos, M. Nedeljkovic, D. Benedikovic, J. S. Penadés, C. G. Littlejohns, A. Z. Khokhar, D. Pérez-Galacho, L. Vivien, P. Cheben, and G. Z. Mashanovich, “Germanium-on-silicon mid-infrared grating couplers with low-reflectivity inverse taper excitation,” Opt. Lett. 41(18), 4324–4327 (2016). [CrossRef]  

16. T.-H. Xiao, Z. Zhao, W. Zhou, C.-Y. Chang, S. Y. Set, M. Takenaka, H. K. Tsang, Z. Cheng, and K. Goda, “Mid-infrared high-Q germanium microring resonator,” Opt. Lett. 43(12), 2885–2888 (2018). [CrossRef]  

17. A. Malik, M. Muneeb, S. Pathak, Y. Shimura, J. V. Campenhout, R. Loo, and G. Roelkens, “Germanium-on-Silicon Mid-Infrared Arrayed Waveguide Grating Multiplexers,” IEEE Photon. Technol. Lett. 25(18), 1805–1808 (2013). [CrossRef]  

18. T.-H. Xiao, Z. Zhao, W. Zhou, M. Takenaka, H. K. Tsang, Z. Cheng, and K. Goda, “Mid-infrared germanium photonic crystal cavity,” Opt. Lett. 42(15), 2882–2885 (2017). [CrossRef]  

19. K. Gallacher, R. W. Millar, U. Griškevičiūtė, M. Sinclair, M. Sorel, L. Baldassarre, M. Ortolani, R. Soref, and D. J. Paul, “Ultra-broadband mid-infrared Ge-on-Si waveguide polarization rotator,” APL Photonics 5(2), 026102 (2020). [CrossRef]  

20. H. Dely, T. Bonazzi, O. Spitz, E. Rodriguez, D. Gacemi, Y. Todorov, K. Pantzas, G. Beaudoin, I. Sagnes, L. Li, A. G. Davies, E. H. Linfield, F. Grillot, A. Vasanelli, and C. Sirtori, “10 Gbit s−1 Free Space Data Transmission at 9 µm Wavelength With Unipolar Quantum Optoelectronics,” Laser Photonics Rev. 16(2), 2100414 (2022). [CrossRef]  

21. M. Joharifar, H. Dely, X. Pang, R. Schatz, D. Gacemi, T. Salgals, A. Udalcovs, Y. Sun, Y. Fan, L. Zhang, E. Rodriguez, S. Spolitis, V. Bobrovs, X. Yu, S. Lourdudoss, S. Popov, A. Vasanelli, O. Ozolins, and C. Sirtori, “High-Speed 9.6-µm Long-Wave Infrared Free-Space Transmission With a Directly-Modulated QCL and a Fully-Passive QCD,” J. Lightwave Technol. 41(4), 1087–1094 (2023). [CrossRef]  

22. T. Li, M. Nedeljkovic, N. Hattasan, W. Cao, Z. Qu, C. G. Littlejohns, J. S. Penades, L. Mastronardi, V. Mittal, D. Benedikovic, D. J. Thomson, F. Y. Gardes, H. Wu, Z. Zhou, and G. Z. Mashanovich, “Ge-on-Si modulators operating at mid-infrared wavelengths up to 8 µm,” Photonics Res. 7(8), 828–836 (2019). [CrossRef]  

23. Y. Zhou, M. Zhang, Z. Guo, L. Miao, S. T. Han, Z. Wang, X. Zhang, H. Zhang, and Z. Peng, “Recent advances in black phosphorus-based photonics, electronics, sensors and energy devices,” Mater. Horiz. 4(6), 997–1019 (2017). [CrossRef]  

24. K. Kim, J.-Y. Choi, T. Kim, S.-H. Cho, and H.-J. Chung, “A role for graphene in silicon-based semiconductor devices,” Nature 479(7373), 338–344 (2011). [CrossRef]  

25. K. S. Novoselov, V. I. Falko, L. Colombo, P. R. Gellert, M. G. Schwab, and K. Kim, “A roadmap for graphene,” Nature 490(7419), 192–200 (2012). [CrossRef]  

26. F. Xia, H. Wang, D. Xiao, M. Dubey, and A. Ramasubramaniam, “Two-dimensional material nanophotonics,” Nat. Photonics 8(12), 899–907 (2014). [CrossRef]  

27. H. Agarwal, B. Terrés, L. Orsini, A. Montanaro, V. Sorianello, M. a. Pantouvaki, K. Watanabe, T. Taniguchi, D. V. Thourhout, M. Romagnoli, and F. H. L. Koppens, “2D-3D integration of hexagonal boron nitride and a high-κ dielectric for ultrafast graphene-based electro-absorption modulators,” Nat. Commun. 12(1), 1070 (2021). [CrossRef]  

28. I. Datta, S. H. Chae, G. R. Bhatt, M. A. Tadayon, B. Li, Y. Yu, C. Park, J. Park, L. Cao, D. N. Basov, J. Hone, and M. Lipson, “Low-loss composite photonic platform based on 2D semiconductor monolayers,” Nat. Photonics 14(4), 256–262 (2020). [CrossRef]  

29. Z. Zhao, Z. Zhang, J. Li, Z. Shang, G. Wang, J. Yin, H. Chen, K. Guo, and P. Yan, “MoS2 hybrid integrated micro-ring resonator phase shifter based on a silicon nitride platform,” Opt. Lett. 47(4), 949–952 (2022). [CrossRef]  

30. L. Huang, B. Dong, Z. G. Yu, J. Zhou, Y. Ma, Y.-W. Zhang, C. Lee, and K.-W. Ang, “Mid-infrared modulators integrating silicon and black phosphorus photonics,” Mater. Today Adv. 12, 100170 (2021). [CrossRef]  

31. H. Lin, Y. Song, Y. Huang, et al., “Chalcogenide glass-on-graphene photonics,” Nat. Photonics 11(12), 798–805 (2017). [CrossRef]  

32. F. Wang, Y. Zhang, C. Tian, C. Girit, A. Zettl, M. Crommie, and Y. R. Shen, “Gate-variable optical transitions in graphene,” Science 320(5873), 206–209 (2008). [CrossRef]  

33. K. F. Mak, L. Ju, F. Wang, and T. F. Heinz, “Optical spectroscopy of graphene: From the far infrared to the ultraviolet,” Solid State Commun. 152(15), 1341–1349 (2012). [CrossRef]  

34. M. Liu, X. Yin, E. Ulin-Avila, B. Geng, T. Zentgraf, L. Ju, F. Wang, and X. Zhang, “A graphene-based broadband optical modulator,” Nature 474(7349), 64–67 (2011). [CrossRef]  

35. M. Liu, X. Yin, and X. Zhang, “Double-layer graphene optical modulator,” Nano Lett. 12(3), 1482–1485 (2012). [CrossRef]  

36. C. T. Phare, Y.-H. Daniel Lee, J. Cardenas, and M. Lipson, “Graphene electro-optic modulator with 30 GHz bandwidth,” Nat. Photonics 9(8), 511–514 (2015). [CrossRef]  

37. H. Dalir, Y. Xia, Y. Wang, and X. Zhang, “Athermal broadband graphene optical modulator with 35 GHz speed,” ACS Photonics 3(9), 1564–1568 (2016). [CrossRef]  

38. V. Sorianello, M. Midrio, G. Contestabile, I. Asselberghs, J. Van Campenhout, C. Huyghebaert, I. Goykhman, A. K. Ott, A. C. Ferrari, and M. Romagnoli, “Graphene-silicon phase modulators with gigahertz bandwidth,” Nat. Photonics 12(1), 40–44 (2018). [CrossRef]  

39. E. Heidari, H. Dalir, F. M. Koushyar, B. M. Nouri, C. Patil, M. Miscuglio, D. Akinwande, and V. J. Sorge, “Integrated ultra-high-performance graphene optical modulator,” Nanophotonics 11(17), 4011–4016 (2022). [CrossRef]  

40. T. Amotchkina, M. Trubetskov, D. Hahner, and V. Pervak, “Characterization of e-beam evaporated Ge, YbF3, ZnS, and LaF3 thin films for laser-oriented coatings,” Appl. Opt. 59(5), A40–A47 (2020). [CrossRef]  

41. M.-S. Kwon, “Discussion of the Epsilon-Near-Zero Effect of Graphene in a Horizontal Slot Waveguide,” IEEE Photonics J. 6(3), 1–9 (2014). [CrossRef]  

42. A. Osman, M. Nedeljkovic, J. S. Penades, Y. Wu, Z. Qu, A. Z. Khokhar, K. Debnath, and G. Z. Mashanovich, “Suspended low-loss germanium waveguides for the longwave infrared,” Opt. Lett. 43(24), 5997–6000 (2018). [CrossRef]  

43. M. Nedeljkovic, J. S. Penades, V. Mittal, G. S. Murugan, A. Z. Khokhar, C. Littlejohns, L. G. Carpenter, C. B. E. Gawith, J. S. Wilkinson, and G. Z. Mashanovich, “Germanium-on-silicon waveguides operating at mid-infrared wavelengths up to 8.5 µm,” Opt. Express 25(22), 27431–27441 (2017). [CrossRef]  

44. K. Gallacher, R. W. Millar, U. Griškevičiūte, L. Baldassarre, M. Sorel, M. Ortolani, and D. J. Paul, “Low loss Ge-on-Si waveguides operating in the 8–14 µm atmospheric transmission window,” Opt. Express 26(20), 25667–25675 (2018). [CrossRef]  

45. Z. Cheng, X. Zhu, M. Galili, L. H. Frandsen, H. Hu, S. Xiao, J. Dong, Y. Ding, L. Oxenløwe, and X. Zhang, “Double-layer graphene on photonic crystal waveguide electro-absorption modulator with 12 GHz bandwidth,” Nanophotonics 9(8), 2377–2385 (2020). [CrossRef]  

46. Y. Hu, M. Pantouvaki, J. V. Campenhout, S. Brems, I. Asselberghs, C. Huyghebaert, P. Absil, and D. V. Thourhout, “Broadband 10 Gb/s operation of graphene electro-absorption modulator on silicon,” Laser Photonics Rev. 10(2), 307–316 (2016). [CrossRef]  

47. Y. Kim and M.-S. Kwon, “Mid-infrared subwavelength modulator based on grating-assisted coupling of a hybrid plasmonic waveguide mode to a graphene plasmon,” Nanoscale 9(44), 17429–17438 (2017). [CrossRef]  

48. M. M. -Ballester, V. Vakarin, J. M. Ramirez, Q. Liu, C. A. -Ramos, X. L. Roux, J. Frigerio, A. Ballabio, A. Barzaghi, L. Deniel, D. Bouville, L. Vivien, G. Isella, and D. M. -Morini, “Optical modulation in Ge-rich SiGe waveguides in the mid-infrared wavelength range up to 11 µm,” Commun. Mater. 1(1), 1–6 (2020). [CrossRef]  

49. M. H. Rezaei and M. Shiri, “High-performance tunable resonant electro-optical modulator based on suspended graphene waveguides,” Opt. Express 29(11), 16299–16311 (2021). [CrossRef]  

50. M. M. -Ballester, L. Deniel, N. Koompai, T. H. N. Nguyen, J. Frigerio, A. Ballabio, V. Falcone, X. L. Roux, C. A. -Ramos, L. Vivien, A. Bousseksou, G. Isella, and D. M. -Morini, “Mid-infrared integrated electro-optic modulator operating up to 225 MHz between 6.4 and 10.7 µm wavelength,” ACS Photonics 9(1), 249–255 (2022). [CrossRef]  

51. A. Barzaghi, V. Falcone, S. Calcaterra, D. M. -Morini, M. Virgilio, and J. Frigerio, “Modelling of an intersubband quantum confined Stark effect in Ge quantum wells for mid-infrared photonics,” Opt. Express 30(26), 46710–46721 (2022). [CrossRef]  

52. T. H. N. Nguyen, N. Koompai, V. Turpaud, M. M. -Ballester, J. Peltier, J. Frigerio, A. Ballabio, R. Giani, J.-R. Coudevylle, C. Villebasse, D. Bouville, C. A. -Ramos, L. Vivien, G. Isella, and D. M. -Morini, “1 GHz electro-optical silicon-germanium modulator in the 5-9 µm wavelength range,” Opt. Express 30(26), 47093–47102 (2022). [CrossRef]  

53. S. A. R. Firoozifar, A. Behjat, E. Kadivar, S. M. B. Ghorashi, and M. B. Zarandi, “A study of the optical properties and adhesion of zinc sulfide anti-reflection thin film coated on a germanium substrate,” Appl. Surf. Sci. 258(2), 818–821 (2011). [CrossRef]  

54. D.-P. Kim, C.-I. Kim, and K.-H. Kwon, “Etching properties of ZnS thin films in Cl2/CF4/Ar plasma,” Thin Solid Films 459(1-2), 131–136 (2004). [CrossRef]  

55. F. Huang, F. Jia, C. Cai, Z. Xu, C. Wu, Y. Ma, G. Fei, and M. Wang, “High- and Reproducible-Performance Graphene/II-VI Semiconductor Film Hybrid Photodetectors,” Sci. Rep. 6(1), 28943 (2016). [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 (9)

Fig. 1.
Fig. 1. Schematic of the proposed EAM. (a) The three-dimensional structure. (b) The cross-section.
Fig. 2.
Fig. 2. (a)-(c) The electric field magnitude |E| distributions of the fundamental TE mode at graphene Fermi level EF = 0 eV in the Ge-on-Si rib, suspended germanium rib, and suspended germanium slot waveguide, respectively. (d) The graphene absorption with different Fermi levels in (a) and (b). (e) The graphene absorption with different Fermi levels in (b) and (c).
Fig. 3.
Fig. 3. (a) Absorption of the EAM for different slot widths as a function of rib width. (b) Absorption and group index of the EAM as a function of slot width of a rib width ∼ 2.8 µm.
Fig. 4.
Fig. 4. Modulation depth of the EAM as a function of graphene Fermi level with (a) different wavelengths at graphene scattering time = 100 fs (b) different graphene scattering times at wavelength λ = 9 µm.
Fig. 5.
Fig. 5. (a) Schematic of the proposed MRM. (b) Transmission spectra of the MRM with different applied voltages. (c) Modulation depth of the MRM with different applied voltages.
Fig. 6.
Fig. 6. (a) Drive voltage of the proposed modulators versus the graphene Fermi level. (b) The equivalent RC circuit of the proposed modulators.
Fig. 7.
Fig. 7. Complex relative permittivity (real part and imaginary part) of graphene versus Fermi level with different wavelengths.
Fig. 8.
Fig. 8. Fabrication steps for the proposed graphene modulators. (a) Ge-on-Si wafer. (b) germanium slot waveguide. (c) ZnS deposition and planarization. (d) Wet etch holes etching. (e) Si trenches etching. (f) Bottom layer graphene transferring and patterning. (g) Bottom layer graphene contact electrode. (h) ZnS dielectric layer deposition and the ZnS on the electrode surface removing. (i) Top layer graphene transferring and patterning. (j) Top layer graphene contact electrode.
Fig. 9.
Fig. 9. Modulation depth of the EAM at wavelength λ = 9 µm with the overlay accuracy consideration. Case a: shifting the bottom layer graphene forward by 50 nm while keeping the top layer graphene unchanged. Case b: shifting the bottom layer graphene backward by 50 nm while keeping the top layer graphene unchanged. Case c: shifting the top layer graphene forward by 50 nm without altering the bottom layer graphene. Case d: shifting the top layer graphene backward by 50 nm without altering the bottom layer graphene. Case e: shifting both the top and bottom layer graphene forward by 50 nm. Case f: shifting both the top and bottom layer graphene backward by 50 nm.

Tables (2)

Tables Icon

Table 1. Parameters of the proposed EAM

Tables Icon

Table 2. Performance comparison of on-chip integrated MIR electro-optic modulators operating at the second atmospheric window (8-12 µm)

Equations (6)

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

T ( λ )  =  | ( 1  -  κ 2  -  e r + i ϕ ( λ ) ) ( 1 1  -  κ 2 e r + i ϕ ( λ ) ) | 2
E F = sgn ( V ) v F π ε 0 ε Z n S | V | t Z n S e
B W = 1 2 π [ ( R C + R S ) C g r a p h e n e ]
R C = 2 R C 1 = 2 ρ C L
R S = 2 R S 1 = 2 R ( w g r a p h e n e w m ) L
C g r a p h e n e = ε 0 ε Z n S w m L t Z n S
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.