Next Article in Journal
Sustainable Development Approaches through Wooden Adhesive Joints Design
Previous Article in Journal
Polymerization of Hexene-1 and Propylene over Supported Titanium–Magnesium Catalyst: Comparative Data on the Polymerization Kinetics and Molecular Weight Characteristics of Polymers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Low Power Consuming Mode Switch Based on Hybrid-Core Vertical Directional Couplers with Graphene Electrode-Embedded Polymer Waveguides

Institute of Advanced Photonics Technology, School of Information Engineering, and Guangdong Provincial Key Laboratory of Information Photonics Technology, Guangdong University of Technology, Guangzhou 510006, China
*
Authors to whom correspondence should be addressed.
Polymers 2023, 15(1), 88; https://doi.org/10.3390/polym15010088
Submission received: 24 November 2022 / Revised: 14 December 2022 / Accepted: 21 December 2022 / Published: 26 December 2022
(This article belongs to the Section Polymer Applications)

Abstract

:
We propose a mode switch based on hybrid-core vertical directional couplers with an embedded graphene electrode to realize the switching function with low power consumption. We designed the device with Norland Optical Adhesive (NOA) material as the guide wave cores and epoxy polymer material as cladding to achieve a thermo-optic switching for the E11, E21 and E12 modes, where monolayer graphene served as electrode heaters. The device, with a length of 21 mm, had extinction ratios (ERs) of 20.5 dB, 10.4 dB and 15.7 dB for the E21, E12 and E11 modes, respectively, over the C-band. The power consumptions of three electric heaters were reduced to only 3.19 mW, 3.09 mW and 2.97 mW, respectively, and the response times were less than 495 µs, 486 µs and 498 µs. Additionally, we applied such a device into a mode division multiplexing (MDM) transmission system to achieve an application of gain equalization of few-mode amplification among guided modes. The differential modal gain (DMG) could be optimized from 5.39 dB to 0.92 dB over the C-band, together with the characteristic of polarization insensitivity. The proposed mode switch can be further developed to switch or manipulate the attenuation of the arbitrary guided mode arising in the few-mode waveguide.

1. Introduction

Mode division multiplexing (MDM) transmission, which exploits different orthogonal modes arising in a few-mode fiber (FMF) as independent transmission channels, plays an important role in improving the optical communication capacity [1,2,3,4]. As for an MDM transmission system, optical components, such as mode division multiplexers and mode selective switch, serve as essential devices [5,6,7,8], where such devices can be realized by multimode interferometers [9], photonic lanterns [10,11], Y-junctions [12,13] and directional couplers (DCs) [14,15,16,17]. DCs, which operate on the mechanism of mode coupling between two parallel waveguides, can process the design flexibility and, therefore, have continually reported a great number of devices in the MDM system. The implementations of those devices allow to spatially manipulate various spatial modes in an FMF. A mode selective switch can dynamically (de)multiplex the spatial modes between waveguides [18,19,20,21]. Such an optical switch can be realized via thermo-optic effect based on the high thermo-optic coefficient polymer material platform or via electric-optic effect based on lithium niobate waveguides [16,18]. By the use of a silicon densely packed waveguide array, an optical switch is reported to manipulate multiple spatial modes [19]. By the use of exploiting a Y-junction and multimode interference structure, a mode switch on a silicon-on-insulator platform is demonstrated [20]. However, most reported mode switch is homogeneous, which makes it hard to only manipulate the fundamental mode. Recently, a thermo-optic mode switch has been presented with a polymer/silica hybrid 3D waveguide [16].
Meanwhile, the power consumption is also a significant issue. By the use of a graphene electrode heater, a mode switch with low power consumption has been demonstrated [21]. With the variety of the refractive index of the Norland Optical Adhesive (NOA) material, the refractive index of the waveguide core can be selected alternatively. Moreover, with the help of a graphene electrode heater, the power consumption can be further reduced. Therefore, hybrid-core directional couplers could be further developed to switch the arbitrary guided mode, including fundamental mode, and present low power consumption as well. For example, in some previously reported arts, such as by integrating a microfiber with a graphene film, the switching power is reported to be 11 mW [22]. By using a Michelson interferometer formed with graphene-coated side-polished twin-core fiber, the pump power can even reach over 140 mW [23]. By the use of a graphene-on-silicon nanobeam cavity, a switch is also demonstrated with a switching power of 47 mW [24]. The polymer thermo-optic switch can obtain a power consumption of 9.5 mW and a switching time of 106 µs (rise)/102 µs (fall) [25]. Meanwhile, the on-chip silicon photonic waveguide switch can reduce the switching time to be 5.4 µs; however, the electric power consumption increases in the meantime, which reaches approximately 22.5 mW [26]. In a recent work, by using a hybrid structure of polymer cladding and silica waveguide core, the power consumption can be reduced to 5.49 mW (TE) and 5.96 mW (TM) [27]. A dual-mode 2 × 2 thermo-optic switch was demonstrated with the switching power of 9.0 mW based on the structure of a Mach–Zehnder Interferometer formed by polymer waveguides [28]. According to the previous study, the switching power can be tremendously reduced by the use of large thermo-optic coefficient (~−3 × 10−4/K) material. Luckily, NOA material with a large thermo-optic coefficient can be used in our study to further reduce the switch power. Thanks to the advantages of NOA material, the integrated device can be further applied to the mode switch with arbitrary guided mode switching and low power consumption.
In this paper, we propose a mode switch based on hybrid-core cascaded vertical directional couplers. The device realizes switching for the E11, E21 and E12 modes via the thermo-optic effect, which is driven by the embedded graphene electrode heaters (GEHs). The power consumptions of GEHs are less than 3.19 mW, and response times of DCs are shorter than 498 µs. The coupling ratios (CRs) for three guided modes are higher than 99.1%, 90.8% and 97.3% over the C-band. Such a device can be further developed to achieve the switching for arbitrary guided modes in the FMF, with the help of a hybrid-waveguide structure and embedded GEHs. The proposed device can be useful in the MDM transmission system where the mode switch or mode tunable attenuation is compulsory.

2. Operating Principle and Design

Figure 1 shows the schematic diagram and the function of the proposed device, which consists of three cascaded vertical asymmetric DCs formed with a few-mode core (FMC) labeled as Core 1, which is located at the lower layer, and three single-mode cores (SMCs) labeled as Core 2, Core 3 and Core 4, which are located at the upper layer. The FMC supports the E11, E21 and E12 modes and each of the SMCs supports only the E11 mode. Three pieces of monolayer graphene, which serve as heating electrodes, are embedded in each SMC at the mode coupling region. Figure 1a shows that the DCs are de-activated with switch states set at OFF and the signal lights stay at Core 1 from the output port. Figure 1b shows that the E11, E21 and E12 modes in Core 1 are demultiplexed into the E11 modes in Core 2, Core 3 and Core 4 with the switch states set at ON. By the use of three GEHs, the local refractive index of waveguides in the mode coupling region can be manipulated with high efficiency via controlling the phase-matching condition of DCs, where GEH 1, GEH 2 and GEH 3 serve to attenuate the power of the E21, E12 and E11 modes in the FMC, respectively. A horizontal linear taper is applied to adjust the waveguide core size to realize a compact device design.
The operation principle of the mode switch is based on the couple mode theory by manipulating the mode coupling between two parallel waveguides. Strong mode coupling happens when the phase-matching condition between two waveguides is satisfied. That is to say, the mode effective refractive indices of the FMC and SMC should be similar. In our design, by the use of NOA material (Edmund Optics Inc., Barrington, USA), the refractive indices of the FMC, the SMC and the cladding are taken as nFMC = 1.566, nSMC = 1.569 and nClad = 1.559, respectively. The effective refractive indices of the modes against the core width are calculated with the finite element method (FEM) based on commercial software (COMSOL Multiphysics 5.5), as shown in Figure 2. Here, we present a typical design, for example, where the heights of the FMC and SMCs are fixed at 8 and 4 µm, respectively. As shown in Figure 2a, before the FMC taper, the FMC width is set as W1 = 13 µm, which can support the E11, E21 and E12 modes. After that, the FMC width is tapered down to the width of W1′ = 8.5 µm, which can better adjust the effective refractive index of the E11 mode. As shown in Figure 2b, the widths for three SMCs (Core 2, Core 3 and Core 4) are chosen to be W2 = 4.98, W3 = 3.65 and W4 = 8.78 µm, respectively. In this design, the E11 modes arising in the SMCs and the E11, E21 and E12 modes arising in the FMC are set at the phase-mismatching point exactly.
Figure 3 shows the phase-matching conditions for the operation mechanism of the proposed mode switch. Figure 3a illustrates the phase-matching process between the effective refractive indices for E21 or E12 modes arising in Core 1 and the effective refractive indices for the E11 modes in Core 2 and Core 3, respectively. With the help of GEH 1 and GEH 2, the raising temperature can reduce the refractive index of the waveguide cores, where the refractive indices variation of the SMCs is higher than the refractive index variation of the FMC. The variations of the refractive index in the FMC and SMCs manipulate the mode coupling effect. Particularly when the effective refractive index of modes in the FMC and SMCs is tuned to a similar value synchronously, strong coupling between modes in the FMC and SMCs occurs because the phase-matching conditions are satisfied. Figure 3b illustrates the condition of the effective refractive index for the E11 mode in Core 1 matching the effective refractive index for the E11 modes in Core 4. Similarly, the coupling of the E11 mode in the FMC and SMCs takes place when the phase-matching condition is satisfied with the help of the thermo-optics effect driven by GEH 3.
The GEHs serve to manipulate the coupling of the corresponding DCs. The dimensions of three DCs with GEHs are shown in Figure 4, which are labeled as DC 1, DC 2 and DC 3. The gap distance between the FMC and SMCs is fixed at 3.0 µm, and the distance between the SMCs and GEHs is fixed at 5.0 µm. There is a core-to-core lateral offset distance of approximately 6.0 µm, 4.0 µm and 4.0 µm for DC 1, DC 2 and DC 3, respectively. As discussed in Figure 2, the FMC width and height are designed to be 13.0 µm and 8.0 µm so that the FMC can support the E11, E21 and E12 modes. A linear taper is designed in FMC for the DC 3 with the core width tapered to be 8.5 µm to adjust the effective refractive index of the E11 mode, where the linear taper induces little loss to the E21 and E12 modes. The height of the SMCs is set to be 4.0 µm and the width of Core 2, Core 3 and Core 4 set to be 5.0 µm, 3.7 µm and 8.8 µm, respectively. The sizes of the SMCs are designed to allow the E11 mode in the SMCs with their effective refractive indices mismatching with the corresponding modes in the FMC. The phase-matching condition can be resumed with a few electric powers applied in the GEHs. The total length of the device, including the coupler lengths, the tapers, the S-bends and the input/output parallel waveguide sections, is 21 mm.

3. Simulation Results and Discussions

The material parameters used in the simulations are shown in Table 1. For the simulation using the finite element method with commercial software (COMSOL), the finite element mesh is set as “Extremely fine” with the minimum element size of 0.0016 µm. We calculate the absorption losses induced by the GEHs, where the graphene film is modeled as a conductive boundary with the chemical potential of µc = 0.3 eV and the complex surface conductivities 6.0792 × 10−5–8.616010−6i for 1550 nm [29,30]. By using the electromagnetic waves equations in the frequency domain as the physics field and mode analysis model, the effective refractive index of the modes can be solved directly. The modal loss is calculated by [31]:
P L o s s ( dB / μ m ) = 10 log 10 [ e 2 k 0 Im ( n e f f ) ] = 8.68 k 0 Im ( n e f f ) ,
where k0 = 2π/λ is the free-space wavenumber and λ is the free-space wavelength, Im(neff) is the imaginary part of the effective refractive index, where the simulation results are shown in Table 2. To study the position of the GEH that affects the signal propagation efficiency, the GEHs are placed at different heights on the SMC surface. Figure 5 shows the variation of the graphene-induced absorption losses to the TE and TM polarized light at the wavelength of 1550 nm for three cores with different core-graphene distances, which is labeled as d1, d2 and d3, respectively. The losses of TE polarized light are dependent on the core-graphene distance. The losses of TE polarized light decreases with the growing core-graphene distance, while the TM polarized light is almost transparent to the graphene. Moreover, the graphene-induced losses to the TE polarized light can be negligible when the core-graphene distance is larger than 4 µm.
We calculate the electric power generated by the GEHs via the 3D finite-difference beam propagation method (3DFD-BPM) with commercial software (Rsoft). By applying the electric power on the GEHs, the change in the heater temperature against the electric power is expressed as [32]:
P = L e W e k ( 1 + 0.88 H W e ) Δ T h ,
where Le, We and H are the length, width and height of the GEH, k is the thermal conductivity, ΔT is the temperature change and h is the natural convection heat transfer coefficient. By the analysis of the couple mode theory, we learn that the change in regional temperature can induce the change in the refractive indices of the FMC and SMCs and further manipulate the coupling of the modes between the FMC and SMCs. The temperature change in the SMC depends on the applied electric power; hence, we use the beam propagation method to study the temperature variation and the response against the electric power by using Equation (2).
To evaluate the performance of this device, we launched the E21, E12 and E11 modes into the input port of Core 1 and monitored the output power at the output port. Coupling ratio (CR) is utilized to characterize the performance of the device. Figure 6 shows the CRs variation of the E21, E12 and E11 modes with the electric power of three GEHs at the different core-graphene distances at 1550 nm, which can be calculated by:
C R m n = P m n i n P m n o u t P m n i n ,
where Pmn-in and Pmn-out are the input and output power (in mW) of the Emn mode from Core 1. The CRs increase with the growing electric power applied to the GEHs, and the increasing speed is dependent on the core-graphene distances during the design. The minimum electric power required becomes larger with the growing core-graphene distance, while the graphene absorption becomes smaller. Therefore, it is important to make a trade-off between the graphene-induced loss and the power consumption of GEHs. As for our design, the core-graphene distance is chosen to be 5 µm to realize low electric power consumption and low graphene-induced loss simultaneously.
Due to the large thermo-optic coefficient (~−3 × 10−4/K) [16], NOA material is used in our study to further reduce the power consumption. The temperature variations against the applied electric power are calculated by BPM based on commercial software (Rsoft). By using the “Perform Simulation” function in Rsoft, Figure 7 shows the optical propagation path of the corresponding mode in core 1, with different electric power applied to the GEHs. Maximal output power can be obtained without electric power applied to the GEHs, where the modal power can be attenuated partially when the electric power is added. Take Figure 7a as an example, the E21 mode can be switched partially when the electric power is set at 1.5 mW. The output power further decreases in case the applied electric power is increased, which can be turned to almost reach 0 mW under 3 mW of electric power, by coupling the optical power to the upper SMC via the DC. The results are similar to the E12 and E11 modes, which are shown in Figure 7b,c, respectively. Although the core 1 guides three different modes, the electric power required to manipulate the mode coupling of three guided modes between the FMC and the SMCs are similar. Figure 8 shows the CRs of E21, E12 and E11 modes with the variable electric power applied to the GEHs, when the core-graphene distance is fixed at 5 µm at 1550 nm. Thus, we can control the CRs for the corresponding modes by applying a variable electric power to the GEHs. The maximal CRs can be obtained for three modes when the electric power is set at 3.19 mW, 3.09 mW and 2.97 mW, respectively.
We also calculate the response times of the device with electric power applied to the GEHs. Here, the response time is defined as the electric power loading time until the desired CR is achieved. To characterize the process of heat transfer, we study the thermal distribution against the time, as shown in Figure 9, where the coefficient of thermal conductivity of the NOA polymer material is set at 0.2 W/m·K [16,33]. The heat transfer in the solids (ht) model is set as the physics field in COMSOL to serve as the temperature model for each section of the device. The “Time Dependent” study (Transient State) is used to calculate the temperature rising and falling. The CR will reach the maximum value at 1000 µs, as shown in Figure 10. The response times for three DCs are summarized in detail in Table 3. The response time of this device, including the rising and the falling time, is shorter than 495 µs, 486 µs and 498 µs, respectively, and the response can be further optimized by the use of other functional polymer materials [34].
To further evaluate the performance of this device, we calculate the CRs for three modes with respect to the operation wavelength, as shown in Figure 11. The CRs are dependent on the applied electric power to the GEHs, and can be turned continually over the C-band. Meanwhile, the CRs for three modes can be higher than 99.1%, 90.8% and 97.3%, respectively, over the C-band, which shows that this device can operate over a large tunable range.

4. Application of Gain Equalization

Mode-dependent loss (MDL) in optical transmission or differential modal gain (DMG) in optical amplification are the challenges in the MDM system, where larger MDL or DMG may cause the failure of multiple-input-multiple-output digital signal processing (MIMO-DSP) at the receiver side [35,36,37,38]. Such a proposed mode switch with the function of MDL or DMG equalization can be used to solve this problem. We demonstrate the DMG mitigation ability of this device via an MDM transmission system. Figure 12 shows an MDM transmission system; the modal power are disparities at the input port after the long-distance transmission and optical amplification. In the system, a uniformly erbium-doped step-index polymer waveguide amplifier is used in the simulation, where the refractive indices of core and cladding are 1.567 and 1.559, respectively. The erbium doping region is the same size as the core with the concentration setting at 5 × 1024 m−3 [39,40]. The width and height of the waveguide core are both fixed at 12 µm so that it can support the E11, E21 and E12 modes. A schematic diagram of the refractive index (RI) profile, the erbium-doped distribution and doping concentration are illustrated in Figure 13. In this model, we can obtain the amplification of the E11, E21 and E12 modes, when the power of the pump laser at 980 nm is 100 mW and the power of the input signal for each mode is 0.1 mW.
In this model, we can also obtain the DMG against the operation wavelength, as shown in Figure 14a,b, respectively. The results show that the average gain of the guided modes is 23.86 dB and the DMG is approximately 5.39 dB over the C-band. Obviously, large DMG will seriously deteriorate the system performance. Therefore, the effect of DMG mitigation is significant with the help of this device. By applying the electric power to the GEH 3 with the value of 1.90 mW, the E11 mode can be attenuated, and as a result, the DMG can be decreased. As shown in Figure 14c,d, the DMG is modified to be lower than 0.92 dB, after the use of proposed mode switch. The tunable design of the mode switch can be used for DMG mitigation and applied to the few-mode amplification system or MDM transmission to achieve the desired value.
The proposed device can be fabricated by the use of standard microfabrication processes, which include spin-coating, optical lithography and RIE etching. The available polymer material for the fabrication has characteristics of low loss and fiber capability [41]. Additionally, the graphene electrodes are demonstrated to be highly efficient with the mature wet transfer method [21].

5. Conclusions

The performance of the proposed mode switch is compared with other reported switches, which is listed in Table 4. We can see that the NOA polymer material-based waveguide devices can somehow reduce the power consumption because of their large thermo-optic coefficient. Because of the merits of the NOA polymer material, the power consumption of the proposed mode switch can be reduced to 3.19 mW. In the meanwhile, thanks to the variety of the refractive index of the NOA polymer material, the refractive index of the waveguide core can be designed alternatively with a linear choice of available refractive indices of material. Taking these advantages into consideration, the hybrid-core structure can be easily realized in the experiment. Such an integrated device can be further applied to the mode switch for arbitrary guided mode switching with the flexible designs of the vertical DCs. However, because the thermal conductivity of the polymer is smaller than silicon, the switching speed of the polymer-based switches is slower than the silicon-based switches. The response time of this work is at the microsecond level, which is similar to the other reported polymer waveguide-based devices. The response time of the mode switch can be further optimized by using a high performance functional polymer material such as EO-polymers.
In this paper, we propose a mode switch based on hybrid-core cascaded vertical DCs. Three GEHs are designed on the top of three vertical DCs, which are formed with an FMC and three SMCs. Such a device can realize modal switching for the E21, E12 and E11 mode by applying various electric power to three GEHs. The maximum power consumptions for three GEHs are approximately 3.19 mW, 3.09 mW and 2.97 mW, respectively. The response times are shorter than 495 µs, 486 µs and 498 µs. The CRs for three modes are higher than 99.1%, 90.8% and 97.3% over the C-band. Moreover, our proposed device also shows great potential for the DMG equalization of FM-EDFA during the long-distance MDM transmission.

Author Contributions

Methodology, Q.H.; Software, L.Z. and J.Z.; Validation, Q.H. and O.X.; Formal analysis, L.Z., Q.H. and J.Z.; Investigation, L.Z. and J.Z.; Resources, L.Z.; Data curation, O.X.; Writing—original draft, L.Z.; Writing—review and editing, Q.H. and O.X.; Supervision, Q.H. and O.X.; Project administration, O.X. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Key Research and Development Program of China (2018YFB1800901) and the Guangdong Provincial Key Laboratory of Photonics Information Technology (2020B121201011).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available upon request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Richardson, D.J.; Fini, J.M.; Nelson, L.E. Space-division multiplexing in optical fibres. Nat. Photon. 2013, 7, 354–362. [Google Scholar] [CrossRef] [Green Version]
  2. Li, G.; Bai, N.; Zhao, N.; Xia, C. Space-division multiplexing: The next frontier in optical communication. Adv. Opti. Photon. 2014, 6, 413–487. [Google Scholar] [CrossRef] [Green Version]
  3. Winzer, P.J.; Neilson, D.T.; Chraplyvy, A.R. Fiber-optic transmission and networking: The previous 20 and the next 20 years [Invited]. Opt. Express 2018, 26, 24190–24239. [Google Scholar] [CrossRef] [PubMed]
  4. Puttnam, B.J.; Rademacher, G.; Luís, R.S. Space-division multiplexing for optical fiber communications. Optica 2021, 8, 1186–1203. [Google Scholar] [CrossRef]
  5. Li, C.; Liu, D.; Dai, D. Multimode silicon photonics. Nanophotonics 2019, 8, 227–247. [Google Scholar] [CrossRef]
  6. Han, X.; Jiang, Y.; Frigg, A.; Xiao, H.; Zhang, P.; Nguyen, T.G.; Boes, A.; Yang, J.; Ren, G.; Su, Y.; et al. Mode and polarization-division multiplexing based on silicon nitride loaded lithium niobate on insulator platform. Laser Photon. Rev. 2022, 16, 2100529. [Google Scholar] [CrossRef]
  7. Liu, Y.; Wang, Z.; Liu, Y.; Wang, X.; Guo, X.; Li, D.; Yao, Y.; Song, Q.; Du, J.; He, Z.; et al. Ultra-compact mode-division multiplexed photonic integrated circuit for dual polarizations. J. Lightw. Technol. 2021, 39, 5925–5932. [Google Scholar] [CrossRef]
  8. Zhao, W.; Liu, R.; Peng, Y.; Yi, X.; Chen, H.; Dai, D. High-performance silicon polarization switch based on a Mach–Zehnder interferometer integrated with polarization-dependent mode converters. Nanophotonics 2022, 11, 2293–2301. [Google Scholar] [CrossRef]
  9. Chang, P.H.; Lin, C.; Helmy, A.S. Integrated photonic functions using anisotropic 2D material structures. J. Lightw. Technol. 2021, 39, 7464–7471. [Google Scholar] [CrossRef]
  10. Douglass, G.; Dreisow, F.; Gross, S.; Withford, M.J. Femtosecond laser written arrayed waveguide gratings with integrated photonic lanterns. Opt. Express 2018, 26, 1497–1505. [Google Scholar] [CrossRef]
  11. Lu, Y.; Liu, W.; Chen, Z.; Jiang, M.; Zhou, Q.; Zhang, J.; Li, C.; Chai, J.; Jiang, Z. Spatial mode control based on photonic lanterns. Opt. Express 2021, 29, 41788–41797. [Google Scholar] [CrossRef]
  12. Chang, W.; Lu, L.; Ren, X.; Li, D.; Pan, Z.; Cheng, M.; Liu, D.; Zhang, M. Ultra-compact mode (de) multiplexer based on subwavelength asymmetric Y-junction. Opt. Express 2018, 26, 8162–8170. [Google Scholar] [CrossRef] [PubMed]
  13. Lin, C.; Hashisaka, M.; Akiho, T.; Muraki, K.; Fujisawa, T. Quantized charge fractionalization at quantum Hall Y junctions in the disorder dominated regime. Nat. Commun. 2021, 12, 131. [Google Scholar] [CrossRef] [PubMed]
  14. Han, L.; Kuo, B.P.P.; Alic, N.; Radic, S. Ultra-broadband multimode 3dB optical power splitter using an adiabatic coupler and a Y-branch. Opt. Express 2018, 26, 14800–14809. [Google Scholar] [CrossRef] [PubMed]
  15. Lin, Y.X.; Younesi, M.; Chung, H.P.; Chiu, H.K.; GEISS, R.; Tseng, Q.H.; Setzpfandt, F.; Pertsch, T.; Chen, Y.H. Ultra-compact, broadband adiabatic passage optical couplers in thin-film lithium niobate on insulator waveguides. Opt. Express 2021, 29, 27362–27372. [Google Scholar] [CrossRef]
  16. Lin, B.; Sun, S.; Yang, K.; Zhu, M.; Gu, Y.; Yu, Q.; Wang, X.; Zhang, D. Polymer/silica hybrid 3D waveguide thermo-optic mode switch based on cascaded asymmetric directional couplers. Appl. Opt. 2021, 60, 6943–6949. [Google Scholar] [CrossRef]
  17. Thomaschewski, M.; Zenin, V.A.; Wolff, C.; Bozhevolnyi, S.I. Plasmonic monolithic lithium niobate directional coupler switches. Nat. Commun. 2020, 11, 748. [Google Scholar] [CrossRef] [Green Version]
  18. Zhang, X.; Chen, H.; Liu, L.; Shang, J.; Yu, L.; Li, K.; Dong, J.; Qiu, W.; Guan, H.; Lu, H.; et al. An optical switch based on electro-optic mode deflection in lithium niobate waveguide. IEEE Photon. Technol. Lett. 2020, 32, 1295–1298. [Google Scholar] [CrossRef]
  19. Chen, K.; Yan, J.; He, S.; Liu, L. Broadband optical switch for multiple spatial modes based on a silicon densely packed waveguide array. Opt. Lett. 2019, 44, 907–910. [Google Scholar] [CrossRef]
  20. Xiong, Y.; Priti, R.B.; Ladouceur, O.L. High-speed two-mode switch for mode-division multiplexing optical networks. Optica 2017, 4, 1098–1102. [Google Scholar] [CrossRef]
  21. Song, Q.Q.; Chen, K.X.; Hu, Z.F. Low-power broadband thermo-optic switch with weak polarization dependence using a segmented graphene heater. J. Lightw. Technol. 2020, 38, 1358–1364. [Google Scholar] [CrossRef]
  22. Gan, X.; Zhao, C.; Wang, Y.; Mao, D.; Fang, L.; Han, L.; Zhao, J. Graphene-assisted all-fiber phase shifter and switching. Optica 2015, 2, 468–471. [Google Scholar] [CrossRef]
  23. Chu, R.; Guan, C.; Bo, Y.; Liu, J.; Shi, J.; Yang, J.; Ye, P.; Li, P.; Yang, J.; Yuan, L. Graphene decorated twin-core fiber Michelson interferometer for all-optical phase shifter and switch. Opt. Lett. 2020, 45, 177–180. [Google Scholar] [CrossRef]
  24. Guo, T.; Gao, S.; Zeng, H.; Tang, L.; Qiu, C. All-optical control of a single resonance in a graphene-on-silicon nanobeam cavity using thermo-optic effect. J. Lightw. Technol. 2021, 39, 4710–4716. [Google Scholar] [CrossRef]
  25. Yang, Y.; Lv, J.; Lin, B.; Cao, Y.; Yi, Y.; Zhang, D. Graphene-assisted polymer waveguide optically controlled switch using first-order mode. Polymers 2021, 13, 2117. [Google Scholar] [CrossRef]
  26. Truong, C.D.; Hang, D.N.T.; Chandrahalim, H.; Trinh, M.T. On-chip silicon photonic controllable 2×2 four-mode waveguide switch. Sci. Rep. 2021, 11, 897. [Google Scholar] [CrossRef]
  27. Xie, Y.; Han, J.; Qin, T.; Ge, X.; Wu, X.; Liu, L.; Wu, X.; Yi, Y. Low power consumption hybrid-integrated thermo-optic switch with polymer cladding and silica waveguide core. Polymers 2022, 14, 5234. [Google Scholar] [CrossRef]
  28. Lin, B.; Sun, S.; Sun, X.; Lian, T.; Zhu, M.; Che, Y.; Wang, X.; Zhang, D. Dual-mode 2 × 2 thermo-optic switch based on polymer waveguide mach-zehnder interferometer. IEEE Photon. Technol. Lett. 2022, 34, 1317–1320. [Google Scholar] [CrossRef]
  29. Mikhailov, S.A.; Ziegler, K. New electromagnetic mode in graphene. Phys. Rev. Lett. 2007, 99, 016803. [Google Scholar] [CrossRef] [Green Version]
  30. Huang, Q.; Zhong, L.; Dong, J.; Xu, O.; Zheng, Z.; Huang, T.; Li, J.; Xiang, M.; Fu, S.; Qin, Y. All-optical light manipulation based on graphene-embedded side-polished fiber. Opt. Lett. 2022, 47, 1478–1481. [Google Scholar] [CrossRef]
  31. Huang, Q.; Ghimire, I.; Yang, J.; Fleer, N.; Chiang, K.S.; Wang, Y.; Gao, S.; Wang, P.; Banerjee, S.; Lee, H. Optical modulation in hybrid antiresonant hollow-core fiber infiltrated with vanadium dioxide phase change nanocrystals. Opt. Lett. 2020, 45, 4240–4243. [Google Scholar] [CrossRef] [PubMed]
  32. He, G.; Ji, L.; Gao, Y.; Liu, R.; Sun, X.; Yi, Y.; Wang, X.; Chen, C.; Wang, F.; Zhang, D. Low power 1×4 polymer/SiO2 hybrid waveguide thermo-optic switch. Opt. Commun. 2017, 402, 422–429. [Google Scholar] [CrossRef]
  33. Yan, Y.F.; Zheng, C.T.; Liang, L.; Meng, J.; Sun, X.Q.; Wang, F.; Zhang, D. Response-time improvement of a 2×2 thermo-optic switch with polymer/silica hybrid waveguide. Opt. Commun. 2012, 285, 3758–3762. [Google Scholar] [CrossRef]
  34. Xu, Q.; Jiang, M.; Niu, D.; Wang, X.; Wang, L.; Chiang, K.S.; Zhang, D. Fast and low-power thermo-optic switch based on organic-inorganic hybrid strip-loaded waveguides. Opt. Lett. 2018, 43, 5102–5105. [Google Scholar] [CrossRef] [PubMed]
  35. Sugawara, N.; Kudo, M.; Fujisawa, T.; Sakamoto, T.; Matsui, T.; Tsujikawa, K.; Nakajima, K.; Saitoh, K. 3-mode PLC-based mode dependent loss equalizer in MDM transmission. In Proceedings of the 2019 24th OptoElectronics and Communications Conference (OECC) and 2019 International Conference on Photonics in Switching and Computing (PSC), Fukuoka, Japan, 7–11 July 2019. Paper TuE2-2. [Google Scholar]
  36. Rademacher, G.; Luís, R.S.; Puttnam, B.J.; Eriksson, T.A.; Ryf, R.; Agrell, E.; Maruyama, R.; Aikawa, K.; Awaji, Y.; Furukawa, H.; et al. High capacity transmission with few-mode fibers. J. Lightw. Technol. 2019, 37, 425–432. [Google Scholar] [CrossRef] [Green Version]
  37. Arikawa, M.; Nakamura, K.; Hosokawa, K.; Hayashi, K. Long-haul WDM/SDM transmission over coupled 4-core fiber with coupled 4-Core EDFA and its mode dependent loss characteristics estimation. J. Lightw. Technol. 2022, 40, 1664–1671. [Google Scholar] [CrossRef]
  38. Fujisawa, T.; Mori, T.; Sakamoto, J.; Yamashita, Y.; Sakamoto, T.; Imada, R.; Ima, R.; Sato, T.; Watanabe, K.; Kasahara, R.; et al. Silica-PLC based mode-dependent-loss equalizer for two LP mode transmission. In Proceedings of the 2022 Optical Fiber Communications Conference and Exhibition (OFC), San Diego, CA, USA, 6–10 March 2022. Paper M4J.4. [Google Scholar]
  39. Bai, N.; Ip, E.; Wang, T.; Li, G. Multimode fiber amplifier with tunable modal gain using a reconfigurable multimode pump. Opt. Express 2011, 19, 16601–16611. [Google Scholar] [CrossRef] [Green Version]
  40. Wang, X.; Zhang, M.; Jiang, M.; Lian, T.; Wang, F.; Zhang, D. Monolithic integrated waveguide device with dual functions of electro-optic modulation and optical amplification. Opt. Lett. 2021, 46, 3536–3539. [Google Scholar] [CrossRef]
  41. Huang, Q.; Wang, X.; Dong, J.; Zheng, Z.; Xu, O.; Fu, S.; Peng, D.; Li, J.; Qin, Y. Ultra-broadband LP11 mode converter with high purity based on long-period fiber grating and an integrated Y-junction. Opt. Express 2022, 30, 12751–12759. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the proposed device, where monolayer graphene is attached on the surface of the SMCs; the operation states of the device with switch are set at (a) OFF state and (b) ON state.
Figure 1. Schematic diagram of the proposed device, where monolayer graphene is attached on the surface of the SMCs; the operation states of the device with switch are set at (a) OFF state and (b) ON state.
Polymers 15 00088 g001
Figure 2. The variation of the effective indices of (a) the E11, E21 and E12 modes with the width of the FMC increasing and (b) the E11 mode with the width of the SMCs increasing.
Figure 2. The variation of the effective indices of (a) the E11, E21 and E12 modes with the width of the FMC increasing and (b) the E11 mode with the width of the SMCs increasing.
Polymers 15 00088 g002
Figure 3. The phase-matching conditions of (a) the E21 and E12 modes in Core 1 and the E11 modes in Core 2 and Core 3; (b) the E11 mode in tapered Core 1 and the E11 mode in Core 4.
Figure 3. The phase-matching conditions of (a) the E21 and E12 modes in Core 1 and the E11 modes in Core 2 and Core 3; (b) the E11 mode in tapered Core 1 and the E11 mode in Core 4.
Polymers 15 00088 g003
Figure 4. The dimensions of three vertical DCs with GEHs for switching (a) the E21, (b) the E12 and (c) the E11 modes.
Figure 4. The dimensions of three vertical DCs with GEHs for switching (a) the E21, (b) the E12 and (c) the E11 modes.
Polymers 15 00088 g004
Figure 5. The variation of the graphene-induced absorption losses to the TE and TM polarized light for (a) Core 2, (b) Core 3, and (c) Core 4 with the different core-graphene distances.
Figure 5. The variation of the graphene-induced absorption losses to the TE and TM polarized light for (a) Core 2, (b) Core 3, and (c) Core 4 with the different core-graphene distances.
Polymers 15 00088 g005
Figure 6. The CRs of (a) the E21, (b) the E12 and (c) the E11 modes with the electric power for different core-graphene distances.
Figure 6. The CRs of (a) the E21, (b) the E12 and (c) the E11 modes with the electric power for different core-graphene distances.
Polymers 15 00088 g006
Figure 7. The propagation path for (a) the E21, (b) the E12 and (c) the E11 modes with different electric power applied to the GEHs.
Figure 7. The propagation path for (a) the E21, (b) the E12 and (c) the E11 modes with different electric power applied to the GEHs.
Polymers 15 00088 g007
Figure 8. The CRs of (a) the E21, (b) the E12 and (c) the E11 modes with different electric power applied to the GEHs.
Figure 8. The CRs of (a) the E21, (b) the E12 and (c) the E11 modes with different electric power applied to the GEHs.
Polymers 15 00088 g008
Figure 9. Schematic diagram for the thermal transfer process.
Figure 9. Schematic diagram for the thermal transfer process.
Polymers 15 00088 g009
Figure 10. The variation of CRs against the response times for three DCs to switch (a) the E21, (b) the E12 and (c) the E11 modes.
Figure 10. The variation of CRs against the response times for three DCs to switch (a) the E21, (b) the E12 and (c) the E11 modes.
Polymers 15 00088 g010
Figure 11. The operating wavelength for switching (a) the E21, (b) the E12 and (c) the E11 modes of the device against the applied electric power to the GEHs.
Figure 11. The operating wavelength for switching (a) the E21, (b) the E12 and (c) the E11 modes of the device against the applied electric power to the GEHs.
Polymers 15 00088 g011
Figure 12. MDM transmission system with DMG mitigation.
Figure 12. MDM transmission system with DMG mitigation.
Polymers 15 00088 g012
Figure 13. The RI profile, erbium−doping distribution and concentration of the few−mode erbium−doped polymer waveguide.
Figure 13. The RI profile, erbium−doping distribution and concentration of the few−mode erbium−doped polymer waveguide.
Polymers 15 00088 g013
Figure 14. The variations of the gains and DMGs versus wavelength over the C−band for three modes: (a,b) before DMG mitigation and (c,d) after DMG mitigation.
Figure 14. The variations of the gains and DMGs versus wavelength over the C−band for three modes: (a,b) before DMG mitigation and (c,d) after DMG mitigation.
Polymers 15 00088 g014
Table 1. Material parameters used in the simulations.
Table 1. Material parameters used in the simulations.
NOA PolymerGraphene
Refractive index1.559/1.566/1.5692.98 + 2.79i
Density ρ (kg·m−3)12001060
Heat capacity at constant pressure Cp (J·kg−1·k−1)1420740
Thermal conductivity k (W·m−1·K−1)0.25300
Thermo-optic coefficient (K−1)−3 × 10−4Not applicable
Table 2. The imaginary part of the effective refractive index for three cores.
Table 2. The imaginary part of the effective refractive index for three cores.
The Core-Graphene Distance (µm)0246
Core 2TE polarized1.52 × 10−42.00 × 10−52.88 × 10−64.38 × 10−7
TM polarized1.08 × 10−61.26 × 10−71.69 × 10−82.46 × 10−9
Core 3TE polarized1.31 × 10−42.02 × 10−53.59 × 10−66.90 × 10−7
TM polarized0.84 × 10−61.03 × 10−71.64 × 10−82.92 × 10−9
Core 4TE polarized1.72 × 10−41.94 × 10−52.27 × 10−62.72 × 10−7
TM polarized1.33 × 10−61.46 × 10−71.68 × 10−81.98 × 10−9
Table 3. Response times for three DCs.
Table 3. Response times for three DCs.
CR (%)20406080Max *
DC 1Rise time (µs)466474474475475
Fall time (µs)491495484486482
DC 2Rise time (µs)477477473475476
Fall time (µs)481486485481483
DC 3Rise time (µs)480471480482483
Fall time (µs)492498492489487
* Max means the maximum CR for each DC.
Table 4. Comparison of the proposed mode switch with the other reported switches.
Table 4. Comparison of the proposed mode switch with the other reported switches.
ReferencesWaveguideWavelength (nm)PC 1RT 2FT 3
[16]Polymer/silica1530–157017.35183259
[21]Polymer1530–16053.12980520
[25]Polymer/silica/silicon980/15509.5106102
[26]Silica/silicon1525–157522.55.46.4
[27]Polymer/silica15505.96121329
[28]Polymer1530–15659.012001420
This workPolymer1530–15653.19483498
1 PC, power consumption (mW); 2 RT, rise time (µs); 3 FT, fall time (µs).
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhong, L.; Huang, Q.; Zhang, J.; Xu, O. Low Power Consuming Mode Switch Based on Hybrid-Core Vertical Directional Couplers with Graphene Electrode-Embedded Polymer Waveguides. Polymers 2023, 15, 88. https://doi.org/10.3390/polym15010088

AMA Style

Zhong L, Huang Q, Zhang J, Xu O. Low Power Consuming Mode Switch Based on Hybrid-Core Vertical Directional Couplers with Graphene Electrode-Embedded Polymer Waveguides. Polymers. 2023; 15(1):88. https://doi.org/10.3390/polym15010088

Chicago/Turabian Style

Zhong, Lixi, Quandong Huang, Jiali Zhang, and Ou Xu. 2023. "Low Power Consuming Mode Switch Based on Hybrid-Core Vertical Directional Couplers with Graphene Electrode-Embedded Polymer Waveguides" Polymers 15, no. 1: 88. https://doi.org/10.3390/polym15010088

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