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

Experimental demonstration of a flexible and high-performance mode-order converter using subwavelength grating metamaterials

Open Access Open Access

Abstract

Mode-order converters, transforming a given mode into the desired mode, have an important implication for the multimode division multiplexing technology. Considerable mode-order conversion schemes have been reported on the silicon-on-insulator platform. However, most of them can only convert the fundamental mode to one or two specific higher-order modes with low scalability and flexibility, and the mode conversion between higher-order modes cannot be achieved unless a total redesign or a cascade is carried out. Here, a universal and scalable mode-order converting scheme is proposed by using subwavelength grating metamaterials (SWGMs) sandwiched by tapered-down input and tapered-up output tapers. In this scheme, the SWGMs region can convert, TEp mode guided from a tapered-down taper, into a TE0-like-mode-field (TLMF) and vice versa. Thereupon, a TEp-to-TEq mode conversion can be realized by a two-step process of TEp-to-TLMF and then TLMF-to-TEq, where input tapers, output tapers, and SWGMs are carefully engineered. As examples, the TE0-to-TE1, TE0-to-TE2, TE0-to-TE3, TE1-to-TE2, and TE1-to-TE3 converters, with ultracompact lengths of 3.436-7.71 µm, are reported and experimentally demonstrated. Measurements exhibit low insertion losses of < 1.8 dB and reasonable crosstalks of < -15 dB over 100-nm, 38-nm, 25-nm, 45-nm, and 24-nm working bandwidths. The proposed mode-order converting scheme shows great universality/scalability for on-chip flexible mode-order conversions, which holds great promise for optical multimode based technologies.

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

1. Introduction

The promising technology termed as mode division multiplexing (MDM), simultaneously exploiting each eigenmode supported by multimode waveguides as an independent channel to encode/carry optical signals, offers a new dimension to significantly enhance transmission capacity requiring only a single operating wavelength for the on-chip photonic integrated circuits (PIC) [1,2]. In MDM systems, mode-order converters are one of the most fundamental components for generating and manipulating guided modes, which are capable of converting a given mode to a specific higher-order mode and vice versa [3]. In addition, mode converters can be also utilized in optical switches [4,5], biosensors [6], and cipher encryptions [7], showing broad applicability in silicon photonics. During the past few years, a considerable amount of structures for converting one mode to another on the silicon-on-insulator (SOI) platform have emerged, including the ones based on asymmetric directional couplers (ADCs) [8,9], multimode interference (MMI) couplers [10], adiabatic tapers [11,12], Mach-Zehnder interferometers (MZIs) [13], hybrid plasmonic waveguides (HPWs) [14], Bragg gratings [15], shallowly etched waveguides (SEWs) or say refractive index perturbations [1620], Y-junctions [21], inverse design (QR-code-like configuration) [22,23], and subwavelength slots [24]. ADCs consisting of two parallel and asymmetric nanowires are most commonly leveraged, owing to the designing simplicity and feasibility of exciting higher-order modes via the phase-matching condition, but their performance is limited by the narrow bandwidth and tight fabrication tolerance [25]. For mode-order converters using structures of Bragg gratings, Y-junctions, and MZI, an obvious drawback of large converting length can be observed. For example, lengths of the TE0-to-TE1 conversion, respectively, are 60-66.8 µm, > 100 µm, and ∼ 50 µm in Refs. [15,21,26], disfavoring the highly integrated density of on-chip PICs. As an alternative approach, inverse design can effectively reduce the converting length and be scalable for arbitrary mode conversions. For instance, the converting length for the TE0-to-TE1 conversion is only 2 µm in Ref. [22]. However, such a pixelated structure normally suffers from a large insertion loss, which is 2.2 dB (experiment) for the TE0-to-TE1 conversion in this report. Moreover, numerical calculations of the inverse design are rather time-consuming. Another alternative approach for realizing compact mode converters is to introduce refractive index perturbations in the mode converting region, but corresponding shallowly-etched trenches with acute angles require an additional etching step as well as a precise fabrication process [1620].

Besides the inherent drawbacks of the above-mentioned mode converters, two main challenges are further presented and highlighted: (1) low scalability for achieving more desired modes by using the same scheme, most converting schemes can be only extended to the TE0-to-TE2 mode conversion [1618]; (2) the lack of a direct mode conversion between higher-order modes, thus, a cascading strategy is required. For the first challenge, ADCs are potential solutions since any target higher-order modes can be excited by the phase-matching condition, but they are sensitive to variations of both fabrication dimension and wavelength, as mentioned above. Converters employing the method of inverse designs [22] can be expanded to realize TE0-to-TE3 but at the cost of a high insertion loss of 1.8 dB (simulated). In Ref. [27], researchers proposed highly scalable mode-order converters by applying tapered metal caps, which can generate any higher-order modes as desired. However, the insertion loss is as large as ∼2.8 dB for the TE0-to-TE4 mode conversion due to the direct absorption of metals, and the loss of the TE0-to-TE5 mode conversion is expected to increase to ∼3.5 dB. Moreover, the corresponding fabrication processes are rather complicated regarding metal depositions and might be incompatible with other components on the same wafer. As to the second challenge, much work so far has focused on the mode conversions from the fundamental mode to higher-order modes. Therefore, the mode conversions between higher-order modes are usually implemented by cascaded schemes [11,16], which means that the mode conversion can be efficiently accomplished in a single component. In this way, corresponding footprints are increased and the performance will deteriorate. For example, a cascade scheme is performed in Ref. [11] to realize the TE2-to-TE3 mode conversion, leading to a large conversion length of ∼55 µm. To tackle these two challenges, a scalable, flexible, and ultracompact mode-order converting scheme, realizing mode conversions from the fundamental mode to higher-order modes as well as mode conversions between higher-order modes, is greatly desired.

In recent years, subwavelength gratings (SWGs) metamaterials, composed of a periodic arrangement of dielectric nanowires with a period much smaller than the working wavelength, can surmount diffraction, function as a homogenous medium, and locally manipulate refractive index [28]. Quite high degrees of freedoms for researchers are provided by SWGs metamaterials, which enables the on-chip devices to be highly functional while occupying an ultracompact footprint and is used in a large range of applications [29], including mode-order converters [30]. Cheng et al. [31] designed a scalable TE0-to-TEq mode converting scheme based on SWGs structures, with q ranging from 1 to 3. The total device length of the TE0-to-TE3 mode converter is as large as 14.54 µm, which can be further reduced. Moreover, mode conversions between higher-order modes cannot be achieved by using the same scheme, showing low flexibility and scalability. Recently, we theoretically proposed a novel concept for an expandable and flexible TEp-to-TEq mode converting scheme (p < q, p = 0, 1, 2, 3&q = 1, 2, 3, 4), realized using a SWGs multimode waveguide and two tapers embedded at its anterior/posterior end [32]. With this mode converting scheme, the TEp mode launched from the input taper can be converted to the desired TEq mode of the output taper. However, the inter-modal crosstalks (CTs) are higher than -20 dB for all converters and insertion losses (ILs) are relatively high with up to 1.17 dB (@ 1.55 µm) for mode conversions between higher-order modes, thus, there is still room for improving performance and reducing device lengths.

In this work, we propose and experimentally demonstrate a significantly improved mode-order converting scheme with sandwiched SWGs and fully etched SWG-arrays at the end of output taper to boost the mode purity and thus improve the insertion loss as well as intermodal crosstalk. Furthermore, the SWGs-wires region is sandwiched by input and output tapers, which introduces a considerable reduction in device lengths. The measurement results, for proofs of concept, show that the fabricated TE0-to-TE1, TE0-to-TE2, TE0-to-TE3, TE1-to-TE2, and TE1-to-TE3 has low ILs/CTs of 1.8/-15 dB over a broadband of ∼ 100 nm, 38 nm, 25 nm, 45 nm, and 24 nm, with ultracompact device lengths ranged from 3.436 to 7.71 µm. The present mode converting scheme exhibits a low insertion loss, a low crosstalk, and an ultracompact footprint, together with high scalability and flexibility. Notably, the fabrication of this scheme is rather simple since only a one-step etching process is required.

2. Design and operating principle

Figures 1(a) and (b) show the three-dimensional (3D) schematic configuration of the proposed mode-order converting scheme, which consists of an input taper named IT, an output taper named OT, and a sandwiched SWG-wires region. The TEp-to-TEq mode-order converters are designed based on the SOI platform with a 220-nm silicon core (nSi = 3.476), a 2-µm buffer oxide layer (nSiO2 = 1.444), and a 2.2-µm silicon oxide up-claddings. Here, the IT is tapered down from wI to wt along a length of LIT, in which the positions of IT and its tip are denoted by Ld and wd, respectively, along lengthwise (z) and crosswise (x) directions. As to the OT started at z = 0, it is tapered up from wt to wO along a length of LOT, whose lateral position is denoted by ws. The SWG-wires region is sandwiched by the IT and OT, with a period number of Npq, a pitch length of Λ1, and a duty cycle of a11. Specifically, q fully etched SWG-arrays are introduced at the end of output taper OT, whose lengthwise and crosswise positions are denoted by L1, …, Lq and w1, …, wq, respectively. For all fully-etched SWG-arrays, they have the same width of wt, a pitch length of Λ2, and a duty cycle of a22, and for the q-th SWG-array, it has a period number of nq.

 figure: Fig. 1.

Fig. 1. 3D schematic configuration of the proposed TEp-to-TEq mode-order converting scheme, together with enlarged views of (a) input taper, namely IT, and (b) fully etched SWGs-arrays and the light propagation profile of TE2-to-TE3 mode conversion as an example. The whole device is covered by upper SiO2 claddings, which are not shown for clarity.

Download Full Size | PDF

Based on the configuration shown in Fig. 1, a mode-order conversion of TEp-to-TEq can be realized by adiabatic mode evolutions of TEp-to-TE0-like-mode-field (TLMF) and then TLMF-to-TEq, where TLMF is the acronym for TE0-like-mode-field, and the operating principle is demonstrated in detail as follows. Figures 2(a) and (b) show the SWGs homogenous behavior and a transverse cross-section of the corresponding taper/SWG-wires waveguide system, where the effective medium index (nswg) of the SWG-wires region for TE modes can be approximately given by well-known Rytov’s formula [28]:

$$n_{\textrm{swg}}^2 \approx \frac{a}{\Lambda } \cdot n_\textrm{1}^2 + \left( {1 - \frac{a}{\Lambda }} \right) \cdot n_\textrm{2}^2$$
where n1 and n2, respectively, are refractive indexes of silicon and silica in our case. For such a waveguide system, the taper width is changing with a constant ws. Here we perform the finite-difference frequency-domain (FDFD) method [33] to calculate the mode field profiles (Ex) for of TE0, TE1, TE2, and TE3 modes which are utilized as examples, as shown in Fig. 2(c). Obviously, one could find that a TLMF is gradually formed in the SWG-wires region as the taper is tapered down, for all modes. Hence, the optical mode TEp launched from the wider side of the taper is converted to a TLMF when the taper is tapered down gradually. As a reverse process, the TLMF can be naturally converted to initial TEp mode as the same taper is inversely tapered up, or converted to a new mode TEq by choosing a different taper in which the TEq mode is supported [32], according to the Lorentz reciprocity theorem [34]. In this way, we use the SWG-wires as a mode exchanging region by sandwiching it between two different tapers IT and OT, which support TEp and TEq modes respectively and effective indexes of these two modes are approximately matched. For the input TEp mode, it is converted to the TLMF from IT to the SWG-wires region via a mode evolution, as IT is tapered down. Next, this TLMF is converted to the desired TEq mode from the SWG-wires region to OT via a reverse mode evolution as OT is tapered up. Overall, a TEp-to-TEq mode conversion can be readily achieved by a two-step process of TEp-to-TLMF in IT/SWG-wires and TLMF-to-TEq in OT/SWG-wires region, as shown in Fig. 2(d), showing excellent flexibility and scalability of the proposed mode converting scheme. In particular, phase differences of modal components of the target TEq mode are further revised by introducing fully etched SWG-arrays at the end of the OT, thereby improving the mode purity with a lower crosstalk as well as insertion loss.

 figure: Fig. 2.

Fig. 2. Working principle of the present mode-order converting scheme: (a) 3D schematic view of the homogenous medium behavior of SWGs-wires waveguide; (b) Cross-section of the taper/SWG-wires waveguide system; (c) Mode field profiles (Ex) of TE0, TE1, TE2, and TE3 modes of the taper/SWG-wires waveguide system; (d) The two-step process of TEp-to-TLMF and then TLMF-to-TEq in the proposed mode-order converting scheme.

Download Full Size | PDF

3. Optimization, results and discussions

Based on the device design investigated in Section 2, the three-dimensional finite-difference time-domain (3D-FDTD) method [35] is performed to simulate/optimize performance and verify the feasibility of the proposed mode-order converting scheme. Here, two key figures of merit (FOM) are calculated, i.e., the insertion loss (IL, say conversion efficiency in some reports) and the inter-modal crosstalk (CT), and they are defined as

$$IL\textrm{ }(dB) ={-} 10 \cdot \log \frac{{P_q^\textrm{O}}}{{P_p^\textrm{I}}}$$
$$CT(\textrm{dB}) = \textrm{max} \{10 \cdot \log \frac{P_{\textrm{ohter}}^{\textrm{O}}}{P_q^{\textrm{O}}}\}$$
where $P_p^\textrm{I}$, $P_q^\textrm{O}$, and $P_{\textrm{other}}^\textrm{O}$ are powers of the given TEp mode at the input port, the desired TEp mode at the output port, and other unexpected modes at the output port, respectively. Following the TEp-to-TLMF-to-TEq mechanism, the particle swarm optimization (PSO) method [36] is particularly adopted here to optimize the designed mode-order converters. During optimization processes, some structural parameters are fixed as h = 220 nm, wt = 100 nm, a1 = 125 nm, Λ1 = 250 nm, a2 = 100 nm, and Λ2 = 200 nm. Meanwhile, searching ranges of 8 parameters of {wI, wO, Ld, LIT, LOT, wd, Npq, ws} for IT, OT, and sandwiched SWG-wires region, and 3 × q parameters of {L1, …, Lq, n1, …, nq, w1, …, wq) for fully etched SWG-arrays, are defined by the given lower and upper limits. Basically speaking, 8 + 3 × q geometry parameters in total need to be optimized for realizing the TEp-to-TEq mode conversion, which means that the optimization of attaining a target mode with a higher mode-order requires more computing time to complete. Fortunately, appropriate searching ranges for these parameters can be determined by the modal analysis of taper/SWG-wires waveguide system shown in Fig. 2(c), which can effectively reduce the computational complexity as well as computing time. Here, we take the TE1-to-TE2 mode converter as an example to show the determination of initial guessing ranges for geometry parameters. Input (IT) and output (OT) parts of the device are first considered, where searching ranges of wI and wO are set to be [0.9, 1.1] and [1.35, 1.5] µm, respectively. Under these width ranges, TE1 and TE2 modes, respectively, can be well supported by input and output waveguides with approximately phase-matching. To yield a compact design, the lower (upper) limit of Ld is chosen to be 0 (1) µm. Next, ws/LIT/LOT is controlled within a searching range of [2, 2.7]/[1.8, 2.8]/[4,5] µm for forming a clear TLMF, where Npq × Λ1 is approximately equal to LOT and thus the range of Npqis set to be [15,22]. Finally, w1, w2, L1, L2, n1, and n2, indicating the positions and period numbers of two SWG-arrays fully etched at the end of OT are studied. As is observed, there are three modal components for the target TE2 mode (can be seen as 3 beams), the phase differences of between two adjacent components can be tuned by introducing a SWG-array to change refractive index distributions for one beam [13], and thus two SWG-arrays are required for three modal components (beams). Generally speaking, q fully etched SWG-arrays are required for target q-th order mode. In this way, the searching ranges of w1 and w2 which denote the crosswise positions of two fully etched SWG-arrays are defined to be [0.35, 0.55] and [0.7, 1.1] µm, around wO/3 and 2wO/3 where two antiphase components are adjacent. Thus, we can find a pattern here indicating that w1, …, wq vary around wO/(q + 1), …, q × wO/(q + 1) for target TEq mode, respectively. Therefore, the initial guessing range for wq can be obtained by q × wO/(q + 1) ± Δw. In parallel, searching ranges of L1, L2, n1, and n2 are chosen based on OT as well as keeping the device compact, i.e., [2,5] µm, [3,5] µm, [1,5], and [1,5], respectively. After the determination of initial searching ranges for the parameters mentioned above, the optimization is conducted, where each generation size is chosen to be 25. Based on the appropriate guessing ranges, the FOM normally becomes saturated with 15 generations. If the FOM is less than satisfactory, a new run of optimization with updated searching ranges will be performed.

We optimize and experimentally demonstrate five design examples of the proposed mode-order converting scheme, i.e., TE0-to-TE1, TE0-to-TE2, TE0-to-TE3, TE1-to-TE2, TE1-to-TE3, and TE2-to-TE5 mode-order converters, and the optimized parameters are listed in Table 1.

Tables Icon

Table 1. Optimized parameters of TE0-to-TE1, TE0-to-TE2, TE0-to-TE3, TE1-to-TE2, TE1-to-TE3 converters, where the length of L and width of w are in units of µm, and n is period number with no units.

Figures 3(a)-(d) show the calculated ILs and CTs of TE0-to-TE1, TE0-to-TE2, TE0-to-TE3, TE1-to-TE2, TE1-to-TE3 mode-order converters, respectively, over a wavelength range from 1450 to 1650 nm. For mode-order conversions from fundamental mode to higher-order modes, it is noteworthy that considerably low CTs of -34.5 dB and -33.7 dB can be achieved around the central wavelength (≈ 1550 nm), respectively, for TE0-to-TE1 and TE0-to-TE2 mode converters, with low ILs of 0.35 dB and 0.27 dB. Meanwhile, IL/CT of 0.45 dB/-21.39 dB is obtained for the TE0-to-TE3 mode converter at the central wavelength of 1550 nm. As to mode-order conversion between the first-order mode and higher-order modes, one could find that performance of TE1-to-TE2 and TE1-to-TE3 mode converters is degraded compared with the TE0-to-TEq (q = 1, 2, 3) mode converters. Fortunately, CTs (ILs) are still lower than -20 dB (0.6 dB) around the central wavelength of 1550 nm for TE1-to-TE2 and TE1-to-TE3 mode converters, which are -25.7 dB (0.53 dB) and -22.1 dB (0.54 dB) @ 1550 nm, respectively. Overall, the bandwidths for achieving CTs of < -15 dB and ILs of < 0.8 dB are as large as 200 nm, 180 nm, 120 nm, 200 nm, and 160 nm, respectively, in theory, for TE0-to-TE1, TE0-to-TE2, TE0-to-TE3, TE1-to-TE2, and TE1-to-TE3 mode-order converters, which shows an ultra-broadband operation for the proposed mode-order converting scheme. Notably, the working bandwidths of the designed TE0-to-TEq (q = 1, 2, 3) mode converters are quite wider than converters using subwavelength metamaterials reported in Ref. [31], with much shorter device lengths in our proposed converters. More importantly, even for a CT of < -20 dB and an IL < 0.55 dB, the present TE0-to-TE1 and TE0-to-TE2 converters can exhibit bandwidths as large as 150 nm (1470-1620 nm) and 100 nm (1510-1610 nm), which are larger than those of previous reported mode-order converters [17], [22], [24], [27]. To clarify the necessity and effectiveness of SWG-arrays, The device performance without fully-etched SWG-arrays is calculated and summarized as follows: IL = 0.22 dB/CT = -22.58 dB, IL = 0.41 dB/CT = -15.6 dB, IL = 0.38 dB/CT = -13.95 dB, IL = 0.87 dB/CT = -10.38 dB, and IL = 1.3 dB/CT = -8.52 dB at 1.55 µm, for the TE0-to-TE1, TE0-to-TE2, TE0-to-TE3, TE1-to-TE2, and TE1-to-TE3 mode-order converters, respectively. One can see that CTs degrade significantly without fully-etched SWG-arrays.

 figure: Fig. 3.

Fig. 3. Simulated IL and CT versus wavelength of (a) TE0-to-TE1, (b) TE0-to-TE2, (c) TE0-to-TE3, (d) TE1-to-TE2, and (e) TE1-to-TE3 mode-order converters.

Download Full Size | PDF

According to the optimized parameters in Table 1, light propagation profiles of TE0-to-TE1, TE0-to-TE2, TE0-to-TE3, TE1-to-TE2, TE1-to-TE3, and TE2-to-TE5 mode-order converters (@ 1.55 µm) are calculated by the 3D-FDTD method and shown in Figs. 4(a)-(e), which verifies the mode converting functionality of the proposed scheme. The total device length for each converter is labeled for easy reading. It can be clearly seen that the input TEp mode (p = 0, 1) is first converted to a TLMF from IT to the sandwiched SWG-wires region via a mode evolution. Afterwards, this TLMF is converted to the target TEq mode (q = 1, 2, 3) from SWG-wires region to OT via a reverse mode evolution, where phase differences of modal components are revised by fully etched SWG-arrays at the end of OT. Essentially, the TLMF here serves as an exchanging mode between given TEp mode and target TEp mode, thereby flexible mode-order conversion can be realized as long as both input TEp mode in IT and output TEq mode in IT can form an effective TLMF in the sandwiched SWG-wires region, with appropriate geometry parameters. In particular, we study and simulate a mode-order converter for TE2-to-TE5 mode conversion to validate the scalability as well as flexibility of the present mode conversion scheme, as shown in Fig. 4(f). It can be seen that the input TE2 mode can be converted to the output TE5 mode by a two-step process of TE2-to-TLMF and then TLMF-to-TE5, with an IL/CT of 0.98 dB/-16.1 dB at 1550 nm. Remarkably, the total length of this converter is only 12.271 µm, indicating an ultracompact mode-order converting scheme.

 figure: Fig. 4.

Fig. 4. Simulated electric filed Ex propagation profiles of (a) TE0-to-TE1, (b) TE0-to-TE2, (c) TE0-to-TE3, (d) TE1-to-TE2, (e) TE1-to-TE3, and (f) TE2-to-TE5 mode-order converters, at the wavelength of 1.55 µm.

Download Full Size | PDF

4. Fabrication and characterization

To verify the proposed theoretical scheme and corresponding numerical simulation results, the TE0-to-TE1, TE0-to-TE2, TE0-to-TE3, TE1-to-TE2, TE1-to-TE3 mode-order converters are fabricated on an SOI platform with a 220-nm-thick Si device layer and a 2-µm buried oxide (BOX) layer by using electron beam lithography (EBL) and reactive ion etching (RIE) processes. The patterns of mode-order converters are defined into a material that is sensitive to electron beam exposure (EBL) and an anisotropic inductively coupled plasma (ICP) RIE etching process is subsequently carried out on the substrate to transfer the patterns into the underlying silicon device layer. As a final step, a SiO2 upper-cladding layer of 2.2-µm is deposited using plasma-enhanced chemical vapor deposition (PECVD).

Figures 5(a)-(e) show the microscope images of the measure schemes for TE0-to-TE1, TE1-to-TE2, TE0-to-TE3, TE1-to-TE3, and TE0-to-TE2 mode conversions, respectively. Meanwhile, Figs. 5(f)-(j) give the corresponding scanning electron microscopy (SEM) images of TE0-to-TE1, TE1-to-TE2, TE0-to-TE3, TE1-to-TE3, and TE0-to-TE2 mode-order converters in five device sets, respectively. For the measurement of each mode-order converter, a device set and an identical reference set without the mode-order converter fabricated on the same chip for comparison and normalization, are cooperatively carried out to characterize all output modes of each mode-order converter. Here, the TEi multiplexer (MUX) can couple, the TE0 mode launched from input port named as I-TEi, to TEi mode in the bus waveguide, and the TEi DeMux can convert the TEi mode to the TE0 mode at output port named as O-TEi. From the reference set, the transmissions of the TEi (De)Mux with a connecting grating coupler can be obtained by feeding the light into I-TEi port, measuring the O-TEi port and then dividing the measurement result by two (i = 0 for attaining the transmission of just the grating coupler), which can be used for normalizations. Thereupon, in the device set, the transmissions for all output modes of each mode-order converter are obtained by feeding light at I-TEi port, measuring the O-TEi port, and then normalizing the transmissions of corresponding TEi (De)Muxs and connecting grating couplers.

 figure: Fig. 5.

Fig. 5. Microscope images of the measure schemes for: (a) TE0-to-TE1, (b) TE1-to-TE2, (c) TE0-to-TE3, (d) TE1-to-TE3, and (e) TE0-to-TE2 mode conversions, respectively. Corresponding pseudocolor SEM images of the (f) TE0-to-TE1, (g) TE1-to-TE2, (h) TE0-to-TE3, (i) TE1-to-TE3, (j) TE0-to-TE2 mode-order converters, respectively.

Download Full Size | PDF

Here, a tunable laser (Santec TSL-710) and an optical power meter (Santec MPM-210) are employed for measuring the transmittance spectra of the fabricated TE0-to-TE1, TE0-to-TE2, TE0-to-TE3, TE1-to-TE2, and TE1-to-TE3 mode-order converters, where the measured and normalized transmittances for all output modes of each mode converter are shown in Figs. 6(a)-(e). For the TE0-to-TE1 converter from Fig. 6(a), one has IL < 1.8 dB and CT < -15.8 dB for the 100 nm bandwidth ranged from 1480 nm to 1580 nm, showing an efficient and broadband mode conversion. At the central wavelength of 1550 nm, the TE0-to-TE1 converter exhibits a low loss and small crosstalk of IL = 1.27 dB and CT = -21.31 dB. From Fig. 6(b), one can obverse that TE0-to-TE2 converter can keep the IL/CT lower than 1.8 dB/-15 dB in the wavelength range of 1523-1561 nm, which shows a 38 nm operating bandwidth. If IL < 1.8 dB/CT < -10 dB is required for a relaxed requirement, the bandwidth can be extended to 81 nm (1480-1561 nm). As to the TE0-to-TE3 converter, one can see that the IL is lower than 1.8 dB for the TE0-to-TE3 mode conversion and the CT is below -15 dB within the wavelength range of 1538-1563 nm, i.e., a 25 nm working bandwidth, as shown in Fig. 6(c). Besides the mode conversions from fundamental mode to higher-order modes, flexible mode conversions between higher-order modes are also realized by using the proposed mode converting scheme. From Figs. 6(d) and (e), the TE1-to-TE2 and TE1-to-TE3 converters, respectively, have 45 nm (1515-1560 nm) and 25 nm (1528-1552 nm) operating bandwidths for simultaneously keeping IL < 1.8 dB and CT < -15 dB. Notably, the TE1-to-TE2 converter can even exhibit a lower crosstalk of CT < -17 dB within that wavelength range, showing a broadband mode conversion between higher-order modes. For the central wavelength 1550 nm, one has IL = 0.73 dB/CT = 21.46 dB for TE1-to-TE2 converter and IL = 1.7 dB/CT = 15.86 dB for TE1-to-TE3 converter, respectively.

 figure: Fig. 6.

Fig. 6. Measured transmittance TTEi spectra for all output modes of the fabricated (a) TE0-to-TE1, (b) TE0-to-TE2, (c) TE0-to-TE3, (d) TE1-to-TE2, and (e) TE1-to-TE3 mode-order converters, respectively.

Download Full Size | PDF

Table 2 gives a brief summary and comparison with previously reported mode-order converters. For those mode-order converters with wide working bandwidths, their footprints are large [16,31], and some of them require an extra etching step for shallowly-etched trenches. As to the mode-order converters with ultracompact device lengths, the device performance is limited, including large insertion losses and high crosstalks. By contrast, the present mode-order converters show large operating bandwidths for achieving low ILs (< 1.8 dB) as well as CTs (< -15 dB), in ultracompact device lengths ranging from 3.436 to 7.71 µm. Furthermore, different from previous reports in which a cascade strategy is needed for performing the mode conversion between higher-order modes, the proposed scheme can realize not only the mode conversion from fundamental mode to a higher-order mode, but also the mode conversion between higher-order modes in a single device, showing great scalability and flexibility.

Tables Icon

Table 2. Comparison of several scalable mode-order converters at the wavelength of 1.55 µm

5. Conclusion

In conclusion, we have proposed a universal and scalable mode-order converting scheme by using SWGs metamaterials which are sandwiched by tapered-down input and tapered-up output tapers. With this design, TEq mode can be transformed into a TLMF and vice versa according to optical reciprocity. By using TLMF as an exchange between input TEp and output TEq modes, the mode-order conversion can be realized through a two-step process of TEp-to-TLMF in IT/SWG-wires and TLMF-to-TEq in OT/SWG-wires region. Based on this mechanism, five proof-of-concept mode-order converters are designed and experimentally demonstrated, including TE0-to-TE1, TE0-to-TE2, TE0-to-TE3, TE1-to-TE2, and TE1-to-TE3 converters with ultracompact lengths ranged from 3.436 to 7.71µm. Compared with previous reports, the proposed scheme not only can perform mode conversion from fundamental mode to higher-order modes but also can implement mode conversions between higher-order modes, showing great universality and scalability. In theory, TE0-to-TE1, TE0-to-TE2, TE0-to-TE3, TE1-to-TE2, and TE1-to-TE3 converters show low ILs (CTs) of 0.35dB (-34.5dB), 0.28dB (-28.6dB), 0.45dB (-21.39dB), 0.53dB (-25.7dB), and 0.54dB (-22.1dB) respectively, at 1550nm. From measurement results, working bandwidths for keeping IL <1.8dB (CT < -15dB) are 100nm, 38nm, 25nm, 45nm, and 24nm, respectively, for TE0-to-TE1, TE0-to-TE2, TE0-to-TE3, TE1-to-TE2, and TE1-to-TE3 mode-order converters. With advantages of the universality, scalability, an ultracompact size, and high performance, we believe that the present mode-order converting scheme will find its applications in ultrahigh-density MDM systems and other multimode based technologies.

Funding

National Natural Science Foundation of China (11574046, 12004092); Natural Science Foundation of Jiangsu Province (BK20211163); Scientific Research Foundation of the Graduate School of Southeast University (YBPY2135).

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. D. Dai and J. E. Bowers, “Silicon-based on-chip multiplexing technologies and devices for Peta-bit optical interconnects,” Nanophotonics 3(4-5), 283–311 (2014). [CrossRef]  

2. D. Dai, C. Li, S. Wang, H. Wu, Y. Shi, Z. Wu, S. Gao, T. Dai, H. Yu, and H. K. Tsang, “10-channel mode (de)multiplexer with dual polarizations,” Laser Photonics Rev. 12(1), 1700109 (2018). [CrossRef]  

3. Z. Chen, Y. Zhu, X. Ruan, Y. Li, Y. Li, and F. Zhang, “Bridged coupler and oval mode converter based silicon mode division (de) multiplexer and terabit WDM-MDM system demonstration,” J. Lightwave Technol. 36(13), 2757–2766 (2018). [CrossRef]  

4. C. Sun, W. Wu, Y. Yu, G. Chen, X. Zhang, X. Chen, D. Thomson, and G. T. Reed, “De-multiplexing free on-chip low-loss multimode switch enabling reconfigurable inter-mode and inter-path routing,” Nanophotonics 7(9), 1571–1580 (2018). [CrossRef]  

5. H. Jia, H. Chen, T. Wang, H. Xiao, G. Ren, A. Mitchell, J. Yang, and Y. Tian, “Multi-Channel Parallel Silicon Mode-Order Converter for Multimode On-Chip Optical Switching,” IEEE J. Select. Topics Quantum Electron. 26(2), 1–6 (2020). [CrossRef]  

6. P. Measor, S. Kühn, E. J. Lunt, B. S. Phillips, A. R. Hawkins, and H. Schmidt, “Multi-mode mitigation in an optofluidic chip for particle manipulation and sensing,” Opt. Express 17(26), 24342–24348 (2009). [CrossRef]  

7. H. Jia, S. Yang, T. Zhou, L. Zhang, T. Wang, H. Chen, J. Yang, and L. Yang, “Mode-Oriented Permutation Cipher Encryption and Passive Signal Switching Based on Multiobjective Optimized Silicon Subwavelength Metastructures,” ACS Photonics 7(8), 2163–2172 (2020). [CrossRef]  

8. J. Wang, S. He, and D. Dai, “On-chip silicon 8-channel hybrid (de)multiplexer enabling simultaneous mode- and polarization-division- multiplexing,” Laser & Photonics Reviews 8(2), L18–L22 (2014). [CrossRef]  

9. Y. Xiong, D. X. Xu, J. H. Schmid, P. Cheben, S. Janz, and W. N. Ye, “Broadband two-mode multiplexer with taper-etched directional coupler on SOI platform,” in 11th International Conference on Group IV Photonics (2014), pp. 39–40.

10. J. Leuthold, J. Eckner, E. Gamper, P.-A. Besse, and H. Melchior, “Multimode interference couplers for the conversion and combining of zero- and first-order modes,” J. Lightwave Technol. 16(7), 1228–1239 (1998). [CrossRef]  

11. D. Chen, X. Xiao, L. Wang, Y. Yu, W. Liu, and Q. Yang, “Low-loss and fabrication tolerant silicon mode-order converters based on novel compact tapers,” Opt. Express 23(9), 11152–11159 (2015). [CrossRef]  

12. Y. Ding, J. Xu, F. Da Ros, B. Huang, H. Ou, and C. Peucheret, “On-chip two-mode division multiplexing using tapered directional coupler-based mode multiplexer and demultiplexer,” Opt. Express 21(8), 10376–10382 (2013). [CrossRef]  

13. Y. Huang, G. Xu, and S. Ho, “An ultracompact optical mode order converter,” IEEE Photon. Technol. Lett. 18(21), 2281–2283 (2006). [CrossRef]  

14. Z. Cheng, J. Wang, Z. Yang, H. Yin, W. Wang, Y. Huang, and X. Ren, “Broadband and high extinction ratio mode converter using the tapered hybrid plasmonic waveguide,” IEEE Photonics J. 11(3), 1–8 (2019). [CrossRef]  

15. R. Xiao, Y. Shi, J. Li, P. Dai, Y. Zhao, L. Li, J. Lu, and X. Chen, “On-chip mode converter based on two cascaded Bragg gratings,” Opt. Express 27(3), 1941–1957 (2019). [CrossRef]  

16. L. Hao, R. Xiao, Y. Shi, P. Dai, Y. Zhao, S. Liu, J. Lu, and X. Chen, “Efficient TE-polarized mode-order converter based on high-index-contrast polygonal slot in a silicon-on-insulator waveguide,” IEEE Photonics J. 11(2), 1–10 (2019). [CrossRef]  

17. H. Wang, Y. Zhang, Y. He, Q. Zhu, L. Sun, and Y. Su, “Compact silicon waveguide mode converter employing dielectric metasurface structure,” Adv. Opt. Mater. 7(4), 1801191 (2018). [CrossRef]  

18. C.-C. Huang and C.-C. Huang, “Theoretical analysis of mode conversion by refractive-index perturbation based on a single tilted slot on a silicon waveguide,” Opt. Express 28(13), 18986–18998 (2020). [CrossRef]  

19. L. Liu, Y. Xu, L. Wen, Y. Dong, B. Zhang, and Y. Ni, “Design of a compact silicon-based TM-polarized mode-order converter based on shallowly etched structures,” Appl. Opt. 58(33), 9075–9081 (2019). [CrossRef]  

20. D. Ohana and U. Levy, “Mode conversion based on dielectric metamaterial in silicon,” Opt. Express 22(22), 27617–27631 (2014). [CrossRef]  

21. J. B. Driscoll, R. R. Grote, B. Souhan, J. I. Dadap, M. Lu, and R. M. Osgood, “Asymmetric Y junctions in silicon waveguides for on-chip mode-division multiplexing,” Opt. Lett. 38(11), 1854–1856 (2013). [CrossRef]  

22. T. Wang, H. Guo, H. Chen, J. Yang, and H. Jia, “Ultra-compact reflective mode converter based on a silicon subwavelength structure,” Appl. Opt. 59(9), 2754–2758 (2020). [CrossRef]  

23. H. Jia, H. Chen, J. Yang, H. Xiao, W. Chen, and Y. Tian, “Ultra-compact dual-polarization silicon mode-order converter,” Opt. Lett. 44(17), 4179–4182 (2019). [CrossRef]  

24. Y. Zhao, X. Guo, Y. Zhang, J. Xiang, K. Wang, H. Wang, and Y. Su, “Ultra-compact silicon mode-order converters based on dielectric slots,” Opt. Lett. 45(13), 3797–3800 (2020). [CrossRef]  

25. Y. He, Y. Zhang, Q. M. Zhu, S. H. An, R. Y. Cao, X. H. Guo, C. Y. Qiu, and Y. K. Su, “Silicon High-Order Mode (De)Multiplexer on Single Polarization,” J. Lightwave Technol. 36(24), 5746–5753 (2018). [CrossRef]  

26. C. Sun, Y. Yu, G. Chen, and X. Zhang, “Integrated switchable mode exchange for reconfigurable mode-multiplexing optical networks,” Opt. Lett. 41(14), 3257–3260 (2016). [CrossRef]  

27. Y. Xu, L. Liu, X. Hu, Y. Dong, B. Zhang, and Y. Ni, “Scalable silicon-based mode-order converters assisted by tapered metal,” Optics & Laser Technology 151, 108028 (2022). [CrossRef]  

28. J. M. Luque-González, A. Sánchez-Postigo, A. Hadij-ElHouati, A. Moñux, J. Pérez, J. H. Schmid, P. Cheben, Í Molina-Fernández, and R. Halir, “A review of silicon subwavelength gratings: building break-through devices with anisotropic metamaterials,” Nanophotonics 10(11), 2765–2797 (2021). [CrossRef]  

29. Z. Li C, M. Xu, H. Xu, Y. Tan, Y. Shi, and D. Dai, “Subwavelength silicon photonics for on-chip mode-manipulation,” PhotoniX. 2(1), 11 (2021). [CrossRef]  

30. M. Lu, C. Deng, Y. Sun, D. Wang, L. Huang, P. Liu, D. Lin, W. Cheng, G. Hu, T. Lin, B. Yun, and Y. Cui, “Compact and broadband silicon mode-order converter using bricked subwavelength gratings,” Opt. Express 30(14), 24655–24666 (2022). [CrossRef]  

31. Z. Cheng, J. Wang, Z. Yang, L. Zhu, Y. Yang, Y. Huang, and X. Ren, “Sub-wavelength grating assisted mode order converter on the SOI substrate,” Opt. Express 27(23), 34434–34441 (2019). [CrossRef]  

32. Z. Guo, S. Wu, and J. Xiao, “Compact and flexible mode-order converter based on mode transitions composed of asymmetric tapers and subwavelength gratings,” J. Lightwave Technol. 39(17), 5563–5572 (2021). [CrossRef]  

33. J. Xiao, H. Ni, and X. Sun, “Full-vector mode solver for bending waveguides based on the finite-difference frequency-domain method in cylindrical coordinate system,” Opt. Lett. 33(16), 1848–1850 (2008). [CrossRef]  

34. D. Jalas, A. Petrov, M. Eich, W. Freude, S. Fan, Z. Yu, R. Baets, M. Popovic, A. Melloni, J. D. Joannopoulos, M. Vanwolleghem, C. R. Doerr, and H. Renner, “What is - and what is not - an optical isolator,” Nat. Photonics 7(8), 579–582 (2013). [CrossRef]  

35. D. M. Sullivan, Electromagnetic Simulation Using the FDTD Method (Wiley, 2013).

36. R. Poli, J. Kennedy, and T. Blackwell, “Particle swarm optimization,” Swarm Intell. 1(1), 33–57 (2007). [CrossRef]  

Data availability

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

Cited By

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

Alert me when this article is cited.


Figures (6)

Fig. 1.
Fig. 1. 3D schematic configuration of the proposed TEp-to-TEq mode-order converting scheme, together with enlarged views of (a) input taper, namely IT, and (b) fully etched SWGs-arrays and the light propagation profile of TE2-to-TE3 mode conversion as an example. The whole device is covered by upper SiO2 claddings, which are not shown for clarity.
Fig. 2.
Fig. 2. Working principle of the present mode-order converting scheme: (a) 3D schematic view of the homogenous medium behavior of SWGs-wires waveguide; (b) Cross-section of the taper/SWG-wires waveguide system; (c) Mode field profiles (Ex) of TE0, TE1, TE2, and TE3 modes of the taper/SWG-wires waveguide system; (d) The two-step process of TEp-to-TLMF and then TLMF-to-TEq in the proposed mode-order converting scheme.
Fig. 3.
Fig. 3. Simulated IL and CT versus wavelength of (a) TE0-to-TE1, (b) TE0-to-TE2, (c) TE0-to-TE3, (d) TE1-to-TE2, and (e) TE1-to-TE3 mode-order converters.
Fig. 4.
Fig. 4. Simulated electric filed Ex propagation profiles of (a) TE0-to-TE1, (b) TE0-to-TE2, (c) TE0-to-TE3, (d) TE1-to-TE2, (e) TE1-to-TE3, and (f) TE2-to-TE5 mode-order converters, at the wavelength of 1.55 µm.
Fig. 5.
Fig. 5. Microscope images of the measure schemes for: (a) TE0-to-TE1, (b) TE1-to-TE2, (c) TE0-to-TE3, (d) TE1-to-TE3, and (e) TE0-to-TE2 mode conversions, respectively. Corresponding pseudocolor SEM images of the (f) TE0-to-TE1, (g) TE1-to-TE2, (h) TE0-to-TE3, (i) TE1-to-TE3, (j) TE0-to-TE2 mode-order converters, respectively.
Fig. 6.
Fig. 6. Measured transmittance TTEi spectra for all output modes of the fabricated (a) TE0-to-TE1, (b) TE0-to-TE2, (c) TE0-to-TE3, (d) TE1-to-TE2, and (e) TE1-to-TE3 mode-order converters, respectively.

Tables (2)

Tables Icon

Table 1. Optimized parameters of TE0-to-TE1, TE0-to-TE2, TE0-to-TE3, TE1-to-TE2, TE1-to-TE3 converters, where the length of L and width of w are in units of µm, and n is period number with no units.

Tables Icon

Table 2. Comparison of several scalable mode-order converters at the wavelength of 1.55 µm

Equations (3)

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

n swg 2 a Λ n 1 2 + ( 1 a Λ ) n 2 2
I L   ( d B ) = 10 log P q O P p I
C T ( dB ) = max { 10 log P ohter O P q O }
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.