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

Theoretical analysis of mode conversion by refractive-index perturbation based on a single tilted slot on a silicon waveguide

Open Access Open Access

Abstract

We propose a compact mode converter operating at the mid-infrared wavelength of 3.4 µm, comprising an etched parallelogram slot filled with silicon nitride on a silicon-on-calcium fluoride platform. The tilted slot introduces transverse and longitudinal index perturbations on the waveguide eigenmodes, achieving mode conversion in the propagation direction. Differing from previous reports using massive parameter sweep, we provide analytical formulas to determine geometry parameters by considering the modified phase-matching condition and the profiles of coupling coefficient of coupled-mode theory. Rigorous 3D numerical examples demonstrate the transverse electric (TE)0-to-TE1, TE0-to-TE2, TE0-to-TE3, and TE0-to-TE4 converters to achieve conversion efficiencies (inter-modal crosstalk [CT] values) of >92.7% (<−27 dB), >91.7% (<−16 dB), >88.2% (<−13 dB), and >75.8% (<−10 dB), respectively, with a total transmitted power of >93%. Converter device lengths range from 16.84 to 24.61 µm for TE0-to-TE1 to TE0-to-TE4, respectively. Over a broadband wavelength of 100 nm, the conversion efficiency, power transmission, and maximum inter-modal CT are almost >80%, >90%, and <−10 dB, respectively. Also, the fabrication tolerance of the proposed structure is addressed. The proposed model can not only realize arbitrary mode-order conversion but extend to other wavelength bands. To validate the feasibility of our model, the numerical results of our device operating at the wavelength of 1.55 µm are also offered and compared with those of other reports. The proposed idea may pave a new approach to designing mode converters with arbitrary geometries.

© 2020 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Mid-infrared (mid-IR) silicon (Si) photonics, spanning wavelengths from 2 to 20 µm [1], has attracted much interest in recent years due to its potential application in various fields such as absorption spectroscopy, chemical and biological sensing, environmental monitoring, and free-space communication [14]. Initial applications have targeted photonic integrated circuits and optical communications around the near-IR wavelength of 1.55 µm. However, driven by the demand for high capacity data transmission in modern networking, the near-IR regime has begun to face the so-called “capacity crunch.” To address this pressing issue, the scope of Si photonics has been expanded to include the short mid-IR range [2,5,6]. In recent years, many Si photonics components, including waveguides [7], electro-optic modulators [8], directional couplers [9], multimode interferometers [10], ring resonators [11], grating couplers [1214], and polarizing beam splitters [15,16], have been proposed and demonstrated. Compared with near-IR photonics using silicon-on-insulator (SOI) structures based on mature complementary metal-oxide semiconductor compatible technology, mid-IR band Si photonics has thus far received much less attention. This is because the silicon dioxide (SiO2) in the SOI platform has strong absorption loss in the mid-IR wavelength range of λ = 2.6–2.9 µm and around λ > 3.6 µm [17,18]. Hence, choosing an appropriate platform for operating in the mid-IR band is essential to building low-loss mid-IR photonic components. At present, several different transparent platforms, including silicon-on-sapphire, silicon-on-nitride, silicon-on-calcium-fluoride (SOCF), and silicon-on-lithium-niobate, have been used in the mid-IR spectrum; the design and fabrication of each platform have been reviewed and discussed previously in [14]. Additionally, the Germanium-based integrated photonics has also reported for the extension of the transparent window up to 15 µm [19]. In addition to extending operating wavelengths, advanced multiplexing technologies [2024] can further increase data traffic volumes for optical information processing and sensing. Among them, mode-division multiplexing (MDM) [2124], which encodes each optical signal on the different spatial modes of multimode waveguides, allows significant transmission capacity scaling. The key component to realizing MDM technology is a mode converter able to transform given spatial modes into desired modes. There are three approaches to mode conversion: Phase matching [2534], beam shaping [35,36], and coherent interference [3739]. The most popular phase-matching technique is to provide the phase difference and field overlap between the given and target modes by perturbing refractive-index profiles along the transverse and propagation directions.

Ding et al. [25] adopted a tapered directional coupler-based TE0-to-TE1 converter with relatively long device length of 50 µm for the operating wavelength of 1550 nm on a SOI platform, and the measured insertion loss (IL) and inter-modal crosstalk (CT) are 0.3 and <−16 dB, respectively. In [26,28], continuous graded and rectangular periodic indices were introduced to a Si waveguide along the transverse and propagation directions, respectively. A transverse electric (TE) mode converter between the TE0 and TE1 modes, with device length of around 23 µm, conversion efficiency of around 96.5%, and power transmissions of around 92%, were numerically and experimentally demonstrated. Chen et al. [27] designed various mode-order converters based on asymmetric tapered structures with device length of around 20 µm. Their simulation results show high conversion efficiency (CE) of around 0.1 dB over a wavelength of 1520 to 1580 nm. In [29], using deposited phase arrays of Si nanorods on lithium niobate waveguides, mode converter device lengths between TE0 and TE1 modes (with power transmission 92.9%) and between TE0 and TE2 modes (with power transmission 82.9%) were 5.4 and 8.4 µm, respectively. However, no clear design principle was presented in this work, and the fabrication precision of the phase arrays can be difficult to control due to size variation between the Si nanorods. A tilted subwavelength periodic slot engraved on a Si waveguide was proposed for converting TE0 to TEj modes, with j in the range 1 to 4 [31]. Device lengths were about 5 to 10 µm, with conversion loss around 1.2 dB and inter-modal crosstalk (CT) around −10 dB. However, the smallest feature size of about 50 nm is difficult to fabricate precisely, requiring moderate relaxation of geometry conditions. In [32], the researchers proposed an on-chip mode converter, achieving forward conversion using two cascaded Bragg reflection processes instead of a general backward conversion process, with the bandwidth and central wavelength adjustable according to their theoretical analysis. However, the power transmission spectrum of the converter between TE0 and TE2 modes were only around 20 nm. TE-polarized mode converters based on deeply etched polygonal slots [33] have been reported based on the theoretical analysis of transformation optics [40,41]. The device length between TE0 and TE1 modes was around 24 µm, with a conversion efficiency of around 97.6% and crosstalk of <−20 dB over the wavelength range of around 100 nm. However, different polygonal slots are required for scaling the design to different mode-order converters. For beam shaping [35,36] and coherent interference [3436] techniques, the main challenges are the large device footprint and limited operating bandwidth. Nevertheless, the mode converters mentioned above [2539] all operate at the telecommunication wavelength of 1.55 µm. To the best of our knowledge, mode converters operating in the mid-IR range have yet to be studied.

In this article, we propose a mode converter operating at the mid-IR wavelength of 3.4 µm, comprising a single etched parallelogram slot filled with silicon nitride (Si3N4) on a Si waveguide. To avoid the SiO2 substrate’s high absorption loss [17], we use a crystalline calcium fluoride (CaF2) substrate because of its transparency in the spectral range from visible to mid-IR wavelengths of 8 µm, and its low refractive index of around 1.4 at a wavelength of 3.4 µm [42]. Additionally, Si photonic devices have been experimentally integrated on a CaF2 substrate [43]. Compared with other mid-IR photonic platforms, the SOCF platform possesses the higher refractive-index contrast, making the proposed device more compact. In addition to using the conventional solid materials, a suspended waveguide platform was reported to build the mid-IR photonic devices operating at longer wavelengths for sensing applications in the molecular fingerprint [44]. Notably, the suspended waveguide can offer an even higher refractive-index contrast [45]. Differing from published reports relying on massive parameter sweeps [2934], we provide analytical formulas to determine the proposed structure’s geometry parameters by considering the phase-matching condition (PMC) and the profiles of the coupling coefficient in coupled-mode theory (CMT) [46]. From a practical perspective, the relatively simple geometry of the proposed design poses easy fabrication challenges, making the fabrication processes feasible and easy. Numerical examples are provided to demonstrate high performance in terms of conversion efficiency, power transmission, and inter-modal crosstalk. The spectral response within the 3.34 to 3.46 µm range and fabrication tolerances are also analyzed. Notably, the numerical results of the present device operating at the wavelength of 1.55 µm are also offered to validate the feasibility.

2. Design principles of the proposed mode converter

Figure 1 shows a 3D schematic diagram and top view of the proposed mode converter, comprising an etched parallelogram slot filled with Si3N4 on a Si waveguide, deposited on a CaF2 substrate. The operating wavelength is at the mid-IR wavelength of λ = 3.4 µm. Instead of using SiO2 in the slot and substrate, as widely used in near-IR systems, we used Si3N4 and CaF2 for the slot and substrate, respectively, because of their transparency at the present operating wavelength [1,4,18]. Along the transverse (x-direction) and propagation (z-direction) directions, the tilted slot with the angle θ and depth het induces dielectric perturbations in the Si waveguide, achieving mode conversion through inter-modal coupling. Before addressing the theoretical aspect of the proposed mode converter, we outline the fabrication steps. A patterned hard mask of the present Si waveguide of width WSi is prepared using electron beam lithography in advance. (1) Depositing a positive photoresist (PR) thin film of height hSi on a CaF2 substrate to define the Si waveguide by the preceding mask, a PR exposure with ultraviolet light, development, and an etching process. (2) Adopting focused ion beam (FIB) milling [47] with high beam currents to directly write the tilted slot region of depth het. (3) A Si3N4 layer of height het is deposited using chemical vapor deposition in the slot. Subsequently, a chemical mechanical polishing is used to flatten the plane of the slot. In the step 2, differing from the optical lithography with larger spot size of few tens of nanometers, a maskless FIB with beam size of <5 nm in diameter can relieve the fabrication tolerance of acute angles A to D of parallelogram region, effectively reducing the length of transverse direction of the acute angles. Compared with the FIB technique, the scanning probe lithography has emerged as an alternative type of lithography with sub-10-nm resolution despite the slow writing speed [48,49]. The coupling mechanism of the proposed structure can be approximately described by the CMT dealing with an electromagnetic wave propagating along a perturbed dielectric structure.

 figure: Fig. 1.

Fig. 1. (a) 3D schematic diagram and (b) top view of the proposed mode converter.

Download Full Size | PDF

For a given dielectric perturbation of Δɛ (x, y, z) of a waveguide, the amplitude conversion between the input mode p and output (converted) mode q for the condition of co-directional coupling is expressed by the following [46]:

$$\frac{d}{{dz}}{A_p}(z) ={-} \sum\limits_q {{\kappa _{pq}}{A_q}(z){\kern 1pt} {e^{i({{\beta_p} - {\beta_q}} )z}}} {\kern 1pt} ,$$
where
$${\kappa _{pq}} = \frac{{\omega {\varepsilon _0}}}{4}\int {{\textbf E}_p^\ast (x,y)} \cdot \Delta \varepsilon (x,y,z) \cdot {{\textbf E}_q}(x,y)\,dxdy.$$
ω is the angular frequency, ɛ0 is the permittivity in vacuum, βp and βq are the propagation constants of modes p and q, respectively, and Ep* (x, y) and Eq(x, y) are the electric profiles of the p-th and q-th modes, respectively, where * denotes the complex conjugate. The term κpq, referred to as the “coupling coefficient,” reflects the coupling strength between the p-th and q-th modes. Practically, only two modes are strongly coupled while near the resonant coupling condition. To satisfy the PMC along the propagation direction, we need to compensate the mismatch (i.e., βpβq) of the propagation constants of modes p and q, which is shown in the exponential term of Eq. (1), to achieve maximum power transfer. Therefore, the PMC can be deduced as follows:
$$\Delta \beta = {\beta _p} - {\beta _q} - {\beta _c} = 0,$$
where βc denotes the compensating propagation constant (i.e., to complement the difference βpβq) that results from perturbations during the conversion. The role of βc is indeed the same as the grating number in conventional codirectional-coupling gratings. Theoretically, the grating number is K = m2π/Λ for periodic perturbations [46], where Λ is a period and m is an integer representing the mth Fourier component of the dielectric perturbation. In this paper, a tilted slot with an aperiodic perturbation was used to accomplish mode-order conversion. We adopted the similar relation of βc = 2π/Lec for the proposed aperiodic perturbation as the grating number for periodic perturbations, where Lec is called the effective coupling length. Note that the Lec is larger for converting to the lower-order mode than that to the higher-order one. On the other hand, choosing a multiple of βc also satisfies the PMC. Accordingly, the relation of Lec = L/q is set in this paper for converting mode p to mode q, where L is the practical device length of TE0q converter and can be expressed by the follows:
$$L = q\frac{{2\pi }}{{{\beta _p} - {\beta _q}}}.$$
Once the device length is determined for any mode conversion, we can determine the analytical formulas of geometry parameters by numerically calculating the κpq.

3. Numerical demonstrations of performance and fabrication tolerance

The relative permittivities of Si, Si3N4, and CaF2 at a wavelength of λ = 3.4 µm are nSi = 3.4293, nSi3N4 = 1.9304, and nCaF2 = 1.4148 [42], respectively. The thickness of the Si waveguide and depth of the etched parallelogram Si3N4 slot chosen were hSi = 220 nm and het = 100 nm, respectively, as shown in Fig. 1(a). In this work, the numerical results were obtained using COMSOL Multiphysics based on a rigorous finite-element method. The computational window was enclosed by perfectly matched layers that effectively absorb the outgoing power. The first step in the design process is to select a suitable width for the Si waveguide, able to support the converted guided modes. For the TE0-to-TE1 mode converter of width WSi = 3.25 µm, the effective refractive indices of TE0 and TE1 modes are neff_0 = 2.0234 and neff_1 = 1.8216, respectively. According to the PMC modified of Eq. (4) with condition q = 1, the total length of the TE01 (TE0-to-TE1) mode converter is L = 16.84 µm. The next step is to find the tilted angle θ and width Wslot of the slot. As shown in Fig. 1(b), the Wslot and θ are related by the trigonometry of tanθ = WSi/(LWslot) = tslot/Wslot. Once tslot is obtained, the θ and Wslot can be determined to completely define geometry parameters for the proposed structure. To maintain a non-zero κ0q(z) through the entire conversion process, the parallelogram geometry must be fulfilled by the condition Wslot < L/2. Therefore, the maxima of Wslot and tslot must be smaller than L/2 and WSi, respectively. Before deriving the parameters, we selected the transverse position x as the independent variable of κ0q because the actual propagation distance z depends on θ. To investigate the effect of κ0q on mode conversion, we examined profiles of κ0q(x) with different values of WSi (3.25, 4.25, 5.25, and 6.25 µm for TE01, TE02, TE03, and TE04, respectively), where q = 1, 2, 3, and 4, along the x-direction from x = 0 (point A in Fig. 1(b)) to x = WSi (point B in Fig. 1(b)), as shown in Fig. 2(a). The choice of WSi is to enable waveguiding for the required modes and to avoid waveguiding for high-order modes. Also, the coupling coefficient profile of κ0q(z) is calculated based on different cross section along propagation distance z from the positions of z = A to z = C with effective-index perturbation for different q’s. In the parallelogram, tslot = WSi is chosen for the TE01 converter, and the last-half triangle ΔBCD accumulates the remaining half-period phase, as per Fig. 2(b), which shows the profiles of κ0q(z) from z = 0 to z = L. Thus, we obtain θ = 21.1° and Wslot = L/2 = 8.42 µm for the TE01 converter.

 figure: Fig. 2.

Fig. 2. Coupling coefficient profiles of (a) κ0q(x) and (b) κ0q(z) versus the transverse position x and propagation distance z of the proposed structure, respectively, at a wavelength of λ = 3.4 µm, with parameters hSi = 220 nm, het = 100 nm, and a fixed WSi for a specific q.

Download Full Size | PDF

For the TE02 converter with WSi = 4.25 µm, the device length is L = 21.06 µm according to the PMC where q = 2. Observing the profile κ02(x) from tslot = WSi/2 (the second point of κ02 = 0) to WSi (the third point of κ02 = 0), maximum coupling strength occurs at tslot = 2WSi/3, as depicted by a red arrow in the profile of κ02(x). This is because the maximum power exchange occurs at Δβ = 0 and |κ0q|L = π/2 in the CMT. The relationship of |κ0q|L = π/2 can be calculated by

$${K_{0q}} = |{{\kappa_{0q}}} |L = \int\limits_0^L {|{\kappa_{0q}^{}{\kern 1pt} (z){\kern 1pt} } |dz} ,$$
where K0q denotes the coupling strength. A larger etched thickness het also makes the mode coupling stronger, reducing the device length. However, the penalty for this is higher insertion loss. Differing from TE01 converter, the input TE0 mode propagating in the TE02 converter also converts a small amount of power to the undesired TE1 mode. To reduce the undesired power coupling to TE1 mode, we can adjust the practical device length. Note that the effective-index difference between TE0 and TE1 modes is smaller than that between TE0 and TE2 modes, making the required L for satisfying the PMC between TE0 and TE2 modes is shorter than that for satisfying the PMC between TE0 and TE1 modes. Therefore, moderately decreasing the device length of TE02 converter can effectively reduce the inter-modal CT of the target TE2 and undesired TE1 modes because the reduced power ratio of TE1 mode is significantly larger than that of TE2 mode. As a result, we express the optimal length of the converter as Lopt = αL, where α is a shrinking factor smaller than 1. To optimize performance, we choose α = 0.97 for the TE02 converter, and thus its actual length is Lopt = 20.43 µm. Selecting tslot = 2WSi/3, we obtain θ = 19.2° and Wslot = 8.17 µm for the TE02 converter. The profile of κ02(z) from z = 0 to Lopt = 20.43 µm is also shown in Fig. 2(b). Using Eq. (5), the calculated coupling strengths of K01 = 1.506 and K02 = 1.504 approximating to π/2 [46] are for the TE01 and TE02 mode converters, respectively, validating the high coupling efficiencies for the two converters. Likewise, the conditions of tslot = 2WSi/4 and tslot = 2WSi/5 shown in Fig. 2(a) are chosen for the TE03 (with WSi = 5.25 µm) and TE04 (with WSi = 6.25 µm) converters, respectively, to achieve maximum coupling strength. In summary, we deduce the geometry formulas of tslot = 2WSi/(q+1) and Wslot = 2WSicotθ/(q+1) from the above principles and numerical calculations. To clearly illustrate the derived results, we indicate the geometry formulas in Fig. 3.

 figure: Fig. 3.

Fig. 3. The obtained geometry relations of the proposed structure with the width WSi of Si waveguide, where q is the mode number and α is shrinking factor.

Download Full Size | PDF

Detailed design parameters for the proposed converters are listed in Table 1. It is clear that device length increases as the converted mode-order increases. Note that a smaller shrinking factor α is chosen for the higher-order mode converter. As stated above, the value of α = 0.97 is sufficient to achieve the deviation from the PMC between the TE0 and TE1 modes, making a significant reduction of inter-modal CT of the target TE2 and undesired TE1 modes for a TE02 converter. As q increases, the number of undesired TEq−1 mode coupling increases because of more guided modes supported in a wider waveguide. To reduce the maximum undesired TEq−1 mode coupling, we need to choose a smaller α (< 0.97) for a TE0q converter with q > 2. This is because the effective-index contrast between TE0 and TEq−1 modes increases as q increases, and therefore, a smaller α is enough to deviate their PMC, reducing the inter-modal CT between the target TEq and the undesired highest-order TEq−1 modes.

Tables Icon

Table 1. Geometry parameters of the TE0 → TEq mode converters operating at a wavelength of λ = 3.4 µm, with parameters hSi = 220 nm and het = 100 nm.

In terms of coupling strength, the calculated Κ0q’s of the TE01 and TE02 converters are close to the theoretical value of π/2, ensuring the proposed design can achieve extremely high-power conversion. However, Κ03 and Κ04 show larger deviations from π/2, resulting in lower power conversion between the input and target modes. The profiles of κ03(z) and κ04(z) are also shown in Fig. 2(b). It can be seen that the profiles of κ0q(z) are symmetrical to the propagation distance of Lopt/2. According to the results in Table 1, we show the simulated electric field (Ex) evolutions of the four converters in Fig. 4 along the propagation distance. The numerical algorithm to simulate the evolutions of electric field (Ex) is based on the COMSOL build-in direct solver of multifrontal massively parallel sparse. It can be clearly observed that the input TE0 modes for different Si waveguide widths launched from the left sides are converted to high-order modes of TEq after the propagating slots. The field profiles of the TE0 mode are first squeezed into the sharp corners because of the lower index of the slot and then scattered across the slot region to form the desired high-order modes. Note that the squeeze order becomes increasingly strong as the converted mode order increases. The numerical results demonstrate that the proposed etched parallelogram slot can effectively accomplish arbitrary mode-order conversion. Notably, considerable feature sizes can be precisely fabricated with modern technologies, and the relatively simple fabrication process ensures the proposed structure is feasible.

 figure: Fig. 4.

Fig. 4. Electric field (Ex) profiles of the (a) TE01, (b) TE02, (c) TE03, and (d) TE04 mode converters at the plane of y = 100 nm, where y = 0 nm is the bottom of the Si waveguide, at a wavelength of λ = 3.4 µm, with parameters hSi = 220 nm and het = 100 nm.

Download Full Size | PDF

To quantify mode converter performance, we assessed three crucial indices, including: (1) Conversion efficiency (CE) - defined as the power ratio between desired output mode and input mode; (2) Inter-modal crosstalk (CT) - defined as the power ratio between output undesired and desired modes (namely to evaluate the mode purity); and (3) Transmission (T) - defined as the total transmitted power. The three indices are formulated by the following:

$$\textrm{CE} = 10{\log _{10}}\left( {\frac{{{P_d}}}{{{P_{in}}}}} \right),\; \textrm{CT} = 10{\log _{10}}\left( {\frac{{{P_{und}}}}{{{P_d}}}} \right),\; \textrm{and}\; \textrm{T} = 10{\log _{10}}\left( {\frac{{{P_t}}}{{{P_{in}}}}} \right), $$
where Pd, Pin, Pund, and Pt are the desired mode, input, undesired mode, and total transmitted powers, respectively. Based on the definitions in Eq. (6), the calculated indices of the proposed mode converters are listed in Table 2. We observe that the CEs of the TE01 and TE02 converters are greater than 90%. Even for the TE04 converter, the CE is still greater than 75%. Maximum CTs of TE01, TE02, TE03, and TE04 achieve less than −27, −16, −13, and −10 dB, respectively. Inevitably, converting to a higher-order mode (TE4) leads to lower CE (around 75.8%) and high CT (−10.88 dB), due to the existence of more guided modes in a wider waveguide, resulting in some of the input mode power is coupled to other undesired lower-order modes. Importantly, the T values for all converters are >93%. It can be seen that the T of the TE01 converter with a slightly larger angle of θ = 21.1° is slightly lower than that of other converters with θ < 20° due to increased energy scattering.

Next, we analyzed the spectral response of the present structure over the wavelength band of 3.34–3.46 µm, and the calculated CE, CT, and T, taking the material dispersion [39] into account, as shown in Fig. 5. Over the broadband range of 100 nm from the wavelengths of 3.35–3.45 µm, the CEs of the TE01 and TE02 converters were >90%, TE03 was >80%, and TE04 was >70%, as shown in Fig. 5(a). We observed transmissions of >90% to be insensitive over the band of 120 nm, as shown in Fig. 5(b). In terms of target mode purity, the CT of the TE01 converter was <−20 dB, as shown in Fig. 5(c), and the maximum CTs for all converters were generally <−15 dB, as shown in Figs. 5(d)–5(e), except one TE3-to-TE4 case with CT of <−10 dB, as shown in Fig. 5(f). These results validate the proposed structure’s ability to support broadband operation of 120 nm with sufficiently high performance.

 figure: Fig. 5.

Fig. 5. (a) Conversion efficiency (CE) and (b) transmission (T) of the proposed mode converter, and the inter-modal crosstalk (CT) of the (c) TE01, (d) TE02, (e) TE03, and (f) TE04 converters versus the wavelength, with conditions of hSi = 220 nm and het = 100 nm.

Download Full Size | PDF

Tables Icon

Table 2. Performance of the TE0 → TEq mode converters operating at a wavelength of λ = 3.4 µm, with parameters hSi = 220 nm and het = 100 nm.

To evaluate device feasibility, we investigated fabrication tolerance by examining the effects of geometry deviation on performance. We first analyzed CE, CT, and T versus variation in angle Δθ of the etched parallelogram slot, as shown in Fig. 6. Within Δθ < ± 2°, the CEs of the converters TE01 to TE03 were >75%, as shown in Fig. 6(a), and all converters achieved T > 80%, as shown in Fig. 6(b).

 figure: Fig. 6.

Fig. 6. (a) Conversion efficiency (CE) and (b) transmission (T) of the proposed mode converter and the inter-modal crosstalk (CT) of the (c) TE01, (d) TE02, (e) TE03, and (f) TE04 converters, versus the variation of slot angle Δθ under conditions of hSi = 220 nm and het = 100 nm.

Download Full Size | PDF

Decreasing the slot angle appears to reduce insertion loss in total transmitted power. The maximum CTs of all converters remained <−10 dB, except for the CT between TE3 and TE4 modes with Δθ < 0. This is because decreasing θ generally reduces the input-to-target mode coupling strength. Nevertheless, applying the condition of Δθ > 0 can effectively improve the CT between TE3 and TE4 modes. The other critical parameter is the etched depth of the slot. The relationship between performance and etched thickness Δhet of the slot and the calculated results are shown in Fig. 7. Decreasing the etched depth appears to moderately increase the CEs and Ts, as shown in Figs. 7(a) and 7(b), respectively, but significantly reduce the modal purity. The increase of transmission as het decreases is also due to the smaller index perturbation having the same effect induced by decreasing the parallelogram angle. In contrast, shallow depth leads to reduced coupling strength, resulting in a higher CT. Within the range Δhet < ± 10 nm, the CEs and Ts of all converters were >80%, as shown in Figs. 7(a) and 7(b), respectively, and the maximum CTs of all converters remained <−10 dB, as shown in Figs. 7(c) to 7(f), respectively. Extending the range to Δhet > ±10, only the CT between the TE3 and TE4 modes exceeded −10 dB. Fortunately, by selecting a deeper etched thickness of Δhet > 0, CT between the TE3 and TE4 modes remain <−10 dB. Overall, we consider the fabrication tolerances from the above results to be satisfactory, making the proposed structure practically feasible using the proposed mathematical formulas for obtaining geometry parameters.

 figure: Fig. 7.

Fig. 7. (a) Conversion efficiency (CE) and (b) transmission (T) of the proposed mode converter and inter-modal crosstalk (CT) of the (c) TE01, (d) TE02, (e) TE03, and (f) TE04 converters, versus the variation in etched thickness Δhet of the slot under conditions of hSi = 220 nm and het = 100 nm.

Download Full Size | PDF

To validate feasibility of the proposed formulations to the optical communication wavelength of λ = 1.55 µm, we investigate the performances of the proposed mode converter on a SOI platform. Therefore, we replaced the CaF2 substrate used at the mid-IR wavelength of λ = 3.4 µm with the SiO2 substrate at the near-IR wavelength of λ = 1.55 µm. The relative permittivities of Si, Si3N4, and SiO2 at a wavelength of λ = 1.55 µm are nSi = 3.480, nSi3N4 = 1.996, and nSiO2 = 1.444 [42], respectively. The thickness of the Si waveguide and depth of the etched parallelogram Si3N4 slot chosen were hSi = 220 nm and het = 100 nm, respectively, the same as those of above example. Detailed design parameters for the converters are listed in Table 3. Note that the device lengths of the TE01 with WSi = 1.10 µm and TE02 with WSi = 1.50 µm are only 5.47 and 7.23 µm, respectively.

Tables Icon

Table 3. Geometry parameters of the TE0 → TEq mode converters operating at a wavelength of λ = 1.55 µm, with parameters hSi = 220 nm and het = 100 nm.

According to the results in Table 3, the simulated electric field (Ex) evolutions of the two converters are in Fig. 8.

 figure: Fig. 8.

Fig. 8. Electric field (Ex) profiles of the (a) TE01and (b) TE02 mode converters at the plane of y = 100 nm, where y = 0 nm is the bottom of the Si waveguide, at a wavelength of λ = 1.55 µm, with parameters hSi = 220 nm and het = 100 nm.

Download Full Size | PDF

Based on the definitions in Eq. (6), the calculated performances of the proposed converter operating at the wavelength of λ = 1.55 µm are listed in Table 4. We observe that the CEs of the TE01 and TE02 converters are greater than 94% and 82%, respectively; the maximum CTs of TE01 and TE02 are less than −18 and −11 dB, respectively; and the T values for the two converters are >96%. Also, the T of the TE01 converter with a slightly larger angle of θ = 21.9° is slightly lower than that of other TE02 one with θ = 19.1° due to increased energy scattering.

Tables Icon

Table 4. Performance of the TE0 → TEq mode converters operating at a wavelength of λ = 1.55 µm, with parameters hSi = 220 nm and het = 100 nm.

The obtained results demonstrate that our proposed idea can be accurately extended to other bands. Comparing our device with reported mode converters only using dielectric materials, the device length and performances are listed in Table 5. Considering both the device length and performances, it can be seen that the proposed device shows superior figure of merit (i.e., [CE or CT] divided by device length) than the previously reported designs. Importantly, the proposed mode converter accomplishes arbitrary mode-order conversion operating at various wavelength bands using a single tilted slot structure.

Tables Icon

Table 5. Comparison of various mode converters operating at the wavelength of λ = 1.55 µm.

4. Summary

We proposed a compact mode converter comprising an etched parallelogram slot filled with Si3N4 on a Si waveguide on a CaF2 substrate, able to achieve any waveguide mode conversion at the mid-IR wavelength of 3.4 µm. We constructed analytical formulas for the design parameters by considering the PMC and maximum coupling strength of CMT. Numerical examples were provided, demonstrating that the TE0-to-TE1, TE0-to-TE2, TE0-to-TE3, and TE0-to-TE4 converters were able to achieve conversion efficiencies of >92.7%, >91.7%, >88.2%, and >75.8%, respectively, with power transmission and inter-modal CT values of >93% and <−10 dB, respectively, for the four converters. Device lengths of the shortest TE0-to-TE1 and longest TE0-to-TE4 converters were just 16.84 µm and 24.61 µm, respectively. Over a broad wavelength range of 100 nm, the CEs of the first three converters were >80% and >70% for the TE0-to-TE4 converter. Analysis of fabrication tolerances showed that the conditions of Δθ > 0° and Δhet > 0 nm achieved a high T of >80% and low CT of <−10 dB. The proposed idea can be scaled not only for arbitrary-order mode conversion but also for other operating bands. In addition, the simulation results of our device operating at the wavelength of 1550 nm show that the device lengths of the TE0-to-TE1 (with the CE 94%) and TE0-to-TE2 (with the CE 82%) converters can be shrunk to 5.47 and 7.23 µm, respectively, demonstrating the feasibility of the proposed idea to other bands. Overall, this work provides a universal model for effectively designing mode converters and other photonic components.

Funding

Ministry of Science and Technology, Taiwan (108-2112-M-005-006).

Acknowledgments

The authors would like to thank Enago (www.enago.tw) for the English language review.

Disclosures

The author declares no conflicts of interest.

References

1. R. Shankar and M. Loncar, “Silicon photonic devices for mid-infrared applications,” Nanophotonics 3(4-5), 329–341 (2014). [CrossRef]  

2. T. Hu, B. Dong, X. Luo, T. Y. Liow, J. Song, C. Lee, and G. Q. Lo, “Silicon photonic platforms for mid-infrared applications,” Photonics Res. 5(5), 417–430 (2017). [CrossRef]  

3. S. A. Miller, M. Yu, X. Ji, A. G. Griffith, J. Cardenas, A. L. Gaeta, and M. Lipson, “Low-loss silicon platform for broadband mid-infrared photonics,” Optica 4(7), 707–712 (2017). [CrossRef]  

4. Y. Zou, S. Chakravarty, C. J. Chung, X. Xu, and R. T. Chen, “Mid-infrared silicon photonic waveguides and devices,” Photonics Res. 6(4), 254–276 (2018). [CrossRef]  

5. G. Z. Mashanovich, F. Y. Gardes, D. J. Thomson, Y. Hu, K. Li, M. Nedeljkovic, J. S. Penades, A. Z. Khokhar, C. J. Mitchell, S. Stankovic, R. Topley, S. A. Reynolds, Y. Wang, B. Troia, V. M. N. Passaro, C. G. Littlejohns, T. D. Bucio, P. R. Wilson, and G. T. Reed, “Silicon photonic waveguides and devices for near- and mid-IR applications,” IEEE J. Sel. Top. Quantum Electron. 21(4), 407–418 (2015). [CrossRef]  

6. H. Lin, Z. Luo, T. Gu, L. C. Kimerling, K. Wada, A. Agarwal, and J. Hu, “Mid-infrared integrated photonics on silicon: a perspective,” Nanophotonics 7(2), 393–420 (2017). [CrossRef]  

7. T. Baehr-Jones, A. Spott, R. Ilic, A. Spott, B. Penkov, W. Asher, and M. Hochberg, “Silicon-on-sapphire integrated waveguides for the mid-infrared,” Opt. Express 18(12), 12127–12135 (2010). [CrossRef]  

8. M. A. V. Camp, S. Assefa, D. M. Gill, T. Barwicz, S. M. Shank, P. M. Rice, T. Topuria, and W. M. J. Green, “Demonstration of electrooptic modulation at 2165 nm using a silicon Mach-Zehnder interferometer,” Opt. Express 20(27), 28009–28016 (2012). [CrossRef]  

9. P. T. Lin, V. Singh, H. G. Lin, T. Tiwald, L. C. Kimerling, and A. M. Agarwal, “Low-stress silicon nitride platform for mid-infrared broadband and monolithically integrated microphotonics,” Adv. Opt. Mater. 1(10), 732–739 (2013). [CrossRef]  

10. M. Nedeljkovic, A. Z. Khokhar, Y. Hu, X. Chen, J. S. Penades, S. Stankovic, H. M. H. Chong, D. J. Thomson, F. Y. Gardes, G. T. Reed, and G. Z. Mashanovich, “Silicon-on-sapphire integrated waveguides for the mid-infrared,” Opt. Mater. Express 3(9), 1205–1214 (2013). [CrossRef]  

11. Y. Xia, C. Qiu, X. Zhang, W. Gao, J. Shu, and Q. Xu, “Suspended Si ring resonator for mid-infrared application,” Opt. Lett. 38(7), 1122–1124 (2013). [CrossRef]  

12. Z. Cheng, X. Chen, C. Y. Wong, K. Xu, C. K. Y. Fung, Y. M. Chen, and H. K. Tsang, “Mid-infrared grating couplers for silicon-on-sapphire waveguides,” IEEE Photonics J. 4(1), 104–113 (2012). [CrossRef]  

13. Y. Zou, H. Subbaraman, S. Chakravarty, X. Xu, A. Hosseini, W. C. Lai, P. Wray, and R. T. Chen, “Grating-coupled silicon-on-sapphire integrated slot waveguides operating at mid-infrared wavelengths,” Opt. Lett. 39(10), 3070–3073 (2014). [CrossRef]  

14. N. Chen, B. Dong, X. Luo, H. Wang, N. Singh, G. Q. Lo, and C. Lee, “Efficient and broadband subwavelength grating coupler for 3.7 µm mid-infrared silicon photonics integration,” Opt. Express 26(20), 26242–26256 (2018). [CrossRef]  

15. J. Wang, C. Lee, B. Niu, H. Huang, Y. Li, M. Li, X. Chen, Z. Sheng, A. Wu, W. Li, X. Wang, S. Zou, F. Gan, and M. Qi, “Silicon-on-sapphire integrated waveguides for the mid-infrared,” Opt. Express 23(11), 15029–15037 (2015). [CrossRef]  

16. B. Kumari, R. K. Varshney, and B. P. Pai, “Design of a silicon-on-calcium-fluoride-based ultracompact and highly efficient polarization splitter for the midinfrared,” Opt. Eng. 58(03), 1 (2019). [CrossRef]  

17. R. kitamura, L. Pilon, and M. Jonasz, “Optical constants of silica glass from extreme ultraviolet to far infrared at near room temperature,” Appl. Opt. 46(33), 8118–8133 (2007). [CrossRef]  

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

19. D. Marris-Morini, V. Vakarin, J. M. Ramirez, Q. Liu, A. Ballabio, J. Frigerio, M. Montesinos, C. Alonso-Ramos, X. L. Roux, S. Serna, D. Benedikovic, D. Chrastina, L. Vivien, and G. Isella, “Germanium-based integrated photonics from near- to mid-infrared applications,” Nanophotonics 7(11), 1781–1793 (2018). [CrossRef]  

20. K. Wada, H. C. Luan, D. R. Lim, and L. C. Kimerling, “On-chip interconnection beyond semiconductor roadmap: silicon microphotonics,” Proc. SPIE 4870(7), 437 (2002). [CrossRef]  

21. P. J. Winzer, “Making spatial multiplexing a reality,” Nat. Photonics 8(5), 345–348 (2014). [CrossRef]  

22. L. W. Luo, N. Ophir, C. P. Chen, L. H. Gabrielli, C. B. Poitras, K. Bergmen, and M. Lipson, “WDM-compatible mode-division multiplexing on a silicon chip,” Nat. Commun. 5(1), 3069 (2014). [CrossRef]  

23. B. Stern, X. Zhu, C. P. Chen, L. D. Tzuang, J. Cardenas, K. Bergman, and M. Lipson, “On-chip mode-division multiplexing switch,” Optica 2(6), 530–535 (2015). [CrossRef]  

24. 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]  

25. Y. H. Ding, J. Xu, F. D. 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]  

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

27. 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]  

28. D. Ohana, B. Desiatov, N. Mazurski, and U. Levy, “Dielectric metasurface as a platform for spatial mode conversion on nanoscale waveguides,” Nano Lett. 16(12), 7956–7961 (2016). [CrossRef]  

29. Z. Li, M. H. Kim, C. Wang, Z. Han, S. Shrestha, A. C. Overvig, M. Lu, A. Stein, A. M. Agarwal, M. Lonar, and N. Yu, “Controlling propagation and coupling of waveguide modes using phase-gradient metasurfaces,” Nat. Nanotechnol. 12(7), 675–683 (2017). [CrossRef]  

30. H. Xu and Y. Shi, “Ultra-sharp multi-mode waveguide bending assisted with metamaterial-based mode converters,” Laser Photonics Rev. 12(3), 1700240 (2018). [CrossRef]  

31. 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(1), 1801191 (2018). [CrossRef]  

32. 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]  

33. 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), 6601210 (2019). [CrossRef]  

34. 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]  

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

36. H. Guan, Y. Ma, R. Shi, A. Novack, J. Tao, Q. Fang, A. E. Lim, G. Q. Lo, T. Baehr-Jones, and M. Hochberg, “Ultracompact silicon-on-insulator polarization rotator for polarization-diversified circuits,” Opt. Lett. 39(16), 4703–4706 (2014). [CrossRef]  

37. V. Liu, D. A. B. Miller, and S. Fan, “Ultra-compact photonic crystal waveguide spatial mode converter and its connection to the optical diode effect,” Opt. Express 20(27), 28388–28397 (2012). [CrossRef]  

38. G. Chen and J. U. Kang, “Waveguide mode converter based on two-dimensional photonic crystals,” Opt. Lett. 30(13), 1656–1658 (2005). [CrossRef]  

39. L. H. Frandsen, Y. Elesin, L. F. Frellsen, M. Mitrovic, Y. Ding, O. Sigmund, and K. Yvind, “Topology optimized mode conversion in a photonic crystal waveguide fabricated in silicon-on-insulator material,” Opt. Express 22(7), 8525–8532 (2014). [CrossRef]  

40. U. Leonhardt, “Optical conformal mapping,” Science 312(5781), 1777–1780 (2006). [CrossRef]  

41. J. B. Pendry, D. Schurig, and D. R. Smith, “Controlling electromagnetic fields,” Science 312(5781), 1780–1782 (2006). [CrossRef]  

42. M. Bass, Handbook of Optics, Volume IV: Optical Properties of Materials, Nonlinear Optics, Quantum Optics, (McGraw-Hill Education, 2009).

43. Y. Chen, H. Lin, J. Hu, and M. Li, “Heterogeneously integrated silicon photonics for the mid-infrared and spectroscopic sensing,” ACS Nano 8(7), 6955–6961 (2014). [CrossRef]  

44. A. Sánchez-Postigo, J. G. Wangüemert-Pérez, J. S. Penadés, A. Ortega-Moñux, M. Nedeljkovic, R. Halir, F. E. M. Mimun, Y. X. Cheng, Z. Qu, A. Z. Khokhar, A. Osman, W. Cao, C. G. Littlejohns, P. Cheben, G. Z. Mashanovich, and Í. Molina-Fernández, “Mid-infrared suspended waveguide platform and building blocks,” IET Optoelectron. 13(2), 55–61 (2019). [CrossRef]  

45. 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]  

46. A. Yariv and P. Yeh, Optical Waves in Crystals: Propagation and Control of Laser Radiation, (Wiley-Interscience, 2002).

47. C. A. Volkert and A. M. Minor, “Focused ion beam microscopy and micromachining,” MRS Bull. 32(5), 389–399 (2007). [CrossRef]  

48. R. Garcia, A. W. Knoll, and E. Riedo, “Advanced scanning probe lithography,” Nat. Nanotechnol. 9(8), 577–587 (2014). [CrossRef]  

49. K. Yu, C. Ryu, D. R. Colin, W. Heiko, S. Martin, B. Samuel, R. Steffen, S. Marilyne, R. K. Subarna, D. B. J. Tevis, and W. K. Armin, “Sub-10 nanometer feature size in silicon using thermal scanning probe lithography,” ACS Nano 11(12), 11890–11897 (2017). [CrossRef]  

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 (8)

Fig. 1.
Fig. 1. (a) 3D schematic diagram and (b) top view of the proposed mode converter.
Fig. 2.
Fig. 2. Coupling coefficient profiles of (a) κ0q(x) and (b) κ0q(z) versus the transverse position x and propagation distance z of the proposed structure, respectively, at a wavelength of λ = 3.4 µm, with parameters hSi = 220 nm, het = 100 nm, and a fixed WSi for a specific q.
Fig. 3.
Fig. 3. The obtained geometry relations of the proposed structure with the width WSi of Si waveguide, where q is the mode number and α is shrinking factor.
Fig. 4.
Fig. 4. Electric field (Ex) profiles of the (a) TE01, (b) TE02, (c) TE03, and (d) TE04 mode converters at the plane of y = 100 nm, where y = 0 nm is the bottom of the Si waveguide, at a wavelength of λ = 3.4 µm, with parameters hSi = 220 nm and het = 100 nm.
Fig. 5.
Fig. 5. (a) Conversion efficiency (CE) and (b) transmission (T) of the proposed mode converter, and the inter-modal crosstalk (CT) of the (c) TE01, (d) TE02, (e) TE03, and (f) TE04 converters versus the wavelength, with conditions of hSi = 220 nm and het = 100 nm.
Fig. 6.
Fig. 6. (a) Conversion efficiency (CE) and (b) transmission (T) of the proposed mode converter and the inter-modal crosstalk (CT) of the (c) TE01, (d) TE02, (e) TE03, and (f) TE04 converters, versus the variation of slot angle Δθ under conditions of hSi = 220 nm and het = 100 nm.
Fig. 7.
Fig. 7. (a) Conversion efficiency (CE) and (b) transmission (T) of the proposed mode converter and inter-modal crosstalk (CT) of the (c) TE01, (d) TE02, (e) TE03, and (f) TE04 converters, versus the variation in etched thickness Δhet of the slot under conditions of hSi = 220 nm and het = 100 nm.
Fig. 8.
Fig. 8. Electric field (Ex) profiles of the (a) TE01and (b) TE02 mode converters at the plane of y = 100 nm, where y = 0 nm is the bottom of the Si waveguide, at a wavelength of λ = 1.55 µm, with parameters hSi = 220 nm and het = 100 nm.

Tables (5)

Tables Icon

Table 1. Geometry parameters of the TE0 → TEq mode converters operating at a wavelength of λ = 3.4 µm, with parameters hSi = 220 nm and het = 100 nm.

Tables Icon

Table 2. Performance of the TE0 → TEq mode converters operating at a wavelength of λ = 3.4 µm, with parameters hSi = 220 nm and het = 100 nm.

Tables Icon

Table 3. Geometry parameters of the TE0 → TEq mode converters operating at a wavelength of λ = 1.55 µm, with parameters hSi = 220 nm and het = 100 nm.

Tables Icon

Table 4. Performance of the TE0 → TEq mode converters operating at a wavelength of λ = 1.55 µm, with parameters hSi = 220 nm and het = 100 nm.

Tables Icon

Table 5. Comparison of various mode converters operating at the wavelength of λ = 1.55 µm.

Equations (6)

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

d d z A p ( z ) = q κ p q A q ( z ) e i ( β p β q ) z ,
κ p q = ω ε 0 4 E p ( x , y ) Δ ε ( x , y , z ) E q ( x , y ) d x d y .
Δ β = β p β q β c = 0 ,
L = q 2 π β p β q .
K 0 q = | κ 0 q | L = 0 L | κ 0 q ( z ) | d z ,
CE = 10 log 10 ( P d P i n ) , CT = 10 log 10 ( P u n d P d ) , and T = 10 log 10 ( P t P i n ) ,
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.