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

Numerical design of a high-performance polarization beam splitter assisted by composite subwavelength gratings

Open Access Open Access

Abstract

We report a compact polarization beam splitter (PBS) consisting of slotted waveguides assisted by composite subwavelength gratings (CSWGs) on a silicon-on-insulator platform. By tailoring the material anisotropy of the CSWGs, coupling strengths of transverse-electric (TE) and transverse-magnetic (TM) polarization coupling strengths are respectively suppressed and enhanced significantly, achieving concurrent improvements in polarization extinction ratio (PER), device footprint, and working bandwidth (BW) compared with purely slotted waveguides. Differing in construction from mono-material SWGs, the CSWGs comprise silicon strips covered with a silicon dioxide (SiO2) layer of the same thickness as the slot layer of the slotted waveguides, simplifying the fabrication process and further reducing device length. Numerical simulations show significant improvement in PERTM from about 15 dB for the purely slotted waveguides to 28 dB for the proposed design, with a 40% reduction in device length at a wavelength of λ = 1550 nm. Within a BW of ∼60 nm, the proposed PBS achieves PERTM ∼25 dB, PERTE >15 dB, and insertion loss (ILs) <0.1 dB for TE and TM modes. Fabrication tolerance investigations are also described and discussed. The proposed idea paves the way for simultaneous improvements in PER, footprint, and working BW for PBSs comprising a variety of coupled-waveguide systems.

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

1. Introduction

Silicon-on-insulator (SOI) is the main material platform for realizing on-chip photonic integrated circuits (PICs) [14], offering two main benefits: mature CMOS-compatible fabrication technology and high refractive index contrast, which plays a significant role in enabling compact devices. However, being highly birefringent, the SOI platform causes strong polarization dependence, which is undesirable in optical-fiber networks. Hence, polarization diversity devices including polarizers, polarization beam splitters (PBSs), and polarization rotators have been proposed to overcome this problem. Of these, PBSs, which separate transverse-electric (TE) and transverse-magnetic (TM) polarizations, are the most widely used. Over the years, many types of PBSs using various splitting mechanisms [518] have been reported, including adiabatic mode-evolution (AME) [5,6], multimode interference (MMI) structures [7,8], photonic crystal (PhC) [9,10], grating couplers [11,12], computationally optimized metamaterials [13], and directional couplers (DC) [1418]. To assess a PBS, we must consider various criteria including the device footprint, polarization extinction ratio (PER), insertion loss (IL), working bandwidth (BW), and fabrication complexity. Although AME-based PBSs [5,6] are long (>200 µm) because of their use of gradually evolving geometry to obtain satisfactory performance, they have offer less stringent fabrication tolerances and broadband operation. The simpler fabrication process of MMI-based PBSs [7,8] with pure rectangular waveguides without any extra design, such as subwavelength gratings (SWGs), makes them more attractive; however, the drawback is their extremely long footprint (>1000 µm), determined by the common multiple of the self-imaging lengths [19] of TE and TM modes. PhC-based [9,10] and grating-based PBSs [11,12] achieve footprints of tens of micrometers, but the former has the disadvantages of fabrication complexity and relatively large scattering loss. Recently, a PBS with an extremely small footprint (2.4 µm × 2.4 µm) was realized using freeform metamaterials, designed using computational optimization algorithms [13]. However, optimizing a desirable device is time-consuming, and a PER of ∼10 dB within a working BW of 32 nm is still limited. By contrast, DC-based PBSs [1418] are the most popular due to their structural simplicity, relatively small footprint, satisfactory performance, and diverse designs.

In principle, DC-based PBSs separate two polarizations by satisfying the phase-matching condition for one polarization coupling to the cross port and deviating from another propagating along the through port with negligible coupling. Accordingly, DC-based PBSs can be formed by assembling various waveguide structures (e.g., silicon wire, hybrid plasmonic, multimode interference, and nano-slot waveguides), depending on priorities in terms of device size, PER, IL, BW, and fabrication complexity. Fukuda et al. [14] reported a DC-based PBS of 7 µm × 16 µm using a coupler comprising two Si wires (strips) fabricated on a SOI wafer, calculating PERTE (PERTM) and ILTE (ILTM) to be ∼15 (10) and 0.5 (0.5) dB, respectively, in the C-band-wavelength range. Guan et al. [15] proposed an asymmetric DC-based PBS with a device footprint of 1.9 µm × 3.7 µm, comprising a Si wire and a hybrid plasmonic waveguide (HPW); the PERs of both polarizations were >12 dB, operating in a 120 nm BW. Instead of using the evanescent wave coupling between a Si wire and an HPW, they adopted an MMI coupler to achieve a small device of 1.8 µm × 2.5 µm [16]; however, the obtained PER was limited to about 10 dB over a BW of 80 nm. In addition to using conventional strip waveguides, Yue et al. [17] proposed a PBS comprising two horizontally slotted [20,21] waveguides to significantly increase coupling of the TM mode and enhance polarization dependence compared with that of a PBS comprising two Si-strip waveguides [14]. The coupling length was thus significantly reduced from ∼350 to 46.7 µm, and the calculated PERs of the two polarizations were >20 dB over only an 18 nm BW [17]. A PBS fabricated on an 8-inch Si wafer with a 2 µm silicon dioxide (SiO2) layer displayed experimental values of 16.8 and 14.1 dB for TE and TM modes, respectively [18].

SWGs [2228] comprising an arrangement of dielectric strips with dimensions much smaller than the wavelength have recently been used to construct various silicon photonic devices, following maturation of nanoscale pattern fabrication technology over the past decades. In principle, SWGs act as homogeneous slab waveguides with equivalent material anisotropy; the refractive index tensor can be flexibly controlled to obtain the required properties. Diffraction effects are suppressed in SWG structures because of the deep subwavelength pitches. In addition to siting SWGs in the waveguide core region [2528], Jahani and Jacob [2931] used them in the waveguide cladding region to reduce crosstalk. Also, Xu et al. [32] reported a similar design as that of [31] but changed two strip waveguides to two SWG-based ones to build a PBS. Throughout the introduction mentioned above, the PERs or ILs all come from simulation except the results of references [26], [31], and [32] come from actual measurements. In this work, we designed a PBS comprising horizontally slotted waveguides assisted by composite subwavelength gratings (CSWGs) between the two slot waveguides. Differing from mono-material SWGs, the CSWGs in our design consist of SiO2 layers of the same thickness as the slot layer of the slotted waveguides. With the addition of SiO2 layers to the top of the Si strips, the fabrication process is simplified, and the device length further reduced. The proposed PBS not only significantly reduces TE mode coupling, but also effectively enhances TM mode coupling, creating extremely strong polarization dependence. Compared with the performance of a PBS comprising a purely slotted waveguide coupler (PSWC) [17], our proposed PBS achieves around 40% reduction in device length, 13 dB improvement in PERTM, and a better working BW.

2. Mode characteristic and coupling effect of anisotropic composite subwavelength gratings

Figure 1(a) shows a 3D-schematic diagram of the proposed PBS, comprising two horizontally slotted waveguides assisted by CSWGs (four composite strips are shown here). The slot waveguide consists of a low-index SiO2 slot layer embedded between two high-index Si layers. A 90° angled slotted waveguide with R connected to the end of a slot waveguide effectively decouples the two waveguides. The substrate and cladding are SiO2 and air, respectively. Zoomed in views of the cross section of the input, TE output, and TM output ports are also shown in Figs. 1(b), 1(c), and 1(d), respectively.

 figure: Fig. 1.

Fig. 1. (a) 3D schematic diagram and zoomed in views of the (b) input, (c) TE output, and (d) TM output ports of the proposed polarization beam splitter.

Download Full Size | PDF

Si widths for the slot waveguides and multiple dielectric strips are WSi and Wcl, respectively, and the heights of Si and SiO2 in the slot waveguides with the edge-to-edge spacing s are hSi and tslot, respectively. The CSWG pitch is set to Λ = Wcl + g with duty cycle ρ = Wcl /Λ, where g is the gap between the strips, and Λ is set below the subwavelength regime [22]. The TE mode with a major electric component in the x-direction is guided along the through bar connected by a curved waveguide with negligible power coupling, while the TM mode with a major electric component in the y-direction is coupled by satisfying the phase-matching condition into the cross bar with a straight geometry throughout the entire propagation. In addition, two of the four dielectric strips follow the curve of the through bar of the 90° angled waveguide for performance leverage, as shown in Fig. 1(c). To ensure the proposed design could be realized experimentally, geometry sizes of the CSWGs consisting of Si strips covered with SiO2 layers were as per those of the slot waveguides. As a result, the required fabrication steps were almost identical to those of the PSWC conventional slot waveguide coupler, making fabrication of the proposed PBS comparatively simple. The processes involved in fabricating the proposed device are shown schematically in Fig. 2. First, the patterned hard masks of the angled TE channel, straight TM channel, and CSWGs are fabricated using high resolution electron beam lithography (EBL), then the following steps are carried out:

 figure: Fig. 2.

Fig. 2. Schematic of the fabrication processes for the proposed PBS.

Download Full Size | PDF

(1) We prepare a SiO2 substrate (blue) for deposition with a negative photoresist (PR) thin film (green) of height hSi to define the lower Si regions of the slotted waveguides and the Si strips of CSWGs by the preceding masks, a PR exposure with ultraviolet (UV) light, development, and an etching process. (2) The angled TE channel, straight TM channel, and the Si strips of CSWGs are formed by etching SiO2 and lifting off the PR film. (3) A Si layer of height hSi is deposited using chemical vapor deposition (CVD) in the wells. Subsequently, we use a chemical mechanical polishing (CMP) process to obtain a flat plane. (4) A SiO2 layer of height tslot is deposited using thermal oxidation. Next, CMP is used to obtain a flat SiO2 surface. (5) After depositing a positive PR film (red), we use the patterned masks, a PR exposure, development, and an etching process to pattern the PR film. (6) The SiO2 slot layer of height tslot is formed using reactive ion etching before lifting off the PR. (7) A Si layer is deposited with the same height hSi as the lower Si layer, then CMP is used to obtain a flat Si surface. (8) After depositing a positive PR film on the Si layer, the masks, a PR exposure, and development are used to pattern the PR film. (9) Si is etched, and the PR film removed. (10) A positive PR film is deposited using the angled TE and straight TM channel masks, a PR exposure, and development. (11) Finally, the proposed device is formed by etching Si and removing the PR film.

The relative Si and SiO2 permittivities used were nSi = 3.480 and nSiO2 = 1.444 [33], respectively, at wavelength λ = 1550 nm, with relevant parameters set as follows: hSi = 150 nm; Wcl = 75 nm; g = 50 nm; ρ = 0.6; WSi = 400 nm; s = 550 nm. Note in particular that a dielectric strip aspect ratio of 2 substantially reduces fabrication difficulty with an SOI platform. Therefore, the geometry sizes of the proposed device can be easily realized for experimentation using modern fabrication technology. Numerical results were obtained using the boundary mode analysis feature of the commercial COMSOL Multiphysics software package, based on a rigorous finite element method. To reduce field reflection from computational window, perfectly matched layers were used to effectively absorb the outgoing light power to obtain more precise simulation results. To demonstrate the merit of the proposed design, TM (LTM) and TE (LTE) mode coupling lengths as a function of slot thickness, tslot, for the PSWC and our structures were obtained, as shown in Fig. 3(a). The coupling length of a coupled waveguide, which determines the PBS device length, can be calculated using Li = π /(βi, symβi, aym), where i is TE or TM, and βi,sym and βi,asym are the propagation constants of the symmetrical and anti-symmetrical modes of the i mode, respectively. The LTM displays a dramatic decrease as tslot increases from 0 to 50 nm. On further increase in tslot, LTM gradually decreases due to slight variations in TM mode confinement.

 figure: Fig. 3.

Fig. 3. (a) Coupling lengths of TM (LTM) and TE (LTE) modes and (b) the coupling-length ratio LTE / LTM versus slot thickness tslot for PSWC and our structure.

Download Full Size | PDF

Comparing coupling lengths, the PBSs comprising two Si strips [14] (with tslot = 0 nm) and PSWC (tslot = 50 nm) have LTMs values of ∼137 µm and ∼35 µm, respectively. The dramatic decrease in LTM occurs due to the stronger polarization dependence of the slotted waveguide coupler, which increases TM mode coupling strength by embedding a low-index slot layer. It can be seen that the TM mode coupling length LTM of the proposed structure is ∼21 µm with a 40% reduction in device length of PSWC. By comparison, TE mode coupling lengths LTE of PSWC and our structure when tslot = 50 nm are ∼1520 and ∼387 µm, respectively. This means that TE mode power coupling of the proposed PBS to cross bar is significantly lower than that of PSWC at the TM mode’s phase matching condition. Another index, LTE / LTM, referred to as the coupling length ratio, as shown in Fig. 3(b) is able to predict the maximum PERTM (defined in the next section) and minimum ILTE of a PBS. This is because the larger value of LTE / LTM implies retention of more TE power in the through bar, while TM power is completely transferred to the cross bar at its phase matching condition. For the PSWC with LTM = 35.71 µm, the maximum value of LTE / LTM = 10.9 occurs with tslot = 50 nm. By contrast, the proposed PBS with LTM = 21.56 µm achieves LTE / LTM = 70.5, which is much larger, at the same tslot.

The design principles of this PBS can be explained as follows. According to the effective-medium theory (EMT) [34], which limits the grating pitch to be much smaller than the wavelength of light, the SWG regions show equivalent material anisotropy as follows:

$$n_{emt}^2 = diag[{n_{xx}^2,n_{yy}^2,n_{zz}^2,} ],$$
$$n_{yy}^2 = n_{zz}^2 = \rho n_{high}^2 + ({1 - \rho } )n_{low}^2,$$
$$\frac{1}{{n_{xx}^2}} = \frac{\rho }{{n_{high}^2}} + \frac{{({1 - \rho } )}}{{n_{low}^2}},$$
where nxx, nyy, and nzz are the equivalent refractive indices in the x-, y-, and z-directions, respectively, and nhigh (here nSi) and nlow (here nair) are the high and low material refractive indices of the SWG. Therefore, the SWG provides additional degrees of freedom to tailor the desired anisotropy for the two polarizations by modifying duty cycle, material constituents, and geometric arrangement. The same design principles were used for the proposed CSWGs. The significantly reduced device length and increased PERTM are a result of introducing the CSWGs. In optical waveguide theory, mode confinement is understood to be determined by the refractive index contrast of the core and cladding. After replacing the air cladding with CSWGs, TE mode contrast between the slot waveguide and the CSWGs is reduced according to Eqs. (1)–(3). Therefore, TE mode coupling strength should be increased. However, the obtained result is counterintuitive. This can be explained by the fact that the decay constant of the evanescent wave of the TE mode, kTE, is controlled by the ratio of the permittivity components $\sqrt {n_{zz}^2/n_{xx}^2}$ [29], and the condition nzz (= nyy) > nxx is always fulfilled according to Eqs. (2) and (3). By contrast, the SWGs can improve TE mode confinement (i.e., reduce crosstalk), substantially increasing the value of LTE / LTM by several times (Fig. 3(b)). For the TM mode, the decay constant of its evanescent wave, kTM, depends on $\sqrt {n_{zz}^2/n_{yy}^2} = 1$. Therefore, TM mode confinement is not altered by SWG material anisotropy, but rather determined by the averaged refractive index of the cladding. The large refractive index of the cladding leads to loose TM mode confinement, increasing coupling strength between slotted waveguides. To clearly illustrate the coupling strengths, field contours containing 50 isolines from normalized amplitude of 0 to 1 for TE (Ex) and TM (Ey) symmetrical modes in the x-y plane for the proposed structure are shown in Figs. 4(a) and 4(b), respectively; those for the PSWC are shown in Figs. 4(c) and 4(d), respectively. Evidently, the strength of the field overlap between the slot waveguides is significantly suppressed (enhanced) for the TE (TM) mode in the proposed structure, due to the SWGs. Relative field amplitudes of the TE (TM) modes of the PSWC and our structure at a height of y = 175 nm (the middle of the slot layer) from the top of the substrate are shown in Fig. 4(e) and 4(f).

 figure: Fig. 4.

Fig. 4. Field profiles of (a) TE (Ex)- and (b) TM (Ey)- symmetrical modes of the proposed PBS, and those of (c) TE- and (d) TM-symmetrical modes of the PSWC, for tslot = 50 nm, WSi = 400 nm, hSi = 150 nm, Wcl = 75 nm, g = 50 nm, and s = 550 nm. At the central plane of the slot layer (i.e., y = 175 nm from the top of the substrate), the field amplitudes along the x-direction of (a) and (c) are shown in (e), and those of (b) and (d) are shown in (f).

Download Full Size | PDF

For the proposed PBS, we also observe that TE field suppression (the downward arrow in Fig. 4(e)) results in weak mode coupling. In contrast, field enhancement (the upward arrow in Fig. 4(f)) of the TM field profile shows strong mode coupling. Overall, according to the index of LTE / LTM shown in Figs. 3(b) and the mode profiles shown in Fig. 4, introduction of the CSWGs significantly reduces device length and increases the PERTM, as verified in the next section.

3. Propagation performance and fabrication tolerance of the proposed PBS

Based on the mode characteristics and coupling length obtained in section 2, propagation characteristics of the proposed PBS were obtained using 3D finite element method (FEM) simulations. With the same parameters as shown in Fig. 4 and R = 3 µm, the propagation Poynting vector flows of the TE and TM modes for the proposed PBS with the modified TM mode coupling length, Lcm = 20.85 µm, are shown in Figs. 5(a) and 5(b), respectively, and those for the PSWC with Lcm = 33.40 µm are shown in Figs. 5(c) and 5(d), respectively. Here, the Lcm optimized for performance differs from the LTM calculated according to coupled waveguide systems. As mentioned above, we connected an angled waveguide at the end of the through bar guiding the TE mode to effectively decouple the two waveguides. Accordingly, the modified coupling length of Lcm is moderately shorter than the corresponding LTM because the coupling effect remains a short distance within the angled waveguide before gradually decreasing. The difference between Lcm and LTM is within 1 µm at R = 3 µm, regardless of slot thickness. This difference varies moderately for different values of R.

 figure: Fig. 5.

Fig. 5. Evolutions of the Poynting vector flow of a) TE and b) TM modes of the proposed PBS with LTM = 21.56 µm (Lcm = 20.85 µm), and those of c) TE and d) TM modes of the PSWC with LTM = 35.71µm (Lcm = 33.4 µm), for tslot = 50 nm, hSi = 150 nm, Wcl = 75 nm, gap = 50 nm, WSi = 400 nm, and s = 550 nm.

Download Full Size | PDF

To assess PBS transmission characteristics, we investigated PER and IL of the TE and TM modes, as defined in Eqs. (4) and (5), respectively [18]:

$$\textrm{PE}{\textrm{R}_{TE(TM)}} = 10\,\;{\log _{10}}\left( {\frac{{{P_{TE(TM),\,TRU(CRO)}}}}{{{P_{TM(TE),\,TRU(CRO)}}}}} \right),$$
$$\textrm{I}{\textrm{L}_{TE(TM)}} = \; - 10\,\;{\log _{10}}\left( {\frac{{{P_{TE(TM),\,TRU(CRO)}}}}{{{P_{in}}}}} \right),$$
where Pin is input power, PTE(TM), THU(CRO) is TE (TM) mode power at the through (cross) bar, and PTM(TE), THU (CRO) is the TM (TE) mode power at the through (cross) bar. For the proposed PBS with Lcm = 20.85 µm, we obtain PERTE = 31.5 dB and PERTM = 26.1 dB, and ILTE = 0.03 dB and ILTM = 0.01 dB. By contrast, the structure [17] with Lcm = 33.4 µm obtains PERTE = 25.7 dB and PERTM = 14.2 dB, and ILTE = 0.05 dB and ILTM = 0.03 dB. The SWGs used in various designs of photonic device are typically mono material. If the SiO2 layers on the top of the Si strips are removed, the device length is Lcm = 26.1 µm. As a result, adding SiO2 layers to the top of the Si strips forming the CSWGs not only makes the fabrication process easier but also further reduces the device length. The results demonstrate that the proposed PBS not only significantly shortens the device length by 40% for the PSWC but also obtains superior PERs and ILs for both TE and TM modes, particularly for PERTM. To assess working BW, PERs and ILs versus wavelength are shown in Figs. 6(a) and 6(b), respectively.

 figure: Fig. 6.

Fig. 6. (a) PER and (b) IL versus the wavelength with tslot = 50 nm for the proposed PBS with Lcm = 20.85 µm and the PSWC with Lcm = 33.4 µm.

Download Full Size | PDF

For PERTM (PERTE), the proposed PBS is about 12 (2–5) dB higher than that of the PSWC in the broadband of 100 (40) nm, while ILTM values are close for the two structures. For ILTE, the proposed structure is smaller than 0.1 dB in the entire band of 100 nm but the maximum ILTE of the PSWC is close to 0.3 dB. It can be seen that PERTE is significantly dependent on wavelength because of the stronger deviation of LTM versus LTE from their own phase-matching conditions. By contrast, PERTM variation is moderate because of the extremely long LTE. The proposed PBS achieves a 100 nm BW when PERTE > 20 dB, but the PERTE of PSWC varies from 10 dB to 18 dB. To achieve PERs >15 dB and ILs <0.1 dB for both modes, the BWs of the proposed PBS and the PSWC are 50 nm (1530–1580 nm) and 20 nm (1530–1550 nm), respectively. The results show the proposed PBS can achieve a BW greater than twice that of the PSWC.

To investigate the effects of geometry parameters on performance, PERs and ILs versus tslot are shown in Fig. 7. The optimal PERs and ILs for both modes are within the range 40 nm < tslot < 60 nm. We know that a thinner tslot (<40 nm) leads to tighter mode confinement for TE (retaining more power in the through bar, resulting in higher PERTM) and TM (coupling less power to the cross bar, resulting in lower PERTE) modes for a slot waveguide. By contrast, the thicker tslot results in looser mode confinement for the TE (resulting in lower PERTM) mode but leads to a little variation in TM mode confinement. Therefore, PERTE for the thicker tslot (>60 nm) should show a slight increase rather than the moderate decrease. This can be explained as follows: The shorter LTM resulting from the thicker tslot, as shown in Fig. 3(a), causes a greater reduction in TM power, resulting in lower power scattering in the entire coupler, compared to that of the longer LTM, although the theoretical power transfer rate is 100% at the LTM. If we choose tslot = 90 nm, the device length is reduced to Lcm = 13.9 µm, providing satisfactory performance of PERTE = 24.3 dB, PERTM = 23.1 dB, ILTE = 0.08 dB, and ILTM = 0.11 dB. By trading-off all PBS assessment criteria, we arrive at tslot = 50 nm as the optimal condition with Lcm = 20.85 µm, PERTE = 31.5 dB, PERTM = 26.1 dB, ILTE = 0.03 dB, and ILTM = 0.01 dB for the PBS design parameters.

 figure: Fig. 7.

Fig. 7. PER (left axis) and IL (right axis) versus slot thickness, tslot, of the proposed PBS.

Download Full Size | PDF

Next, the performances and Lcm versus thickness of CSWGs, hSi, are shown in Fig. 8(a) and 8(b), respectively. We observe that thicker hSi leads to tighter mode confinement for TE mode, resulting in higher PERTM and lower ILTE for a slot waveguide. For the TM mode, thicker hSi also leads to tighter mode confinement due to effect of the slot waveguide mode [20], [21], resulting in lower PERTE, the result is also observed from Fig. 8(b) with longer Lcm. Although the confinement of TM mode is weaker for thinner hSi; however, it also obtains lower PERTE. This is because the crosstalk of the TE mode increases as hSi decreases [2931].

 figure: Fig. 8.

Fig. 8. (a) PER (left axis) and IL (right axis) and (b) modified coupling length of TM mode, Lcm, versus thickness of CSWGs, hSi, of the proposed PBS.

Download Full Size | PDF

On the other hand, duty cycle ρ is essential to tailoring the material anisotropy, and also influences performance of integrated photonic devices. We therefore calculated PERs and ILs for both modes versus ρ, as shown in Fig. 9(a). As mentioned in section 2, the maximum kTE indicates tightest confinement of the TE mode occurring at ρ = 0.5, according to Eqs. (2) and (3) for single SWGs. However, in our numerical investigations, optimal PERs occur at ρ = 0.6 or 0.7, deviating a little from the theoretical prediction in [29] for SWGs. Note that PERs and ILs for TE and TM modes have opposing trends with variation of ρ. This means that PER is significantly dependent on IL. In addition, variation in TM mode coupling strength with ρ is insignificant because the major electric field, Ey, of the TM mode is parallel to the plane of the CSWGs. Therefore, the highest PERTE is mainly dominated by the lowest ILTE, rather than the remaining TM power in the through bar. The proposed PBS can be reduced to Lcm = 18.3 µm with moderately lower accordingly if we choose ρ = 0.8. We know that large ρ values alleviate fabrication difficulties but moderately decrease the performance to PERTE = 28.9 dB, PERTM = 23.3 dB, ILTE = 0.19 dB, and ILTM = 0.11 dB. Practically, this level of performance is satisfactory for a commercial PBS. If the duty cycle increases to 0.9, the gap between dielectric strips is about 13 nm, making the fabrication more difficult. In addition, the Lcm versus the duty cycle is also shown in Fig. 9(b) to offer the trade-off between the performance and footprint.

 figure: Fig. 9.

Fig. 9. (a) PER (left axis) and IL (right axis) and (b) modified coupling length of TM mode, Lcm, versus duty cycle, ρ, of the proposed PBS.

Download Full Size | PDF

The other critical SWG parameter is number of strips. PER and IL values versus number of strips are shown in Fig. 10(a); the practical coupling length, Lcm, versus number of strips are shown in Fig. 10(b). It can be seen that PERs and ILs increase when number of strips increases from two to four. By further increasing the number of strips to six (the strip has an aspect ratio of 3), PERs and ILs show slight variation because of their saturated mode confinements. Increasing the number of strips to six (Lcm = 20.5 µm) is not useful for reducing device length.

 figure: Fig. 10.

Fig. 10. (a) PER (left axis) and IL (right axis) and (b) modified coupling length of TM mode, Lcm, versus number of strips of the CSWGs

Download Full Size | PDF

Finally, we evaluate fabrication tolerances of the proposed PBS by investigating performance versus slot thickness variation, Δtslot, cladding width, ΔWcl, and thickness of CSWGs, ΔhSi, as shown in Figs. 11(a), 11(b), and 11(c), respectively. We note that PERTM shows only slight variation with the three types of fabrication error, owing to the extremely long TE mode coupling length (LTE = 1.5 mm). By contrast, PERTE varies significantly with fabrication error. This is because the short LTM leads to a large deviation from the TM mode phase-matching condition. Within the variation range Δtslot = ±5 (10) nm, PERTE achieves a level of >20 (15) dB and the ILs of the two modes are lower than 0.2 (0.25) dB (Fig. 11(a)).

 figure: Fig. 11.

Fig. 11. PER (left axis) and IL (right axis) versus variation in (a) slot thickness, Δtslot, (b) cladding width, ΔWcl, and (c) thickness of CSWGs, ΔhSi, of the proposed PBS.

Download Full Size | PDF

Thanks to modern semiconductor fabrication technology, the surface roughness of a SiO2 layer deposited on a Si substrate can be precisely controlled to under 5 nm by deposition techniques such as plasma-enhanced chemical vapor deposition (PECVD), revealing the highest surface roughness of 4 nm, as well as physical-vapor deposition and ion-beam deposition, revealing the lowest surface roughness of 0.2 nm [35]. In fabrication processes, the width variations of SWGs and slot waveguides are correlated. Therefore, we analyze performances versus ΔWcl = ±5 (10) nm while simultaneously altering WSi by the same range of ΔWSi = ±5 (10) nm. The obtained PERTE is around 20 (12) dB and the ILs of both modes are lower than 0.2 (0.3) dB (Fig. 11(b)). In the variation range of ΔhSi = ±5 (10) nm, only PERTE varies moderately and achieves a level of >25 (20) dB, and the ILs of the two modes are lower than 0.18 (0.2) dB (Fig. 11(c)). Achieving ΔWcl <5 nm is expected if we require PERTE >20 dB. Fortunately, the sidewall roughness of a Si strip can be reduced to <1 nm by adopting the mixed inductively coupled plasma-reactive ion etching (ICP-RIE) process and hydrogen annealing [36], in which the word “mixed” means that the etching and passivation processes of the conventional ICP-RIE work synchronously. Overall, the fabrication tolerance analyses for the three geometry parameters implies that Δtslot, ΔWcl, and ΔhSi have similar fabrication tolerances.

4. Summary

A compact, high-performance PBS is proposed, consists of two horizontally slotted waveguides with composite subwavelength gratings (CSWGs) comprising a SiO2 slot layer over a Si strip. Introduction of the CSWGs leads to significant coupling strength suppression and enhancement of the TE and TM modes, respectively, simultaneously achieving improved PER, reduced device length, and expanded working BW. Differing from mono-material SWGs, CSWGs comprising silicon strips covered with a silicon dioxide (SiO2) layer not only simplify the fabrication process, but further reduce the device length. Using numerical optimization, the present PBS with device length of Lcm = 20.85 µm achieved PERs of >31 dB and >26 dB for TE and TM modes, respectively, and ILs of <0.03 dB for both modes, at a wavelength of λ = 1550 nm. Working BWs of ∼50 nm for PERTM ∼ 30 dB, PERTE > 20 dB, and ILs < 0.1 dB were obtained for TE and TM modes. The proposed PBS can be further reduced to Lcm = 18.3 µm by selecting a duty cycle of 0.8, which moderately decreases the performance to PERTE = 28.9 dB, PERTM = 23.3 dB, ILTE = 0.19 dB, and ILTM = 0.11 dB. In terms of fabrication tolerances, PERs and ILs of >20 dB and <0.3 dB, respectively, were obtained with moderate variation in slot thickness (Δtslot = ±5 nm), CSWG width (ΔWcl = ±5 nm), and CSWG thickness (ΔhSi = ±5 nm). The proposed idea has application in various coupled-waveguide systems, to improve their overall performance and reduce device length for building dense photonic integrated circuits.

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. D. Dai, J. Bauters, and J. E. Bowers, “Passive technologies for future large-scale photonic integrated circuits on silicon: polarization handling, light non-reciprocity and loss reduction,” Light: Sci. Appl. 1(12), e1 (2012). [CrossRef]  

2. D. Dai, L. Liu, S. Gao, D. X. Xu, and S. He, “Polarization management for silicon photonic integrated circuits,” Laser Photonics Rev. 7(3), 303–328 (2013). [CrossRef]  

3. A. Rahim, T. Spuesens, R. Baets, and W. Bogaerts, “Open-access silicon photonics: current status and emerging initiatives,” Proc. IEEE 106(12), 2313–2330 (2018). [CrossRef]  

4. W. Bogaerts and L. Chrostowski, “Silicon photonics circuit design: method, tools, and challenges,” Laser Photonics Rev. 12(4), 1700237 (2018). [CrossRef]  

5. M. R. Watts, H. A. Haus, and E. P. Ippen, “Integrated mode-evolution-based polarization splitter,” Opt. Lett. 30(9), 967–969 (2005). [CrossRef]  

6. Y. Yin, Z. Li, and D. Dai, “Ultra-broadband polarization splitter-rotator based on the mode evolution in a dual-core adiabatic taper,” J. Lightwave Technol. 35(11), 2227–2233 (2017). [CrossRef]  

7. J. M. Hong, H. H. Ryu, S. R. Park, J. W. Jeong, S. G. Lee, E. H. Lee, S. G. Park, D. Woo, S. Kim, and O. B. Hoan, “Design and fabrication of a significantly shortened multimode interference coupler for polarization splitter application,” IEEE Photonics Technol. Lett. 15(1), 72–74 (2003). [CrossRef]  

8. L. S. Han, S. Liang, H. L. Zhu, C. Zhang, and W. Wang, “A High extinction ratio polarization beam splitter with MMI couplers on InP substrate,” IEEE Photonics Technol. Lett. 27(7), 782–785 (2015). [CrossRef]  

9. D. R. Solli, C. F. McCormick, and J. M. Hickmann, “Polarization-dependent reflective dispersion relations of photonic crystals for waveplate mirror construction,” J. Lightwave Technol. 24(10), 3864–3867 (2006). [CrossRef]  

10. J. M. Park, S. G. Lee, H. R. Park, and M. H. Lee, “High-efficiency polarization beam splitter based on a self-collimating photonic crystal,” J. Opt. Soc. Am. B 27(11), 2247–2254 (2010). [CrossRef]  

11. J. Feng and Z. Zhou, “Polarization beam splitter using a binary blazed grating coupler,” Opt. Lett. 32(12), 1662–1664 (2007). [CrossRef]  

12. Y. B. Tang, D. Dai, and S. He, “Proposal for a grating waveguide serving as both a polarization splitter and an efficient coupler for silicon-on-insulator nanophotonic circuits,” IEEE Photonics Technol. Lett. 21(4), 242–244 (2009). [CrossRef]  

13. B. Shen, P. Wang, R. Polson, and R. Menon, “An integrated-nanophotonics polarization beamsplitter with 2.4 × 2.4 µm2 footprint,” Nat. Photonics 9(6), 378–382 (2015). [CrossRef]  

14. H. Fukuda, K. Yamada, T. Tsuchizawa, T. Watanabe, H. Shinojima, and S. Itabashi, “Ultrasmall polarization splitter based on silicon wire waveguides,” Opt. Express 14(25), 12401–12408 (2006). [CrossRef]  

15. X. Guan, H. Wu, Y. Shi, L. Wosinski, and D. Dai, “Ultracompact and broadband polarization beam splitter utilizing the evanescent coupling between a hybrid plasmonic waveguide and a silicon nanowire,” Opt. Lett. 38(16), 3005–3008 (2013). [CrossRef]  

16. X. Guan, H. Wu, Y. Shi, and D. Dai, “Extremely small polarization beam splitter based on a multimode interference coupler with a silicon hybrid plasmonic waveguide,” Opt. Lett. 39(2), 259–262 (2014). [CrossRef]  

17. Y. Yue, L. Zhang, J. Y. Yang, R. G. Beausoleil, and A. E. Willner, “Silicon-on-insulator splitter using two horizontally slotted waveguides,” Opt. Lett. 35(9), 1364–1366 (2010). [CrossRef]  

18. H. Zhang, Y. Huang, S. Das, C. Li, M. Yu, P. G. Lo, M. Hong, and J. Thong, “Polarization splitter using horizontal slot waveguide,” Opt. Express 21(3), 3363–3369 (2013). [CrossRef]  

19. L. B. Soldano and E. C. M. Pennings, “Optical multi-mode interference devices based on self-imaging: principles and applications,” J. Lightwave Technol. 13(4), 615–627 (1995). [CrossRef]  

20. V. R. Almeida, Q. Xu, C. A. Barrios, and M. Lipson, “Guiding and confining light in void nanostructure,” Opt. Lett. 29(11), 1209–1211 (2004). [CrossRef]  

21. Q. Xu, V. R. Almeida, R. R. Panepucci, and M. Lipson, “Experimental demonstration of guiding and confining light in nanometer-size low-refractive-index material,” Opt. Lett. 29(14), 1626–1628 (2004). [CrossRef]  

22. R. Halir, P. J. Bock, P. Cheben, A. Ortega-Moñux, C. Alonso-Ramos, J. H. Schmid, J. Lapointe, D. X. Xu, J. G. Wangüemert-Pérez, Í. Molina-Fernández, and S. Janz, “Waveguide sub-wavelength structures: a review of principles and applications,” Laser Photonics Rev. 9(1), 25–49 (2015). [CrossRef]  

23. R. Halir, A. Ortega-Moñux, D. Benedikovic, G. Z. Mashanovich, J. G. Wangüemert-Pérez, J. H. Schmid, Í. Molina-Fernández, and P. Cheben, “Subwavelength-grating metamaterial structures for silicon photonic devices,” Proc. IEEE 106(12), 2144–2157 (2018). [CrossRef]  

24. P. Cheben, R. Halir, J. H. Schmid, H. A. Atwater, and D. R. Smith, “Subwavelength integrated photonics,” Nature 560(7720), 565–572 (2018). [CrossRef]  

25. Y. Xiong, J. G. Wangüemert-Pérez, D. X. Xu, J. H. Schmid, P. Cheben, and W. N. Ye, “Polarization splitter and rotator with subwavelength grating for enhanced fabrication tolerance,” Opt. Lett. 39(24), 6931–6934 (2014). [CrossRef]  

26. R. Halir, P. Cheben, J. M. Luque-González, J. D. Sarmiento-Merenguel, and J. H. Schmid, “Ultra-broadband nanophotonic beamsplitter using an anisotropic sub-wavelength metamaterial,” Laser Photonics Rev. 10(6), 1039–1046 (2016). [CrossRef]  

27. Y. Xu and J. Xiao, “Compact and high extinction ratio polarization beam splitter using subwavelength grating couplers,” Opt. Lett. 41(4), 773–776 (2016). [CrossRef]  

28. L. Liu, Q. Deng, and Z. Zhou, “Manipulation of beat length and wavelength dependence of a polarization beam splitter using a subwavelength grating,” Opt. Lett. 41(21), 5126–5129 (2016). [CrossRef]  

29. S. Jahani and Z. Jacob, “Transparent subdiffraction optics: nanoscale light confinement without metal,” Optica 1(2), 96–100 (2014). [CrossRef]  

30. S. Jahani and Z. Jacob, “Photonic skin-depth engineering,” J. Opt. Soc. Am. B 32(7), 1346–1353 (2015). [CrossRef]  

31. S. Jahani, S. Kim, J. Atkinson, J. C. Wirth, F. Kalhor, A. A. Noman, W. D. Newman, P. Shekhar, K. Han, V. Van, R. G. Decorby, L. Chrostowski, M. Qi, and Z. Jacob, “Controlling evanescent waves using silicon photonic all-dielectric metamaterials for dense integration,” Nat. Commun. 9(1), 1893 (2018). [CrossRef]  

32. H. Xu, D. Dai, and Y. Shi, “Ultra-broadband and ultra-compact on-chip silicon polarization beam splitter by using hetero-anisotropic metamaterials,” Laser Photonics Rev. 13(4), 1800349 (2019). [CrossRef]  

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

34. P. Lalanne, S. Astilean, P. Chavel, E. Cambril, and H. Launois, “Design and fabrication of blazed binary diffractive elements with sampling periods smaller than the structure cutoff,” J. Opt. Soc. Am. A 16(5), 1346–1353 (1999). [CrossRef]  

35. M. R. Amirzada, A. Tatzel, V. Viereck, and H. Hillmer, “Surface roughness analysis of SiO2 for PECVD, PVD and IBD on different substrates,” Appl. Nanosci. 6(2), 215–222 (2016). [CrossRef]  

36. F. Gao, Y. Wang, G. Cao, X. Jia, and F. Zhang, “Reduction of sidewall roughness in silicon-on-insulator rib waveguides,” Appl. Surf. Sci. 252(14), 5071–5075 (2006). [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 (11)

Fig. 1.
Fig. 1. (a) 3D schematic diagram and zoomed in views of the (b) input, (c) TE output, and (d) TM output ports of the proposed polarization beam splitter.
Fig. 2.
Fig. 2. Schematic of the fabrication processes for the proposed PBS.
Fig. 3.
Fig. 3. (a) Coupling lengths of TM (LTM) and TE (LTE) modes and (b) the coupling-length ratio LTE / LTM versus slot thickness tslot for PSWC and our structure.
Fig. 4.
Fig. 4. Field profiles of (a) TE (Ex)- and (b) TM (Ey)- symmetrical modes of the proposed PBS, and those of (c) TE- and (d) TM-symmetrical modes of the PSWC, for tslot = 50 nm, WSi = 400 nm, hSi = 150 nm, Wcl = 75 nm, g = 50 nm, and s = 550 nm. At the central plane of the slot layer (i.e., y = 175 nm from the top of the substrate), the field amplitudes along the x-direction of (a) and (c) are shown in (e), and those of (b) and (d) are shown in (f).
Fig. 5.
Fig. 5. Evolutions of the Poynting vector flow of a) TE and b) TM modes of the proposed PBS with LTM = 21.56 µm (Lcm = 20.85 µm), and those of c) TE and d) TM modes of the PSWC with LTM = 35.71µm (Lcm = 33.4 µm), for tslot = 50 nm, hSi = 150 nm, Wcl = 75 nm, gap = 50 nm, WSi = 400 nm, and s = 550 nm.
Fig. 6.
Fig. 6. (a) PER and (b) IL versus the wavelength with tslot = 50 nm for the proposed PBS with Lcm = 20.85 µm and the PSWC with Lcm = 33.4 µm.
Fig. 7.
Fig. 7. PER (left axis) and IL (right axis) versus slot thickness, tslot, of the proposed PBS.
Fig. 8.
Fig. 8. (a) PER (left axis) and IL (right axis) and (b) modified coupling length of TM mode, Lcm, versus thickness of CSWGs, hSi, of the proposed PBS.
Fig. 9.
Fig. 9. (a) PER (left axis) and IL (right axis) and (b) modified coupling length of TM mode, Lcm, versus duty cycle, ρ, of the proposed PBS.
Fig. 10.
Fig. 10. (a) PER (left axis) and IL (right axis) and (b) modified coupling length of TM mode, Lcm, versus number of strips of the CSWGs
Fig. 11.
Fig. 11. PER (left axis) and IL (right axis) versus variation in (a) slot thickness, Δtslot, (b) cladding width, ΔWcl, and (c) thickness of CSWGs, ΔhSi, of the proposed PBS.

Equations (5)

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

n e m t 2 = d i a g [ n x x 2 , n y y 2 , n z z 2 , ] ,
n y y 2 = n z z 2 = ρ n h i g h 2 + ( 1 ρ ) n l o w 2 ,
1 n x x 2 = ρ n h i g h 2 + ( 1 ρ ) n l o w 2 ,
PE R T E ( T M ) = 10 log 10 ( P T E ( T M ) , T R U ( C R O ) P T M ( T E ) , T R U ( C R O ) ) ,
I L T E ( T M ) = 10 log 10 ( P T E ( T M ) , T R U ( C R O ) 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.