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

Polarization-division and spatial-division shared-aperture nanopatch antenna arrays for wide-angle optical beam scanning

Open Access Open Access

Abstract

Chip-based optical beam scanners hold promise for future compact high-speed light detection and ranging (LIDAR) systems. Many of the demonstrated chip-based optical beam scanners are designed based on diffraction-based waveguide gratings as on-chip antennas. The waveguide grating antenna, however, only provides a typical field-of-view (FOV) of roughly 10° by tuning the input light wavelength. In this paper, polarization-division and spatial-division multiplexed nanoantenna arrays are proposed to expand the FOV of on-chip antennas. The proposed device, based on silicon-on-insulator (SOI) platform, consists of three nanoantenna groups which are densely packed and fed by a common silicon nanostrip. It is demonstrated that the combination of the optical mode-multiplexing technique and the antenna engineering allows independent controls over the interactions between multiple nanoantenna groups and the waveguide. By proper engineering of the antenna dimensions, the proposed device achieves a FOV of over 40° within a 100 nm wavelength tuning range, almost tripling that of the conventional waveguide grating antenna.

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

1. Introduction

Optical beam scanners (OBSs), capable of dynamically steering light beams, constitute a key building block for numerous technologies like sensing [12], imaging [34], image projections [5], displays [6], light detection and ranging (LiDAR) [78] and autonomous driving [9]. Traditionally, optical beam scanning can be achieved by mechanically rotating reflective mirrors or using liquid-crystal spatial light modulators. However, their low response speed and bulky structure have posed severe limitations on their practical applications.

For agile and compact OBSs, it has emerged over the past decade that solid-state OBSs, which allow for a dense integration of thousands of optical antennas into an area of fingernail sizes [10] and a beam scanning speed over GHz [11], could be a practical solution. Tremendous efforts have been made in the developments of solid-state OBSs, and different light-beam-steering mechanisms have been explored, including those based on micro-electro-mechanical systems (MEMSs) [12], those using on-chip phase modulators to align the phases of on-chip antennas [10,1317] and those scanning the beam by controlling the input light wavelength [1824]. Among them, MEMS-actuated OBSs provide simultaneously a relatively shorter response time of microsecond level and a large aperture of over 1mm2 [12] for ultra-small beam divergence. However, the MEMS-based OBSs suffer from limited field-of-views (FOVs) [12], because the multi-wavelength sizes of each grating element give rise to grating lobes in the far-field zone. Meanwhile, the MEMS-based OBSs are driven by external out-of-plane far-field illuminations, which means the device cannot be fully integrated. As a result, the inherent flexibility of the on-chip OBS is reduced. Comparatively, phased OBSs deal with in-plane propagating light, which means the antennas, the phase modulators and even the optical sources could be integrated on the same wafer [25]. Moreover, because no mechanical actuations are involved in phase modulations, phased OBSs are totally inertial-free with potentially ultra-high scan rate above GHz using electro-optic modulators [11,26]. Unfortunately, a main drawback of OBSs is the difficulty in constructing large-aperture beam scanners. As the scale of the antenna array and the number of control units increase, the coherent controls over hundreds and even thousands of on-chip antennas become challenging [13]. Meanwhile, the crosstalk between on-chip waveguides prevents the spacing between adjacent on-chip antennas from being reduced to sub-wavelength scale [1516,27]. With an inter-element spacing of over one wavelength, there are more than one emission angle that could fulfill the constructive interference condition of light and then grating lobes appear, significantly reducing the FOV and efficiency of the device.

Alternatively, optical beam scanning can also be achieved passively using on-chip antennas with wavelength-dependent beams, e.g. waveguide grating antennas, in conjunction with wavelength-tunable laser sources [1824]. Compared to MEMS-based and phased OBSs, the wavelength-controlled OBSs can also be fast and show easier implementations and lower cost, because they do not require additional electronic control units and are free from complex power distribution networks. Meanwhile, in the wavelength-controlled manner, large-aperture small-beam-divergence scanners can be achieved by engineering the antenna structures to control the light emission rates along the antenna apertures [1819], which are not as challenging as scaling up the phased on-chip antenna array. As shown in [18], an ultra-small beam divergence with a full-width-half-maximum below 0.1° can be achieved using silicon nitride as overlays. More importantly, the spacing between the grating elements or the nanoantenna elements in the wavelength-controlled beam scanners can be made significantly smaller than the wavelength even below ideal half-wavelength, which means the light beam steering is inherently grating-lobe-free and thus it provides higher device efficiency than those with grating lobes. The simple implementations, grating-lobe-free beam steering and the reach to narrow beam-divergence make wavelength-controlled OBSs very suitable for applications that are cost-conscious and require ultra-fine spatial resolutions like LIDAR systems on unmanned vehicles. As an example, the OBSs formed by a linear array of wavelength-agile gratings, controlling the beam direction by wavelength in one dimension while using the phased approach to steer the beam along the other dimension, have provided a practical solution for large-aperture two-dimensional scanning demanded by LIDAR [11,14,17,2729], after intensive research efforts devoted. Two-dimensional-scanning beam scanner controlled purely by wavelength has also been demonstrated [24]. Except for silicon, wavelength-controlled OBSs are widely adopted in LIDAR researches involving various integrated photonic platforms including silicon nitride [30], plasmonic [2223] and III-V materials [11].

For practical applications of OBSs, a large FOV is always pursued. Unfortunately, without special engineering, typical FOV of the wavelength-controlled OBS, e.g., the waveguide grating antenna, is only roughly 10° with a wavelength variation of 100 nm (0.1°/nm) [14,20,2829]. Although further increasing the wavelength range could widen the scanning range, the wavelength tuning range of over 100 nm is not commonly available and could be challenging for on-chip laser sources. As a result, most of the existing 2-D solid-state OBSs feature a significantly smaller scanning range in their wavelength-controlled dimension compared to that in the phased dimension [14,17,27,29]. To solve this problem, lattice-shifted photonic crystal waveguides have been proposed to make the emission angle sensitivity against wavelength increase from 0.1°/nm of the standard waveguide grating antenna to 1.0°/nm with a steering range of 23° demonstrated [20]. However, the photonic bandgap can only be formed within a relatively narrow wavelength range (30 nm in [20]), preventing further expansion of the scanning range.

Nanoantennas, with their unique abilities to concentrate light into deep-subwavelength scale and tailor the emission pattern of light, have received great attention in the past decade. For chip-based applications, silicon photonic devices with different radiation characteristics have been demonstrated by taking full advantages of the flexible design freedoms provided by nanoantennas. These nanoantenna-based devices have enabled a lot of fascinating applications including reconfigurable on-chip interconnects [31], ultra-compact lab-on-a-chip microflow cytometers [32], wireless on-chip communications and networks [3336]. However, the majority of these researches, focuses on the manipulation of in-plane light emission. The potential of using nanoantenna array as the on-chip OBS has largely remained unexplored.

In this paper, we report a nanoantenna-based wavelength-controlled OBS, achieving a FOV over 40°, significantly higher than the previously reported grating-based designs. The proposed device is based on the silicon-on-insulator (SOI) platform and the principles described here can also be applied to other integrated photonic platforms. The proposed device leverages the design flexibilities provided by the aperture-coupled nanopatch antennas and the integrated photonic mode multiplexing technique to engineer the waveguide-antenna couplings of multiple shared-aperture nanoantenna groups simultaneously, thereby enabling the independent controls over the light emissions from these nanoantenna groups. We show that with proper designs of the antenna structures, a large FOV can be obtained by combining the scanning ranges of all the constituent nanoantenna groups. Compared to the previous designs aiming to expand the FOV of the on-chip OBS [15,16], the proposed device is significantly different from the following two aspects. Firstly, light emission of the proposed device is vertical to the chip surface while the designs in [15,16] deal with in-plane light emission. The capability of achieving wide-angle vertical light beam steering is crucial for photonic integrated circuit applications in terms of bridging the gap between out-of-plane free space light and in-plane guided waves [37]. Secondly, beam scanning in the proposed device is achieved by the wavelength-controlled manner which does not require active control units and thus has the advantages of easy implementation and low cost. Regarding to wavelength-controlled OBSs, the already reported designs achieve 23° and 28° FOVs within 30 [20] and 200 nm [11] wavelength tuning range, respectively. In comparison, the proposed device achieves a 41.2° FOV within 100 nm wavelength tuning range.

2. Device overview and working principles

Figure 1(a) depicts a schematic overview of the proposed device. It consists of the optical multiplexing section, the antenna section and the transition section between them. The launched light from different input ports will excite different modes in the bus waveguide after the optical multiplexing section which is composed of a polarization beam splitter (PBS) and an asymmetrical directional coupler (ADC), as shown in Fig. 1(b). There are three nanoantenna groups arranged on top of the silicon nanostrip waveguide. There is only one nanoantenna group to interact with one mode launched in the bus waveguide to convert the guided waves into free space waves efficiently. Each of the nanoantenna group is designed to cover a specific angular range in free space and a large FOV can be obtained by combining the scanning ranges of all the three nanoantenna groups. Meanwhile, the covered angular range of each nanoantenna group must be continuous for blind-spot-free beam scanning and without overlapped angles for efficient scanning range expansion.

 figure: Fig. 1.

Fig. 1. Schematic and structure of the proposed device. (a) Schematic overview of the proposed device consisting of the optical multiplexer section, the transition section and the antenna section. The inset shows part of the aperture of the antenna section. (b) Top-view of the optical multiplexing section. (c) 3-D view of the antenna section of the proposed device. (d) Side-view of the antenna section of the proposed device.

Download Full Size | PDF

The above described functionalities necessitate the design flexibilities on the independent controls of the interactions between each launched mode in the bus waveguide and different nanoantenna groups. Conventional antennas used in the wavelength-controlled OBSs, e.g., waveguide grating antennas, however, fall short of this freedom. Here, the aperture-coupled nanopatch antenna based on the structure in [38] is adopted and engineered to control the interactions between the waveguide and different nanoantenna groups. Figures 1(c) and 1(d) show the 3-dimensional view and the cross-section view of the antenna structure. It is formed by stacking two additional metallic layers on top of the standard SOI nanostrip waveguide with low-index dielectrics, e.g., SU-8 photoresists, as the spacers. The first metallic layer is formed by etching a nanoslot array in the silver thin film, while the second layer is composed of a nanopatch array. The silicon nanostrip, the SU-8 cladding and the silver thin film form a hybrid plasmonic waveguide. The nanopatches couple with the waveguide through the nanoslots. Considering the practical fabrication process of the structure, the nanoslots are also filled with SU-8 in the simulations.

Depending on how the interactions between the waveguide modes and antennas are controlled, the working principles are described in the following two parts, i.e., the polarization-division multiplexed nanopatch arrays (PDMNA) and the spatial-division multiplexed nanopatch arrays (SDMNA). Finally, the combination of the two kinds of arrays is explored to achieve the wide-angle beam scanning. Here, the PDMNA is referring to two multiplexed nanopatch groups composed of two different antenna elements whose couplings with the waveguide are sensitive to the polarization of the launched mode, e.g., TE and TM modes. The SDMNA is referring to two multiplexed nanopatch groups with identical antenna elements but giving different response to waveguide modes with different orders and the same polarization, e.g., TE0 and TE2 modes.

The rest of the paper is organized as follows. In section 2.1, the design of the antenna elements is presented. Sections 2.2 and 2.3 respectively introduce PDMNA and SDMNA, explaining how two nanopatch groups can be multiplexed with low crosstalk. Then, based on the principles described in Section 2, the design of the proposed device in Fig. 1 is presented in Section 3, including the design of the antenna section, the optical multiplexing section and the transition section. Finally, the fabrication tolerance is discussed in Section 4.

2.1 Antenna element

Figures 2(a) and 2(b) show the models of the x- and y-polarized elements. The nanopatch in the x-polarized element has its longitudinal dimension (size along the y-direction) remarkably smaller than its lateral one (size along the x-direction), making its x- polarized fundamental mode resonate at around 1550 nm and the y-polarized one resonate at a much shorter wavelength. The x-polarized nanopatch couples with the silicon nanostrip waveguide via the y-oriented nanoslot (Wsa << Lsa) etched on the silver thin film. In contrast, the y-polarized element is excited by the x-oriented nanoslot (Lsb << Wsb). The parameters of the x- and the y-polarized elements are Wpa = 320, Lpa = 150, Lsa = 350, Wsa = 80 and Wpb = 150, Lpb = 320, Lsb = 80, Wsb = 250, respectively, all in nm. The other parameters are W = 1200, H = 240, T1 = 80, H1 = 200 and T2 = H2 = 100, all in nm. The x-polarized element is designed to interact with the TE0 mode efficiently. The simulated coupling coefficients of the x-polarized element with the TE0 and TM0 mode are −28.67 and −46.83 dB at the wavelength of 1620 nm, respectively, with a difference of 18.16 dB. The coupling coefficient here is determined by doing the Poynting vector integral on the nanoslot upper surface, which denotes the power that goes up into the cavity formed by the nanopatch and the silver film. The simulations are performed using High-Frequency Structure Simulator (HFSS) from ANSYS which is based on finite element method [39]. The antenna radiation characteristics are obtained by near-to-far-field projections on a closed box surrounding the antenna. In the simulations, one pair of the excitation ports are added on both sides of the waveguide to excite and terminate the waveguide. The refractive indices of the silicon nanostrip, the silica substrate and the SU-8 cladding are 3.48, 1.444 and 1.575, respectively. The Drude model is employed to fit the permittivity of silver with εinf= 5, ωp= 13.4×1015 rad/s and Γ = 1.12×1014 1/s. The polarization-dependent couplings with the waveguide are achieved by engineering the nanoslot dimensions. Because the nanopatch is fed by etching the nanoslot on the silver thin film, which performs perturbations to the waveguide fields by cutting the currents flowing on the silver thin film, the way how the currents are disturbed results in different antenna-waveguide couplings. Figures 2(c) and 2(d) show the electric field distributions on the xoz plane and the induced surface currents on the bottom surface of the silver thin film for both the TE0 and the TM0 waveguide modes. It is seen that the induced currents of the TE0 mode are mostly flowing along the x-direction, i.e., perpendicular to the wave propagation direction. In contrast, the currents of the TM0 mode are dominated by their y-components which are in parallel to the wave propagation direction. For the x-polarized element, when the TE0 mode is launched, the induced surface currents along the x-direction are cut off by the nanoslot. As a result, displacement currents polarized along the x-direction will be induced within the nanoslot to fulfill the current continuity formula. On the other hand, when the TM0 mode is launched, the y-direction surface currents of the TM0 mode experience a much weaker disturbance by the same y-oriented nanoslot, considering its small lateral dimension Wsa (80 nm). Following the similar principle, the y-polarized element couples with the TM0 mode 35.9 dB stronger than with the TE0 mode and the corresponding coupling coefficients between the TE0 and the TM0 mode are −55.12 and −19.22 dB, respectively. The coupling coefficients of the two types of elements versus the TE0 and TM0 modes are summarized in Table 1. It can be seen when the TE0 and TM0 modes are launched, the dominating element, i.e., the element exhibiting stronger interaction with the waveguide, is the x- and y-polarized one, respectively, with coupling coefficient differences of 26.45 and 27.61 dBc. In the following discussions, the parameters of the antenna elements are the same as those in Fig. 2. The dependences of the element coupling coefficient upon wavelength are investigated and shown in Fig. 2(e). The coupling coefficients of the two types of elements do not experience significant variation within a 100 nm wavelength range. More importantly, the polarization-sensitive couplings are well maintained within the 100 nm wavelength range.

 figure: Fig. 2.

Fig. 2. Antenna elements. (a) and (b). Models of the x- and y-polarized elements. To clearly show the nanoslot, the SU-8 spacer is set to be transparent. The parameters are Wpa = 320, Lpa = 150, Lsa = 350, Wsa = 80, Wpb = 150, Lpb = 320, Lsb = 80, Wsb = 250, W = 1200, H = 240, T1 = 80, H1 = 200 and T2 = H2 = 100, all in nm. (c) and (d). Electric field distributions on the xoz plane and the induced current distributions on the bottom surface of the silver film of the hybrid plasmonic waveguide under TE0 (c) or TM0 (d) mode excitations at the wavelength of 1620 nm. The parameters of the waveguide are the same as those in (a) and (b). (e) Element coupling coefficient versus wavelength.

Download Full Size | PDF

Tables Icon

Table 1. Summary of the coupling coefficients of the x- and y-polarized antenna elementsa

2.2 PDMNA

Figure 3(a) shows the structure of the PDMNA, which is used to illustrate the principles of the polarization-division shared-aperture technique. It consists of two nanopatch groups, respectively composed of the x- and y-polarized elements. In the designed PDMNA, both two nanopatch groups have 10 elements with an identical element spacing of 650 nm (Dy1p = Dy2p = 650 nm) and are arranged along the y-axis with Dx1p = Dx2p = 0, i.e., the central line of the waveguide on the xoy plane, as depicted in the top panel of Fig. 3(b). Since the element coupling coefficient is stable within the 100 nm range [Fig. 2(e)], the 1550 nm telecommunication wavelength is selected as the operating wavelength to illustrate the principles. As the dominant component of the cavity mode supported by the nanopatches [4041], Ez below the nanopatches is monitored when the TE0 or TM0 mode is launched. As shown in the middle panel of Fig. 3(b), when the TE0 mode is launched, the fundamental mode of Group P1, i.e., the x-polarized group, is efficiently excited with remarkable Ez distributed on the nanopatch short edges. On the contrary, the Ez intensities beneath Group P2, i.e., the y-polarized group, is negligible. In the case of TM0-mode excitation, remarkable Ez distributed on the short edges of the y-polarized group is observed while Ez distributed beneath the x-polarized group becomes much weaker. The Ez distributions below the y-polarized group coincide with that of the nanopatch fundamental cavity mode polarized along the y-direction. The field distributions in Fig. 3(b) indicate that the element design make each waveguide mode interacts efficiently with only one nanopatch group (the x-polarized group for the TE0 mode and the y-polarized group for the TM0 mode). The far-field emission patterns of the PDMNA in Fig. 3(b) are plotted in Fig. 3(c). In Fig. 3(c), Dθ and Dφ denote the directivity of an antenna for the θ and φ field components and are expressed as: ${D_\theta } = 4\pi \frac{{{U_\theta }(\theta ,\varphi )}}{{{P_{rad}}}}$ and ${D_\varphi } = 4\pi \frac{{{U_\varphi }(\theta ,\varphi )}}{{{P_{rad}}}}$, where Uθ(θ, φ), Uφ(θ, φ) and Prad represent the radiated power density contained in θ and φ field components and the total radiated power [42]. Gθ and Gφ denote the gain of an antenna for the θ and φ field components and are expressed as: ${G_\theta } = 4\pi \frac{{{U_\theta }(\theta ,\varphi )}}{{{P_{accepted}}}}$ and ${G_\varphi } = 4\pi \frac{{{U_\varphi }(\theta ,\varphi )}}{{{P_{accepted}}}}$, respectively, where Paccepted represents the accepted power of the antenna [42]. It is seen that the main beams are steered to 26° and −8° when the TE0 and TM0 mode are launched, respectively. Meanwhile, the cross-polarization levels are both below −50 dB and the sidelobe levels (SLLs) are both below −10 dB. The corresponding radiation efficiencies under TE0 and TM0 excitations are 63.58% and 57.78%, respectively.

 figure: Fig. 3.

Fig. 3. (a) Illustration of the PDMNA layer by layer. (b) Top panel: the model of the designed PDMNA. Other parameters are the same as those in Fig. 2. Middle and bottom panels: Ez intensity distributions on a cut-plane 10 nm below the nanopatches at 1550 nm when the TE0 (the middle panel) or TM0 mode (the bottom panel) is launched into the designed PDMNA, showing the polarization-division excitations of the two nanopatch groups. To clearly show the difference between the excitation states of two groups, the Ez intensity is normalized in dB scale. (c) The emission patterns of the designed PDMNA on the yoz plane when the TE0 or TM0 mode is launched into the waveguide at 1550 nm.

Download Full Size | PDF

2.3 SDMNA

The modal properties of the modes with different orders are exploited to construct the SDMNA. The electric field distributions of the TE0, TE1 and TE2 modes and their corresponding induced currents in the silver thin film are depicted in Fig. 4(a). Depending on the mode order, the positions of the zeros and poles are different for three modes. As explained in Section 2.1, if y-oriented nanoslots are etched at the zeros of Jx, the interactions between the antenna and the waveguide is suppressed. The same holds true for x-oriented nanoslots etched at the zeros of Jy. Figure 4(b) plots the simulated coupling coefficients between the x-polarized element and the TE0, TE1 and TE2 modes versus the element lateral positions. The Jx intensity along Reference line 1 on the inset of Fig. 4(b) of the TE0, TE1 and TE2 modes are plotted in Fig. 4(c). As seen from Fig. 4(b), sharp dips of the coupling coefficient are observed at Dx = 0 and 200 nm for TE1 and TE2 modes, respectively, which are consistent with the positions of the minimum Jx in Fig. 4(c). The coupling coefficients of the TE0 mode, in comparison, gradually decrease with the increasing Dx due to the evanescent decay of the field intensity. The spatial dependence of the antenna-waveguide coupling implies two possible feeding mode combinations for the SDMNA, i.e., TE0 and TE1 modes or TE0 and TE2 modes. Taking the latter one as the example, Fig. 4(c) shows the element arrangement for the SDMNA fed by the TE0 and TE2 modes. One pair of the x-polarized elements are placed at Dx1s = ±200 nm (Group S1), i.e., the minimum Jx of the TE2 mode, and thus will couple with only TE0 mode efficiently. Besides, there is another pair of the x-polarized elements placed at Dx2s = ±700 nm (Group S2), flanking Group S1. When the TE0 mode is launched, due to the evanescent decay of the TE0 mode modal field along the lateral direction, Group S1 will receive stronger coupled fields than Group S2. The simulations indicate that the difference is 19.71 dB, as marked out by the blue circles in Fig. 4(b). On the other hand, when the TE2 mode is launched, only Group S2 can couple with the waveguide efficiently, because the lateral positions of Group S1 coincide with the minimum Jx of the TE2 mode. As marked out by the red circles in Fig. 4(b), the elements of Group S2 couple with the TE2 mode by 14.52 dB stronger than those of Group S1. It is also found that the elements of Group S2 show stronger couplings with the TE2 mode than with the TE0 mode, due to the two poles of Jx of the TE2 mode at 500 nm away from the center.

 figure: Fig. 4.

Fig. 4. (a) Upper panel: Jx and Jy distributions of the TE0, TE1, TE2, TM0, and TM1 modes in the silver thin film. The black solid lines denote the geometrical profile of the silver thin film and the silicon waveguide. For each mode, the intensities are normalized by the maximum value of the current component with higher intensity so that the relative relations of the two components can be revealed. Bottom panel: Ex distributions of the TE0, TE1 and TE2 modes and Ez distributions of the TM0 and TM1 modes at the wavelength of 1620 nm. All on the xoz cross section. (b) Coupling coefficients of the x-polarized element between the TE0, TE1 and TE2 modes versus the element lateral position at 1620 nm. (c) Upper panel: normalized Jx intensity along Reference line 1 in (b) of the TE0, TE1 and TE2 modes at 1620 nm. Bottom panel: schematic showing the corresponding antenna element positions for the designed SDMNA. Other parameters are the same as those in Fig. 2.

Download Full Size | PDF

Based on the above described method, a SDMNA is designed to illustrate the principles of the spatial-division shared-aperture technique. The structure of the designed SDMNA are shown in Figs. 5(a) and 5(b). The parameters are Dx1s = ±200, Dy1s = 650, Dx2s = ±700 and Dy2s = 771, all in nm. Each nanopatch group has 10 elements along the y-direction. The Ez intensity distributions below the nanopatches when the TE0 or TE2 modes are launched at 1550 nm are shown in the bottom panel of Fig. 5(b). It is seen that they are concentrated below Groups S1 and S2 when the TE0 and TE2 modes are launched, respectively. The corresponding far-field emission patterns are plotted in Fig. 5(c). The main beam directions are 28° and 15° under TE0 and TE2 excitations, respectively. Moreover, the SLLs for both cases are below −9 dB. The SLL could be further optimized by engineering the nanoslot dimensions to control the amplitude distribution along the antenna aperture.

 figure: Fig. 5.

Fig. 5. (a) Illustration of the SDMNA layer by layer. (b) Top panel: model of the designed SDMNA. Other parameters are the same as those in Fig. 2. Bottom panel: Ez intensity distributions on the cut-plane 10 nm below the nanopatches at 1550 nm when the TE0 (the bottom-left panel) or TE2 mode (the bottom-right one) is launched into the SDMNA, showing the mode-division excitation behaviors. The Ez intensity is normalized in dB scale to clearly show the difference between the excitation states of two groups. (c) Emission patterns of the designed SDMNA on the yoz plane at 1550 nm when the TE0 or TE2 mode is launched. G and D represent antenna gain and directivity, respectively. (d) neff of the TE0 and TE2 modes. (e) Element coupling coefficients of the SDMNA versus wavelength.

Download Full Size | PDF

The radiation efficiencies under TE0 and TE2 excitations are 69% and 35%, respectively. The antenna radiation efficiency η can be expressed as: η = Prad/(Prad+ La+ Lwg), where La and Lwg denote the dissipated loss in the antenna and waveguide, respectively. The lower radiation efficiency under TE2 excitation is caused by two reasons: the higher Lwg and the lower Prad. Figure 5(d) depicts the imaginary part of the effective mode indies neff= neff - jneff of the TE2 and TE0 modes. The waveguide attenuation constant α is expressed as α = neff”k0. neff can be obtained by the eigenmode analysis of the waveguide structure. As seen from Fig. 5(d), neff of the TE2 mode is at least 1.73 times higher than that of the TE0 mode, giving rise to higher Lwg and the reduced radiation efficiency. Figure 5(e) depicts the antenna element coupling coefficients of the two groups. The element coupling coefficient of the dominant group under TE2 excitation (Group S2) is at least 4.36 dB lower than that under TE0 excitation (Group S1). Because Prad is proportional to the antenna-waveguide coupling coefficient, the lower element coupling coefficients under TE2 excitation also give rise to the reduced radiation efficiency.

Note that, in principle, TE0 and TE1 modes can also be exploited to construct the SDMNA. In that case, Group S1 is replaced with a x-polarized group at the center of the waveguide, i.e., the position of the minimum Jx of the TE1 mode. Group S2 can be retained to couple with the TE1 mode. However, it is observed in Fig. 4(a) that the phases of Jx of the TE1 mode are opposite with respect to the symmetrical plane of the waveguide. It means the emitted light of the elements of Group S2 will unfortunately cancel each other in the yoz plane, i.e., the scanning plane. Therefore, TE0 and TE1 modes are not an ideal combination to construct the SDMNA.

3. Design and results

3.1 Antenna section: wide-angle beam scanning enabled by combination of PDMNA and SDMNA

The combination of the PDMNA and the SDMNA provides more design freedoms to expand the device FOV, but proper engineering of the element types (x- or y-polarized) and positions of the nanopatch groups is required to insure their compatibility. Figure 6(a) depicts the structure of a shared-aperture nanopatch array consisting of three multiplexed nanopatch groups. Based upon the design principles of the SDMNA, one can multiplex two nanopatch groups fed by TE0 and TE2 modes by controlling the lateral positions of the two nanopatch groups. The PDMNA in Fig. 3, in contrast, does not require staggering two nanopatch groups laterally, facilitating its integration. The design goal is to insure that, for each of the launched mode, there is only one nanopatch group interacting efficiently with the launched mode. Hence, as depicted in Case I in Fig. 6(b), an intuitive way to multiplex three nanopatch groups is directly adding a TM0-mode-fed y-polarized group into the SDMNA in Fig. 5(b), at the same lateral positions of Group S1. Table 2 lists the coupling coefficients of the elements of the three nanopatch groups in Case I, i.e., the x-polarized group at Dx1 = ±200 nm (Group 1), the y-polarized one at Dx2 = ±200 nm (Group 2) and the x-polarized one at Dx3 = ±700 nm (Group 3), calculated at 1620 nm. It is seen that when the TM0 mode is launched, the elements of Group 2 indeed exhibit a coupling over 10 dB stronger than the elements of the other two nanopatch groups. The same holds true for Group 1 when the TE0 mode is launched. Nonetheless, the TE2 mode, which is expected to couple with Group 3, shows unwanted high coupling coefficients (−22.72 dB, the bold number in Table 2) with the elements of Group 2. This unwanted coupling results from the disturbance of Jy of the TE2 mode [Fig. 4(a)]. Although the phases of Jy of the TE2 mode are opposite with respect to the symmetrical plane of the waveguide, the unwanted coupling still reduces the device efficiency.

 figure: Fig. 6.

Fig. 6. (a) Illustration of the structure of the shared-aperture nanopatch array with three multiplexed nanopatch groups. (b) Schematics of two cases investigated. (c) Upper panel: Jx and Jy intensity distributions along Reference line 1 in the inset of Fig. 4(b) for the TE0, TM0 and TM1 modes, calculated at 1620 nm. Lower panel: schematic showing the antenna element positions in Case II. Other parameters are the same as those in Fig. 2.

Download Full Size | PDF

Tables Icon

Table 2. Summary of the element coupling coefficients of the three nanopatch groups in Case I in Fig. 6(b)

To solve this issue, an alternative method is proposed. The design is shown in Figs. 6(b) and 6(c), referred to as Case II. The three nanopatch groups are fed by TE0, TM0 and TM1 modes. Table 3 lists the element coupling coefficients of the nanopatch groups in Case II between the TE0, TM0 and TM1 modes at 1620 nm. In Case II, Groups 1 and 2 are composed of the x-polarized elements at Dx1 = ±350 nm and the y-polarized elements at Dx2 = 0 nm, respectively. Groups 1 and 2 are multiplexed in the polarization-division manner. As seen in Table 3, the coupling coefficient of the TE0 mode between Group 1 is 23.81 dB higher than that between Group 2. The coupling coefficient of the TM0 mode between Group 2 is 16.59 dB higher than that between Group 1. Figure 6(c) depicts the Jx and Jy intensity distributions of the TE0, TM0 and TM1 modes along Reference line 1 in the inset of Fig. 4(b). For each mode, the intensities are normalized by the maximum value of the current component with higher intensity so that the relative relations of the two components can be revealed. Groups 1 and 2 are placed at the minimum points of Jx [the green curve with dots in Fig. 6(c)] and Jy [the orange curve in Fig. 6©] of the TM1 mode, respectively. As a result, the coupling coefficients between Groups 1 and 2 and the TM1 mode are both below −45 dB (the 3rd row of Table 3). Note that similar strategy can also be applied to the TE2 mode, whose Jx and Jy distributions both feature at least one minimum [Fig. 4(a)]. Nonetheless, the minimum Jx of the TE2 mode are distributed at 200 nm away from the center [Fig. 4(c)]. Considering the sizes of the nanopatch element, it would be challenging to accommodate two nanopatch groups within the 200 nm space. Group 3 is composed of the x-polarized elements at Dx3 = ±700 nm. It is excited by cutting Jx of the TM1 mode [the green curve with dots in Fig. 6(c)]. The coupling coefficient between the TM1 mode and the element of Group 3 is −29.87 dB. Because the couplings between the TM1 mode and Groups 1 and 2 are suppressed by taking advantages of the current distributions, the coupling of the TM1 mode between Group 3 is 16.19 and 20.98 dB stronger than between Groups 1 and 2, respectively. In addition, in spatial-division manner, the coupling coefficients of the TE0 and TM0 modes between Group 3 are 17.72 and 16.91 dB lower than between Groups 1 and 2, respectively. The coupling coefficients of the dominant groups, i.e., the group exhibiting the strongest interaction with the waveguide, are shown in bold in Table 3. It is seen that under TE0, TM0 and TM1 mode excitations, the dominant groups are Groups 1, 2 and 3, respectively, with coupling coefficients at least 17.72, 16.59 and 16.19 dB higher than those of other two groups, respectively.

Tables Icon

Table 3. Summary of the coupling coefficients of the antenna elements of the proposed OBS [Case II in Fig. 6(b)] fed by TE0, TM0 and TM1 modes

The beam direction θi of the ith nanopatch group can be calculated according to: ${k_0}{D_{yi}}\sin ({\theta _i}) = {\rho _i}$, where Dyi and ρi are element spacing of the ith group and the element feeding phase step of the ith group, respectively. ρi is determined by:${\rho _i} = 2\pi (\frac{{n_{eff}^i{D_{yi}}}}{{{\lambda _0}}} - 1)$, where nieff and λ0 represent the waveguide mode index and the free space wavelength, respectively. For continuous and blind-spot-free beam scanning, the beam scanning ranges of the three nanopatch groups must satisfy: θU1 ≤ θL2 and θU2 ≤ θL3. Here, θU1 and θU2 represent the beam directions of Groups 1 and 2 at the upper edge wavelength of the available wavelength tuning range, respectively, while θL2 and θL3 represent the beam directions of Groups 2 and 3 at the lower edge wavelength of the available wavelength tuning range, respectively. When θU1 = θL2 and θU2 = θL3, there are no overlapped angles and the total beam scanning range is the sum of their respective ones. In this design, the wavelength tuning range is set to be 100 nm (1520-1620 nm), which is within the range available with commercial C-L band laser sources, as adopted in literatures relating to wavelength-controlled OBSs [2728]. Moreover, the 100 nm bandwidth is a reasonable one achievable for optical multiplexers.

Based on the above described principles and the three nanopatch groups in Case II, a wavelength-controlled OBS is designed. The mode indices of the TE0, TM0 and TM1 modes are 2.8632, 2.2722, 2.0762 at 1520 nm, respectively and 2.8078, 2.1729 and 1.9566 at 1620 nm, respectively. Groups 1, 2 and 3 in Case II are designed to cover angular range from 11.2° to 24.3°, from −2.1° to 11.5° and from −16.9° to −1.8°, respectively. The corresponding element spacings of Groups 1, 2 and 3 are Dy1 = 620 nm, Dy2 = 733 nm and Dy3 = 721 nm, respectively. In this case, θU1, θL2, θU2 and θL3 are 11.2°, 11.5°, −2.1° and −1.8°, respectively. Figure 7(a) shows the model of the proposed OBS, in which each nanopatch group has 16 elements along the y-direction. From left to right, Fig. 7(b) gives the Ez intensity distributions below the nanopatches at 1550 nm when the TE0, TM0 and TM1 modes are launched into the waveguide, respectively. Consistent with the above analysis, when TE0, TM0 and TM1 modes are launched, the Ez fields are concentrated below Groups 1, 2 and 3, respectively. The simulated emission patterns of the proposed OBS at different wavelengths are shown in Fig. 7(c). It is observed that as the wavelength increases from 1520 to 1620 nm, the main beam is steered from 24.5° to 11.3°, from 11.3° to −2.2° and from −1.8° to −16.7° when the TE0, TM0 and TM1 modes are launched into the waveguide, respectively. The simulated beam scanning ranges show reasonable agreements with the calculated ones. Meanwhile, for all the patterns, the SLLs are below −10 dB. Note that the three nanopatch groups in the proposed OBS are with different element spacings along y-direction. It means if the couplings of the undesired nanopatch groups are not suppressed, the emissions contributed by other groups will be pointed to a direction different from the desired one and parasitic beams appear. Hence, the low SLLs observed in Fig. 7(c) further demonstrate the effectiveness of the proposed design. By combining the beam scanning ranges of the three nanopatch groups, a total scanning range of 41.2° is obtained, almost tripling that of the waveguide grating antenna within the same 100 nm wavelength tuning range. To evaluate the radiation performance under different feeding modes, the ratios of the radiated power, the antenna loss and the waveguide loss to the total power loss are calculated and shown in Fig. 7(d). The radiation efficiencies under TE0, TM0 and TM1 mode excitations are 64%, 65% and 22% at 1550 nm, respectively. The relatively low radiation efficiency under TM1 mode excitation is mainly caused by the dissipated loss in the waveguide, as seen from the bottom panel of Fig. 7(d). One possible way to improve the radiation efficiency is to control the antenna-waveguide coupling to increase Prad by engineering the nanoslot dimensions.

 figure: Fig. 7.

Fig. 7. Results of the proposed wavelength-controlled OBS. (a) Model of the proposed OBS. (b) Ez intensity distributions on the cut-plane 10 nm below the nanopatches when the TE0, TM0 or TM1 mode is launched into the waveguide at 1550 nm. (c) Emission patterns of the proposed OBS on the yoz plane at different wavelengths showing the wavelength-controlled optical beam scanning and the expansion of the device FOV. The patterns under TM0 excitation are plotted by the solid curves with dots while those under TE0 and TM1 excitations are plotted by the solid and dashed curves without dots, respectively. The numbers in the legend denote the wavelengths of the input light in nm. (d) Ratios of the radiated power, dissipated loss in waveguide and dissipated loss in antenna to the total power loss under TE0, TM0 and TM1 excitations.

Download Full Size | PDF

A procedure of the proposed OBS design is suggested as follows.

Step 1) determine the initial dimensions of the nanopatch to make it resonate within the required bandwidth. The typical resonate length of the nanopatch is roughly λ/5 [22,43];

Step 2) optimize the nanoslot dimensions to achieve the polarization-sensitive antenna-waveguide coupling for both the x- and y-polarized elements;

Step 3) determine the feeding mode combination and the corresponding Dx1, Dx2 and Dx3. The suggested method utilizes the combination of two zeroth-order modes, i.e., the TE and TM0 modes, and one high-order mode, i.e., the TM1 mode;

Step 4) verify the design by listing the coupling coefficients of all the groups. The criteria is that for each of the launched mode, there is only one nanopatch group interacting efficiently with the launched mode. Special attention needs to be paid on the phase distribution of the current. An anti-symmetrical phase distribution will give rise to canceled light emission at the yoz plane, for example, Jx of the TE1 mode [Fig. 4(a)];

Step 5) determine Dy1, Dy2 and Dy3 based on the targeted beam scanning ranges for each group.

3.2 Optical multiplexer

The optical multiplexer is designed to excite the desired TE0, TM0 and TM1 modes in the bus waveguide within the 1520 to 1620 nm wavelength range required by the antenna section. Specifically, the PBS is used to excite the TE0 and TM0 modes in the bus waveguide while the ADC is used to excite the TM1 mode in the bus waveguide. The silicon thickness in the optical multiplexer section is the same with that of the antenna section (240 nm) and the whole structure is cladded by the SU-8 photoresist with the same thickness to the antenna section (440 nm). Here, the PBS and the ADC are designed based upon the waveguide coupler systems introduced in [4445], which feature a simple structure, a high flexibility and a broadband property [45] and thus are suitable for our design. Taking the ADC as an example, to efficiently excite the TM1 mode in the 1200 nm wide bus waveguide of the antenna section, the width of the narrow input waveguide is designed according to the phase matching conditions [45], which is 501.8 nm in our case. When the coupling length between the input waveguide and the bus waveguide is optimized, a high coupling efficiency can be obtained. The left-sided inset of Fig. 8 shows the structure and the field distributions of the designed ADC when the TM0 mode is launched at the I3 port at 1580 nm. It is seen that with a coupling length of 17 µm, the TM0 mode in the input waveguide is efficiently coupled to the desired TM1 mode in the bus waveguide with a coupling efficiency of −0.14 dB. The simulated coupling efficiencies versus wavelength is plotted in Fig. 8. Within the desired 1520 to 1620 nm wavelength range, an insertion loss below 0.99 dB is obtained. The right-sided inset of Fig. 8 shows the simulated field distributions of the PBS formed by cascading two ADCs. It can be seen when the TM0 mode is launched at the I2 port, the TM0 mode in the bus waveguide is efficiently excited with the aid of the middle wide waveguide. In contrast, when the TE0 mode is launched at the I1 port, light propagates directly through the straight waveguide with low crosstalk between the adjacent waveguide. Figure 8 gives the wavelength dependence of the transmission/coupling efficiencies of the designed PBS. Within the desired 1520 to 1620 nm wavelength range, the transmission efficiency between the TE0 mode launched at the I1 port and the TE0 mode in the bus waveguide is above −0.03 dB, and the coupling efficiency between the TM0 mode launched at the I2 port and the TM0 mode in the bus waveguide is better than −1.82 dB with a maximum of −0.19 dB.

 figure: Fig. 8.

Fig. 8. Efficiencies of the optical multiplexer. The purple/blue/green solid curve shows the simulated transmission/coupling/coupling efficiency between the TE0/TM0/TM0 mode launched from the I1/I2/I3 port in Fig. 1(a) and the TE0/TM0/TM1 mode in the bus waveguide, respectively. The left-sided inset shows the structure of the ADC and the Hx distributions on the cut plane 120 nm above the silica substrate at 1580 nm. The right-sided inset shows the structure of the PBS and the Ex/Hx distributions on the cut plane 120 nm above the silica substrate when the TE0/TM0 mode is launched from I1/I2 port at 1580 nm, respectively. The parameters are as follows: W1 = W3 = W5 = 0.5018, G1= G2 = G3 = 0.3, W2 = W4 = 1.2, Lc1 = 17, Lc2 = Lc3 = 18 and L1 = L2 = 5, all in µm.

Download Full Size | PDF

3.3 Transition section

The coupling efficiency of the transition between the bus waveguide and the hybrid plasmonic waveguide (HPW) in the antenna section is studied, as shown in the inset of Fig. 9(a). Figures 9(a) and 9(b) investigate the dependences of the coupling efficiencies and the reflections of the transition upon H1, respectively. For all the three modes, the coupling efficiencies increase with H1 while the reflections decrease with H1. This is not counter-intuitive because the perturbations introduced by the silver thin film increase with the decrease of H1. In the design of the antenna section in Fig. 7, H1 is chosen as 200 nm in order to reach the compromise between the strength of the antenna-waveguide interaction and the waveguide loss. The coupling efficiencies and reflections of the transition within the required 1520-1620 nm wavelength range when H1 = 200 nm are shown in Figs. 9(c) and 9(d). Within the 100 nm wavelength range, the coupling efficiencies for all the three modes are above −1 dB and the reflections are all below −25 dB.

 figure: Fig. 9.

Fig. 9. Efficiencies of the transition section. (a) Coupling efficiency versus H1 at 1550 nm. (b) Reflection versus H1 at 1550 nm. (c) Coupling efficiency versus wavelength with H1 = 200 nm. (d) Reflection versus wavelength with H1 = 200 nm. The parameters of the waveguide are the same as those in Fig. 2.

Download Full Size | PDF

3.4 Device overall efficiency

The loss of the system is mainly caused by the hybrid plasmonic waveguide. Figure 10(a) shows how the waveguide loss varies with H1. Remarkable plasmonic loss is observed for H1 below 50 nm. The plasmonic loss will result in reduced efficiencies of the antenna and the transition sections, whose structures are involved with metal. Figure 10(b) shows how the radiation efficiency of the antenna section in Fig. 7 varies with H1. When H1 is below 50 nm, the high waveguide losses give rises to low radiation efficiencies for all the three modes. The efficiencies of the transition section are depicted in Fig. 9(a). A transition loss over 1.72 dB is observed for the TM0 and TM1 modes when H1= 100 nm.

 figure: Fig. 10.

Fig. 10. (a) Waveguide loss versus H1. (b) Antenna section radiation efficiency versus H1. (c) Device overall efficiency versus H1. The legends denote the input ports and the corresponding excited modes in the bus waveguide. (d) Element coupling coefficients of the dominant group under different excitation modes. The parameters are the same as those in Fig. 7. Each group has 8 elements along the y-direction. The wavelength is 1550 nm.

Download Full Size | PDF

Figure 10(c) depicts the device overall efficiency versus H1 when light is launched from I1, I2 and I3 ports of the device. The corresponding excited modes in the bus waveguide are TE0, TM0 and TM1 modes, respectively. The device overall efficiency is the product of the efficiencies of the three sections, i.e., the optical multiplexer, the antenna and the transition sections. The transmission/coupling efficiencies for launching the TE0, TM­0 and TM1 modes in the bus waveguide are 0.99, 0.90 and 0.92 at 1550 nm, respectively. By comparing Figs. 10(b) and 10(c), it is found that the high transition loss results in a drop of the device overall efficiency at H1= 100 nm, especially for the TM­0 mode. Figure 10(d) shows how the element coupling coefficient varies with H1. The element coupling coefficients monotonically decrease as H1 increases. When H1 exceeds 300 nm, the element coupling coefficient under TE­0 excitation is below −40 dB, making the antenna-waveguide interaction inefficient. Therefore, H1 is chosen as 200 nm in our design to provide an efficient antenna-waveguide coupling and minimize the impact of metal loss. For small H1, the high waveguide/transition loss will give rise to a reduced device overall efficiency while for large H1, the antenna-waveguide coupling becomes inefficient.

4. Fabrication tolerance

Figure 11(a) depicts a schematic showing part of the antenna aperture when misalignments between the nanoslot and nanopatch layers occur in the OBS in Fig. 7. In Fig. 11(a), Mx and My denote the misalignments along the x- and y-directions, respectively. The element number along the -direction is 8 for all the three groups for ease of simulation. Figures 11(b) and 11(c) show the emission patterns for different feeding modes when up to 100 nm misalignment occur along the y- and x-direction, respectively. Figures 11(d) and 11(e) show how the peak directivity and radiation efficiency vary with My and M, respectively. It is observed that both the emission patterns and the radiation efficiencies do not exhibit significant variations within the 100 nm misalignment for all the three modes. The maximum directivity drops are 0.62 and 0.90 dB for the x- and y-direction misalignments, respectively.

 figure: Fig. 11.

Fig. 11. (a) Schematic showing the misalignment. The dashed boxes indicate the nanopatch positions without misalignment. (b) and (c) Emission patterns at the yoz plane at 1550 nm when y- (b) and x-direction (c) misalignments occur, respectively. The numbers in the legend denote the value of My in (b) and Mx in (c) in nm, respectively. (d) and (e) Peak directivity and radiation efficiency versus My (d) and Mx (e).

Download Full Size | PDF

Figure 12(a) shows the emission patterns of the OBS in Fig. 7 when ± 100 nm errors are applied to the nominal sizes of the nanopatches. Figures 12(b)–12(d) investigate the dependences of the peak directivity and radiation efficiency upon the nanopatch fabrication error for TE­0, TM0 and TM1 excitations, respectively. For sake of brevity, only the emission patterns under TE­0 excitation are shown. The main beam directions are robust against fabrication errors while drops of the peak directivities are observed. This is mainly caused by the mutual coupling between different nanopatch groups. To distinguish the effects of the mutual coupling, the directivities when only the dominant group is presented are simulated. Taking TE0 excitation as an example, the blue curve in Fig. 12(b) represents the directivity of the nanopatch array having the same parameters as those in Fig. 7 but with Groups 2 and 3 removed. As shown by comparing the directivity curves in two cases, the deviation between the two curves becomes more evident with the increase of the nanopatch sizes, because of the reduced distance between the adjacent nanopatches.

 figure: Fig. 12.

Fig. 12. (a) Emission patterns at the yoz plane at 1550 nm with different fabrication errors of the nanopatches. The numbers in the legend denote the error value in nm. (b) (c) and (d) Peak directivity and radiation efficiency versus nanopatch size error for TE0 (b), TM0 (c) and TM1 (d) excitations. (e) Peak directivity and radiation efficiency versus nanoslot size error.

Download Full Size | PDF

Figure 12(e) investigates the dependences of the peak directivity and radiation efficiency upon the nanoslot fabrication error. Because the nominal size of the nanoslot is below 100 nm, the lower limit of the fabrication error is set to be −50 nm. The maximum directivity drop is 2.47 dB, observed under TE0 excitation. It is also observed that the radiation efficiencies are increasing with the increase of the nanoslot sizes for all the three modes. This is because the increased nanoslot sizes result in enhanced element coupling coefficient and Prad increases correspondingly, as discussed in Section 2.3.

In general, the structure is robust against the misalignment error while the mutual coupling effect results in a reduced tolerance against the nanopatch fabrication errors, as shown by comparing the directivity curves of the shared-aperture array and the single group array in Figs. 12(b)–12(d). This could be improved by slightly reducing the nominal sizes of the nanopatch layer because the structure is more robust against negative errors, as observed in Figs. 12(b) and 12(c). In addition, the structure could be fabricated using modern electron beam lithography technique which provides the capability of high-precision fabrication for multi-layer metal structure, as demonstrated in [46]. Furthermore, it is observed that the main beam directions are very robust against the fabrication errors, which is crucial for the expansion of the FOV.

5. Conclusion

In this paper, the polarization-division and spatial-division multiplexed nanopatch arrays are studied to expand the FOV of the wavelength-controlled OBS. The design principles of both the polarization-division and spatial-division nanopatch arrays are investigated in detail. The PDMNA is discussed in the first place. It is found that by engineering the antenna dimensions, the couplings of the antenna element between the waveguide can be made sensitive to the polarizations of the launched waveguide modes. Based on the antenna elements with the polarization-dependence coupling to the waveguide, a PDMNA consists of two nanopatch groups is designed. The simulated field distributions clearly show the polarization-division excitations of the two multiplexed nanopatch groups. Secondly, the design principles of the SDMNA are investigated. It is demonstrated that the first- and third-order TE modes can be exploited to feed two laterally staggered x-polarized nanopatch groups. Finally, with the simultaneous implementations of the polarization-division and spatial-division nanopatch arrays, a wavelength-controlled OBS composed of three shared-aperture nanopatch groups and achieving a scanning range over 40°, is successfully demonstrated. In addition, the SLLs of all the nanopatch groups are below −10 dB. Furthermore, the presented design methodology allows independent control over the interactions between multiple nanoantenna groups and the waveguide. Thus it could be useful for various optical devices with different functions, e.g., dual-polarization OBSs with low polarization-dependent gain loss, polarization splitters with controllable beam projection directions and free-space optical (de)multiplexers.

Funding

Fundamental Research Funds for the Central Universities (ZYGX2019Z005); National Natural Science Foundation of China (61721001).

Disclosures

The authors declare no conflicts of interest

References

1. Y. Li, J. Zhu, M. Duperron, P. O’Brien, R. Schüler, S. Aasmul, M. Melis, M. Kersemans, and R. Baets, “Six-beam homodyne laser Doppler vibrometry based on silicon photonics technology,” Opt. Express 26(3), 3638–3645 (2018). [CrossRef]  

2. W. D. Sacher, X. Liu, F. D. Chen, H. Moradi-Chameh, I. F. Almog, T. Lordello, M. Chang, A. Naderian, T. M. Fowler, E. Segev, T. Xue, S. Mahallati, T. A. Valiante, L. C. Moreaux, J. K. S. Poon, and M. L. Roukes, “Beam-Steering Nanophotonic Phased-Array Neural Probes,” in Conference on Lasers and Electro-Optics (CLEO), OSA Technical Digest (Optical Society of America, 2019), paper ATh4I. 4.

3. B. Abiri, R. Fatemi, and A. Hajimiri, “A 1-D heterodyne lens-free optical phased array camera with reference phase shifting,” IEEE Photonics J. 10(5), 1–12 (2018). [CrossRef]  

4. Y. Kohno, K. Komatsu, R. Tang, Y. Ozeki, Y. Nakano, and T. Tanemura, “Ghost imaging using a large-scale silicon photonic phased array chip,” Opt. Express 27(3), 3817–3823 (2019). [CrossRef]  

5. M. Raval, A. Yaacobi, and M. R. Watts, “Integrated visible light phased array system for autostereoscopic image projection,” Opt. Lett. 43(15), 3678–3681 (2018). [CrossRef]  

6. G. Yang, W. Han, T. Xie, and H. Xie, “Electronic holographic three-dimensional display with enlarged viewing angle using non-mechanical scanning technology,” OSA Continuum 2(6), 1917–1924 (2019). [CrossRef]  

7. C. V. Poulton, A. Yaacobi, D. B. Cole, M. J. Byrd, M. Raval, D. Vermeulen, and M. R. Watts, “Coherent solid-state LIDAR with silicon photonic optical phased arrays,” Opt. Lett. 42(20), 4091–4094 (2017). [CrossRef]  

8. H. Choi, N. C. Park, and W. C. Kim, “Optical system design for light detection and ranging with ultra-wide field-of-view using liquid lenses,” Microsyst. Technol. 26(1), 121 (2020). [CrossRef]  

9. H. Yoo, N. Druml, D. Brunner, C. Schwarzl, T. Thurner, M. Hennecke, and G. Schitter, “MEMS-based lidar for autonomous driving,” e & i Elektrotechnik und Informationstechnik 135(6), 408–415 (2018). [CrossRef]  

10. J. Sun, E. Timurdogan, A. Yaacobi, E. S. Hosseini, and M. R. Watts, “Large-scale nanophotonic phased array,” Nature 493(7431), 195–199 (2013). [CrossRef]  

11. W. Xie, T. Komljenovic, J. Huang, M. Tran, M. Davenport, A. Torres, P. Pintus, and J. Bowers, “Heterogeneous silicon photonics sensing for autonomous cars,” Opt. Express 27(3), 3642–3663 (2019). [CrossRef]  

12. Y. Wang, G. Zhou, X. Zhang, K. Kwon, P. Blanche, N. Triesault, K. Yu, and M. C. Wu, “2D broadband beamsteering with large-scale MEMS optical phased array,” Optica 6(5), 557–562 (2019). [CrossRef]  

13. S. W. Chung, H. Abediasl, and H. Hashemi, “A monolithically integrated large-scale optical phased array in silicon-on-insulator CMOS,” IEEE J. Solid-State Circuits 53(1), 275–296 (2018). [CrossRef]  

14. C. V. Poulton, M. J. Byrd, P. Russo, E. Timurdogan, M. Khandaker, D. Vermeulen, and M. R. Watts, “Long-Range LiDAR and free-space data communication with high-performance optical phased arrays,” IEEE J. Sel. Top. Quantum Electron. 25(5), 1–8 (2019). [CrossRef]  

15. C. T. Phare, M. C. Shin, S. A. Miller, B. Stern, and M. Lipson, “Silicon optical phased array with high-efficiency beam formation over 180 degree field of view,” https://arxiv.org/abs/1802.04624.

16. W. Xu, L. Zhou, L. Lu, and J. Chen, “Aliasing-free optical phased array beam-steering with a plateau envelope,” Opt. Express 27(3), 3354–3368 (2019). [CrossRef]  

17. S. A. Miller, C. T. Phare, Y.-C. Chang, X. Ji, O. A. J. Gordillo, A. Mohanty, S. P. Roberts, M. C. Shin, B. Stern, M. Zadka, and M. Lipson, “512-element actively steered silicon phased array for low-power lidar,” in Conference on Lasers and Electro-Optics, (Optical Society of America, 2018), p. JTh5C.2.

18. M. Zadka, Y. C. Chang, A. Mohanty, C. T. Phare, S. P. Roberts, and M. Lipson, “On-chip platform for a phased array with minimal beam divergence and wide field-of-view,” Opt. Express 26(3), 2528–2534 (2018). [CrossRef]  

19. K. Shang, C. Qin, Y. Zhang, G. Liu, X. Xiao, S. Feng, and S. J. B. Yoo, “Uniform emission, constant wavevector silicon grating surface emitter for beam steering with ultra-sharp instantaneous field-of-view,” Opt. Express 25(17), 19655–19658 (2017). [CrossRef]  

20. K. Kondo, T. Tatebe, S. Hachuda, H. Abe, F. Koyama, and T. Baba, “Fan-beam steering device using a photonic crystal slow-light waveguide with surface diffraction grating,” Opt. Express 42(23), 4990–4993 (2017). [CrossRef]  

21. G. Liu, Q. Lu, and W. Guo, “Ultrafast speed, large angle, and high resolution optical beam steering using widely tunable lasers,” OSA Continuum 2(5), 1746–1753 (2019). [CrossRef]  

22. S. W. Qu and Z. P. Nie, “Plasmonic nanopatch array for optical integrated circuit applications,” Sci. Rep. 3(1), 3172 (2013). [CrossRef]  

23. M. A. Panahi, L. Yousefi, and M. Shahabadi, “Highly directive hybrid plasmonic leaky-wave optical antenna with controlled side-lobe level,” J. Lightwave Technol. 33(23), 4791–4798 (2015). [CrossRef]  

24. Z. Du, C. Hu, G. Cao, H. Lin, B. Jia, S. Yang, M. Chen, and H. Chen, “Integrated wavelength beam emitter on silicon for 2-dimensional optical scanning,” IEEE Photonics J. 11(6), 1–10 (2019). [CrossRef]  

25. J. Notaros, N. Li, C. V. Poulton, Z. Su, M. J. Byrd, E. S. Magden, E. Timurdogan, C. Baiocco, N. M. Fahrenkopf, and M. R. Watts, “CMOS-Compatible optical phased array powered by a monolithically-integrated erbium laser,” J. Lightwave Technol. 37(24), 5982–5987 (2019). [CrossRef]  

26. Y. Hirano, Y. Motoyama, K. Tanaka, K. Machida, T. Yamada, A. Otomo, and H. Kikuchi, “Demonstration of an optical phased array using electro-optic polymer phase shifters,” Jpn. J. Appl. Phys. 57(3S2), 03EH09 (2018). [CrossRef]  

27. D. N. Hutchison, J. Sun, J. K. Doylend, R. Kumar, J. Heck, W. Kim, C. T. Phare, A. Feshali, and H. Rong, “High-resolution aliasing-free optical beam steering,” Optica 3(8), 887–890 (2016). [CrossRef]  

28. J. K. Doylend, M. J. R. Heck, J. T. Bovington, J. D. Peters, L. A. Coldren, and J. E. Bowers, “Two-dimensional free-space beam steering with an optical phased array on silicon-on-insulator,” Opt. Express 19(22), 21595–21604 (2011). [CrossRef]  

29. J. C. Hulme, J. K. Doylend, M. J. R. Heck, J. D. Peters, M. L. Davenport, J. T. Bovington, L. A. Coldren, and J. E. Bowers, “Fully integrated hybrid silicon two-dimensional beam scanner,” Opt. Express 23(5), 5861–5874 (2015). [CrossRef]  

30. C. V. Poulton, M. J. Byrd, M. Raval, Z. Su, N. Li, E. Timurdogan, D. Coolbaugh, D. Vermeulen, and M. R. Watts, “Large-scale silicon nitride nanophotonic phased arrays at infrared and visible wavelengths,” Opt. Lett. 42(1), 21–24 (2017). [CrossRef]  

31. C. García-Meca, S. Lechago, A. Brimont, A. Griol, S. Mas, L. Sánchez, L. Bellieres, N. S. Losilla, and J. Martí, “On-chip wireless silicon photonics: from reconfigurable interconnects to lab-on-chip devices,” Light: Sci. Appl. 6(9), e17053 (2017). [CrossRef]  

32. S. Lechago, C. García-Meca, N. Sánchez-Losilla, A. Griol, and J. Martí, “High signal-to-noise ratio ultra-compact lab on-a-chip microflow cytometer enabled by silicon optical antennas,” Opt. Express 26(20), 25645–25656 (2018). [CrossRef]  

33. G. Bellanca, G. Calò, A. E. Kaplan, P. Bassi, and V. Petruzzelli, “Integrated Vivaldi plasmonic antenna for wireless on-chip optical communications,” Opt. Express 25(14), 16214–16227 (2017). [CrossRef]  

34. G. Calò, G. Bellanca, B. Alam, A. E. Kaplan, P. Bassi, and V. Petruzzelli, “Array of plasmonic Vivaldi antennas coupled to silicon waveguides for wireless networks through on-chip optical technology - WiNOT,” Opt. Express 26(23), 30267–30277 (2018). [CrossRef]  

35. Y. Yang, Q. Li, and M. Qiu, “Broadband nanophotonic wireless links and networks using on-chip integrated plasmonic antennas,” Sci. Rep. 6(1), 19490 (2016). [CrossRef]  

36. J. M. Merlo, N. T. Nesbitt, Y. M. Calm, A. H. Rose, L. D’Imperio, C. Yang, J. R. Naughton, M. J. Burns, K. Kempa, and M. J. Naughton, “Wireless communication system via nanoscale plasmonic antennas,” Sci. Rep. 6(1), 31710 (2016). [CrossRef]  

37. H. Huang, H. Li, W. Li, A. Wu, X. Chen, X. Zhu, Z. Sheng, S. Zou, X. Wang, and F. Gan, “High-efficiency vertical light emission through a compact silicon nanoantenna array,” ACS Photonics 3(3), 324–328 (2016). [CrossRef]  

38. Y. S. Zeng, S. W. Qu, C. Wang, B. J. Chen, and C. Chan, “Efficient unidirectional and broadband vertical-emitting optical coupler assisted by aperture-coupled nanopatch antenna array,” Opt. Express 27(7), 9941–9954 (2019). [CrossRef]  

39. J. Jian-Ming, The finite element method in electromagnetics (John Wiley & Sons, 2015).

40. F. Wang, A. Chakrabarty, F. Minkowski, K. Sun, and Q. H. Wei, “Polarization conversion with elliptical patch nanoantennas,” Appl. Phys. Lett. 101(2), 023101 (2012). [CrossRef]  

41. F. Minkowski, F. Wang, A. Chakrabarty, and Q. H. Wei, “Resonant cavity modes of circular plasmonic patch nanoantennas,” Appl. Phys. Lett. 104(2), 021111 (2014). [CrossRef]  

42. C. A. Balanis, Antenna theory: analysis and design (John wiley & sons, 2016), Chap. 2.

43. G. S. Unal and M. I. Aksun M, “Bridging the gap between RF and optical patch antenna analysis via the cavity model,” Sci. Rep. 5(1), 15941 (2015). [CrossRef]  

44. D. Dai, “Silicon polarization beam splitter based on an asymmetrical evanescent coupling system with three optical waveguides,” J. Lightwave Technol. 30(20), 3281–3287 (2012). [CrossRef]  

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

46. C. Pfeiffer, N. K. Emani, A. M. Shaltout, A. Boltasseva, V. M. Shalaev, and A. Grbi, “Efficient Light Bending with Isotropic Metamaterial Huygens,” Nano Lett. 14(5), 2491–2497 (2014). [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 (12)

Fig. 1.
Fig. 1. Schematic and structure of the proposed device. (a) Schematic overview of the proposed device consisting of the optical multiplexer section, the transition section and the antenna section. The inset shows part of the aperture of the antenna section. (b) Top-view of the optical multiplexing section. (c) 3-D view of the antenna section of the proposed device. (d) Side-view of the antenna section of the proposed device.
Fig. 2.
Fig. 2. Antenna elements. (a) and (b). Models of the x- and y-polarized elements. To clearly show the nanoslot, the SU-8 spacer is set to be transparent. The parameters are Wpa = 320, Lpa = 150, Lsa = 350, Wsa = 80, Wpb = 150, Lpb = 320, Lsb = 80, Wsb = 250, W = 1200, H = 240, T1 = 80, H1 = 200 and T2 = H2 = 100, all in nm. (c) and (d). Electric field distributions on the xoz plane and the induced current distributions on the bottom surface of the silver film of the hybrid plasmonic waveguide under TE0 (c) or TM0 (d) mode excitations at the wavelength of 1620 nm. The parameters of the waveguide are the same as those in (a) and (b). (e) Element coupling coefficient versus wavelength.
Fig. 3.
Fig. 3. (a) Illustration of the PDMNA layer by layer. (b) Top panel: the model of the designed PDMNA. Other parameters are the same as those in Fig. 2. Middle and bottom panels: Ez intensity distributions on a cut-plane 10 nm below the nanopatches at 1550 nm when the TE0 (the middle panel) or TM0 mode (the bottom panel) is launched into the designed PDMNA, showing the polarization-division excitations of the two nanopatch groups. To clearly show the difference between the excitation states of two groups, the Ez intensity is normalized in dB scale. (c) The emission patterns of the designed PDMNA on the yoz plane when the TE0 or TM0 mode is launched into the waveguide at 1550 nm.
Fig. 4.
Fig. 4. (a) Upper panel: Jx and Jy distributions of the TE0, TE1, TE2, TM0, and TM1 modes in the silver thin film. The black solid lines denote the geometrical profile of the silver thin film and the silicon waveguide. For each mode, the intensities are normalized by the maximum value of the current component with higher intensity so that the relative relations of the two components can be revealed. Bottom panel: Ex distributions of the TE0, TE1 and TE2 modes and Ez distributions of the TM0 and TM1 modes at the wavelength of 1620 nm. All on the xoz cross section. (b) Coupling coefficients of the x-polarized element between the TE0, TE1 and TE2 modes versus the element lateral position at 1620 nm. (c) Upper panel: normalized Jx intensity along Reference line 1 in (b) of the TE0, TE1 and TE2 modes at 1620 nm. Bottom panel: schematic showing the corresponding antenna element positions for the designed SDMNA. Other parameters are the same as those in Fig. 2.
Fig. 5.
Fig. 5. (a) Illustration of the SDMNA layer by layer. (b) Top panel: model of the designed SDMNA. Other parameters are the same as those in Fig. 2. Bottom panel: Ez intensity distributions on the cut-plane 10 nm below the nanopatches at 1550 nm when the TE0 (the bottom-left panel) or TE2 mode (the bottom-right one) is launched into the SDMNA, showing the mode-division excitation behaviors. The Ez intensity is normalized in dB scale to clearly show the difference between the excitation states of two groups. (c) Emission patterns of the designed SDMNA on the yoz plane at 1550 nm when the TE0 or TE2 mode is launched. G and D represent antenna gain and directivity, respectively. (d) neff of the TE0 and TE2 modes. (e) Element coupling coefficients of the SDMNA versus wavelength.
Fig. 6.
Fig. 6. (a) Illustration of the structure of the shared-aperture nanopatch array with three multiplexed nanopatch groups. (b) Schematics of two cases investigated. (c) Upper panel: Jx and Jy intensity distributions along Reference line 1 in the inset of Fig. 4(b) for the TE0, TM0 and TM1 modes, calculated at 1620 nm. Lower panel: schematic showing the antenna element positions in Case II. Other parameters are the same as those in Fig. 2.
Fig. 7.
Fig. 7. Results of the proposed wavelength-controlled OBS. (a) Model of the proposed OBS. (b) Ez intensity distributions on the cut-plane 10 nm below the nanopatches when the TE0, TM0 or TM1 mode is launched into the waveguide at 1550 nm. (c) Emission patterns of the proposed OBS on the yoz plane at different wavelengths showing the wavelength-controlled optical beam scanning and the expansion of the device FOV. The patterns under TM0 excitation are plotted by the solid curves with dots while those under TE0 and TM1 excitations are plotted by the solid and dashed curves without dots, respectively. The numbers in the legend denote the wavelengths of the input light in nm. (d) Ratios of the radiated power, dissipated loss in waveguide and dissipated loss in antenna to the total power loss under TE0, TM0 and TM1 excitations.
Fig. 8.
Fig. 8. Efficiencies of the optical multiplexer. The purple/blue/green solid curve shows the simulated transmission/coupling/coupling efficiency between the TE0/TM0/TM0 mode launched from the I1/I2/I3 port in Fig. 1(a) and the TE0/TM0/TM1 mode in the bus waveguide, respectively. The left-sided inset shows the structure of the ADC and the Hx distributions on the cut plane 120 nm above the silica substrate at 1580 nm. The right-sided inset shows the structure of the PBS and the Ex/Hx distributions on the cut plane 120 nm above the silica substrate when the TE0/TM0 mode is launched from I1/I2 port at 1580 nm, respectively. The parameters are as follows: W1 = W3 = W5 = 0.5018, G1= G2 = G3 = 0.3, W2 = W4 = 1.2, Lc1 = 17, Lc2 = Lc3 = 18 and L1 = L2 = 5, all in µm.
Fig. 9.
Fig. 9. Efficiencies of the transition section. (a) Coupling efficiency versus H1 at 1550 nm. (b) Reflection versus H1 at 1550 nm. (c) Coupling efficiency versus wavelength with H1 = 200 nm. (d) Reflection versus wavelength with H1 = 200 nm. The parameters of the waveguide are the same as those in Fig. 2.
Fig. 10.
Fig. 10. (a) Waveguide loss versus H1. (b) Antenna section radiation efficiency versus H1. (c) Device overall efficiency versus H1. The legends denote the input ports and the corresponding excited modes in the bus waveguide. (d) Element coupling coefficients of the dominant group under different excitation modes. The parameters are the same as those in Fig. 7. Each group has 8 elements along the y-direction. The wavelength is 1550 nm.
Fig. 11.
Fig. 11. (a) Schematic showing the misalignment. The dashed boxes indicate the nanopatch positions without misalignment. (b) and (c) Emission patterns at the yoz plane at 1550 nm when y- (b) and x-direction (c) misalignments occur, respectively. The numbers in the legend denote the value of My in (b) and Mx in (c) in nm, respectively. (d) and (e) Peak directivity and radiation efficiency versus My (d) and Mx (e).
Fig. 12.
Fig. 12. (a) Emission patterns at the yoz plane at 1550 nm with different fabrication errors of the nanopatches. The numbers in the legend denote the error value in nm. (b) (c) and (d) Peak directivity and radiation efficiency versus nanopatch size error for TE0 (b), TM0 (c) and TM1 (d) excitations. (e) Peak directivity and radiation efficiency versus nanoslot size error.

Tables (3)

Tables Icon

Table 1. Summary of the coupling coefficients of the x- and y-polarized antenna elementsa

Tables Icon

Table 2. Summary of the element coupling coefficients of the three nanopatch groups in Case I in Fig. 6(b)

Tables Icon

Table 3. Summary of the coupling coefficients of the antenna elements of the proposed OBS [Case II in Fig. 6(b)] fed by TE0, TM0 and TM1 modes

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.