Optical phased arrays (OPAs) are crucial in beam-steering applications, particularly as transmitters in light detection and ranging and free-space communication systems. In this paper, we demonstrate a on-chip OPA that emits multiple orbital angular momentum (OAM) beams in different directions, each carrying unique topological charges. By superimposing a forked |
1.IntroductionDeveloped in recent decades, the optical phased array (OPA) technology represents a significant breakthrough in photonics, providing a versatile and dynamic method of manipulating light.1–9 Similar to the phased arrays utilized in radar, an OPA is an array of closely spaced optical emitters, such as lasers or grating couplers. This technology enables precise control over the phase and amplitude of light waves, achieving the steering, shaping, and focusing of light beams without mechanical movement.10–13 OPAs find applications in various domains, including communications, light detection and ranging (lidar) systems, medical imaging, and holographic displays.14–25 Aiming for enhanced surveillance, imaging, and communication capabilities, multibeam OPA schemes have been proposed. Current implementations typically rely on two main approaches: (1) Dammann gratings,26–28 which divide the input beam into multiple diffraction levels, producing multiple beams, and (2) sparse emitter arrays,11,29–37 where techniques such as thermal tuning or liquid crystals adjust the phase of individual subarrays, enabling each to emit in a different direction. Compared with the single-beam OPA scheme, multibeam OPA schemes can effectively expand the field of view (FOV), improve the scanning speed, and increase the number of simultaneously handled targets.28 To reduce the interference among various echoes in multibeam systems, a practical approach is coding the light with orthogonal waveforms, which facilitates the identification of different beams in lidar systems.38,39 In addition, wavelength and time division multiplexing technologies can help discern the directions of the light beams.40 However, these approaches require complex digital signal processing (DSP) algorithms to mitigate cross talk among echoes. Introducing orbital angular momentum (OAM) beams into lidar transmitters can simplify or even eliminate the need for complex DSP in systems, thereby reducing the complexity of electronic chips.41 OAM beams, characterized by rotating wavefronts of ( is the azimuthal angle and is the topological charge) and spatial orthogonality among infinite channels,42–44 introduce a novel dimension of beam control into the OPAs of lidar systems.45–53 When an obstacle is illuminated by an OAM beam, a spatially varied echo carries more detailed information about the obstacle, providing higher-resolution imaging beyond the Rayleigh limit54,55 and efficient detection of rotational Doppler shifts.56 Moreover, as an OAM light beam travels through the atmosphere, its phase and intensity distribution can be affected by atmospheric turbulence. Upon reflection, the echo experiences the same phase and intensity distortions in reverse order, effectively canceling out the atmosphere disturbances imposed on the propagation path of the beam.57 Thus, as the transmitter in a lidar system, using OAM beams in the OPA can provide a robust and accurate method of enhancing the imaging resolution and maintaining signal integrity despite atmospheric interferences.58–60 The existing schemes generate multiple indistinguishable Gaussian beams or single OAM beams43–47,61–64 for ranging or communication purposes. To date, no OPA scheme exists that can simultaneously diffract OAM beams that carry different topological charges for the purpose of beam identification. To address this issue, we propose an on-chip multi-OAM-beam OPA scheme. The OPA consists of a pair of phase shifter arrays and a forked Dammann grating array, which effectively couple the in-plane guided mode to the out-of-plane OAM beams carrying different topological charges. We fabricated the proposed multibeam OAM OPA chip on a 220-nm-thick commercial silicon-on-insulator (SOI) platform. By injecting light beams from both the left and right sides of the Dammann grating, six () diffracted beams can cover a 140-deg FOV in azimuth within a wavelength range of 120 nm. These beams can also be precisely deflected using thermal phase shifters within in elevation. The experimental results show that the proposed scheme enables an ultralarge FOV of , significantly surpassing traditional two-dimensional (2D) beam-steering approaches. Our design serves as a practical integrated OPA for OAM multibeam generation and steering. Compared with other multibeam OPA schemes, this approach enables the identification of multiple beams by simply addressing multiple topological charges. By strategically selecting the diffraction orders and angles, we also achieved the largest FOV documented in this field, to the best of our knowledge. 2.ResultsFigure 1(a) shows the designed multiple-OAM OPA chip, which enables 2D multibeam forming and steering. This functionality is achieved by tuning the wavelength of light to steer beams in the dimension and by heating phase shifters to rotate beams in the dimension. The key feature of the design is the forked Dammann grating array, which is superimposed on a binary holographic fork grating. The forked Dammann grating array consists of two structures: a Dammann grating array and a waveguide–grating directional coupler (WG-DC) array. This configuration is essential for uniformly diffracting in-plane guided-mode light into three distinct OAM beams in free space. In Fig. 1(b), the distribution of the fork Dammann grating array is given by where represents the topological charge of the holograph [ is an integer number, and in Fig. 1(a)], and is the propagation constant of the silicon waveguide mode. to are the optimized structural parameters of the Dammann gratings, as shown in Fig. 2(a).Fig. 1Working principle of emitting OAM beams with different topological charges using an on-chip OPA. (a) Schematic of the proposed OAM OPA chip, detailing the essential components and interconnection structures. (b) Method of generating the proposed multiple-OAM OPA. The etching pattern of the forked grating is achieved using a spiral phase distribution () and a Dammann grating array. The layout of the multiple-OAM OPA is the intersection of the forked grating and the waveguide array. ![]() Fig. 2(a) Schematic of the designed Dammann grating. The grating pitch () is , and the lengths of the various segments within the grating are , , , , , and . (b) Emitted diffraction orders , , and of far-field patterns with the left input as a function of wavelength. The red regions indicate the blind zone. (c) Far-field patterns with left and right inputs as a function of wavelength. The dashed box indicates that six diffractions (orders through ) with bidirectional inputs cover a 180-deg FOV in the wavelength range of 1510 to 1630 nm. (d) Spacing between the waveguide and grating as a function of distance before and after PSO. (e) Schematic of the WG-DC structure. (f) Optical power distributions within the proposed WG-DC structure at 1510, 1570, and 1630 nm. The power distributions at these wavelengths resemble Gaussian distributions. ![]() For the ’th-order diffraction, the phase difference among the light diffracted by neighboring cells in a grating is multiplied times . Therefore, each ’th-order vortex diffracted by the forked Dammann grating array carries a topological charge of . As depicted in Fig. 1(a), the forked Dammann grating array design enables the uniform diffraction of bidirectional input lights into six diffraction orders (, , and ). Consequently, six OAM beams, carrying different topological charges (, , and ), effectively sweep six sub-FOVs. 2.1.1 × 3 Dammann GratingWe employed a Dammann grating and binary-phase Fourier hologram to generate an array of point sources, thereby simultaneously generating multiple beams. Because on-chip surface-emitting Dammann gratings have not been extensively discussed in the literature, we ran a parameter sweep to obtain the far-field intensities for various grating periods. The etching parameters are determined through subsequent optimization, as detailed in the Supplementary Material. Figure 2(a) shows a three-dimensional (3D) schematic of the optimized Dammann grating. We chose a grating width and period , which radiates diffraction orders through into free space. The red regions in Fig. 2(b) indicate blind zones, which remain despite the large bandwidth of 160 nm. To address the blind zones, two beams are input from both sides, causing the grating to diffract three OAM pairs of opposite orders, as illustrated in Fig. 2(c). In addition, the low efficiencies of the three diffractions at 1510 nm are attributed to the exact fulfillment of the high-order Bragg condition. 2.2.Waveguide–Grating Directional CouplerWe propose a WG-DC structure [Fig. 2(e)] comprising a Dammann grating and a strip waveguide to extend the antenna aperture beyond and to improve the mode purity of the OAM beams. Due to the exponential attenuation of power, the practical length of a 70-nm-etched grating is constrained to less than . In our design, we place a symmetrically bent waveguide adjacent to a straight grating to establish a DC structure. By precisely manipulating the gap between the grating and waveguide, the power distribution in the grating conforms to a Gaussian distribution. This approach enables the flexible enlargement of the grating aperture and concurrently boosts the mode purities of OAM (Laguerre–Gaussian mode) beams. Because the waveguide is symmetric along the axis, the spacing between the grating and the waveguide is a real, even function whose Fourier series is also a real, even function. The spacing between the grating and the waveguide is then apodized as a set of cosine series, , as shown in Fig. 2(d). Neglecting the bend loss, the coefficients are optimized using the particle swarm optimization (PSO) algorithm in MATLAB to achieve a Gaussian distribution in the power intensity (detailed in Supplementary Material S2). The power in the strip waveguide and Dammann grating can be expressed as where and are the complex amplitudes of the waveguide and grating, respectively; and are conjugated mode-coupling coefficients, which describe the interaction between the waveguide and the grating; and is the propagation constant of the guided mode, which equals that in the Dammann grating owing to the phase-matching condition. The loss factor can be obtained from the 3D finite-difference time-domain (FDTD) simulation of the Dammann grating (see Supplementary Material S3 for the fitting of the loss factor of a grating and further discussion).We define the figure of merit (FoM) to evaluate the similarity between the power distribution in the grating and the Gaussian distribution where represents the Gaussian distribution characterized by the parameters and , and XCorr represents the normalized cross-correlation of and with a unity peak. We repeated the PSO algorithm 10 times, each with 200 iterations. All iterations converged to identical structural parameters, as shown in Fig. 2(d). The power distributions of the WG-DC at 1510, 1570, and 1630 nm are shown in Fig. 2(f).3.Simulation and Experimental DemonstrationTo validate and improve the performance of the proposed OAM OPA, we simulated a quasi-square OPA with dimensions of . The simulated far-field intensities, amplitudes, and phase profiles, illustrated in Fig. 3(a), demonstrate the efficient conversion of guided modes into OAM modes with topological charges of , , and , respectively. As the wavelength increases from 1480 to 1640 nm, vortices that carry , , and topological charges sweep angles from to , to , and to , respectively, indicating that an FOV of 154 deg can be achieved by injecting light from two sides of the OPA. Furthermore, the divergence angles of the , , and beams are , 4, and 5 deg, respectively. Fig. 3Simulated and measured far-field intensities for vertical emission (). (a) Simulated far-field patterns, amplitudes, and phase profiles of three orders (, , and ) at 1565 nm with left input. (b) Measured far-field intensities and interference patterns of the emitted OAM lights. (c) Measured far-field intensity profiles of the -polarized at 1565 nm following a rotation of the polarizer (0, 45, and 90 deg). The arrows indicate the angles of the polarizer. (d) Zoomed-in and microscopic images of the packaged device, consisting of the multiple OAM OPA chip wire bonded to a silicon interposer, along with left and right optical fiber inputs. (e) Experimental setup used to characterize the far-field profiles of the multibeam OPA (laser: Keysight 81960A). PC, polarization controller; VOA, variable optical attenuator; HWP, half-wave plate; P1 and P2, polarizers; BS, beam splitter; , , and , lenses; IR-CCD, infrared charge-coupled device. Inset: FoM versus iteration and far-field profiles before and after phase alignment. ![]() We fabricated an OAM OPA chip on a commercial 220-nm-thick SOI platform. Figure 3(d) shows the zoomed-in and microscopic images of the fabricated silicon OAM OPA chip. A fork structure can be observed at the center of the grating array. The fabricated chip was packaged using edge coupling and wire bonding and connected to a 64-channel voltage source via three 40-pin FPC connectors on the printed circuit board (PCB). The host computer controlled the 64-channel voltage source using a general-purpose interface bus (GPIB) connector. The detailed device fabrication and packing processes are described in Sec. 6. Figure 3(e) shows a schematic of the experimental setup used to characterize the fabricated OAM OPA chip. A tunable laser source (Keysight 81960A) illuminated the fabricated device from 1505 to 1630 nm. The laser output was split into two branches using a 50:50 coupler. In the top branch, we used a polarization controller (PC) and variable optical attenuator (VOA) to control the polarization state and power of the light. After edge coupling, the guided mode was split into 64 copies using a six-stage MMI tree, and 56 copies were sent to the waveguide array. These lights in the waveguide array were coupled to the Dammann grating array (superposed holographic fork gratings) and emitted into free space as single-polarization OAM modes (). The other branch served as a reference Gaussian beam, whose power and polarization were adjusted using a PC and a VOA. The generated OAM modes and reference Gaussian beams with the same polarization, proper relative powers, and spot sizes were combined using a beam splitter (BS) to obtain the interferogram. We used an infrared charge-coupled device (IR-CCD, Luster) to monitor the intensity profiles of the generated OAM modes and corresponding interferograms. The phase mismatch of the input lights caused random intensity profiles of the OAM beams. We used a genetic algorithm (GA) to align the input phases of the OPA according to , which is the coefficient of determination for fitting the intensity profiles to the ideal OAM intensity in MATLAB. Here, we take the generation of at 1565 nm as an example. As shown in the inset of Fig. 3(e), the measured intensity profile exhibits a donut shape owing to the phase singularity after 100 iterations of GA optimization (see Supplementary Material S3). We recorded the intensity profiles of the beams, as depicted in Fig. 3(b). These profiles closely match the corresponding far-field patterns shown in Fig. 3(a), demonstrating the effectiveness of the GA in aligning the phase profiles. The topological charges of the generated OAM beams can be confirmed by counting the number of twists and twist directions of the interferograms. To verify the polarization state of light, we removed the reference light and inserted a polarizer before the IR-CCD [not shown in Fig. 3(e)]. Figure 3(c) presents the intensity profiles with the polarizer axis rotated to 0 (horizontal), 45, and 90 deg (vertical), where the power of the generated -polarized is gradually enhanced. A discussion of the GA used to control the voltages of the thermal phase shifters to reconstruct the OAM light is provided in the Supplementary Material. 4.Sweeping in and DirectionsWe measured the emission angles of the OAM beams to obtain the FOV along the axis of the fabricated device. Figure 4(b) presents the intensities measured at different wavelengths. The intensities of the far fields initially increase and then decrease with increasing wavelength, in contrast to the behavior of the simulated far-field shown in Fig. 4(a). This variation is due to the center wavelength of the MMI, which is . Figure 4(c) illustrates emission angles of to , to , and to for , , and beams, respectively, for a 1480- to 1640-nm input with left input. Specifically, within the wavelength range of 1505 to 1625 nm, the diffracted beams of the OAM OPA with bidirectional inputs covered an FOV of approximately to , as shown in the bottom images in Figs. 4(a) and 4(b). The FOV coverage and wavelength range in Fig. 4(c) agree well with the simulation results for the Dammann grating. Fig. 4Simulated and measured far-field intensities of emitted multiple vortices with different wavelength inputs. (a) and (b) Simulated and measured far-field emission patterns in a wavelength range from 1480 to 1640 nm with left input (up) and left and right inputs (down), respectively. (c) Angle sweeping of emitted diffraction orders , , and with bidirectional inputs. (d) Measured far fields of multiple vortices at different emission angles with respect to the axis. ![]() In the direction, the maximum deflection angle of the 0’th-order far field, quantified as , was approximately ( represents the period of the OPA in the direction). If the phase differences among the neighboring waveguides are close to 180 deg, the far field will experience attenuation and distortion significantly; therefore, we limited the deflection angle of the chip to values from to . We adjusted the tilt angle of the OPA chip by , , 0, , and with respect to the axis and repeated the fitting and optimization processes for 13 wavelengths in the range of 1505 to 1625 nm. The far-field profiles for the beams with the left and right input, illustrated in Fig. 4(d), indicate the beam deflection of multiple OAM beams. Using a tunable laser and thermal phase shifters to steer the light beams, the angular resolution is determined by the divergence angle of the OAM beam. The theoretical resolution is ; however, for convenience and time efficiency, we used a resolution in our experiments. Details of the far-field patterns are provided in the Supplementary Material. 5.DiscussionThe experimental demonstration of a wide FOV, multibeam OPA marks a significant advancement in multibeam emission technology. This system uses different topological charges for precise beam steering. The direction of OAM beams with positive and negative topological charges can be controlled by adjusting the left and right thermal phase shifter arrays, respectively. Another key feature of our approach is the use of waveguide–grating directional couplers, which substantially increase the antenna aperture of the silicon grating compared with those of the conventional structures. Theoretically, three OPA chips can provide a complete 360-deg FOV, significantly reducing the number of optical components. Our multibeam strategy accelerates angular scanning and maintains a wide FOV, enabling the use of fewer OPA chips in high-speed applications. Moreover, the proposed design is scalable; optimization of the grating size and reduction in the far-field spot diameter enhance adaptability. Including a nonredundant OPA using a Costas array can further expand the footprint of the aperture.9 This scalability, along with the simplicity of the proposed WG-DC structure, makes the wide FOV, multi-OAM OPA a versatile and promising solution for high-precision optical beam steering in broadcast systems,65–69 secure quantum communications,55,70–74 applications in environments with complex atmospheric turbulence,45,75 and detection of flying vehicle propellers.76,77 6.Appendix: MethodsDetailed descriptions of the optical simulation and optimization, characterization of the far field, and phase modulation of the thermal phase shifter are presented in the Supplementary Material. 6.1.Numerical SimulationsThe numerical results presented in Figs. 2–4 were obtained through 3D FDTD simulations using the commercial software ANSYS Lumerical. In the simulation setup, the -thick substrate layer and the -thick cladding layer were with a refractive index of 1.44, and the material of the sandwiched waveguide was silicon with a refractive index of 3.47. The simulation region implemented a perfectly matched layer as the boundary condition, and a mesh accuracy of 2 was specified. Fifty-six mode sources were set to the fundamental transverse electric () mode, and a frequency domain power monitor was positioned above the interface between the cladding and air to capture the outgoing field. 6.2.Device Fabrication and PackagingBased on the structure design, theory, and simulation results, the devices were fabricated on an SOI platform with a 220-nm-thick silicon top layer and a -thick buried oxide layer. The waveguides and the Dammann grating array were patterned via electron beam lithography (Vistec EBPG-5200+). Full and partial etching was performed using an inductively coupled plasma (ICP, SPTS) etching process. After that, a -thick cladding layer was uniformly deposited on top of the devices using plasma-enhanced chemical vapor deposition (Oxford). A 400-nm-thick Ti layer functioned as a thermal phase shifter, and a 300-nm-thick gold layer was deposited using an e-beam evaporator (Denton Electron Beam Evaporator). Each phase shifter had a resistance of . A total of 112 gold wires were bonded from the gold pads on the silicon chip to a PCB using F&S BONDTEC 53XXBDA. 6.3.Optical Performance Characterization SetupTo characterize the optical performance of the fabricated device, we used a continuous-wave tunable semiconductor laser source (Santec TSL-770) to generate light ranging from 1480 to 1640 nm. A 50:50 coupler was utilized to split the light into two beams, which were sent to two PCs. One of the beams was coupled to the fabricated chip, and the other was collimated using a collimator. The diffracted light and collimated beam were subsequently combined using a BS and captured using an IR-CCD equipped with a four-lens system. Code and Data AvailabilityData underlying the results presented in this paper may be obtained from the authors upon reasonable request. AcknowledgementsThis work was supported by the National Natural Science Foundation of China (Grant Nos. 62341508 and 62105203) and the Shanghai Municipal Science and Technology Major Project (Grant No. BH0300071). The authors thank the Center for Advanced Electronic Materials and Devices of Shanghai Jiao Tong University for its support in device fabrication. ReferencesR. Collis,
“LiDAR,”
Appl. Opt., 9
(8), 1782
–1788 https://doi.org/10.1364/AO.9.001782 APOPAI 0003-6935
(1970).
Google Scholar
U. Wandinger, M. McCormick, C. Weitkamp, LiDAR: Range-Resolved Optical Remote Sensing of the Atmosphere, 1
–18 Springer Berlin/Heidelberg(
(2005). Google Scholar
O. Wolf et al.,
“Phased-array sources based on nonlinear metamaterial nanocavities,”
Nat. Commun., 6
(1), 7667 https://doi.org/10.1038/ncomms8667 NCAOBW 2041-1723
(2015).
Google Scholar
P. Hyde et al.,
“Exploring LiDAR–radar synergy—predicting aboveground biomass in a southwestern ponderosa pine forest using LiDAR, SAR and InSAR,”
Remote Sens. Environ., 106
(1), 28
–38 https://doi.org/10.1016/j.rse.2006.07.017
(2007).
Google Scholar
T. Raj et al.,
“A survey on lidar scanning mechanisms,”
Electronics, 9
(5), 741 https://doi.org/10.3390/electronics9050741 ELECAD 0013-5070
(2020).
Google Scholar
P. Dong and Q. Chen, LiDAR Remote Sensing and Applications, CRC Press(
(2017). Google Scholar
J. K. Doylend et al.,
“Two-dimensional free-space beam steering with an optical phased array on silicon-on-insulator,”
Opt. Express, 19
(22), 21595
–21604 https://doi.org/10.1364/OE.19.021595 OPEXFF 1094-4087
(2011).
Google Scholar
J. Sun et al.,
“Large-scale nanophotonic phased array,”
Nature, 493
(7431), 195
–199 https://doi.org/10.1038/nature11727
(2013).
Google Scholar
T. Fukui et al.,
“Non-redundant optical phased array,”
Optica, 8
(10), 1350
–1358 https://doi.org/10.1364/OPTICA.437453
(2021).
Google Scholar
S. A. Miller et al.,
“Large-scale optical phased array using a low-power multi-pass silicon photonic platform,”
Optica, 7
(1), 3
–6 https://doi.org/10.1364/OPTICA.7.000003
(2020).
Google Scholar
P. Lu et al.,
“Integrated multi-beam optical phased array based on a 4 × 4 Butler matrix,”
Opt. Lett., 46
(7), 1566
–1569 https://doi.org/10.1364/OL.419828 OPLEDP 0146-9592
(2021).
Google Scholar
D. Kwong et al.,
“On-chip silicon optical phased array for two-dimensional beam steering,”
Opt. Lett., 39
(4), 941
–944 https://doi.org/10.1364/OL.39.000941 OPLEDP 0146-9592
(2014).
Google Scholar
K. Van Acoleyen et al.,
“Off-chip beam steering with a one-dimensional optical phased array on silicon-on-insulator,”
Opt. Lett., 34
(9), 1477
–1479 https://doi.org/10.1364/OL.34.001477 OPLEDP 0146-9592
(2009).
Google Scholar
D. J. Lum,
“Ultrafast time-of-flight 3D LiDAR,”
Nat. Photonics, 14
(1), 2
–4 https://doi.org/10.1038/s41566-019-0568-2 NPAHBY 1749-4885
(2020).
Google Scholar
Y. Jiang, S. Karpf and B. Jalali,
“Time-stretch LiDAR as a spectrally scanned time-of-flight ranging camera,”
Nat. Photonics, 14
(1), 14
–18 https://doi.org/10.1038/s41566-019-0548-6 NPAHBY 1749-4885
(2020).
Google Scholar
T. Inomata et al.,
“Origins and spread of formal ceremonial complexes in the Olmec and Maya regions revealed by airborne LiDAR,”
Nat. Hum. Behav., 5
(11), 1487
–1501 https://doi.org/10.1038/s41562-021-01218-1
(2021).
Google Scholar
M. Brydegaard et al.,
“LiDAR reveals activity anomaly of malaria vectors during pan-African eclipse,”
Sci. Adv., 6
(20), eaay5487 https://doi.org/10.1126/sciadv.aay5487 STAMCV 1468-6996
(2020).
Google Scholar
J. Tachella et al.,
“Real-time 3D reconstruction from single-photon LiDAR data using plug-and-play point cloud denoisers,”
Nat. Commun., 10
(1), 4984 https://doi.org/10.1038/s41467-019-12943-7 NCAOBW 2041-1723
(2019).
Google Scholar
J. Park et al.,
“All-solid-state spatial light modulator with independent phase and amplitude control for three-dimensional LiDAR applications,”
Nat. Nanotechnol., 16
(1), 69
–76 https://doi.org/10.1038/s41565-020-00787-y NNAABX 1748-3387
(2021).
Google Scholar
R. Tobin et al.,
“Robust real-time 3D imaging of moving scenes through atmospheric obscurant using single-photon LiDAR,”
Sci. Rep., 11
(1), 11236 https://doi.org/10.1038/s41598-021-90587-8 SRCEC3 2045-2322
(2021).
Google Scholar
C.-P. Hsu et al.,
“A review and perspective on optical phased array for automotive LiDAR,”
IEEE J. Sel. Top. Quantum Electron., 27
(1), 1
–16 https://doi.org/10.1109/JSTQE.2020.3022948 IJSQEN 1077-260X
(2020).
Google Scholar
C. V. Poulton et al.,
“Long-range LiDAR and free-space data communication with high-performance optical phased arrays,”
IEEE J. Sel. Top. Quantum Electron., 25
(5), 1
–8 https://doi.org/10.1109/JSTQE.2019.2908555 IJSQEN 1077-260X
(2019).
Google Scholar
W. S. Rabinovich et al.,
“Free space optical communication link using a silicon photonic optical phased array,”
Proc. SPIE, 9354 93540B https://doi.org/10.1117/12.2077222 PSISDG 0277-786X
(2015).
Google Scholar
S. Royo and M. Ballesta-Garcia,
“An overview of LiDAR imaging systems for autonomous vehicles,”
Appl. Sci., 9
(19), 4093 https://doi.org/10.3390/app9194093
(2019).
Google Scholar
J. Choi et al.,
“Multi-target tracking using a 3D-LiDAR sensor for autonomous vehicles,”
in 16th Int. IEEE Conf. Intell. Transp. Syst. (ITSC 2013),
881
–886
(2013). https://doi.org/10.1109/ITSC.2013.6728343 Google Scholar
Y. Guo et al.,
“Integrated optical phased arrays for beam forming and steering,”
Appl. Sci., 11
(9), 4017 https://doi.org/10.3390/app11094017
(2021).
Google Scholar
Y. Liu et al.,
“A single-chip multi-beam steering optical phased array: design rules and simulations,”
Opt. Express, 29
(5), 7049
–7059 https://doi.org/10.1364/OE.417821 OPEXFF 1094-4087
(2021).
Google Scholar
Y. Liu et al.,
“On-chip multi-beam emitting optical phased array for wide-angle LiDAR,”
in CLEO: Appl. and Technol.,
JTu2G.25
(2020). https://doi.org/10.1364/CLEO_AT.2020.JTu2G.25 Google Scholar
C. Liu et al.,
“High-resolution 3D imaging with a low-pixel APD array through multi-beam scanning of fiber-type OPA,”
in Sixteenth Natl. Conf. Laser Technol. and Optoelectron.,
750
–754
(2021). https://doi.org/10.1117/12.2603243 Google Scholar
Y. Wu et al.,
“Multi-beam optical phase array for long-range LiDAR and free-space data communication,”
Opt. Laser Technol., 151 108027 https://doi.org/10.1016/j.optlastec.2022.108027 OLTCAS 0030-3992
(2022).
Google Scholar
G. He et al.,
“A review of multibeam phased array antennas as LEO satellite constellation ground station,”
IEEE Access, 9 147142
–147154 https://doi.org/10.1109/ACCESS.2021.3124318
(2021).
Google Scholar
A. Jacomb-Hood and E. Lier,
“Multibeam active phased arrays for communications satellites,”
IEEE Microwave Mag., 1
(4), 40
–47 https://doi.org/10.1109/6668.893245 1527-3342
(2000).
Google Scholar
G. Toso, P. Angeletti and C. Mangenot,
“Multibeam antennas based on phased arrays: an overview on recent ESA developments,”
in 8th Eur. Conf. Antennas and Propag. (EuCAP 2014),
178
–181
(2014). https://doi.org/10.1109/EuCAP.2014.6901721 Google Scholar
S. Prasad et al.,
“mmWave multibeam phased array antenna for 5G applications,”
J. Electromagn. Wave, 35
(13), 1802
–1814 https://doi.org/10.1080/09205071.2021.1917004 JEWAE5 0920-5071
(2021).
Google Scholar
Z. Zhou et al.,
“Butler matrix enabled multi-beam optical phased array for two-dimensional beam-steering,”
in 27th OptoElectron. and Commun. Conf. (OECC) and Int. Conf. Photonics in Switch. and Comput. (PSC),
1
–3
(2022). https://doi.org/10.23919/OECC/PSC53152.2022.9850221 Google Scholar
Y. Wu, S. Shao and D. Che,
“Fast and low grating lobe multi-beam steering with a subarray level unequally spaced optical phased array,”
JOSA B, 38
(11), 3417
–3424 https://doi.org/10.1364/JOSAB.434347
(2021).
Google Scholar
K. Yoneda et al.,
“Vehicle localization using 76 GHz omnidirectional millimeter-wave radar for winter automated driving,”
in IEEE Intell. Veh. Symp. (IV),
971
–977
(2018). https://doi.org/10.1109/IVS.2018.8500378 Google Scholar
J. De Wit, W. Van Rossum and A. De Jong,
“Orthogonal waveforms for FMCW MIMO radar,”
in IEEE RadarCon (RADAR),
(2011). https://doi.org/10.1109/RADAR.2011.5960625 Google Scholar
G. Chang et al.,
“Orthogonal waveform with multiple diversities for MIMO radar,”
IEEE Sens. J., 18
(11), 4462
–4476 https://doi.org/10.1109/JSEN.2018.2824824 ISJEAZ 1530-437X
(2018).
Google Scholar
D. Wu et al.,
“Multi-beam single-photon LiDAR with hybrid multiplexing in wavelength and time,”
Opt. Laser Technol., 145 107477 https://doi.org/10.1016/j.optlastec.2021.107477 OLTCAS 0030-3992
(2022).
Google Scholar
C. Weimer et al.,
“LiDARs utilizing vortex laser beams,”
Proc. SPIE, 10631 106310Q https://doi.org/10.1117/12.2310042 PSISDG 0277-786X
(2018).
Google Scholar
N. Bozinovic et al.,
“Terabit-scale orbital angular momentum mode division multiplexing in fibers,”
Science, 340
(6140), 1545
–1548 https://doi.org/10.1126/science.1237861 SCIEAS 0036-8075
(2013).
Google Scholar
J. Wang et al., Optical Fiber Telecommunications VIB: Chapter 12. Multimode Communications Using Orbital Angular Momentum,
(2013). Google Scholar
J. Wang et al.,
“Terabit free-space data transmission employing orbital angular momentum multiplexing,”
Nat. Photonics, 6
(7), 488
–496 https://doi.org/10.1038/nphoton.2012.138 NPAHBY 1749-4885
(2012).
Google Scholar
Y. Shen et al.,
“Optical vortices 30 years on: OAM manipulation from topological charge to multiple singularities,”
Light: Sci. Appl., 8
(1), 90 https://doi.org/10.1038/s41377-019-0194-2
(2019).
Google Scholar
S. Zhang et al.,
“Broadband detection of multiple spin and orbital angular momenta via dielectric metasurface,”
Laser Photonics Rev., 14
(9), 2000062 https://doi.org/10.1002/lpor.202000062
(2020).
Google Scholar
D. R. Gozzard et al.,
“Optical vortex beams with controllable orbital angular momentum using an optical phased array,”
OSA Contin., 3
(12), 3399
–3406 https://doi.org/10.1364/OSAC.412607
(2020).
Google Scholar
Y. Weng and Z. Pan,
“Orbital angular momentum based sensing and their applications: a review,”
J. Lightwave Technol., 41
(7), 2007
–2016 https://doi.org/10.1109/JLT.2022.3202184 JLTEDG 0733-8724
(2022).
Google Scholar
B. Cochenour et al.,
“The detection of objects in a turbid underwater medium using orbital angular momentum (OAM),”
Proc. SPIE, 10186 1018603 https://doi.org/10.1117/12.2264626 PSISDG 0277-786X
(2017).
Google Scholar
N. Dostart et al.,
“Serpentine optical phased arrays for scalable integrated photonic LiDAR beam steering,”
Optica, 7
(6), 726
–733 https://doi.org/10.1364/OPTICA.389006
(2020).
Google Scholar
K. Sayyah et al.,
“Two-dimensional pseudo-random optical phased array based on tandem optical injection locking of vertical cavity surface emitting lasers,”
Opt. Express, 23
(15), 19405
–19416 https://doi.org/10.1364/OE.23.019405 OPEXFF 1094-4087
(2015).
Google Scholar
S. 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 https://doi.org/10.1109/JSSC.2017.2757009 IJSCBC 0018-9200
(2017).
Google Scholar
K. Van Acoleyen, H. Rogier and R. Baets,
“Two-dimensional optical phased array antenna on silicon-on-insulator,”
Opt. Express, 18
(13), 13655
–13660 https://doi.org/10.1364/OE.18.013655 OPEXFF 1094-4087
(2010).
Google Scholar
O. A. Saraereh,
“Design and performance evaluation of OAM-antennas: a comparative review,”
IEEE Access, 11 27992
–28013 https://doi.org/10.1109/ACCESS.2023.3259732
(2023).
Google Scholar
R. Chen et al.,
“Orbital angular momentum waves: generation, detection, and emerging applications,”
IEEE Commun. Surv. Tutor., 22
(2), 840
–868 https://doi.org/10.1109/COMST.2019.2952453
(2019).
Google Scholar
K. Liu et al.,
“Microwave-sensing technology using orbital angular momentum: overview of its advantages,”
IEEE Veh. Technol. Mag., 14
(2), 112
–118 https://doi.org/10.1109/MVT.2018.2890673
(2019).
Google Scholar
K. Liu et al.,
“Orbital-angular-momentum-based electromagnetic vortex imaging,”
IEEE Antennas Wireless Propag. Lett., 14 711
–714 https://doi.org/10.1109/LAWP.2014.2376970 IAWPA7 1536-1225
(2014).
Google Scholar
K. Liu et al.,
“Generation of OAM beams using phased array in the microwave band,”
IEEE Trans. Antennas Propag., 64
(9), 3850
–3857 https://doi.org/10.1109/TAP.2016.2589960 IETPAK 0018-926X
(2016).
Google Scholar
J. Long et al.,
“High-power mode-programmable orbital angular momentum beam emitter with an internally sensed optical phased array,”
Chin. Opt. Lett., 22
(2), 021402 https://doi.org/10.3788/COL202422.021402 CJOEE3 1671-7694
(2024).
Google Scholar
K. Yang et al.,
“Modulating and identifying an arbitrary curvilinear phased optical vortex array of high-order orbital angular momentum,”
Opt. Laser Technol., 168 109984 https://doi.org/10.1016/j.optlastec.2023.109984 OLTCAS 0030-3992
(2024).
Google Scholar
Y. Chen et al.,
“Integrated phased array for scalable vortex beam multiplexing,”
J. Lightwave Technol., 41
(7), 2070
–2078 https://doi.org/10.1109/JLT.2022.3217976 JLTEDG 0733-8724
(2022).
Google Scholar
R. M. Fouda et al.,
“Experimental BER performance of quasi-circular array antenna for OAM communications,”
IEEE Antennas Wireless Propag. Lett., 19
(8), 1350
–1354 https://doi.org/10.1109/LAWP.2020.3000925 IAWPA7 1536-1225
(2020).
Google Scholar
P.-Y. Feng, S.-W. Qu and S. Yang,
“OAM-generating transmitarray antenna with circular phased array antenna feed,”
IEEE Trans. Antennas Propag., 68
(6), 4540
–4548 https://doi.org/10.1109/TAP.2020.2972393 IETPAK 0018-926X
(2020).
Google Scholar
N. Zhou et al.,
“Ultra-compact broadband polarization diversity orbital angular momentum generator with footprint,”
Sci. Adv., 5
(5), eaau9593 https://doi.org/10.1126/sciadv.aau9593 STAMCV 1468-6996
(2019).
Google Scholar
P. Hall and S. Vetterlein,
“Review of radio frequency beamforming techniques for scanned and multiple beam antennas,”
in IEE Proc. H (Microwaves, Antennas and Propag.),
293
–303
(1990). Google Scholar
Y. Aslan et al.,
“Orthogonal versus zero-forced beamforming in multibeam antenna systems: review and challenges for future wireless networks,”
IEEE Journal of Microwaves, 1
(4), 879
–901 https://doi.org/10.1109/JMW.2021.3109244
(2021).
Google Scholar
H. Wolf et al.,
“Satellite multibeam antennas at airbus defence and space: state of the art and trends,”
in 8th Eur. Conf. Antennas and Propag. (EuCAP 2014),
182
–185
(2014). Google Scholar
D. Striccoli, G. Piro and G. Boggia,
“Multicast and broadcast services over mobile networks: a survey on standardized approaches and scientific outcomes,”
IEEE Commun. Surv. Tutor., 21
(2), 1020
–1063 https://doi.org/10.1109/COMST.2018.2880591
(2018).
Google Scholar
N. Chukhno et al.,
“Models, methods, and solutions for multicasting in 5G/6G mmwave and sub-THz systems,”
IEEE Commun. Surv. Tutor., 26
(1), 119
–159 https://doi.org/10.1109/COMST.2023.3319354
(2023).
Google Scholar
I. B. Djordjevic,
“Multidimensional OAM-based secure high-speed wireless communications,”
IEEE Access, 5 16416
–16428 https://doi.org/10.1109/ACCESS.2017.2735994
(2017).
Google Scholar
M. Zhang et al.,
“Physical layer key generation for secure OAM communication systems,”
IEEE Trans. Veh. Technol., 71
(11), 12397
–12401 https://doi.org/10.1109/TVT.2022.3194660
(2022).
Google Scholar
M. A. B. Abbasi et al.,
“Physical layer secure communication using orbital angular momentum transmitter and a single-antenna receiver,”
IEEE Trans. Antennas Propag., 68
(7), 5583
–5591 https://doi.org/10.1109/TAP.2020.2981732 IETPAK 0018-926X
(2020).
Google Scholar
J. Wang et al.,
“Orbital angular momentum and beyond in free-space optical communications,”
Nanophotonics, 11
(4), 645
–680 https://doi.org/10.1515/nanoph-2021-0527
(2022).
Google Scholar
A. E. Willner et al.,
“Optical communications using orbital angular momentum beams,”
Adv. Opt. Photonics, 7
(1), 66
–106 https://doi.org/10.1364/AOP.7.000066 AOPAC7 1943-8206
(2015).
Google Scholar
S. Li et al.,
“Atmospheric turbulence compensation in orbital angular momentum communications: advances and perspectives,”
Opt. Commun., 408 68
–81 https://doi.org/10.1016/j.optcom.2017.09.034 OPCOB8 0030-4018
(2018).
Google Scholar
L. Marrucci,
“Spinning the Doppler effect,”
Science, 341
(6145), 464
–465 https://doi.org/10.1126/science.1242097 SCIEAS 0036-8075
(2013).
Google Scholar
M. P. Lavery et al.,
“Detection of a spinning object using light’s orbital angular momentum,”
Science, 341
(6145), 537
–540 https://doi.org/10.1126/science.1239936 SCIEAS 0036-8075
(2013).
Google Scholar
BiographyZhen Wang received his PhD from the Department of Electronic Engineering, Shanghai Jiao Tong University, Shanghai, China, in 2024. He is currently a postdoctoral fellow at the Shanghai Institute of Optics and Fine Mechanics, Chinese Academy of Sciences, Shanghai, China. His current research interests include silicon photonics, optical phased arrays, and mode multiplexing technologies. Yikai Su received his PhD in electrical engineering from Northwestern University, Evanston, Illinois, United States, in 2001. He served at Crawford Hill Laboratory of Bell Laboratories and joined Shanghai Jiao Tong University, Shanghai, China, as a full professor in 2004. His research areas cover silicon photonic devices, optical transmission, and signal processing. He has published over 600 publications and has given >70 invited talks at conferences including OFC, CLEO, and IPS. He is a senior member of IEEE and a fellow of Optica. |