Next Article in Journal
One-Step Synthesis of Ultra-Small RhNPs in the Microreactor System and Their Deposition on ACF for Catalytic Conversion of 4–Nitrophenol to 4–Aminophenol
Previous Article in Journal
Investigation of Viscoelastic Properties of Macrophage Membrane–Cytoskeleton Induced by Gold Nanorods in Leishmania Infection
Previous Article in Special Issue
Metasurface-Enhanced Infrared Photodetection Using Layered van der Waals MoSe2
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

On-Chip OPA: Progress and Prospects in Liquid Crystal, Lithium Niobate, and Silicon Material Platforms

1
College of Advanced Interdisciplinary Studies, National University of Defense Technology, Changsha 410073, China
2
Key Laboratory of Semiconductor Display Materials and Chips & i-Lab, Suzhou Institute of Nano-Tech and Nano-Bionics, Chinese Academy of Sciences, Suzhou 215123, China
3
Nanhu Laser Laboratory, National University of Defense Technology, Changsha 410073, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2025, 15(17), 1374; https://doi.org/10.3390/nano15171374
Submission received: 30 June 2025 / Revised: 26 August 2025 / Accepted: 2 September 2025 / Published: 5 September 2025

Abstract

Non-mechanical beam steering is required for holographic displays, free-space optical communication, and chip-scale LiDAR. Optical phased arrays (OPAs), which allow for inertia-free, high-speed beam control via electronic phase control, are an important research topic. The present study investigates the primary material platform for on-chip OPAs: Liquid crystal OPAs (LC-OPAs) employ electrically tunable refractive indices for low-voltage operation; lithium niobate OPAs (LN-OPAs) utilize high electro-optic coefficients for high-speed, low-power consumption, and large-bandwidth operation; and silicon-based OPAs (Si-OPAs) apply mature photonic integration to achieve high integration density and GHz-range steering. The paper thoroughly examines OPA basics, recent material-specific advancements, performance benchmarks, outstanding issues, and future prospects.

1. Introduction

Beam steering agility is essential for detection and communication systems, especially for light detection and ranging (LiDAR) [1,2,3,4,5,6,7] and free-space optical communication (FSO) [8,9,10,11]. Phased array technology [12] satisfies these requirements through coherent interference of waves from individual radiating elements, enabling inertia-free beam steering [13,14], dynamic beamforming [15,16], and multi-beam generation [17,18,19] by electronically, precisely, and independently modulating the phase and amplitude of each element, all without relying on mechanical beam control mechanisms. Compared to mechanical systems, phased array technology provides improved multi-target tracking [20,21,22,23], anti-jamming robustness, and compact designs [24,25]. Advances in semiconductor integration and signal processing have extended their application in satellite communications, radio astronomy, and 5G/6G networks [26,27,28,29,30,31], eventually establishing itself as an important core technology underpinning modern electronic systems.
Tunable phase shifters are responsible for the main functioning of a phased array. In microwave and millimeter wave systems, accurate electronic phase modulation is accomplished using materials with customized electromagnetic properties. Ferrite phase shifters, with magnetically tunable permeability, provide direct control over phase delay in waveguide or microstrip transmission lines, acting as the basic building blocks for typical phased array architectures [32,33,34]. Semiconductor phase shifters use their inherent electrical tunability to enable dynamic phase control. These solid-state devices modulate junction capacitances using applied bias voltages or toggling between distinct transmission route lengths, allowing for exact phase shift calibration through entirely electronic means [35,36]. Ongoing research has investigated liquid crystal phase shifters as a means to enhance integration density and performance. These devices use field-induced changes in dielectric anisotropy, where molecular reorientation under applied electric fields alters the effective permittivity, allowing for unique phase-shifting processes with high accuracy [37,38,39]. The basic similarity between these various material techniques is their ability to directly and efficiently convert externally applied electrical control signals (voltage or current) into precise and dynamic phase modulation of microwave or millimeter signals. This method of direct phase modulation via electrical control signals is well established in microwave and millimeter frequency bands. To attain higher spatial resolution and detection performance, beam steering systems must use shorter wavelengths and highly directional light sources. Lasers, with their remarkable coherence and collimation capabilities, emerge as the most effective alternative to achieve these stringent requirements.
The optical phased array (OPA) [40,41,42,43] operates on principles analogous to radio-frequency phased arrays. Its core functionality relies critically on a material’s capacity for rapid, reversible, and low-loss optical modulation via external stimuli. Key modulation mechanisms include the electro-optic effect (e.g., in Lithium Niobate, LN) [13,14], the thermo-optic effect (e.g., in Silicon, Si) [44,45], the plasma dispersion effect (e.g., in Indium Phosphide, InP) [46,47], and liquid crystal (LC) birefringence tuning [48,49]. Material-dependent properties, such as electro-optic coefficients, thermo-optic coefficients, and response times, fundamentally dictate OPA performance metrics like beam steering range, ranging capability, and ranging resolution [50]. Consequently, as illustrated in Figure 1, distinct material platforms have naturally found specialized application domains based on these inherent properties.
LC-OPAs exploit tunable birefringence, achieving broadband coverage from visible to infrared wavelengths through scalable element arrays with high resolution. These attributes make them particularly suitable for wearable displays, holographic imaging, light field control, satellite communications, and optical communication systems. LN-OPAs, leveraging high electro-optic coefficients, power handling capabilities, high modulation bandwidth, and high power tolerance, operate primarily in the near-infrared (NIR) spectrum. These characteristics establish them as the preferred technology for demanding applications such as high-power lasers and high-performance 3D LiDAR systems. Si-OPAs capitalize on CMOS compatibility, enabling highly integrated architectures and material versatility. This allows precise engineering of OPA wavelengths across the visible, NIR, and mid-infrared (MIR) spectral ranges. Consequently, Si-OPAs serve as the workhorse for solid-state LiDAR (enabling tiny, low-power systems), free-space optical communication, and light field control.
This paper presents a systematic review of tunable materials for on-chip OPAs. To thoroughly explain the current research environment and the developmental trajectory within this topic, the structure of this review is organized as follows: Starting with fundamental principles, it digs into the main working processes of OPAs, emphasizing crucial performance factors including beam steering, angular resolution (or corresponding beam width), and beam steering range. Second, we systematically categorize and compare the major material systems, such as LC, LN, and Si materials, examining their distinct physical properties and corresponding application scenarios when used in OPA construction, as well as analyzing each material’s inherent advantages and limitations. Finally, based on a study of the existing technological state-of-the-art, this paper analyzes the future problems confronting the OPA area, as well as potential development directions and application opportunities.

2. Principle and Characteristics

The far-field radiation pattern of an OPA can be derived from coherent superposition of emitter contributions. For an N × N rectangular array with uniform spacing (Figure 2a), the diffraction plane ( x , y ) and observation plane ( ξ , η ) are separated by distance L. The center-to-center spacing between array elements is d. Assuming each element emits a Gaussian beam with waist diameter w 0 , the position of the ( m , n ) -th element is ( x m , y n ) for m , n = 1 , 2 , , N . The amplitude of the beam is a m , n , and its electric field distribution on the diffraction plane ( x , y ) is expressed as [51,52]:
E m , n ( x , y ) = a m , n exp ( x x m ) 2 + ( y y n ) 2 w 0 2
The total electric field distribution in the diffraction plane ( x , y ) for the N × N OPA is:
E 0 ( x , y ) = m = 1 N n = 1 N a m , n exp ( x x m ) 2 + ( y y n ) 2 w 0 2 exp j ϕ m , n
where ϕ m , n = ϕ m , n in + ϕ m , n s is the phase of the ( m , n ) -th element. Here, ϕ m , n in is the initial phase and ϕ m , n s is the additional phase required for beam steering to angles ( θ x s , θ y s ) :
ϕ m , n s = ( m 1 ) k d sin θ x s + ( n 1 ) k d sin θ y s
where k = 2 π / λ and λ is the laser wavelength in vacuum.
Based on Fraunhofer diffraction theory [53], the electric field distribution on the observation plane ( ξ , η ) located at distance L from the diffraction plane is given by:
E ( ξ , η ) = exp ( j k L ) j λ L + + E 0 ( x , y ) exp j k x ξ + y η L d x d y = exp ( j k L ) j λ L π w 0 2 exp k 2 w 0 2 ξ 2 + η 2 2 L 2 × m = 1 N n = 1 N a m , n exp j ϕ m , n exp j k L ( x m ξ + y n η )
The intensity distribution on the observation plane ( ξ , η ) is given by:
I ( ξ , η ) = | E ( ξ , η ) | 2 = π 2 w 0 4 ( λ L ) 2 exp k 2 w 0 2 ξ 2 + η 2 L 2 × m = 1 N n = 1 N a m , n exp j ϕ m , n exp j k L ( x ξ + y η ) 2 = I EN ( ξ , η ) × I SP ( ξ , η )
where * denotes the complex conjugate operator. Here, I EN ( ξ , η ) is the diffraction factor of a single beam, which determines the envelope of the far-field intensity distribution, and I SP ( ξ , η ) is the interference factor, which determines the coherence characteristics of the far-field intensity distribution:
I EN ( ξ , η ) = π 2 w 0 4 ( λ L ) 2 exp k 2 w 0 2 ξ 2 + η 2 L 2
I SP ( ξ , η ) = m = 1 N n = 1 N a m , n exp j ϕ m , n exp j k L ( x ξ + y η ) 2
When the distance L between the diffraction plane ( x , y ) and observation plane ( ξ , η ) is sufficiently large, the diffraction angles can be approximated as θ x ξ / L and θ y η / L . Equation (5) can be rewritten as:
I ( θ x , θ y ) = π 2 w 0 4 ( λ L ) 2 exp k 2 w 0 2 2 ( θ x 2 + θ y 2 ) ×   m = 1 N n = 1 N a m , n exp j ϕ m , n exp j k ( x θ x + y θ y ) 2
Under ideal conditions, neglecting phase noise introduced by the system or environment ( ϕ m , n in = 0 ), and assuming the polarization states remain unchanged during propagation, with all beam amplitudes a m , n = 1 , Equation (8) simplifies to:
I ( ξ , η ) = π 2 w 0 4 ( λ L ) 2 sin 1 2 N k d ( sin θ x sin θ x s ) sin 1 2 k d ( sin θ x sin θ x s ) 2 × sin 1 2 N k d ( sin θ y sin θ y s ) sin 1 2 k d ( sin θ y sin θ y s ) 2 × exp k 2 w 0 2 2 ( θ x 2 + θ y 2 )
where ( θ x s , θ y s ) are the beam steering angles.

2.1. Beam Steering

Consider the x-axis component. The maximum intensity occurs when:
sin 1 2 N k d ( sin θ x sin θ x s ) sin 1 2 k d ( sin θ x sin θ x s ) = 1
This condition is satisfied when:
sin θ x sin θ x s = 0 sin θ x = sin θ x s
The relationship between steering angle and phase difference is:
θ x s = sin 1 λ 2 π d Δ φ
where Δ φ is the phase difference between adjacent array elements.

2.2. Beam Steering Range

The main lobe scanning is governed by:
2 π λ d sin θ M 2 π λ d sin θ x s = ± 2 i π , i = 0 , ± 1 , ± 2 ,
where θ M is the scanning angle of the beam maximum.
When the main lobe scanning angle θ x s = 0 :
sin θ M = ± λ d i , i = 0 , ± 1 , ± 2 ,
Since | sin θ | 1 , grating lobes appear when d λ .
When the main lobe scans to its maximum angle θ max :
sin θ M = ± λ d i + sin θ max , i = 0 , ± 1 , ± 2 ,
To avoid grating lobes at maximum scan angle:
d λ 1 + | sin θ max |
According to Equation (16), multi-beam formation occurs when the element spacing exceeds half the wavelength ( d 1 > 0.5 λ ), as illustrated in Figure 2b, where the annotated steering range highlights the angular coverage. In contrast, a single-beam pattern emerges when d 1 < 0.5 λ (Figure 2c). This stringent spacing requirement introduces significant fabrication challenges for LiDAR application [54].

2.3. Beam Width

The beam width θ FWHM is defined as the full width at half maximum (FWHM). According to the sinc function characteristics, the half-power point occurs when:
1 2 N k d ( sin θ x sin θ x s ) 1.39
Thus:
sin θ x sin θ x s = 1.39 N π · λ d
Setting θ x = θ x s + 1 2 θ FWHM and using the small-angle approximation, which holds strictly when θ FWHM is sufficiently small.
sin θ x s + 1 2 θ FWHM sin θ x s + 1 2 θ FWHM cos θ x s
Substituting into Equation (18) yields:
θ FWHM 0.88 λ N d cos θ x s
Here, we demonstrate the influence of element count on the full-width at half-maximum (FWHM) of the beam pattern, as calculated using Equation (20). A comparison between Figure 2c,d reveals two antenna arrays with identical element spacing but different aperture sizes. The results clearly show that a larger aperture yields a narrower FWHM, highlighting the direct relationship between array size and angular resolution.
It should be explicitly noted that, due to inconsistencies in parameter terminology across different disciplines and the fact that certain parameters reflect the unique advantages of specific materials, this paper adopts the following unified definitions for key OPA parameters to ensure clarity and precision: beam steering rate refers to the intrinsic response time of the material (denoted as Response time); beam steering range corresponds to the total angular coverage achievable by the beam steering system, i.e., the field-of-view (FOV); and beam width represents the minimum resolvable angular separation between adjacent beams, which is equivalent to the angular resolution.

3. Liquid Crystal Optical Phased Arrays

A spatial light modulator (SLM) is a diffractive optical element (DOE) device [37,38,39,55,56]. SLMs have shown great potential in widespread applications, and have been crucial in quantum optics [57,58], microscopy [59], imaging [60,61], optical trapping and tweezers [62,63], materials processing [64], and holography [65]. In this section, we focus on the OPA application of SLM and provide an overview of device scale, deflection angle, response time, and power tolerance capability.

3.1. Device Scale

The device scale of an LC-OPA is defined as the number, arrangement, and physical size of independently controllable basic modulation units within it, typically measured in “pixels” or “units”. It is closely related to regulation accuracy, scanning range, and pointing accuracy. In 1989, Raytheon Company in the United States developed the first one-dimensional LC-OPA device, which included 43,000 electrodes. With the help of LC, periodic scintillation gratings were generated to achieve 4 discrete angle deflection, and the highest beam efficiency was 85 % . In 1991, Raytheon developed an LC-OPA with an aperture of 4.3 cm × 4.1 cm, which divided 43,000 array elements into 168 sub arrays, and each array’s 256 units could be independently addressed. They implemented control of 43,000 array elements through 256 electrodes. In 1996, Raytheon reported a reflective LC-OPA with an aperture of 2 cm × 2 cm and 5000 elements, which achieved a deflection of 10.6 μm infrared beam within ±5° range, with a diffraction efficiency of 81 % at 4.5° [12]. In 2000, BNS Corporation in the United States adopted LC on silicon technology, achieving 8000 quasi continuous deflection angles within ±3° FOV, covering a wavelength range of 633–1550 nm, and an effective optical aperture of 7.4 mm × 6 mm [66]. In 2004, BNS developed an LC-OPA, which had 12,288 pixel units, electrode width of 1 μm, electrode spacing reduced to 0.6 μm, effective aperture increased to 19.66 mm × 19.66 mm, driving voltage increased to 13.2 V, and deflection range at 1550 μm wavelength increased to ±7° [67].
LC-OPAs benefit from semiconductor manufacturing compatibility, enabling scalable mass production. Their strengths in miniaturization and high integration of pixel arrays make them highly promising for aerospace, electronics, and advanced industrial applications. Crucially, as reflected in Figure 3 [68], continuous advances in LC materials and manufacturing technologies have driven significant improvements in LC-OPA performance, characterized by progressively smaller pixel pitch (reduced aperture) and concurrently increasing pixel density (higher number of elements).

3.2. Deflection Angle and Diffraction Efficiency

The diffraction efficiency of LC-OPAs is defined as the ratio of the diffracted light intensity in a specific direction to the incident light intensity. The diffraction efficiency serves as an indicator that is intimately associated with the device’s inherent space-bandwidth product, material properties, and structural characteristics. BNS successively launched one-dimensional and two-dimensional LC-OPAs. The one-dimensional LC-OPA has 12,288 independently controllable array elements, an LC-OPA aperture of 19.66 mm × 19.66 mm, a zero order light diffraction efficiency of 80–95%, a beam deflection angle range of ±4–7°, an LC response time of 5–30 ms, and a working wavelength range of 635 nm to 1.55 μm. The independent controllable array number of the two-dimensional LC-OPA is 512 × 512, with a working wavelength range of 532 nm to 1.55 μm. In 2008, Jay et al. produced a 1024-element LC-OPA device, which achieved scanning at 60 angles [69]. In 2018, Zhuo Rusheng from the University of Electronic Science and Technology of China proposed a large aperture array design with a PA in PA structure, with a single sub aperture of 10 mm and a total aperture of 100 mm (Figure 4a) [70]. In 2019, He et al. proposed an LC-OPA structure with expandable aperture, which maintained good main lobe focusing performance even when expanded to four times the original optical aperture [71].
In the late 20th and early 21st centuries, Raytheon and BNS in the United States were the first to conduct research on LC-OPAs. Due to technological limitations, the developed devices could only achieve an angle deviation of ±3° to ±7°. In 2008, North Carolina State University successfully developed an infrared LC-OPA that supported ±40° continuous scanning, achieving a large angle diffraction efficiency of over 80% at a wavelength of 1550 nm [75]. In 2012, Wenben et al. achieved the first cascade of LC-OPAs and Wollaston prism, achieving a large angle deflection of ±13.25° [76]. MIT researchers have pioneered an LC-OPA capable of visible-light beam forming and steering, achieving a 7.2° beam-steering range within ±3.4 V driving voltage (Figure 4c) [73]. This advancement represents a critical miniaturization milestone, enabling ultra-compact visible-wavelength systems for applications spanning free-space communications, LiDAR, and holographic displays. The architecture leverages cascaded phase-shifting elements with LC claddings to overcome traditional limitations in modulation efficiency and wavelength compatibility at visible spectra. In 2025, S. L. Zhou et al. revolutionized omnidirectional scanning by integrating an LC-OPA with conical mirrors, achieving 360° × 2.1° continuous FOV and reducing scanning periods to 1/4800 of conventional systems, a pivotal advance for LiDAR applications (Figure 4d) [74].

3.3. Response Time

The response time of LC-OPAs refers to the duration required for the device to complete phase modulation switching, encompassing both the rise time (transition from 10% to 90% of maximum phase shift). Such temporal characteristics ultimately determine the achievable beam steering refresh rate and precision in beam control applications. From 2006 to 2007, BNS successively launched one-dimensional and two-dimensional LC-OPAs. The one-dimensional LC-OPA had a response time of 5–30 ms. In 2009, David Engström et al. from the American company Display Technology demonstrated a one-dimensional ferroelectric LC-OPA made of highly tilted ferroelectric LC materials, which could provide 91% phase modulation between 0 and 2 π , achieve more than 700 distinguishable deflection angles, and had a response time of less than 200 μs [77]. In 2013, Mengjiao et al. used parallel arranged LC cells as an LC-OPA model for research and obtained the relationship between LC cell thickness and voltage response time [78]. In 2022, He et al. proposed a new three-layer stacked LC-OPA with a deflection angle coverage of ±100 μrad. The emission efficiency in all experiments was higher than 90%, the shortest switching time was only 1.52 ms, and the maximum deflection angle could reach ±260 μrad [79].

3.4. Power Tolerance Capability

The power tolerance of LC devices is mainly determined by power density, which is defined as the power value that can be accommodated per unit area. It determines the lifespan and power stability of the equipment. The laser tolerance of LC devices has long been constrained by the limited transmittance of their electrodes within specific wavelength bands. Since the turn of the century, researchers worldwide have conducted extensive research on electrode materials and heat dissipation techniques. For instance, the Transcon 1 device developed by Rockwell and Boeing in 2016 achieved an absorption rate of only 0.2% with active cooling [80]. In 2017, a reflective LC-OPA developed by the University of Electronic Science and Technology of China demonstrated a high-power laser tolerance of 272.4 W/cm2. However, even under active cooling its temperature rise reached 14 °C and its reflectivity was limited to 89.87% [81].
As laser technology advances towards higher energy levels, greater average power, and multi-wavelength applications, the demand for optical field modulation devices, especially LC-OPAs capable of broader wavelength operation, higher repetition rates, and enhanced laser damage thresholds, has grown significantly. To address these challenges, in 2020 Xing et al. [82] developed an LC-OPA employing Si- and Mg-doped GaN as transparent electrodes. Their experiments confirmed that GaN-based LC devices exhibit a high laser damage threshold exceeding 1 J/cm2 across both visible and near-infrared regions, demonstrating the potential for high-power GaN-based LC light valves [83] in demanding applications like spatial light modulation for LC-OPAs (Figure 5a,b). Subsequently, in 2022, Du et al. [84] employed laser damage modeling to investigate the damage characteristics of LC devices with ITO versus GaN electrodes under high-power laser irradiation. Analysis of thermal distribution and thermal stress revealed that GaN-electrode LC devices offer superior laser damage resistance and thermal stability compared to ITO-based devices (Figure 5c,d). These advantages are critical for the reliable operation of LC-OPAs in high-power laser systems.
Table 1 shows the performance comparison of the LC-OPA. The development trajectory of LC-OPA technology demonstrates wider steering angles (from ±2° to ±45°) and improved diffraction efficiency (reaching 99.5%). These advancements, coupled with expanding wavelength coverage (450–1550 nm) and recent breakthroughs in high-efficiency designs, indicate the technology is maturing toward practical applications. The ongoing improvements in FOV and response speed are laying essential technical foundations for LC-OPA’s future implementation in real-world systems such as LiDAR and optical communications, where large-angle, high-speed beam steering capabilities are critical requirements.

4. Lithium Niobate Optical Phased Arrays

LN [87] has low absorption loss in a wide transparent window. Its Pockels electro-optic (EO) coefficient and second-order nonlinear optical coefficient are relatively high [88]. The electro-optical characteristics of LN endow it with efficient phase modulation capability and high modulation speed [89]. The development of photonic devices on the thin film LN (TFLN) platform has attracted widespread attention from both the industrial and academic communities [90,91,92]. Compared with traditional bulk LN devices, it greatly reduces the footprint and improves performance [93,94]. TFLN modulators have demonstrated pure phase modulation with CMOS-compatible drive voltages and ultra-high bandwidths [95]. Although their fabrication is not yet fully established, this platform still enables the breakthrough potential for achieving systematic, large-scale, and multifunctional photonic chips on a single material platform [96,97,98].

4.1. Separate Lithium Niobate OPA

The application of OPA based on LN can be divided into two paths. One is to use bulk LN to form discrete phase modulation devices. Another approach is to use TFLN to form integrated on-chip devices. With the gradual maturity of LN electro-optic modulation devices, which have replaced acousto-optic modulation devices, researchers have proposed various phase control methods and achieved the basic prerequisite for active phase control. It is also a period of rapid development of such OPA in coherent synthesis applications.
Liu et al. presented the outcomes of phase locking attained through multi-channel spliced fiber optic arrays. Despite the potential for phase distortion in each channel to reach 180 Hz, constructive interference can be attained through the phase locking of the fiber array [99]. T. M. Shay et al. presented an electronic phase-locked optical array utilizing an LN electro-optic modulator platform, independent of an external reference beam. The phase locking of a passive fiber 3 × 3 array and a 6-element fiber amplifier array was achieved, as shown in Figure 6a,b. These optical phase-locked technologies are simple and can resist mechanical vibrations and thermal disturbances. Finally, using these techniques to achieve phase locking in a 22 W fiber amplifier [100], Xiao et al. demonstrated coherent beam combining using 2- and 3-element fiber amplifier arrays in the MOPA configuration. An LN phase modulator with a 500 MHz 3 dB bandwidth was employed in the system [101]. In 2018, Xidian University demonstrated a computational ghost imaging system using an optical fiber phased array employing LN phase modulators, as illustrated in Figure 6c [51]. In 2020, H. C. Chang et al. presented a 107-element coherent fiber array systems using LN phase modulators to modulate phase for the first time (Figure 6d) [102].

4.2. Thin Film Lithium Niobate OPA

The current approach to utilizing a single phased array antenna for two-dimensional scanning involves employing a phase shifter array to manage scanning in one dimension, while adjusting the wavelength facilitates scanning in the orthogonal dimension. The predominant platform for phase shifter arrays is TFLN. In 1974, Tien et al. first reported epitaxial TFLN waveguides based on OPA [103]. This article is among the earliest contributions to the field. Yue et al. presented an integrated LN OPA functioning in the near-infrared region. The system exhibits a two-dimensional beam steering range of 24° × 8°, a full width at half maximum of the far-field beam point measuring 2° × 0.6°, and a sidelobe suppression level of 10 dB [104]. Expanding the FOV of OPA fundamentally depends on the efficient reduction of crosstalk among array elements. In 2022, Tao Li et al. proposed a port-selective OPA based on a metasurface-enhanced LN integrated platform. It provided new degrees of freedom for attaining non-aliasing beam scanning and enhancing the FOV. The metasurface layer above the OPA was engineered to enhance the wide steering range. Consequently, they performed beam scanning throughout an FOV of 41.04° × 7.06° (Figure 7a,b) [105]. In 2023, Zhejiang University considered an on-chip 16-channel OPA based on TFLN phase modulators with traveling-wave electrodes. The OPA achieved an FOV of 50° × 8.6° and a beam width of 0.73° × 2.8° in the phase tuning direction and the wavelength scanning direction, respectively (Figure 7c,d) [106]. Moreover, OPA utilizing sparse array technology to theoretically disrupt the periodicity of waveguide array spacing, resulting in the high-order grating lobes in the far field no longer conforming to the constructive interference condition, thereby achieving the objective of broadening the beam scanning FOV. Shi et al. conceived and tested a two-dimensional scanning OPA utilizing a TFLN phase modulator and a traveling wave electrode. The modulation bandwidth was around 2.5 GHz. A mix of non-periodic arrays and flat grating antennas is employed to mitigate the grating lobes of far-field beams, resulting in an extensive FOV and narrow beam width. A 16-channel OPA has an FOV of 50° × 8.6° and a beam width of 0.73° × 2.8° in the phase tuning and wavelength scanning directions, respectively [106]. Li et al. examined the LN-OPA from a design standpoint. The OPA, consisting of various waveguide arrays, has been adjusted to reduce sidelobes to below −10 dB at a wavelength of 1.55 μm. Simultaneously, OPAs equipped with grating emitters or supplementary silicon dioxide cavities enhance the quality of beam deflection. Subsequently, the one-dimensional beam steering at a 50° angle, governed by an LN-OPA, was demonstrated [107].
The FOV in wavelength modulation is primarily constrained by the tunable wavelength range of the source and the wavelength tuning efficiency of the antenna. The potential for enhancing wavelength tuning efficiency by optimizing the structure of OPA antennas is significant, primarily through mode multiplexing and multi-line techniques. In 2024, Wang et al. presented a multi-waveguide channel integrated OPA based on the TFLN platform, illustrated in Figure 8a,b. Two LN-OPA chips utilize 32 and 48 channel LN-OPA waveguides, respectively, to achieve an FOV of 62.2° × 8.8° and a beam divergence of 2.4° × 1.2° for the 32-channel configuration through electro-optic modulation, while the 48-channel configuration provides an FOV of 40° × 8.8° and a beam divergence of 0.33° × 1.8° [108]. Furthermore, researchers have suggested a theoretical design of cascaded structures aimed at enhancing the performance of OPAs. Li et al. simulated an OPA architecture based on cascaded periodically poled LN sequences and only two control electronics to program the 2D beam-steering trajectory with a range of θ y = ±20° and θ z = ±16° [109]. Li et al. designed a multi-layer cascaded domain engineering structure within the LN waveguide, integrated with wavelength tuning, to facilitate two-dimensional beam scanning through a single electrode controlled OPA, as illustrated in Figure 8c. Simulation results indicated a two-dimensional beam steering capability of 42° × 9.2° [110]. Li et al. proposed a cascaded domain engineering OPA structure. A six-layer cascaded domain amplifier model was designed, comprising 32 array elements. A single electronic control device managed all components of the array. It showed a rapid response utilizing LN electro-optic crystals, with beam control speed reaching up to 3 MHz [111].
The ongoing advancement of OPA technology is anticipated to contribute significantly to the practical implementation of LiDAR. Shi et al. introduced and experimentally validated a hybrid system comprising a silicon integrated OPA and an optical frequency micro comb for parallel laser radar applications, as illustrated in Figure 9 [112]. Each OPA channel can attain 2D steering with an FOV of 80° while avoiding grating lobes in the horizontal direction. Table 2 shows the performance comparison of the LN-OPA.

5. Silicon Optical Phased Arrays

The notable difference in refractive indices between silicon and silica facilitates effective optical field confinement, providing a cost-efficient and stable approach for phase modulation. Beyond CMOS compatibility, silicon enable advanced functionalities like optical frequency combs and quantum light sources. Figure 10 illustrates that the advancement of silicon photonics has led to the emergence of various material platforms for OPA, each offering unique benefits. The silicon-on-insulator (SOI) platform [115] (Figure 10a) is notable for its superior optical confinement and seamless integration with established CMOS processes, facilitating the development of monolithically integrated active devices [116,117,118,119] and low-loss passive components [120,121,122,123]. In addition to SOI, various silicon-based platforms have developed to meet particular requirements. S i N waveguides (Figure 10b) can achieve ultra-low transmission loss. Compared with silicon materials, the nonlinear loss of S i N is significantly reduced, supporting high-power laser operation. Its optical transparency window can cover the visible to near-infrared band, making it an ideal platform for visible light OPA, rendering it suitable for long-distance applications, whereas the hybrid III/V silicon platform (Figure 10c) [124] integrates effective light generation with the processing benefits of silicon. Moreover, the hybrid III/V silicon platform (Figure 10c) [124] integrates effective light generation with the processing benefits of silicon, offering complementary advantages. Advanced methodologies such as the S i - S i N multilayered platform enhance light transmission via heterogeneous integration. The inherent CMOS compatibility of these silicon-based solutions provides OPA devices with low cost, low power consumption and rapid scanning capabilities, facilitating their adoption in solid-state LiDAR applications [125,126,127]. The performance of these systems is fundamentally linked to the unique optical properties of each material, including refractive index, thermal-optic coefficients, and loss characteristics. These properties ultimately determine critical OPA metrics such as array scale, ranging resolution, and beam steering rates.

5.1. Si-Based OPA

The evolution of OPA technology has progressed through distinct material platforms. Silicon materials exhibit limitations in electro-optic modulation attributed to their low electro-optic coefficients. Nevertheless, their thermal-optic coefficient, approximately 1.86 × 10 4 K 1 at a wavelength of 1.55 μm, enables the potential application of the thermal-optic effect.
In 2009, Ghent University demonstrated a one-dimensional 16-element OPA utilizing thermal-optic phase shifters, achieving a beam scanning range of 14.1° × 2.3° [125]. In 2011, a study was conducted on a 16-channel, independently tuned waveguide surface grating OPA in silicon for two dimensional beam steering with a total FOV of 20° × 14° and beam width of 0.6° × 1.6° (Figure 11a) [128]. In 2010, a two-dimensional OPA was proposed, capable of steering beams in both the θ and ϕ directions, utilizing delay lines and grating couplers [127]. The Massachusetts Institute of Technology (MIT) advanced the development of silicon-based OPA for high-density integration and system-level applications. In 2013, researchers presented a large-scale two-dimensional nanophotonic phased array that integrated 64 × 64 optical nanoantennas on a silicon chip. The compact system, with dimensions of 576 μm × 576 μm, successfully demonstrated dynamic beam steering and shaping capabilities through an 8 × 8 sub-array configuration (Figure 11b) [116]. In 2015, the University of Southern California presented the initial demonstration of monolithically integrated 2D OPA featuring photodetectors on a single SOI platform. This study combined the driving circuit with an 8 × 8 two-dimensional OPA on a single SOI wafer, where each array element was fitted with a thermo-optic phase shifter and attenuator. This configuration facilitated arbitrary beamforming through independently adjustable amplitude and phase [129]. In 2016, Intel technologies introduced a non-uniform one-dimensional 128-element periodic Si-OPA. This study mitigated the formation of grating lobes in the phased array by disrupting the periodicity of the antenna array, enabling beam steering across a 100° × 17° FOV via thermo-optic phase shifting and wavelength tuning (Figure 11c) [130]. In 2017, MIT developed a prototype of a solid-state OPA LiDAR that integrated beam transmission and reception functions. The integration of heterodyne coherent detection technology enabled a 2 m range with a 50° FOV, signifying the practical application of OPA technology in exploration [117]. The team consistently achieved significant advancements and engineering-level upgrades from 2017 to 2020. An OPA system with 512 dimensions incorporated a custom-designed ASIC driver chip, resulting in an FOV of 56° × 15°, a beam width of 0.04° × 0.02°, and a ranging distance of 200 m [131,132]. In 2020, the Columbia University demonstrated a low-power, 2D, 512-element large-scale OPA enabled by a novel multi-pass phase shifter structure [133] (Figure 11d). This structure achieved an FOV of 70° × 6° and consumed power of 1.9 W. In 2020, researchers presented the highest-performance 8192-dimensional ultra-large-scale one-dimensional array, utilizing flip-chip CMOS driver technology. This array features a radiation aperture of 8 × 5 mm2, enabling a two-dimensional scan of 100° × 17° and achieving a beam accuracy of 0.01° × 0.039°. This represents a significant industrial advancement in scale and performance for silicon-based OPAs [119].

5.2. SiN-Based OPA

Silicon materials are fundamental for effective OPA on SOI platforms, attributed to their elevated refractive index, robust optical confinement, and compatibility with fabrication processes. Additionally, device performance can be improved through the incorporation of low-loss silicon nitride (SiN) materials. Since 2017, SiN has developed into a novel material platform for OPAs due to its low-loss characteristics. MIT has pioneered a CMOS-compatible 4 × 4 mm2 SiN phased array operating in the visible spectrum, achieving a significant milestone with 400 mW high-power handling and a beam width of 0.021°. However, the thermo-optic modulation efficiency remains constrained by the material properties [131].
In 2018, researchers at MIT proposed a novel architecture known as ‘lens-assisted optical path with wavelength tuning’. This design utilizes Mach–Zehnder interferometer (MZI) switches and on-chip integrated lenses to dynamically select optical paths, marking the first demonstration of 38.8° × 12° two-dimensional beam steering on a SiN platform [121]. The Interuniversity Microelectronics Center (IMEC) has developed a dual-layer OPA chip through a SiN/Si bilayer fabrication process. The implementation of a power distribution network featuring SiN edge couplers and cascaded multimode interference (MMI) enhanced the chip’s power tolerance. Thermo-optic phase modulation was subsequently achieved through low-loss interlayer transition structures on silicon waveguides, followed by etching on SiN waveguides to create weakly perturbed dual-layer radiating antenna gratings for off-chip light emission [134]. In 2021, the Shanghai Institute of Microsystem and Information Technology created a 64-dimensional OPA chip utilizing a multilayer SiN/Si process, resulting in an FOV of 35.5° × 22.7° and a beam width of 0.69° × 0.075° [135]. Shanghai Jiao Tong University has reported a LiDAR transmitter that integrates a hybrid tunable external cavity laser with a high-resolution two-dimensional OPA beam steering system. This transmitter, constructed on a three-layer SiN/Si platform, features a widely tunable (100 nm) narrow-linewidth (2.8 kHz) external cavity laser and a 256-dimensional non-uniform OPA with low insertion loss. The chip dimensions are 15.6 mm × 11.7 mm, facilitating an FOV of 140° × 16° and a beam width of 0.051° × 0.016° [136]. The Institute of Semiconductors, Chinese Academy of Sciences, has achieved a breakthrough in passive architecture design by utilizing 27 μm/54 μm delay lines, resulting in scanning densities of 0.44°/nm × 0.07°/nm and 0.87°/nm × 0.07°/nm, respectively. This advancement marks a significant improvement in low phase noise far-field quality. This technical route emphasizes the iterative capabilities of SiN materials in low-power, high-precision OPA systems [123]. In 2024, Jilin University propose a 64-channel Si-SiN dual-layer OPA chip for optical communication applications (Figure 12a) [137]. The system exhibited 6.256 m detection range and 52 Kbps data. In 2025, the Chinese Academy of Sciences presented a 128-channel aperiodic OPA based on a dual-layer emission grating on a SiN-on-SOI platform (Figure 12b) [138]. It achieved an FOV of 150° × 16° with a divergence angle of 0.022° × 0.060°. The OPA was used in a frequency modulation continuous wave system to achieve a distance measurement of 40 m.

5.3. Si-SiN-Based OPA

Another important method to increase the output power of the OPA is to input a higher-power laser from the outside [135,136,139,140,141]. Silicon waveguides and devices feature compact dimensions and support diverse functionalities, enabling realization of densely integrated array elements and large-scale OPA chips. However, their performance is limited by nonlinear effects including two-photon absorption, which restricts the operational power threshold (with nonlinear loss typically occurring around 100 mW and damage threshold at approximately 2 W). In contrast, SiN materials exhibit superior characteristics of low optical loss and high nonlinear threshold (where weak nonlinear effects are only observable in high-quality-factor resonant structures, and the damage threshold can reach 35 W), while maintaining full CMOS compatibility. Consequently, a silicon–SiN multilayer integration strategy effectively combines the advantages of both material systems, providing an optimal solution for achieving high main lobe power output—in this hybrid approach, passive components such as power splitters are implemented in SiN to minimize loss, while active thermo-optic or electro-optic phase shifting/modulation elements are fabricated in silicon to maintain functionality.
The transmitting unit (antenna) may employ a single material or integrate multiple approaches. Interlayer coupling is facilitated by evanescent field coupling [142,143], with coupling loss kept under 0.15 dB [144]. Ref. [145] depicts a large-scale OPA utilizing a multi-layer SiN on silicon photonic platform. The design integrates a hybrid I I I / V - S i N external cavity laser (ECL) with a 256-channel OPA, resulting in a significant 100 nm wavelength tuning range, a linewidth of 2.8 kHz, and an FOV of 50° × 16° for accurate beam steering [145]. In 2020, the Chinese Academy of Sciences developed a dual-layer OPA utilizing SiN and silicon. This design leverages the low-loss properties of SiN alongside the superior modulation capabilities of silicon, resulting in improved performance of single-layer silicon-based OPA while effectively reducing nonlinear effects associated with high-power optical input in silicon. The system exhibited an FOV measuring 96° × 14° [120]. In 2021, Jilin University developed a 128-dimensional PN junction phase-shifting OPA on the Si-SiN platform, utilizing dual-layer SiN waveguide grating antennas to attain a wide FOV. This configuration achieves a detection range of 100 m when utilized in frequency modulated continuous wave (FMCW) LiDAR systems. As showm in Figure 13 [146].
Table 3 presents a comparison of the performance of various materials utilized in silicon-based OPA. With the maturation of silicon-based OPA technology, commercial OPA LiDAR products are starting to appear on the market. Advancing high-performance OPA systems requires material innovation, specifically through the optimization of thermal management materials on the silicon platform to reduce thermal-optic interference and addressing the material processing challenges linked to III-V/silicon hetero-integration. These advancements will result in compact, low-power, high-speed, and low-loss phased array chips.

6. Other Materials for Optical Phased Arrays

Beyond the primary material platforms discussed, several alternative approaches have demonstrated significant potential for OPAs. The early development of AlGaAs-based OPAs laid critical foundations. Hobbs et al. demonstrated a 5-element GaAs OPA, achieving ±20° beam steering at 500 kHz for LiDAR applications. Subsequent advances included a 50-element integrated AlGaAs OPA attaining ±0.41° FOV [147]. In 1997, Xidian University conducted extensive theoretical and experimental work on AlGaAs OPA, leading to the successful fabrication of multiple OPA prototypes [148,149]. However, limited by the low integration capability of AlGaAs materials at the time, this approach failed to advance further. III-V platforms, particularly InP, now enable monolithic integration of active/passive components in the 1.3–1.6 μm band [150], featuring semiconductor optical amplifiers for high-power operation, GHz-speed phase shifters via current injection, and nanosecond-scale tunable lasers. The demonstrated InP OPAs achieve fast beam steering with 10° × 6.5° FOV [151]. III-V/Si hybrid platforms address silicon’s thermal limitations by leveraging quantum-confined Stark/Pockels effects [152,153], achieving breakthrough performance: heterogeneous phase shifters exhibit ultra-low V 2 π voltage (0.35–1.4 V) and minimal residual amplitude modulation (0.1–0.15 dB) across 200 nm [154], while 32-channel arrays with 4μm pitch enable 22° × 28° FOV with 0.78° × 0.02° beamwidth at <3 nW/2π shift [153]. Fully integrated 2D OPAs (164 components) achieve 23° × 3.6° FOV on 6 × 11.5 mm2 chips [155], and sub-2μm pitch designs demonstrate 51° × 28° FOV [152]. These platforms collectively expand OPA capabilities beyond conventional approaches through high-speed active integration (InP) and voltage-efficient heterogeneous systems (III-V/Si).

7. Conclusions and Outlook

This review provides a comprehensive analysis of the progress, advantages, and inherent limitations of OPA technology across three key material platforms: LC, LN, and Si. LC-OPAs exhibit excellent diffraction efficiency and the ability to integrate a large number of phase-shifting elements, enabling high-resolution beam steering and wavefront shaping. These characteristics make them particularly suitable for applications requiring precise optical field manipulation and laser communications. However, their relatively slow response time remains a major constraint for dynamic applications. LN-OPAs offer superior high-speed modulation capabilities (with bandwidths reaching up to 2.5 GHz), low optical loss, and high optical power tolerance, making them ideal candidates for high-power optical communication and quantum optics applications. Nevertheless, challenges remain in improving power efficiency and simplifying the fabrication process. Si-OPAs benefit from mature CMOS compatibility and high integration density, establishing them as the leading platform for photonic integrated circuits and automotive LiDAR systems. However, their application is limited by constraints in optical power handling capacity.
Future progress in OPA technology hinges on addressing platform-specific performance limitations through material innovation. For LC-OPAs, the slow response time represents the most critical limitation for dynamic applications. Therefore, substantial research efforts are required to develop new LC materials, driving schemes, or device architectures capable of achieving microsecond-level or faster switching. For LN-OPAs, despite their excellent modulation speed and power handling capabilities, efforts must focus on reducing fabrication complexity to improve yield and lower production costs. For Si-OPAs, which excel in integration density, further advancements are needed to enhance optical power handling and phase control fidelity and range in order to broaden their applicability in high-power or complex beam manipulation scenarios.
From a performance perspective, as shown in Table 4, the practical deployment of OPAs hinges critically on achieving wide FOV scanning and high-speed beam steering. Realizing a wide FOV fundamentally requires element spacing approaching or even sub-half the operating wavelength ( 0.5 λ ) to suppress grating lobes. However, current material platforms face inherent limitations in attaining 0.5 λ spacing: LC-OPAs exhibit significant difficulty in achieving sub-wavelength spacing. Nevertheless, techniques such as multi-beam scanning and device cascading have been employed to effectively extend the effective FOV despite this spatial limitation. LN-OPAs, fabricated via semiconductor processing routes, demonstrate potential for reduced element pitch. However, realizing consistent sub- λ spacing demands further breakthroughs in fabrication precision. Similar to LC-OPAs, multi-beam architectures offer a viable pathway to enhance their angular coverage. Si-OPAs, while theoretically capable of 0.5 λ designs, practically achieve at best 1.5–2 λ spacing in scalable implementations. Aggressive scaling below λ induces severe inter-element crosstalk. This crosstalk degrades phase fidelity across the array, exacerbates sidelobe levels, and ultimately compromises beamforming accuracy.
Recent studies on non-uniform, strongly coupled waveguide arrays, particularly in SiN platforms, demonstrate significant potential for simultaneous control over both optical power distribution and phase profiles across the array [156,157]. This capability substantially expands the functional scope and application potential of OPAs. Additionally, emerging dual-layer platforms that integrate materials such as silicon and SiN offer enhanced flexibility for shaping complex phase profiles [137]. These developments signify a shift toward leveraging innovative photonic designs, moving beyond conventional periodic arrays to achieve functionalities that were previously difficult or unattainable using traditional single-material OPA approaches.
The integration of computational intelligence into OPA design and control presents another promising avenue. Artificial intelligence (AI) algorithms can expedite the optimization of complex antenna layouts, such as non-uniform arrays, improve beam steering accuracy and speed through adaptive control mechanisms, and potentially enable real-time compensation for optical aberrations or environmental disturbances, thereby enhancing system robustness and adaptability.
Table 4. The performance comparison of the OPA.
Table 4. The performance comparison of the OPA.
TypePower Con-SumptionFootprintScaleFOV
θ x (°) × θ y (°)
Response
Times
Angular Resolution
θ x (°) × θ y (°)
YearRef.
LC-OPA19.66 mm × 19.66 mm1 × 12,2886.95 × –24.8 ms2006[69]
80 mm × 80 mm24 × 2417.2 ms2022[85]
20 mm × 20 mm22.9 × –70 μs2022[158]
15.36 cm × 9.6 cm1900 × 1200360 × 2.114.0 ms0.04 × –2025[74]
LN-OPA1.11 nJ/π48 μm × 100 μm3262.2 × 8.814 ns2.4 × 1.22024[108]
18.5 mW/π12880 × 50.1 × 0.0372025[112]
6421.22 × –1.37 μs0.07 × –2025[113]
41.6 pJ/π1662 × 7.626 ns3.2 × 1.42025[114]
Si-OPA8.5 mW/π576 μm × 576 μm40966 × 62013[116]
8 mm × 5 mm8192100 × 1730 μs0.01 × 0.042022[119]
1.8 mm × 2.5 mm25614.8 × 1500.066 × 0.0682023[141]
6 mW/π256160 × –0.16 × 0.042024[159]
While LC-, LN-, and Si-based OPA currently form the technological foundation with distinct advantages for specific applications, the future development of OPA technology will be driven by targeted material and platform optimizations, the potential of hybrid systems, and, critically, the design of novel waveguide architectures leveraging materials like SiN and heterogeneous integration. When combined with intelligent computational control, these advancements are expected to enable unprecedented levels of light manipulation, positioning OPA technology to play a transformative role across a rapidly expanding range of scientific and technological applications.

Funding

This research was supported by the Jiangsu Province Key R&D Program (Grant No. BE2023009-5), the Open Fund of Key Laboratory for Intelligent Nano Materials and Devices of the Ministry of Education NJ2023002 (Grant No. INMD-2023M06), the Suzhou Science and Technology Plan projects (Grant No. ZXL2024375), and the Postgraduate Scientific Research Innovation Project of Hunan Province (Grant No. QL20230006). The support from the Vacuum Interconnected Nanotech Workstation (Nano-X) of Suzhou Institute of Nano-tech and Nano-bionics (SINANO), Chinese Academy of Sciences, and SINANO Excellent Young Talents Program are also acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Beer, M.; Haase, J.F.; Ruskowski, J.; Kokozinski, R. Background Light Rejection in SPAD-Based LiDAR Sensors by Adaptive Photon Coincidence Detection. Sensors 2018, 18, 4338. [Google Scholar] [CrossRef]
  2. Hao, Y.; Zhu, Q.; Han, Y.; Zang, Z.; Ye, Z.; Chang-Hasnain, C.; Fu, H.Y. Hybrid FMCW LiDAR System With Nonlinearity-Corrected MEMS-VCSEL for Extended 3D Imaging Range. J. Light. Technol. 2025, 43, 4042–4052. [Google Scholar] [CrossRef]
  3. Lin, Y.X.; Ahmad, Z.; Ou, S.Y.; Su, W.C.; Chang, Y.C.; Chen, J.H.; Hung, Y.J.; Chang, Y.C.; Wei, C.C.; Horng, T.S.; et al. A 4-D FMCW LiDAR With Ultra-High Velocity Sensitivity. J. Light. Technol. 2023, 41, 6664–6674. [Google Scholar] [CrossRef]
  4. Li, C.; Zhang, F.; Qu, X. High-Resolution Frequency-Modulated Continuous-Wave LiDAR Using Multiple Laser Sources Simultaneously Scanning. J. Light. Technol. 2023, 41, 367–373. [Google Scholar] [CrossRef]
  5. Behroozpour, B.; Sandborn, P.A.M.; Quack, N.; Seok, T.J.; Matsui, Y.; Wu, M.C.; Boser, B.E. Electronic-Photonic Integrated Circuit for 3D Microimaging. IEEE J. Solid-State Circuits 2017, 52, 161–172. [Google Scholar] [CrossRef]
  6. Dong, Y.; Zhu, Z.; Tian, X.; Qiu, L.; Ba, D. Frequency-Modulated Continuous-Wave LIDAR and 3D Imaging by Using Linear Frequency Modulation Based on Injection Locking. J. Light. Technol. 2021, 39, 2275–2280. [Google Scholar] [CrossRef]
  7. Chen, G.; Liu, J.; Liu, M.; Qu, X.; Zhang, F. Non-Line-of-Sight Ranging and 3D Imaging Using Vector Enhanced Sensitive FMCW LiDAR. J. Light. Technol. 2025, 43, 4119–4126. [Google Scholar] [CrossRef]
  8. Liu, Q.; Benedikovic, D.; Smy, T.; Atieh, A.; Cheben, P.; Ye, W.N. Circular Optical Phased Arrays with Radial Nano-Antennas. Nanomaterials 2022, 12, 1938. [Google Scholar] [CrossRef]
  9. Chow, C.W.; Chang, Y.C.; Kuo, S.I.; Kuo, P.C.; Wang, J.W.; Jian, Y.H.; Ahmad, Z.; Fu, P.H.; Shi, J.W.; Huang, D.W.; et al. Actively Controllable Beam Steering Optical Wireless Communication (OWC) Using Integrated Optical Phased Array (OPA). J. Light. Technol. 2023, 41, 1122–1128. [Google Scholar] [CrossRef]
  10. Zou, F.; Pan, Z.; Liu, J.; Li, Z.; Pan, L.; Yang, R.; Jiang, J.; Li, F.; Geng, C.; Li, X. Turbulence Compensation for Receiving and Coherent Combining via Phased Fiber Laser Array. IEEE Photonics Technol. Lett. 2023, 35, 733–736. [Google Scholar] [CrossRef]
  11. Jayaweera, V.L.; Peiris, C.; Darshani, D.; Edirisinghe, S.; Dharmaweera, N.; Wijewardhana, U. Visible Light Communication for Underwater Applications: Principles, Challenges, and Future Prospects. Photonics 2025, 12, 593. [Google Scholar] [CrossRef]
  12. McManamon, P.F.; Dorschner, T.A.; Corkum, D.L.; Friedman, L.J.; Hobbs, D.S.; Holz, M.; Liberman, S.; Nguyen, H.Q.; Resler, D.P.; Sharp, R.C.; et al. Optical phased array technology. Proc. IEEE 1996, 84, 268–298. [Google Scholar] [CrossRef]
  13. Huang, W.R.; Montoya, J.; Kansky, J.E.; Redmond, S.M.; Turner, G.W.; Sanchez-Rubio, A. High speed, high power one-dimensional beam steering from a 6-element optical phased array. Opt. Express 2012, 20, 17311–17318. [Google Scholar] [CrossRef]
  14. Wang, X.; Liu, C.; Cao, Y.; Liu, R.; Zhang, L.; Zhao, X.; Lu, F.; Miao, Z.; Li, Q.; Han, X. High-precision two-dimensional beam steering with a 64-element optical fiber phased array. Appl. Opt. 2021, 60, 10002–10008. [Google Scholar] [CrossRef] [PubMed]
  15. Vorobyov, S.; Gershman, A.; Luo, Z.Q. Robust adaptive beamforming using worst-case performance optimization: A solution to the signal mismatch problem. IEEE Trans. Signal Process. 2003, 51, 313–324. [Google Scholar] [CrossRef]
  16. Ni, G.; Song, Y.; Chen, J.; He, C.; Jin, R. Single-Channel LCMV-Based Adaptive Beamforming With Time-Modulated Array. IEEE Antennas Wirel. Propag. Lett. 2020, 19, 1881–1885. [Google Scholar] [CrossRef]
  17. Xin’gang, Z.; Dongsheng, Z.; Pengzhan, H.; Men, G.; Zhengxin, S. Optimized design of high throughput satellite antenna based on differential evolution algorithm. Chin. J. Radio Sci. 2024, 39, 1154–1159. [Google Scholar] [CrossRef]
  18. Barad, D.; Dash, J.C.; Sarkar, D.; Srinivasulu, P. Dual-band dual-polarized sub-6 GHz phased array antenna with suppressed higher order modes. Sci. Rep. 2024, 14, 6139. [Google Scholar] [CrossRef] [PubMed]
  19. Xiao, F.; Kong, L. Optical multi-beam forming method based on a liquid crystal optical phased array. Appl. Opt. 2017, 56, 9854–9861. [Google Scholar] [CrossRef]
  20. Ding, L.; Shi, C.; Zhou, J. Joint Beampattern Design and Online Route Planning for Multitarget Tracking in Airborne Radar System. IEEE Trans. Aerosp. Electron. Syst. 2024, 60, 774–788. [Google Scholar] [CrossRef]
  21. Yan, J.; Pu, W.; Liu, H.; Zhou, S.; Bao, Z. Cooperative target assignment and dwell allocation for multiple target tracking in phased array radar network. Signal Process. 2017, 141, 74–83. [Google Scholar] [CrossRef]
  22. Zhang, W.; Shi, C.; Zhou, J.; Lv, R. Joint Aperture and Transmit Resource Allocation Strategy for Multitarget Localization in the Phased Array Radar Network. IEEE Trans. Aerosp. Electron. Syst. 2023, 59, 1551–1565. [Google Scholar] [CrossRef]
  23. Yuan, Y.; Yi, W.; Hoseinnezhad, R.; Varshney, P.K. Robust Power Allocation for Resource-Aware Multi-Target Tracking With Colocated MIMO Radars. IEEE Trans. Signal Process. 2021, 69, 443–458. [Google Scholar] [CrossRef]
  24. Zhao, D.; Gu, P.; Zhong, J.; Peng, N.; Yang, M.; Yi, Y.; Zhang, J.; He, P.; Chai, Y.; Chen, Z.; et al. Millimeter-Wave Integrated Phased Arrays. IEEE Trans. Circuits Syst. I Regul. Pap. A Publ. IEEE Circuits Syst. Soc. 2021, 68, 3977–3990. [Google Scholar] [CrossRef]
  25. Sadhu, B.; Paidimarri, A.; Liu, D.; Yeck, M.; Ozdag, C.; Tojo, Y.; Lee, W.; Gu, K.X.; Plouchart, J.O.; Baks, C.W.; et al. A 24–30-GHz 256-Element Dual-Polarized 5G Phased Array Using Fast On-Chip Beam Calculators and Magnetoelectric Dipole Antennas. IEEE J. Solid-State Circuits 2022, 57, 3599–3616. [Google Scholar] [CrossRef]
  26. Akaishi, A.; Iguchi, M.; Hariu, K.; Shimada, M.; Kuroda, T.; Yajima, M. Ka-band Active Phased Array Antenna for WINDS Satellite. In Proceedings of the 21st International Communications Satellite Systems Conference and Exhibit, Yokohama, Japan, 15 April–19 April 2003. [Google Scholar]
  27. Pang, J.; Li, Z.; Kubozoe, R.; Luo, X.; Wu, R.; Wang, Y.; You, D.; Fadila, A.A.; Saengchan, R.; Nakamura, T.; et al. A 28-GHz CMOS Phased-Array Beamformer Utilizing Neutralized Bi-Directional Technique Supporting Dual-Polarized MIMO for 5G NR. IEEE J. Solid-State Circuits 2020, 55, 2371–2386. [Google Scholar] [CrossRef]
  28. Pang, J.; Li, Z.; Luo, X.; Alvin, J.; Saengchan, R.; Fadila, A.A.; Yanagisawa, K.; Zhang, Y.; Chen, Z.; Huang, Z.; et al. A CMOS Dual-Polarized Phased-Array Beamformer Utilizing Cross-Polarization Leakage Cancellation for 5G MIMO Systems. IEEE J. Solid-State Circuits 2021, 56, 1310–1326. [Google Scholar] [CrossRef]
  29. Pascale, V.; Maiarelli, D.; D’Agristina, L.; Gatti, N. Design and qualification of Ku-band-radiating chains for receive active array antennas of flexible telecommunication satellites. Int. J. Microw. Wirel. Technol. 2020, 12, 487–503. [Google Scholar] [CrossRef]
  30. Roper, D.H.; Babiec, W.E.; Hannan, D.D. WGS phased arrays support next generation DoD SATCOM capability. In Proceedings of the IEEE International Symposium on Phased Array Systems and Technology, Boston, MA, USA, 14–17 October 2003. [Google Scholar]
  31. Yajima, M.; Hasegawa, T.; Kuroda, T.; Shimada, M. On-board Evaluation Results of Active Phased Array Antenna for WINDS Satellite. Trans. Jpn. Soc. Aeronaut. Space Sci. Aerosp. Technol. Jpn. 2011, 8, Pj13–Pj18. [Google Scholar] [CrossRef]
  32. Boyd, C. A Dual-Mode Latching Reciprocal Ferrite Phase Shifter. IEEE Trans. Microw. Theory Tech. 1970, 18, 1119–1124. [Google Scholar] [CrossRef]
  33. Che, W.; Junding, W.; Shu, K.; Wen, Y. Phase-adjustment technique of the digital-latching ferrite phase shifter. IEEE Trans. Microw. Theory Tech. 1999, 47, 1125–1127. [Google Scholar] [CrossRef]
  34. Ghaffar, F.A.; Shamim, A. A Partially Magnetized Ferrite LTCC-Based SIW Phase Shifter for Phased Array Applications. IEEE Trans. Magn. 2015, 51, 4003108. [Google Scholar] [CrossRef]
  35. Simon, K.; Schindler, M.; Mieczkowski, V.; Newman, P.; Goldfarb, M.; Reese, E.; Small, B. A production ready, 6-18-GHz, 5-b phase shifter with integrated CMOS compatible digital interface circuitry. IEEE J. Solid-State Circuits 1992, 27, 1452–1456. [Google Scholar] [CrossRef]
  36. Hao, Z.; Yuan, W. A 6-18GHz 6-bit Phase Shifter for Broadband Phased Array Applications. In Proceedings of the 2022 International Conference on Microwave and Millimeter Wave Technology (ICMMT), Harbin, China, 12–15 August 2022; pp. 1–3. [Google Scholar] [CrossRef]
  37. Hu, W.; Ismail, M.; Cahill, R.; Gamble, H.; Dickie, R.; Fusco, V.; Linton, D.; Rea, S.; Grant, N. Tunable liquid crystal reflectarray patch element. Electron. Lett. 2006, 42, 509–511. [Google Scholar] [CrossRef]
  38. Karabey, O.H.; Gaebler, A.; Strunck, S.; Jakoby, R. A 2-D Electronically Steered Phased-Array Antenna With 2 × 2 Elements in LC Display Technology. IEEE Trans. Microw. Theory Tech. 2012, 60, 1297–1306. [Google Scholar] [CrossRef]
  39. Aghabeyki, P.; Cai, Y.; Deng, G.; Frølund Pedersen, G.; Zhang, S. Beam-Scanning Reflectarray With Switchable Polarization Based on Liquid Crystal at Millimeter Wave. IEEE Trans. Antennas Propag. 2025, 73, 1894–1899. [Google Scholar] [CrossRef]
  40. Wang, Y.; Wang, Y.; Yang, G.; Li, Q.; Zhang, Y.; Yan, S.; Wang, C. All-Solid-State Optical Phased Arrays of Mid-Infrared Based Graphene-Metal Hybrid Metasurfaces. Nanomaterials 2021, 11, 1552. [Google Scholar] [CrossRef]
  41. Wang, Z.; Feng, J.; Li, H.; Zhang, Y.; Wu, Y.; Hu, Y.; Wu, J.; Yang, J. Ultra-Compact and Broadband Nano-Integration Optical Phased Array. Nanomaterials 2023, 13, 2516. [Google Scholar] [CrossRef]
  42. Zhao, S.; Chen, J.; Shi, Y. All-Solid-State Beam Steering via Integrated Optical Phased Array Technology. Micromachines 2022, 13, 894. [Google Scholar] [CrossRef] [PubMed]
  43. Hsu, C.P.; Li, B.; Solano-Rivas, B.; Gohil, A.R.; Chan, P.H.; Moore, A.D.; Donzella, V. A Review and Perspective on Optical Phased Array for Automotive LiDAR. IEEE J. Sel. Top. Quantum Electron. 2021, 27, 8300416. [Google Scholar] [CrossRef]
  44. Yi, Y.; Wu, D.; Kakdarvishi, V.; Yu, B.; Zhuang, Y.; Khalilian, A. Photonic Integrated Circuits for an Optical Phased Array. Photonics 2024, 11, 243. [Google Scholar] [CrossRef]
  45. Heck, M.J. Highly integrated optical phased arrays: Photonic integrated circuits for optical beam shaping and beam steering. Nanophotonics 2017, 6, 93–107. [Google Scholar] [CrossRef]
  46. Jiao, Y.; Nishiyama, N.; van der Tol, J.; van Engelen, J.; Pogoretskiy, V.; Reniers, S.; Kashi, A.A.; Wang, Y.; Calzadilla, V.D.; Spiegelberg, M.; et al. InP membrane integrated photonics research. Semicond. Sci. Technol. 2020, 36, 013001. [Google Scholar] [CrossRef]
  47. Isaac, B.J.; Song, B.; Pinna, S.; Coldren, L.A.; Klamkin, J. Indium Phosphide Photonic Integrated Circuit Transceiver for FMCW LiDAR. IEEE J. Sel. Top. Quantum Electron. 2019, 25, 8000107. [Google Scholar] [CrossRef]
  48. Guo, H.; Li, J.; Wang, X.; Shang, J. Research review on the development of liquid crystal optical phased array device on the application of free space laser communication. Proc. SPIE 2018, 10841, 69–89. [Google Scholar] [CrossRef]
  49. Mcmanamon, P.F.; Ataei, A. Progress and opportunities in the development of nonmechanical beam steering for electro-optical systems. Opt. Eng. 2019, 58, 120901. [Google Scholar] [CrossRef]
  50. Xu, W.; Yuan, Q.; Yang, Y.; Lu, L.; Chen, J.; Zhou, L. Progress and prospects for LiDAR-oriented optical phased arrays based on photonic integrated circuits. npj Nanophotonics 2025, 2, 14. [Google Scholar] [CrossRef]
  51. Liu, C.B.; Chen, J.Q.; Liu, J.X.; Han, X.E. High frame-rate computational ghost imaging system using an optical fiber phased array and a low-pixel APD array. Opt. Express 2018, 26, 10048–10064. [Google Scholar] [CrossRef]
  52. Liu, C.B.; Lu, F.; Wu, C.; Han, X.E. Spatial correlation properties of coherent array beams modulated by space-time random phase. Opt. Commun. 2015, 346, 26–33. [Google Scholar] [CrossRef]
  53. Born, M.; Wolf, E. Principles of Optics: 60th Anniversary Edition, 7th ed.; Cambridge University Press: Cambridge, UK, 2019. [Google Scholar] [CrossRef]
  54. Wang, X. Beams Optimization and Steering via Optical Phased Array. Ph.D. Thesis, Xidian University, Xi’an, China, 2022. [Google Scholar] [CrossRef]
  55. Soref, R.A.; Rafuse, M.J. Electrically Controlled Birefringence of Thin Nematic Films. J. Appl. Phys. 1972, 43, 2029–2037. [Google Scholar] [CrossRef]
  56. Mailer, H.; Likins, K.L.; Taylor, T.R.; Fergason, J.L. Effect of Ultrasound on A Nematic Liquid Crystal. Appl. Phys. Lett. 1971, 18, 105–107. [Google Scholar] [CrossRef]
  57. Yao, E.; Franke-Arnold, S.; Courtial, J.; Padgett, M.J.; Barnett, S.M. Observation of quantum entanglement using spatial light modulators. Opt. Express 2006, 14, 13089–13094. [Google Scholar] [CrossRef]
  58. Kong, L.J.; Sun, Y.; Zhang, F.; Zhang, J.; Zhang, X. High-Dimensional Entanglement-Enabled Holography. Phys. Rev. Lett. 2023, 130, 053602. [Google Scholar] [CrossRef]
  59. Maurer, C.; Jesacher, A.; Bernet, S.; Ritsch-Marte, M. What spatial light modulators can do for optical microscopy. Laser Photonics Rev. 2011, 5, 81–101. [Google Scholar] [CrossRef]
  60. Shapiro, J.H. Computational ghost imaging. Phys. Rev. A PRA 2008, 78, 061802. [Google Scholar] [CrossRef]
  61. Moreau, P.A.; Toninelli, E.; Gregory, T.; Padgett, M.J. Ghost Imaging Using Optical Correlations. Laser Photonics Rev. 2018, 12, 1700143. [Google Scholar] [CrossRef]
  62. Padgett, M.; Bowman, R. Tweezers with a twist. Nat. Photonics 2011, 5, 343–348. [Google Scholar] [CrossRef]
  63. Grier, D.G. A revolution in optical manipulation. Nature 2003, 424, 810–816. [Google Scholar] [CrossRef]
  64. Sun, B.; Salter, P.S.; Roider, C.; Jesacher, A.; Strauss, J.; Heberle, J.; Schmidt, M.; Booth, M.J. Four-dimensional light shaping: Manipulating ultrafast spatiotemporal foci in space and time. Light Sci. Appl. 2018, 7, 17117. [Google Scholar] [CrossRef]
  65. Jesacher, A.; Maurer, C.; Schwaighofer, A.; Bernet, S.; Ritsch-Marte, M. Near-perfect hologram reconstruction with a spatial light modulator. Opt. Express 2008, 16, 2597–2603. [Google Scholar] [CrossRef]
  66. Resler, D.P.; Hobbs, D.S.; Sharp, R.C.; Friedman, L.J.; Dorschner, T.A. High-efficiency liquid-crystal optical phased-array beam steering. Opt. Lett. 1996, 21, 689–691. [Google Scholar] [CrossRef]
  67. Stockley, J.; Serati, S. Advances in liquid crystal beam steering. In Proceedings of the Optical Science and Technology, SPIE 49th Annual Meeting, Denver, CO, USA, 2–6 August 2004; SPIE: Bellingham, WA, USA, 2004; Volume 5550. [Google Scholar] [CrossRef]
  68. Yang, Y.; Forbes, A.; Cao, L. A review of liquid crystal spatial light modulators: Devices and applications. Opto-Electron. Sci. 2023, 2, 230026-1–230026-29. [Google Scholar] [CrossRef]
  69. Linnenberger, A.; Serati, S.; Stockley, J. Advances in optical phased array technology. In Free-Space Laser Communications VI; SPIE Optics + Photonics; SPIE: Bellingham, WA, USA, 2006; Volume 6304. [Google Scholar] [CrossRef]
  70. Xu, L.; Zhang, J.; Wu, L.Y. Influence of phase delay profile on diffraction efficiency of liquid crystal optical phased array. Opt. Laser Technol. 2009, 41, 509–516. [Google Scholar] [CrossRef]
  71. He, X.; Wang, X.; Wu, L.; Liu, X.; Guo, H.; Huang, X.; Xie, X.; Tan, Q.; Cao, J. Aperture scalable liquid crystal optically duplicated array of phased array. Opt. Commun. 2019, 451, 174–180. [Google Scholar] [CrossRef]
  72. Zhang, Y.; Wang, C.; Wang, Q.; Mu, Q.; Peng, Z.; Dong, K.; Song, Y.; Liu, Y.; Jiang, H. Dual-wavelength synchronous control method for liquid crystal optical phased array. Opt. Express 2025, 33, 4280–4292. [Google Scholar] [CrossRef] [PubMed]
  73. Notaros, M.; DeSantis, D.M.; Raval, M.; Notaros, J. Liquid-crystal-based visible-light integrated optical phased arrays and application to underwater communications. Opt. Lett. 2023, 48, 5269–5272. [Google Scholar] [CrossRef]
  74. Zhou, S.; Cao, J.; Hao, Q.; Li, W.; Sheng, Y.; Chen, H.; Gao, Z.H.; Ji, P. All-solid-state omnidirectional fast scanning using liquid crystal optical phased array and conical mirror. Opt. Express 2025, 33, 18124–18135. [Google Scholar] [CrossRef]
  75. Kim, J.; Oh, C.; Escuti, M.; Hosting, L.; Serati, S. Wide-angle nonmechanical beam steering using thin liquid crystal polarization gratings. In Advanced Wavefront Control: Methods, Devices, and Applications VI; Optical Engineering + Applications; SPIE: Bellingham, WA, USA, 2008; Volume 7093. [Google Scholar] [CrossRef]
  76. Wenben, X.; Yongmei, H.; Qiongyan, W.; Cai, Z. A wide-angle beam steering system based on Wollaston prisms. Opt. Tech. 2012, 38, 588–592. [Google Scholar] [CrossRef]
  77. Engström, D.; O’Callaghan, M.J.; Walker, C.; Handschy, M.A. Fast beam steering with a ferroelectric-liquid-crystal optical phased array. Appl. Opt. 2009, 48, 1721–1726. [Google Scholar] [CrossRef]
  78. Mengjiao, Q.; Qidong, W.; Quanquan, M.; Yonggang, L.; Lishuang, Y.; Zhaoliang, C.; Hongsheng, Z.; Chengliang, Y.; Lifa, H.; Li, X. Study of Response Time Depending on Driving Voltage of Liquid Crystal Spatial Light Modulator. Laser Optoelectron. Prog. 2013, 50, 178–183. [Google Scholar] [CrossRef]
  79. He, X.; Li, M.; Liang, Z.; Wang, X.; Liu, X.; Wang, Y.; Zhang, L. A liquid crystal stackable phased array to achieve fast and precise nonmechanical laser beam deflection. Opt. Commun. 2022, 506, 127598. [Google Scholar] [CrossRef]
  80. Gu, D.; Wen, B.; Mahajan, M.; Taber, D.; Winker, B.; Guthals, D.; Campbell, B.; Sox, D. High power liquid crystal spatial light modulators. In Advanced Wavefront Control: Methods, Devices, and Applications IV; SPIE Optics + Photonics; SPIE: Bellingham, WA, USA, 2006; Volume 6306. [Google Scholar]
  81. Xiangru, W.; Zhuangqi, Z. Research progress of liquid crystal optical phased array in high power laser applications (Invited). Infrared Laser Eng. 2018, 47, 103006. [Google Scholar] [CrossRef]
  82. Xing, Z.; Fan, W.; Huang, D.; Cheng, H.; Xia, G. High laser damage threshold liquid crystal optical switch based on a gallium nitride transparent electrode. Opt. Lett. 2020, 45, 3537–3540. [Google Scholar] [CrossRef] [PubMed]
  83. Du, T.; Huang, D.; Cheng, H.; Fan, W.; Xing, Z.; Zhu, J.; Liu, W. Research on High Power Laser Damage Resistant Optically Addressable Spatial Light Modulator. Photonics 2022, 9, 811. [Google Scholar] [CrossRef]
  84. Xing, Z.; Fan, W.; Huang, D.; Cheng, H.; Du, T. High laser damage threshold reflective optically addressed liquid crystal light valve based on gallium nitride conductive electrodes. High Power Laser Sci. Eng. 2022, 10, e35. [Google Scholar] [CrossRef]
  85. Liang, Z.; Huang, Y.; Xiaoxian, H.; Wen, D.; Wang, Y.; Wang, X. Four-access, 80 mm aperture all phase-controlled liquid crystal laser antenna. Opt. Eng. 2022, 61, 105113. [Google Scholar] [CrossRef]
  86. Cao, J.; Zhou, S.; Cheng, Y.; Hao, Q. Omnidirectional all-solid-state aliasing-free scanning using liquid crystal optical phased array and conical mirror. Opt. Lasers Eng. 2025, 184, 108664. [Google Scholar] [CrossRef]
  87. Wei, Z.; Wu, J.; Chen, C.; Li, H.; Wan, W.; Chen, X.; Chen, Y. Erbium-doped lithium niobate waveguide amplifier enhanced by an inverse-designed on-chip reflector. Opt. Lett. 2025, 50, 3624–3627. [Google Scholar] [CrossRef]
  88. Kong, Y.; Bo, F.; Wang, W.; Zheng, D.; Liu, H.; Zhang, G.; Rupp, R.; Xu, J. Recent Progress in Lithium Niobate: Optical Damage, Defect Simulation, and On-Chip Devices. Adv. Mater. 2020, 32, 1806452. [Google Scholar] [CrossRef]
  89. Boes, A.; Chang, L.; Langrock, C.; Yu, M.; Zhang, M.; Lin, Q.; Lončar, M.; Fejer, M.; Bowers, J.; Mitchell, A. Lithium niobate photonics: Unlocking the electromagnetic spectrum. Science 2023, 379, eabj4396. [Google Scholar] [CrossRef]
  90. Wang, Z.; Zhang, B.; Wang, Z.; Zhang, J.; Kazansky, P.G.; Tan, D.; Qiu, J. 3D Imprinting of Voxel-Level Structural Colors in Lithium Niobate Crystal. Adv. Mater. 2023, 35, 2303256. [Google Scholar] [CrossRef]
  91. Zhang, J.; Wang, Z.; Zhang, B.; Qiu, J. Single-Pulse-Driven Frame Printing of Chromatic Pixels in Lithium Niobate Crystal. Laser Photonics Rev. 2024, 18, 2400054. [Google Scholar] [CrossRef]
  92. Qi, Y.; Li, Y. Integrated lithium niobate photonics. Nanophotonics 2020, 9, 1287–1320. [Google Scholar] [CrossRef]
  93. Levy, M.; Osgood, R.M., Jr.; Liu, R.; Cross, L.E.; Cargill, G.S., III; Kumar, A.; Bakhru, H. Fabrication of single-crystal lithium niobate films by crystal ion slicing. Appl. Phys. Lett. 1998, 73, 2293–2295. [Google Scholar] [CrossRef]
  94. Lin, J.; Yao, N.; Hao, Z.; Zhang, J.; Mao, W.; Wang, M.; Chu, W.; Wu, R.; Fang, Z.; Qiao, L.; et al. Broadband Quasi-Phase-Matched Harmonic Generation in an On-Chip Monocrystalline Lithium Niobate Microdisk Resonator. Phys. Rev. Lett. 2019, 122, 173903. [Google Scholar] [CrossRef] [PubMed]
  95. Kharel, P.; Reimer, C.; Luke, K.; He, L.; Zhang, M. Breaking voltage–bandwidth limits in integrated lithium niobate modulators using micro-structured electrodes. Optica 2021, 8, 357–363. [Google Scholar] [CrossRef]
  96. Wang, C. Breaking anisotropy limitations in thin-film lithium niobate arrayed waveguide gratings. Light Sci. Appl. 2024, 13, 209. [Google Scholar] [CrossRef]
  97. Snigirev, V.; Riedhauser, A.; Lihachev, G.; Churaev, M.; Riemensberger, J.; Wang, R.N.; Siddharth, A.; Huang, G.; Möhl, C.; Popoff, Y.; et al. Ultrafast tunable lasers using lithium niobate integrated photonics. Nature 2023, 615, 411–417. [Google Scholar] [CrossRef]
  98. Poberaj, G.; Hu, H.; Sohler, W.; Günter, P. Lithium niobate on insulator (LNOI) for micro-photonic devices. Laser Photonics Rev. 2012, 6, 488–503. [Google Scholar] [CrossRef]
  99. Liu, L.; Vorontsov, M.A. Phase-locking of tiled fiber array using SPGD feedback controller. SPIE Proc. 2005, 5895, 138–146. [Google Scholar] [CrossRef]
  100. Shay, T.M.; Benham, V.; Baker, J.T.; Ward, B.; Sanchez, A.D.; Culpepper, M.A.; Pilkington, D.; Spring, J.; Nelson, D.J.; Lu, C.A. First experimental demonstration of self-synchronous phase locking of an optical array. Opt. Express 2006, 14, 12015–12021. [Google Scholar] [CrossRef] [PubMed]
  101. Xiao, R.; Hou, J.; Liu, M.; Jiang, Z.F. Coherent combining technology of master oscillator power amplifier fiber arrays. Opt. Express 2008, 16, 2015–2022. [Google Scholar] [CrossRef]
  102. Chang, H.; Chang, Q.; Xi, J.; Hou, T.; Su, R.; Ma, P.; Wu, J.; Li, C.; Jiang, M.; Ma, Y.; et al. First experimental demonstration of coherent beam combining of more than 100 beams. Photonics Res. 2020, 8, 1943–1948. [Google Scholar] [CrossRef]
  103. Tien, P.K.; Riva-Sanseverino, S.; Ballman, A.A. Light beam scanning and deflection in epitaxial LiNbO3 electro-optic waveguides. Appl. Phys. Lett. 1974, 25, 563–565. [Google Scholar] [CrossRef]
  104. Yue, G.; Li, Y. Integrated lithium niobate optical phased array for two-dimensional beam steering. Opt. Lett. 2023, 48, 3633–3636. [Google Scholar] [CrossRef]
  105. Wang, Z.; Song, W.; Chen, Y.; Fang, B.; Ji, J.; Xin, H.; Zhu, S.; Li, T. Metasurface empowered lithium niobate optical phased array with an enlarged field of view. Photonics Res. 2022, 10, B23. [Google Scholar] [CrossRef]
  106. Li, W.; Zhao, X.; He, J.; Yan, H.; Pan, B.; Guo, Z.; Han, X.; Chen, J.; Dai, D.; Shi, Y. High-speed 2D beam steering based on a thin-film lithium niobate optical phased array with a large field of view. Photonics Res. 2023, 11, 1912–1918. [Google Scholar] [CrossRef]
  107. Li, Y.; Chen, H.; Liu, R.; Deng, S.; Zhu, J.; Hu, Y.; Yang, T.; Guan, H.; Lu, H. Design of optical phased array with low-sidelobe beam steering in thin film lithium niobate. Opt. Laser Technol. 2024, 171, 110432. [Google Scholar] [CrossRef]
  108. Wang, Z.; Li, X.; Ji, J.; Sun, Z.; Sun, J.; Fang, B.; Lu, J.; Li, S.; Ma, X.; Chen, X.; et al. Fast-speed and low-power-consumption optical phased array based on lithium niobate waveguides. Nanophotonics 2024, 13, 2429–2436. [Google Scholar] [CrossRef]
  109. Li, J.; Zheng, H.; He, Y.; Liu, Y.; Wei, X.; Xu, Z. Cascaded domain engineering optical phased array for 2D beam steering. Nanophotonics 2023, 12, 4017–4030. [Google Scholar] [CrossRef]
  110. Li, J.; Zheng, H.; Han, Y.; Li, B.; Hao, W.; Qiu, L.; Liu, Y.; He, Y.; Wei, X.; Xu, Z. Design and analysis of single-electrode integrated lithium niobate optical phased array for two-dimensional beam steering. Opt. Lasers Eng. 2025, 184, 108617. [Google Scholar] [CrossRef]
  111. Li, J.; He, Y.; Zheng, H.; Luo, S.; Liu, X.; Hu, Q.; Chen, H.; Liang, W.; Liu, J.; Chen, H.; et al. Cascaded domain engineering optical phased array for beam steering. Appl. Phys. Rev. 2023, 10, 031413. [Google Scholar] [CrossRef]
  112. Chen, J.; Li, W.; Kang, Z.; Lin, Z.; Zhao, S.; Lian, D.; He, J.; Huang, D.; Dai, D.; Shi, Y. Single soliton microcomb combined with optical phased array for parallel FMCW LiDAR. Nat. Commun. 2025, 16, 1056. [Google Scholar] [CrossRef] [PubMed]
  113. Sun, Z.; Deng, X.; Yu, J.; Wu, C.; Chen, H.; Zhu, W.; Ma, W.; Chen, Z.; Pan, W.; Zou, X.; et al. 50Gbps Mobile Optical Wireless Communication Enabled by Lithium-Niobate-Based Beam Steering. J. Light. Technol. 2025, 43, 8016–8028. [Google Scholar] [CrossRef]
  114. Liu, J.; Wang, B.; Wang, J.; Zhang, L.; Gan, X.; Qiao, D.; Li, Y. Taylor Amplitude Distribution for Low Sidelobe in Thin-Film Lithium Niobate Optical Phased Array. J. Light. Technol. 2025, 43, 8286–8292. [Google Scholar] [CrossRef]
  115. Zhao, S.; Chen, J.; Shi, Y. Dual Polarization and Bi-Directional Silicon-Photonic Optical Phased Array With Large Scanning Range. IEEE Photonics J. 2022, 14, 6620905. [Google Scholar] [CrossRef]
  116. Sun, J.; Timurdogan, E.; Yaacobi, A.; Hosseini, E.S.; Watts, M.R. Large-scale nanophotonic phased array. Nature 2013, 493, 195–199. [Google Scholar] [CrossRef]
  117. Poulton, C.V.; Yaacobi, A.; Cole, D.B.; Byrd, M.J.; Raval, M.; Vermeulen, D.; Watts, M.R. Coherent solid-state LIDAR with silicon photonic optical phased arrays. Opt. Lett. 2017, 42, 4091–4094. [Google Scholar] [CrossRef]
  118. Hodgson, M.E.; Jensen, J.R.; Schmidt, L.; Schill, S.; Davis, B. An evaluation of LIDAR- and IFSAR-derived digital elevation models in leaf-on conditions with USGS Level 1 and Level 2 DEMs. Remote Sens. Environ. 2003, 84, 295–308. [Google Scholar] [CrossRef]
  119. Poulton, C.V.; Byrd, M.J.; Russo, P.; Moss, B.; Shatrovoy, O.; Khandaker, M.; Watts, M.R. Coherent LiDAR With an 8,192-Element Optical Phased Array and Driving Laser. IEEE J. Sel. Top. Quantum Electron. 2022, 28, 6100508. [Google Scholar] [CrossRef]
  120. Wang, P.; Luo, G.; Xu, Y.; Li, Y.; Su, Y.; Ma, J.; Wang, R.; Yang, Z.; Zhou, X.; Zhang, Y.; et al. Design and fabrication of a SiN-Si dual-layer optical phased array chip. Photonics Res. 2020, 8, 912–919. [Google Scholar] [CrossRef]
  121. López, J.J.; Skirlo, S.A.; Kharas, D.; Sloan, J.; Sorace-Agaskar, C. Planar-lens Enabled Beam Steering for Chip-scale LIDAR. In Proceedings of the CLEO: Science and Innovations, San Jose, CA, USA, 13–18 May 2018. [Google Scholar] [CrossRef]
  122. Li, C.; Cao, X.; Wu, K.; Li, X.; Chen, J. A Switch-based Integrated 2D Beam-steering device for Lidar Application. In Proceedings of the 2019 Conference on Lasers and Electro-Optics (CLEO), San Jose, CA, USA, 5–10 May 2019; IEEE: Piscataway, NJ, USA, 2019. [Google Scholar] [CrossRef]
  123. Yu, L.; Wang, P.; Ma, P.; Cui, L.; Wang, Z.; Yang, Y.; Zhang, Y.; Pan, J. Two-Dimensional Beam Scanning of Passive Optical Phased Array Based on Silicon Nitride Delay Line. J. Light. Technol. 2023, 41, 2756–2764. [Google Scholar] [CrossRef]
  124. Chen, J.; Zhao, S.; Li, W.; Wang, X.; Han, X.; Shi, Y. Silicon Optical Phased Array Hybrid Integrated with III–V Laser for Grating Lobe-Free Beam Steering. Photonics 2024, 11, 952. [Google Scholar] [CrossRef]
  125. Van Acoleyen, K.; Bogaerts, W.; Jágerská, J.; Le Thomas, N.; Houdré, R.; Baets, R. Off-chip beam steering with a one-dimensional optical phased array on silicon-on-insulator. Opt. Lett. 2009, 34, 1477–1479. [Google Scholar] [CrossRef]
  126. Van Acoleyen, K.; Komorowska, K.; Bogaerts, W.; Baets, R. One-Dimensional Off-Chip Beam Steering and Shaping Using Optical Phased Arrays on Silicon-on-Insulator. J. Light. Technol. 2011, 29, 3500–3505. [Google Scholar] [CrossRef]
  127. Van Acoleyen, K.; Rogier, H.; Baets, R. Two-dimensional optical phased array antenna on silicon-on-insulator. Opt. Express 2010, 18, 13655–13660. [Google Scholar] [CrossRef]
  128. Doylend, J.K.; Heck, M.J.R.; Bovington, J.T.; Peters, J.D.; Coldren, L.A.; Bowers, J.E. Two-dimensional free-space beam steering with an optical phased array on silicon-on-insulator. Opt. Express 2011, 19, 21595–21604. [Google Scholar] [CrossRef]
  129. Abediasl, H.; Hashemi, H. Monolithic optical phased-array transceiver in a standard SOI CMOS process. Opt. Express 2015, 23, 6509–6519. [Google Scholar] [CrossRef] [PubMed]
  130. Hutchison, D.N.; Sun, J.; Doylend, J.K.; Kumar, R.; Heck, J.; Kim, W.; Phare, C.T.; Feshali, A.; Rong, H.S. High-resolution aliasing-free optical beam steering. Optica 2016, 3, 887–890. [Google Scholar] [CrossRef]
  131. Poulton, C.V.; Byrd, M.J.; Raval, M.; Su, Z.; Li, N.X.; Timurdogan, E.; Coolbaugh, D.; Vermeulen, D.; Watts, M.R. Large-scale silicon nitride nanophotonic phased arrays at infrared and visible wavelengths. Opt. Lett. 2017, 42, 21–24. [Google Scholar] [CrossRef]
  132. Poulton, C.V.; Byrd, M.J.; Russo, P.; Timurdogan, E.; Khandaker, M.; Vermeulen, D.; Watts, M.R. Long-Range LiDAR and Free-Space Data Communication With High-Performance Optical Phased Arrays. IEEE J. Sel. Top. Quantum Electron. 2019, 25, 7700108. [Google Scholar] [CrossRef]
  133. Miller, S.A.; Chang, Y.C.; Phare, C.T.; Shin, M.C.; Zadka, M.; Roberts, S.P.; Stern, B.; Ji, X.; Mohanty, A.; Gordillo, O.A.J.; et al. Large-scale optical phased array using a low-power multi-pass silicon photonic platform. Optica 2020, 7, 3–6. [Google Scholar] [CrossRef]
  134. Marinins, A.; Dwivedi, S.P.; Kjellman, J.Ø.; Kerman, S.; David, T.; Figeys, B.; Jansen, R.; Tezcan, D.S.; Rottenberg, X.; Soussan, P. Silicon photonics co-integrated with silicon nitride for optical phased arrays. Jpn. J. Appl. Phys. 2020, 59, SGGE02. [Google Scholar] [CrossRef]
  135. Wang, Q.; Wang, S.; Jia, L.; Cai, Y.; Yue, W.; Yu, M. Silicon nitride assisted 1×64 optical phased array based on a SOI platform. Opt. Express 2021, 29, 10509–10517. [Google Scholar] [CrossRef] [PubMed]
  136. Xu, W.; Guo, Y.; Li, X.; Liu, C.; Lu, L.; Chen, J.; Zhou, L. Fully Integrated Solid-State LiDAR Transmitter on a Multi-Layer Silicon-Nitride-on-Silicon Photonic Platform. J. Light. Technol. 2023, 41, 832–840. [Google Scholar] [CrossRef]
  137. Li, Y.; Wang, Z.; Du, H.; Chen, B.; Song, J.; Tao, M. Integrated communication and sensing system based on Si-SiN dual-layer optical phased array. Opt. Express 2024, 32, 33222–33231. [Google Scholar] [CrossRef] [PubMed]
  138. Wang, J.; Hu, H.; Yang, D.; Wang, W.; Wang, S.; Wang, C.; Wang, Q.; Yue, W.; Cai, Y. Aperiodic optical phased array with wide field of view and high sidelobe suppression ratio based on SiN-on-SOI platform. Opt. Express 2025, 33, 4499–4509. [Google Scholar] [CrossRef] [PubMed]
  139. Li, Y.; Chen, B.; Na, Q.; Xie, Q.; Tao, M.; Zhang, L.; Zhi, Z.; Li, Y.; Liu, X.; Luo, X.; et al. Wide-steering-angle high-resolution optical phased array. Photonics Res. 2021, 9, 2511–2518. [Google Scholar] [CrossRef]
  140. Zhang, Z.; Yu, H.; Chen, B.; Xia, P.; Zhou, Z.; Huang, Q.; Dai, T.; Wang, Y.; Yang, J. A Tri-Layer SiN-on-Si Optical Phased Array With High Angular Resolution. Photonics Technol. Lett. IEEE 2022, 34, 4. [Google Scholar] [CrossRef]
  141. Wang, Z.; Yu, L.; Yang, Y.; Ma, P.; Cui, L.; Luo, S.; Ji, Z.; Song, Z.; Su, Y.; Pan, J.; et al. Wide field of view optical phased array with a high-directionality antenna. Opt. Express 2023, 31, 21192–21199. [Google Scholar] [CrossRef]
  142. Sacher, W.D.; Huang, Y.; Lo, G.Q.; Poon, J.K.S. Multilayer Silicon Nitride-on-Silicon Integrated Photonic Platforms and Devices. J. Light. Technol. 2015, 33, 901–910. [Google Scholar] [CrossRef]
  143. Sacher, W.D.; Mikkelsen, J.C.; Dumais, P.; Jiang, J.; Poon, J.K.S. Tri-layer silicon nitride-on-silicon photonic platform for ultra-low-loss crossings and interlayer transitions. Opt. Express 2017, 25, 30862–30875. [Google Scholar] [CrossRef] [PubMed]
  144. Li, X.; Gao, W.; Lu, L.; Chen, J.; Zhou, L. Ultra-low-loss multi-layer 8 × 8 microring optical switch. Photonics Res. 2023, 11, 712–723. [Google Scholar] [CrossRef]
  145. Lu, L.; Xu, W.; Guo, Y.; Liu, C.; Chen, J.; Zhou, L. Large-Scale Optical Phased Array Based on a Multi-Layer Silicon-Nitride-on-Silicon Photonic Platform. In Proceedings of the 2024 Optical Fiber Communications Conference and Exhibition (OFC), San Diego, VA, USA, 24–28 March 2024; pp. 1–3. [Google Scholar]
  146. Zhang, L.; Li, Y.; Hou, Y.; Wang, Y.; Tao, M.; Chen, B.; Na, Q.; Li, Y.; Zhi, Z.; Liu, X.; et al. Investigation and demonstration of a high-power handling and large-range steering optical phased array chip. Opt. Express 2021, 29, 29755–29765. [Google Scholar] [CrossRef] [PubMed]
  147. Vasey, F.; Reinhart, F.K.; Houdré, R.; Stauffer, J.M. Spatial optical beam steering with an AlGaAs integrated phased array. Appl. Opt. 1993, 32, 3220–3232. [Google Scholar] [CrossRef] [PubMed]
  148. Dong, H.Z.; Shi, S.X.; Li, J.L. Novel unitary optical mode converter. Opt. Laser Technol. 2007, 39, 282–284. [Google Scholar] [CrossRef]
  149. Li, J.; Shi, S.; Liu, J.; Sun, Y.; Dong, H.; Liang, H. Electromagnetism theory of waveguide array electro-optical scanner: Model and characteristics of optical field distribution in the waveguide array. J. Appl. Phys. 2006, 100, 023105. [Google Scholar] [CrossRef]
  150. Nagarajan, R.; Kato, M.; Pleumeekers, J.; Evans, P.; Corzine, S.; Hurtt, S.; Dentai, A.; Murthy, S.; Missey, M.; Muthiah, R.; et al. InP Photonic Integrated Circuits. IEEE J. Sel. Top. Quantum Electron. 2010, 16, 1113–1125. [Google Scholar] [CrossRef]
  151. Guo, W.; Binetti, P.R.A.; Althouse, C.; Mašanović, M.L.; Ambrosius, H.P.M.M.; Johansson, L.A.; Coldren, L.A. Two-Dimensional Optical Beam Steering With InP-Based Photonic Integrated Circuits. IEEE J. Sel. Top. Quantum Electron. 2013, 19, 6100212. [Google Scholar] [CrossRef]
  152. Xie, W.; Komljenovic, T.; Huang, J.; Bowers, J. Dense III-V/Si Phase Shifters for Optical Phased Arrays. In Proceedings of the 2018 Asia Communications and Photonics Conference (ACP), Hangzhou, China, 26–29 October 2018; pp. 1–3. [Google Scholar] [CrossRef]
  153. Xie, W.; Bowers, J. High-performance III-V/Si Phase Shifter Arrays for Beam Steering. In Proceedings of the 2019 IEEE Photonics Society Summer Topical Meeting Series (SUM), Ft. Lauderdale, FL, USA, 8–10 July 2019; pp. 1–2. [Google Scholar] [CrossRef]
  154. Xie, W.; Komljenovic, T.; Huang, J.; Tran, M.; Davenport, M.; Torres, A.; Pintus, P.; Bowers, J. Heterogeneous silicon photonics sensing for autonomous cars [Invited]. Opt. Express 2019, 27, 3642–3663. [Google Scholar] [CrossRef]
  155. Hulme, J.C.; Doylend, J.K.; Heck, M.J.R.; Peters, J.D.; Davenport, M.L.; Bovington, J.T.; Coldren, L.A.; Bowers, J.E. Fully integrated hybrid silicon two dimensional beam scanner. Opt. Express 2015, 23, 5861–5874. [Google Scholar] [CrossRef] [PubMed]
  156. Haines, J.; Vitali, V.; Bottrill, K.; Naik, P.U.; Gandolfi, M.; Angelis, C.D.; Franz, Y.; Lacava, C.; Petropoulos, P.; Guasoni, M. Fabrication of 1 × N integrated power splitters with arbitrary power ratio for single and multimode photonics. Nanophotonics 2024, 13, 339–348. [Google Scholar] [CrossRef] [PubMed]
  157. Franz, Y.; Guasoni, M. Compact 1 × N power splitters with arbitrary power ratio for integrated multimode photonics. J. Opt. 2021, 23, 095802. [Google Scholar] [CrossRef]
  158. Fang, Y.; Qinggui, T.; Guangyao, W.; Rui, Y.; Wei, H. Fast and large-angle optical beam deflection based on liquid crystal polarization grating. Chin. J. Liq. Cryst. Displays 2022, 37, 1411–1419. [Google Scholar] [CrossRef]
  159. Chen, B.S.; Li, Y.Z.; Xie, Q.J.; Na, Q.X.; Tao, M.; Wang, Z.M.; Zhi, Z.H.; Hu, H.M.; Li, X.T.; Qu, H.; et al. SiN-On-SOI Optical Phased Array LiDAR for Ultra-Wide Field of View and 4D Sensing. Laser Photonics Rev. 2024, 18. [Google Scholar] [CrossRef]
Figure 1. Categories of OPA and related applications.(a) LC-OPA for holographic imaging, light field control, and satellite communications; (b) LN-OPA for high-power lasers and 3D LiDAR; (c) Si-OPA for solid-state LiDAR, free-space optical communication, and light field control.
Figure 1. Categories of OPA and related applications.(a) LC-OPA for holographic imaging, light field control, and satellite communications; (b) LN-OPA for high-power lasers and 3D LiDAR; (c) Si-OPA for solid-state LiDAR, free-space optical communication, and light field control.
Nanomaterials 15 01374 g001
Figure 2. (a) Diagram of array beam and its far-field diffraction pattern; (b) Multi-beam formation due to element spacing d 1 > 0.5 λ with annotated steering range; (c) Single-beam regime d 2 < 0.5 λ : Beamwidth (FMHW) for baseline array size; (d) Beamwidth narrowing with increased element count at fixed d 2 < 0.5 λ .
Figure 2. (a) Diagram of array beam and its far-field diffraction pattern; (b) Multi-beam formation due to element spacing d 1 > 0.5 λ with annotated steering range; (c) Single-beam regime d 2 < 0.5 λ : Beamwidth (FMHW) for baseline array size; (d) Beamwidth narrowing with increased element count at fixed d 2 < 0.5 λ .
Nanomaterials 15 01374 g002
Figure 3. Advances in materials and manufacturing technology have driven significant LC-OPA performance improvements, achieving micrometer-scale pixel pitch with tens of millions of pixels. Reproduced from ref. [68]. Opto-Electronic Journals Group, 2023.
Figure 3. Advances in materials and manufacturing technology have driven significant LC-OPA performance improvements, achieving micrometer-scale pixel pitch with tens of millions of pixels. Reproduced from ref. [68]. Opto-Electronic Journals Group, 2023.
Nanomaterials 15 01374 g003
Figure 4. (a) Structure profile of LC-OPA. (b) Schematic diagram of the phase modulation principle of LC-OPA. (c) Photograph of the on-chip LC OPA and the beam steering result. (d) Schematic of omnidirectional steering. (a) Reproduced from ref. [70]. Copyright©2008 Elsevier Ltd. (b) Reproduced from ref. [72]. Optica Publishing Group, 2025. (c) Reproduced from ref. [73]. Optica Publishing Group, 2025. (d) Reproduced from ref. [74]. Optica Publishing Group, 2025.
Figure 4. (a) Structure profile of LC-OPA. (b) Schematic diagram of the phase modulation principle of LC-OPA. (c) Photograph of the on-chip LC OPA and the beam steering result. (d) Schematic of omnidirectional steering. (a) Reproduced from ref. [70]. Copyright©2008 Elsevier Ltd. (b) Reproduced from ref. [72]. Optica Publishing Group, 2025. (c) Reproduced from ref. [73]. Optica Publishing Group, 2025. (d) Reproduced from ref. [74]. Optica Publishing Group, 2025.
Nanomaterials 15 01374 g004
Figure 5. (a) SLM structure illustration; (b) Temperature of sapphire on both sides of the SLM device as a function of laser power density and the changes in contrast ratio with power intensity; (c) The basic structure of the reflective SLM; (d) Temperature of sapphire on both sides of the SLM device as a function of laser power density and the changes in contrast ratio with power intensity. (a,b) Reproduced from ref. [83]. MDPI, 2022. (c,d) Reproduced from ref. [84]. Cambridge University Press, 2022.
Figure 5. (a) SLM structure illustration; (b) Temperature of sapphire on both sides of the SLM device as a function of laser power density and the changes in contrast ratio with power intensity; (c) The basic structure of the reflective SLM; (d) Temperature of sapphire on both sides of the SLM device as a function of laser power density and the changes in contrast ratio with power intensity. (a,b) Reproduced from ref. [83]. MDPI, 2022. (c,d) Reproduced from ref. [84]. Cambridge University Press, 2022.
Nanomaterials 15 01374 g005
Figure 6. (a) Beam combining diagram (PM: phase modulators). (b) The electronics based on LN for a 2-element array; (c) A computational ghost imaging system using LN phase modulators; (d) A High power laser system using LN phase modulators. (a,b) Reproduced from ref. [100]. Optica Publishing Group, 2006. (c) Reproduced from ref. [51]. Optica Publishing Group, 2018. (d) Reproduced from ref. [102]. Optica Publishing Group, 2020.
Figure 6. (a) Beam combining diagram (PM: phase modulators). (b) The electronics based on LN for a 2-element array; (c) A computational ghost imaging system using LN phase modulators; (d) A High power laser system using LN phase modulators. (a,b) Reproduced from ref. [100]. Optica Publishing Group, 2006. (c) Reproduced from ref. [51]. Optica Publishing Group, 2018. (d) Reproduced from ref. [102]. Optica Publishing Group, 2020.
Nanomaterials 15 01374 g006
Figure 7. (a) Design of the OPA based on LN platform. (b) Angle dependent optical power in θ y direction (±21.27° FOV). (c) Image of the fabricated chip and electrical and optical packaged chip; (d) Measured optical power distribution versus angle in the phase tuning direction. (a,b) Reproduced from ref. [105]. Optica Publishing Group, 2022. (c,d) Reproduced from ref. [106]. Optica Publishing Group, 2023.
Figure 7. (a) Design of the OPA based on LN platform. (b) Angle dependent optical power in θ y direction (±21.27° FOV). (c) Image of the fabricated chip and electrical and optical packaged chip; (d) Measured optical power distribution versus angle in the phase tuning direction. (a,b) Reproduced from ref. [105]. Optica Publishing Group, 2022. (c,d) Reproduced from ref. [106]. Optica Publishing Group, 2023.
Nanomaterials 15 01374 g007
Figure 8. (a) Schematic illustration of the OPA chip employing LNOI for two-dimensional beam steering. (b) Scanning electron microscopy image of the 32-channel lithium niobate waveguide array. (c) Schematic illustration of the four-layer cascaded domain-engineered LNOI-based OPA. (a,b) Reproduced from ref. [108]. De Gruyter Brill, 2024. (c) Reproduced from ref. [110]. Copyright©2025 Elsevier Ltd.
Figure 8. (a) Schematic illustration of the OPA chip employing LNOI for two-dimensional beam steering. (b) Scanning electron microscopy image of the 32-channel lithium niobate waveguide array. (c) Schematic illustration of the four-layer cascaded domain-engineered LNOI-based OPA. (a,b) Reproduced from ref. [108]. De Gruyter Brill, 2024. (c) Reproduced from ref. [110]. Copyright©2025 Elsevier Ltd.
Nanomaterials 15 01374 g008
Figure 9. Conceptual illustration of LiDAR system based on integrated TFLN OPA. Reproduced from ref. [112]. Springer Nature Limited, 2025.
Figure 9. Conceptual illustration of LiDAR system based on integrated TFLN OPA. Reproduced from ref. [112]. Springer Nature Limited, 2025.
Nanomaterials 15 01374 g009
Figure 10. Si materials for OPA. (a) SOI. (b) S i N . (c) Hybrid III/V silicon. (a) Reproduced from ref. [115]. IEEE, 2022. (b) Reproduced from ref. [120]. Optica Publishing Group, 2020. (c) Reproduced from ref. [124]. MDPI, 2024.
Figure 10. Si materials for OPA. (a) SOI. (b) S i N . (c) Hybrid III/V silicon. (a) Reproduced from ref. [115]. IEEE, 2022. (b) Reproduced from ref. [120]. Optica Publishing Group, 2020. (c) Reproduced from ref. [124]. MDPI, 2024.
Nanomaterials 15 01374 g010
Figure 11. (a) A 16-element OPA with independently phase tuned for 2D beam steering. (b) A large-scale 64-element 2D OPA. (c) A 1D OPA with aperiodic high-resolution. (d) Photo of a 512-element OPA. (a) Reproduced from ref. [128]. Optica Publishing Group, 2011. (b) Reproduced from ref. [116]. Copyright©2013 Springer Nature Limited. (c) Reproduced from ref. [130]. Optica Publishing Group, 2016. (d) Reproduced from ref. [133]. Optica Publishing Group, 2020.
Figure 11. (a) A 16-element OPA with independently phase tuned for 2D beam steering. (b) A large-scale 64-element 2D OPA. (c) A 1D OPA with aperiodic high-resolution. (d) Photo of a 512-element OPA. (a) Reproduced from ref. [128]. Optica Publishing Group, 2011. (b) Reproduced from ref. [116]. Copyright©2013 Springer Nature Limited. (c) Reproduced from ref. [130]. Optica Publishing Group, 2016. (d) Reproduced from ref. [133]. Optica Publishing Group, 2020.
Nanomaterials 15 01374 g011
Figure 12. Schematic illustration of a Si-SiN OPA for (a) optical communication; (b) FMCW LiDAR. (a) Reproduced from ref. [137]. Optica Publishing Group, 2024. (b) Reproduced from ref. [138]. Optica Publishing Group, 2025.
Figure 12. Schematic illustration of a Si-SiN OPA for (a) optical communication; (b) FMCW LiDAR. (a) Reproduced from ref. [137]. Optica Publishing Group, 2024. (b) Reproduced from ref. [138]. Optica Publishing Group, 2025.
Nanomaterials 15 01374 g012
Figure 13. (a) Optical microscope picture of Si-SiN OPA chip; (b) OPA chip packaging. (a,b) Reproduced from ref. [146]. Optica Publishing Group, 2021.
Figure 13. (a) Optical microscope picture of Si-SiN OPA chip; (b) OPA chip packaging. (a,b) Reproduced from ref. [146]. Optica Publishing Group, 2021.
Nanomaterials 15 01374 g013
Table 1. Performance comparison of the LC-OPA.
Table 1. Performance comparison of the LC-OPA.
PlatformScaleFOV
θ x (°) × θ y (°)
Response Time (ms)Angular Resolution θ x (°) × θ y (°)Wavelength (nm)Diffractive Efficiency (%)YearRef.
LC1 × 409610 × –10015502004[67]
LC1 × 12,2886.95 × –24.815502006[69]
LC256 × 2564 × –633752008[70]
LC24 × 2417.20.001 × 0.00115502022[85]
LC1900 × 1200360 × 1.10.2 × 0.03632.880–912025[86]
LC1900 × 1200360 × 2.10.04 × –632.880–912025[74]
Table 2. The performance comparison of the demonstrated (TF) LN-OPA.
Table 2. The performance comparison of the demonstrated (TF) LN-OPA.
PlatformScaleFOV
θ x (°) × θ y (°)
Response Time (ns)Angular Resolution
θ x (°) × θ y (°)
Wavelength (nm)Bandwidth (GHz)YearRef.
LN70.24 × –250.03 × –980102014[13]
LN640.57 × 0.570.07 × 0.0715500.42021[14]
LN21.22 × –13761550402025[113]
TFLN1624 × 80.232 × 0.61550–16004.22023[104]
TFLN1650 × 8.60.40.73 × 2.815502.52023[106]
TFLN3262.2 × 8.814.42.4 × 1.215502024[108]
TFLN4840 × 8.814.40.33 × 1.815502024[108]
TFLN1662 × 7.6263.2 × 1.415502025[114]
TFLN12880 × 50.01 × 0.037155032025[112]
Table 3. Comparison of the demonstrated Si-OPA.
Table 3. Comparison of the demonstrated Si-OPA.
PlatformScaleFOV θ x (°) × θ y (°)Response Time (μs)Angular Resolution θ x (°) × θ y (°)Wavelength (nm)YearRef.
Si1614.1 × –2.7 × 2.515502009[125]
Si409615502013[116]
Si8192100 × –300.01 × 0.0392022[119]
SiN1024146 × –0.064 × 0.0746352017[131]
SiN6435.5 × 22.70.69 × 0.07515502021[135]
SiN128150 × 160.022 × 0.06015502025[138]
Si-SiN25616.1 × 45.60.044 × 0.15415502022[140]
Si-SiN25614.8 × 1500.068 × 0.06615502023[141]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, X.; Guo, J.; Yang, Z.; Zhang, Y.; Leng, J.; Yu, Q.; Wu, J. On-Chip OPA: Progress and Prospects in Liquid Crystal, Lithium Niobate, and Silicon Material Platforms. Nanomaterials 2025, 15, 1374. https://doi.org/10.3390/nano15171374

AMA Style

Wang X, Guo J, Yang Z, Zhang Y, Leng J, Yu Q, Wu J. On-Chip OPA: Progress and Prospects in Liquid Crystal, Lithium Niobate, and Silicon Material Platforms. Nanomaterials. 2025; 15(17):1374. https://doi.org/10.3390/nano15171374

Chicago/Turabian Style

Wang, Xiaobin, Junliang Guo, Zixin Yang, Yuqiu Zhang, Jinyong Leng, Qiang Yu, and Jian Wu. 2025. "On-Chip OPA: Progress and Prospects in Liquid Crystal, Lithium Niobate, and Silicon Material Platforms" Nanomaterials 15, no. 17: 1374. https://doi.org/10.3390/nano15171374

APA Style

Wang, X., Guo, J., Yang, Z., Zhang, Y., Leng, J., Yu, Q., & Wu, J. (2025). On-Chip OPA: Progress and Prospects in Liquid Crystal, Lithium Niobate, and Silicon Material Platforms. Nanomaterials, 15(17), 1374. https://doi.org/10.3390/nano15171374

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop