Next Article in Journal
Optimized Multi-Antenna MRC for 16-QAM Transmission in a Photonics-Aided Millimeter-Wave System
Previous Article in Journal
The Design and Validation of an Intensity-Modulated Multipoint Fiber-Optic Liquid-Level Sensor
Previous Article in Special Issue
High-Density Arrayed Spectrometer with Microlens Array Grating for Multi-Channel Parallel Spectral Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review on Recent Advances in Signal Processing in Interferometry

Tsinghua Shenzhen International Graduate School, Tsinghua University, Shenzhen 518055, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Sensors 2025, 25(16), 5013; https://doi.org/10.3390/s25165013
Submission received: 1 July 2025 / Revised: 7 August 2025 / Accepted: 8 August 2025 / Published: 13 August 2025

Abstract

Optical interferometry provides high-precision displacement and angle measurement solutions for a wide range of cutting-edge industrial applications. One of the key factors to achieve such precision lies in highly accurate optical encoder signal processing, as well as the calibration and compensation techniques customized for specific measurement principles. Optical interferometric techniques, including laser interferometry and grating interferometry, are usually classified into homodyne and heterodyne systems according to their working principles. In homodyne interferometry, the displacement is determined by analyzing the phase variation of amplitude-modulated signals, and common demodulation methods include error calibration methods and ellipse parameter estimation methods. Heterodyne interferometry obtains displacement information through the phase variation of beat-frequency signals generated by the interference of two light beams with shifted frequencies, and its demodulation techniques include pulse-counting methods, quadrature phase-locked methods, and Kalman filtering. This paper comprehensively reviews the widely used signal processing techniques in optical interferometric measurements over the past two decades and conducts a comparative analysis based on the characteristics of different methods to highlight their respective advantages and limitations. Finally, the hardware platforms commonly used for optical interference signal processing are introduced.

1. Introduction

As a core technology in modern industrial production and scientific research, displacement measurement has consistently played a vital role [1,2,3,4]. From rudimentary mechanical tools to today’s high-precision measurement systems that integrate optics, electronics, and other interdisciplinary innovations, demands on measurement accuracy have evolved from the micrometer scale to the nanometer, sub-nanometer, and even picometer levels [5,6,7]. Driven by advances in sensor technologies and digital manufacturing [8], displacement measurement methods continue to undergo innovation and breakthroughs, offering critical support to a broad range of sectors, including optics [9,10,11,12,13,14,15], biomedical engineering [16,17,18], electronics [19], and manufacturing industry [20,21].
In mechanical manufacturing, precise displacement measurement underpins the machining and assembly of high-precision components, directly influencing product quality and performance [22,23]. For instance, in computer numerical control machining, the characterization of surface parameters, the accurate positioning of cutting tools, and defect detection all depend on continuous improvements in measurement precision [24,25,26]. In the electronics industry, particularly in semiconductor fabrication, displacement measurement in lithography processes must reach nanometer-level accuracy, which is essential to ensuring high integration density and performance of microchips. Specifically, the wafer stage in lithography systems, which is responsible for ultra-precise and high-speed motion, must achieve extremely accurate positioning and motion control to meet overlay and throughput requirements [27,28]. For example, ASML’s TWINSCAN NXE: 3600D EUV lithography system achieves a resolution of 13 nm and an overlay accuracy of 1.1 nm, enabling volume production at 5 nm and 3 nm technology nodes, with sub-nanometer positioning precision for both the wafer and mask stages [29].
With ongoing technological advancements, the requirements for measurement precision, speed, and reliability are becoming increasingly stringent. A variety of displacement measurement methods have emerged and continue to evolve. While traditional contact-based mechanical approaches remain in use [30], non-contact measurement techniques are gaining traction in high-precision manufacturing. These include capacitive sensing [31,32,33], eddy current sensing [34,35,36], ultrasonic methods [37,38], optical triangulation [39,40,41], confocal measurement [42,43,44,45,46,47,48], laser interferometry [49,50,51,52,53], and grating interferometry [54,55,56,57,58,59]. Among them, laser interferometry and grating interferometry, both based on optical interference principles, stand out due to their high precision, resolution, non-contact nature, and strong resistance to interference. As a result, they have become essential tools in ultra-precision machining, metrology, microelectronics manufacturing, and scientific research, where extreme measurement accuracy is required [60,61]. In contrast, laser triangulation, while offering non-contact measurement and rapid response, suffers from resolution degradation over extended ranges due to optical divergence. Capacitive and inductive sensors, relying on field variation, provide stable performance in harsh environments but are limited by short measurement ranges within 10 mm and susceptibility to electromagnetic interference.
Laser interferometry and grating interferometry both utilize optical coherence to extract displacement information [49,62]. They can achieve multi-degree-of-freedom measurement with sub-nanometer or even picometer-level precision [63,64,65,66,67,68], as well as absolute positioning [69,70,71,72,73,74]. A laser interferometer detects target displacement by analyzing phase shifts in the reflected light from a measurement surface relative to a reference beam, enabling a wide dynamic range from sub-nanometer to several kilometers [75]. However, because its measurement reference is the laser wavelength, it is susceptible to fluctuations in the air’s refractive index under non-vacuum conditions and is thus sensitive to environmental factors such as temperature and humidity, requiring compensation and calibration [76,77,78].
In contrast, grating interferometers measure target displacement by detecting the phase shift between diffracted beams from a measurement grating and a reference grating [79,80,81,82,83,84]. Unlike laser interferometers that rely on the wavelength of light as the measurement standard, grating interferometers use the grating pitch as a reference, which renders them inherently less sensitive to environmental variations such as temperature and pressure fluctuations. This characteristic makes them particularly well-suited for high-precision measurements in industrial environments where robustness and stability are critical [85,86,87,88]. The measurement accuracy of grating interferometers, however, is intrinsically dependent on the precision and uniformity of the grating itself. Laser lithography has become the key fabrication method for such gratings [89,90], offering high resolution and patterning accuracy over large areas [91,92,93]. The advancement of high-precision fabrication techniques for two-dimensional gratings [94,95,96,97,98] provides the foundational metrology infrastructure for integrated in-plane and out-of-plane multi-axis measurement systems, enabling sub-nanometer positioning accuracy essential to motion control applications such as wafer stage positioning in semiconductor lithography equipment [99,100]. By optimizing exposure parameters, mask design, and etching processes, it is possible to produce gratings with exceptional pitch accuracy, minimal defects, and high cost-efficiency, thus providing a robust metrological foundation for achieving sub-nanometer measurement precision [101,102,103].
Further improvements in measurement precision increasingly rely on the signal subdivision and processing capabilities of demodulation algorithms [104,105,106]. Additionally, in multi-degree-of-freedom measurements, laser interferometers typically require multiple subsystems arranged around the target, whereas grating interferometers can achieve up to six degrees of freedom within a compact space [107,108,109,110].
Based on their operating principles, optical interferometers are generally categorized into homodyne and heterodyne systems [53,111,112,113]. Homodyne interferometers use a single-frequency laser to generate direct-current (DC) amplitude-modulated signals, from which displacement information is extracted through phase demodulation [107,114,115,116]. These systems are structurally simple and often employed at low speed and high resolution. However, they are highly sensitive to light source stability and environmental noise, which makes them vulnerable to intensity fluctuations and DC drift [62,117,118,119]. In contrast, heterodyne interferometers employ two laser beams with a small frequency offset to produce a beat-frequency signal [56,120,121,122]. The low-frequency envelope enables the extraction of high-frequency phase information that would otherwise be difficult to measure directly. Heterodyne systems are more resistant to interference, offer higher sensitivity, and are better suited for multi-axis synchronous measurements that involve high resolution and speed, which places them among the dominant architectures in contemporary interferometry [123,124,125,126]. Currently, both measurement methods can achieve displacement measurement resolutions at the sub-nanometer or even picometer level [62,68,127,128].
Displacement demodulation algorithms are critical to achieving high-precision measurements in interferometers and directly influence system accuracy and performance. In practical applications, interference signals often suffer from periodic nonlinearity due to mixed-frequency sources and polarization crosstalk, requiring algorithmic compensation and correction during phase extraction [129,130,131,132].
For homodyne interferometers, commonly used demodulation algorithms include error calibration methods [133,134,135,136] and ellipse parameter estimation methods [51,137,138,139,140,141,142,143,144]. Error calibration methods sequentially correct various nonlinear errors to obtain standard orthogonal sine and cosine signals, but they involve relatively high computational complexity. In contrast, ellipse parameter estimation methods fit and estimate the parameters of the Lissajous figure formed by the two orthogonal signals to extract displacement information.
Heterodyne systems adopt more diverse and complex phase demodulation algorithms, including pulse counting [145], quadrature phase-locked techniques [146,147,148,149,150], and Kalman filtering [151]. Pulse-counting methods determine the phase by quantifying the time difference between two signals. They are simple and easy to implement but highly dependent on signal quality and frequency stability and perform poorly under low signal-to-noise conditions. Quadrature phase-locked methods employ digital phase-locked loop techniques to extract amplitude and phase information through mixing and filtering, overcoming the cycle-counting limitations of pulse methods and offering stronger noise resistance [146]. However, their dynamic range and noise performance are influenced by the cutoff frequency of the low-pass filter. Kalman filtering directly estimates the phase parameters from the interference signal, providing high real-time performance and exceptional measurement accuracy, with broad application potential [151].
Alongside algorithmic design, the choice of hardware platform plays an essential role in the overall implementation of optical systems. Common computing platforms include personal computers (PCs) [152,153,154,155,156,157], microcontroller units/digital signal processors (MCUs/DSPs) [158,159,160,161,162,163], field-programmable gate arrays (FPGAs) [133,135,136,151,164,165,166,167,168,169], and application-specific integrated circuits (ASICs). PCs offer powerful computational capabilities and rich software resources, facilitating algorithm development and debugging, but they tend to be bulky and lack real-time responsiveness. MCUs and DSPs have excellent real-time performance, compact size, and low power consumption, which makes them ideal for embedded applications, though their computational power may limit the implementation of complex algorithms. FPGAs provide high parallelism and flexibility for real-time signal processing and can be reconfigured for specific tasks. ASICs, with their high integration, low power consumption, and superior performance, are well-suited for high-end applications requiring compact systems, low latency, and fast measurement speeds.
This review aims to provide a systematic and comprehensive overview of signal processing algorithms for both homodyne and heterodyne optical interferometers, along with their associated computational hardware platforms, as the overall writing framework shows in Figure 1. Section 2 introduces the operating principles and signal characteristics of optical interferometers, laying the foundation for the following algorithmic discussion. Section 3 focuses on demodulation methods for homodyne and heterodyne systems, including their principles, advantages, limitations, and real-world applications. Section 4 surveys commonly used computational hardware platforms, including PCs, MCUs/DSPs, FPGAs, and ASICs, highlighting their architectural features, performance advantages, and practical roles in displacement demodulation and offering comparative insights to guide system selection and design. Through this review, we aim to provide readers with a comprehensive understanding of signal conditioning techniques for optical interferometers and offer valuable guidance for both academic research and engineering practice in related fields.

2. Patterns of Optical Interferometric Signals

In the field of optical interferometry, homodyne and heterodyne signals, as the primary signal patterns in coherent detection, play a crucial role in precision measurement systems such as grating interferometers and laser interferometers. A homodyne signal is generated by the interference between signal light carrying information and a local oscillator (LO) light that is strictly matched in frequency and phase [172,173,174,175,176,177,178,179,180]. The resulting signal directly reflects the amplitude and phase of the original optical field, offering ultra-high sensitivity and phase resolution [134,136,167,181]. These characteristics make homodyne detection well-suited for weak optical signal sensing and nanometer-scale displacement measurements. However, this technique imposes stringent requirements on light source stability and optical path consistency, as slight frequency drifts may lead to signal distortion [106,182].
By contrast, a heterodyne signal [183,184,185,186,187,188] is produced by interfering the signal light with an LO light of a different frequency, resulting in an electrical signal at a fixed intermediate frequency (IF), which facilitates subsequent filtering, amplification, and demodulation. It exhibits strong anti-interference capability, and the IF processing can effectively suppress the influence of environmental noise and optical path fluctuations, making it applicable to dynamic measurements in complex environments [148,151,189,190]. Although the detection sensitivity of heterodyne detection is slightly lower than that of homodyne detection, it has lower requirements for frequency synchronization, providing better system robustness and engineering feasibility [191,192,193,194]. Nevertheless, the generation of heterodyne interference signals requires a dual-frequency laser source, which significantly increases the cost.

2.1. Homodyne Signal

Homodyne grating interferometry employs a single-frequency laser source. A Michelson-type homodyne grating interferometer is illustrated in Figure 2a. A laser beam of frequency f first passes through a non-polarizing beam splitter (NPBS), where it is split into two beams directed orthogonally, which are incident on the measurement and reference gratings, respectively. After being diffracted by the gratings, the beams are collimated and then recombined at the NPBS to produce interference. To determine the direction of motion, the interference beam is further split into two beams by an additional NPBS. One beam passes through a polarizer and is detected by a photodiode (PD), while the other beam passes through a quarter-wave plate (QWP), followed by a polarizer, and is then detected by a second PD. The two detected signals exhibit a 90° phase difference, enabling quadrature signal analysis.
The single-frequency laser beam emitted by the laser source can be expressed as
E 0 = A 0 c o s ( 2 π f t + ϕ 0 )
where f denotes the laser frequency, A 0 represents the amplitude, and ϕ 0 is its initial phase. After diffraction by the reference grating, the beam can be written as
E r = A r c o s ( 2 π f t + ϕ r )
where A r is the amplitude and ϕ r is the phase. The frequency of the diffracted light remains unchanged at f. Both the amplitude and phase undergo alterations due to change in the optical path and attenuation by optical components.
When the measurement grating is in motion, the diffracted light from the measurement grating experiences a frequency shift because of the optical Doppler effect. This phenomenon can be mathematically described as
E m = A m c o s [ 2 π ( f + Δ f k ) t + ϕ m ]
= A m c o s [ 2 π f t + Δ ϕ m + ϕ m ]
Here, A m and ϕ m denote the amplitude and phase of the measurement beam, respectively. Δ f k represents the frequency variation caused by the displacement in the measurement grating. This frequency change can be equivalently characterized as a phase change Δ ϕ m , which inherently encodes the displacement signal of the measurement grating and exhibits a linear relationship with the displacement. The displacement is calculated using the formula
L = k Δ ϕ m
where L is the displacement in the measurement grating and k is a proportionality coefficient determined by the grating parameters and the optical encoder structure.
When the measurement beam and reference beam interfere, the wave equation of the resulting interference light is given by E m + E r . Owing to bandwidth limitations, a PD can only respond to low-frequency signals. The signals detected by PD and PD90° can be expressed as follows:
I c o s A r 2 + A m 2 + 2 A r A m c o s ( Δ ϕ m )
I s i n A r 2 + A m 2 + 2 A r A m s i n ( Δ ϕ m )
By deriving the phase difference Δ ϕ m from I c o s and I s i n , the displacement in the measurement grating can be calculated using Formula (5). Therefore, the precise calculation of the phase difference using quadrature signals is critical to homodyne signal processing.
The homodyne laser interferometer also generates homodyne signals, as shown in Figure 2b. Compared with the grating interferometer, the laser interferometer uses a measurement mirror instead of a measurement grating. When the measurement mirror moves, its reflected beam undergoes a frequency shift due to the Doppler effect. The homodyne signal pattern generated by the homodyne laser interferometer is essentially consistent with that of the grating interferometer, as expressed by Formulas (6) and (7). However, the measurement reference in laser interferometry is the wavelength of the light, which makes it more susceptible to environmental influences. In comparison, grating interferometers use the grating pitch as the measurement reference, making them less sensitive to environmental variations.

2.2. Heterodyne Signal

Heterodyne grating interferometry employs a dual-frequency laser source. Figure 3a illustrates a typical heterodyne grating interferometer. The dual-frequency laser emits two orthogonally linearly polarized beams with frequencies f 1 and f 2 , which are separated into two beams by an NPBS. The reflected beam passes through a polarizer, interferes, and is detected by PDr as the reference signal. The transmitted beam is further split into two beams by a polarizing beam splitter (PBS), which are then diffracted by the reference and measurement gratings, respectively. The diffracted beams are then recombined at the PBS. After passing through a polarizer, the recombined beams interfere and are detected by PDm as the measurement signal. The frequency of the reference signal remains constant, while the frequency of the measurement signal shifts due to the Doppler effect when the measurement grating moves.
The two light sources with frequencies f 1 and f 2 are expressed as
E 1 = A 1 c o s ( 2 π f 1 t + ϕ 01 )
E 2 = A 2 c o s ( 2 π f 2 t + ϕ 02 )
respectively, where A 1 and A 2 are the amplitudes, and ϕ 01 and ϕ 02 are the initial phases. Due to bandwidth limitations, a PD can only respond to low-frequency signals. The signals detected by PDr and PDm can be expressed as follows:
I r A 1 2 + A 2 2 + 2 A 1 A 2 c o s [ 2 π ( f 1 f 2 ) t + ϕ r 2 ϕ r 1 ]
I m A 1 2 + A 2 2 + 2 A 1 A 2 c o s [ 2 π ( f 2 f 1 + Δ f k ) t + ϕ m 2 ϕ m 1 ]
A 1 2 + A 2 2 + 2 A 1 A 2 c o s [ 2 π ( f 2 f 1 ) t + Δ ϕ k + ϕ m 2 ϕ m 1 ]
Here, ϕ r 1 , ϕ r 2 , ϕ m 1 , and ϕ m 2 denote the phases of the reference and measurement diffracted beams, respectively. Δ f k denotes the frequency shift induced by the Doppler effect during grating movement, and Δ ϕ k represents the phase variation converted from the frequency shift. The relationship between the measured grating displacement L and the phase variation ϕ is given by
L = k Δ ϕ k
where k is the proportionality coefficient, which is related to the grating parameters and the optical system structure.
The heterodyne laser interferometer also produces heterodyne signals, as illustrated in Figure 3b. Compared with the heterodyne grating interferometer, it replaces the measurement grating with a measurement mirror. When the measurement mirror moves, the reflected beam experiences a frequency shift due to the Doppler effect. The resulting heterodyne signal pattern is essentially identical to that of the grating interferometer and is described by Formulas (10) and (12).

2.3. Comparison of Homodyne Signals and Heterodyne Signals

In the field of optical interferometric metrology, homodyne and heterodyne interferometric signals differ significantly in their underlying principles, signal characteristics, and applications.
In terms of signal characteristics, homodyne interferometry utilizes a single-frequency laser source. The laser beam is split into a reference beam and a measurement beam by a beam splitter. Their interference at the detector generates a DC-modulated signal, whose intensity varies sinusoidally with the displacement in the measurement grating or mirror, as illustrated in Figure 4a. The phase change of the signal is directly proportional to the displacement. In contrast, heterodyne interferometry employs a dual-frequency laser source, for example, a dual-frequency helium–neon laser, that emits two orthogonally linearly polarized beams at frequencies f 1 and f 2 . The interference of these beams produces an alternating-current (AC) beat signal at a frequency of | f 1 f 2 | . When the measurement grating or mirror moves, the Doppler effect induces a frequency shift in the measurement beam, causing a corresponding phase shift in the beat-frequency signal, as shown in Figure 4b. The displacement is then determined by measuring this phase change. Compared with heterodyne detection, homodyne signal processing is relatively simple but susceptible to DC drift. In contrast, the AC signal generated in heterodyne interferometry is better suited for processing with AC amplification circuits and more effectively suppresses DC-related noise.
Regarding noise immunity, homodyne interferometric systems are highly sensitive to environmental disturbances. Due to their DC-modulated signals, fluctuations in laser power, variations in detector dark current, and ambient stray light can cause baseline drift, thereby compromising measurement accuracy. For example, temperature-induced fluctuations in the air’s refractive index can alter the optical path length, introducing measurement errors. Mechanical vibrations may also cause jitter in the interference fringes, further complicating phase demodulation. In contrast, heterodyne interferometric systems offer significant advantages in noise immunity. The AC beat signal allows DC noise to be effectively suppressed using AC-coupled circuits, making the system largely immune to low-frequency disturbances such as laser power instability and stray light. Furthermore, the dual-frequency nature of heterodyne interferometry provides inherent common-mode rejection of environmental changes. For instance, temperature variations affect both frequency components similarly, leaving the beat frequency and phase largely unchanged. As a result, heterodyne interferometry demonstrates greater stability in complex environments such as industrial settings.
Measurement precision is a critical metric for evaluating interferometric system performance. Under ideal conditions, homodyne interferometry can achieve high precision through fringe subdivision techniques, theoretically reaching nanometer- or even sub-nanometer-scale resolution. However, in practical applications, its precision is often limited by sensitivity to environmental disturbances. For example, humidity or temperature fluctuations can introduce errors in high-precision displacement measurements. Heterodyne interferometry, by virtue of its unique signal characteristics and robust noise immunity, typically achieves higher measurement precision and stability. The phase-to-displacement relationship in heterodyne signals is more linear, and AC signal demodulation techniques such as lock-in amplification enable high-precision phase detection even in noisy environments. Furthermore, increasing the beat signal frequency (e.g., to tens of megahertz) enhances measurement resolution. When paired with high-precision counters, this allows for sub-nanometer-scale or even picometer-scale displacement measurements.
From the perspective of system complexity and cost, homodyne interferometric systems are relatively simple in structure, requiring only a single-frequency laser source and basic beam-splitting and -reflecting components. Consequently, they are cost-effective and suitable for cost-sensitive laboratory measurement scenarios with relatively stable environments. Heterodyne interferometric systems, however, require dual-frequency laser sources and more complex optical paths and signal processing circuits, resulting in higher system costs. Nonetheless, their performance advantages make them the preferred choice for high-precision measurements in harsh environments.

3. Optical Interferometric Signal Processing Techniques

3.1. Phase Estimation Methods for Homodyne Signals

The displacement information is embedded in the phase of the homodyne signal, so the homodyne signal phase calculation accuracy directly determines the displacement measurement accuracy of the optical system. The ideal homodyne signal is expressed as Formulas (6) and (7). However, in practical homodyne optical systems, due to factors such as assembly deviations and nonlinear errors of optical components, the detected signals usually contain DC drift, amplitude drift, phase error, and high-frequency noise, as shown in Figure 5.
An actual homodyne signal with non-ideal components can be expressed as
I c o s ( t , L ) = A 1 ( t , L ) c o s ( Δ ϕ m ) + D 1 ( t , L ) + n 1 ( t , L )
I s i n ( t , L ) = A 2 ( t , L ) s i n ( Δ ϕ m + ϕ 2 ( t , L ) ) + D 2 ( t , L ) + n 2 ( t , L )
where A 1 ( t , L ) and A 2 ( t , L ) are the amplitudes of the two homodyne signals, which vary with time and the displacement in the measurement target; ϕ 2 ( t , L ) is the phase error between the two signals, which destroys the orthogonality of the homodyne signals; D 1 ( t , L ) and D 2 ( t , L ) are the DC components of the signals, which also change with time and the displacement in the measurement target; and n 1 ( t , L ) and n 2 ( t , L ) are high-frequency noise components.
These errors introduce significant deviations into phase estimation, severely affecting the displacement measurement accuracy of the optical system. Therefore, these error factors must be considered when calculating the phase of homodyne signals. Currently, the calculation methods for homodyne signals are mainly divided into two categories. The first comprises the error calibration methods, which design a corresponding correction algorithm for each type of error, progressively correct the four main errors mentioned above, finally obtain sine–cosine signals close to the ideal state, and then calculate the phase through the arctangent function or other phase linearization functions. The second includes the ellipse parameter estimation methods, which use the two signals as the x-axis and y-axis coordinates to construct a Lissajous figure. Non-ideal errors cause the Lissajous figure to take an elliptical shape, and the signal phase can be obtained by estimating the ellipse parameters.

3.1.1. Error Calibration Methods

Error correction methods represent a straightforward approach to suppressing noise in homodyne signals. Y. Han first digitized homodyne signals by using an ADC and performed subsequent processing with an FPGA [133], as shown in Figure 6a. Then, FIR filters were applied to filter the signals, removing high-frequency noise and DC components. Next, signal normalization was conducted by acquiring one period of the signal, calculating its maximum value, and dividing all signal points in that period by this maximum. Subsequently, a pulse-counting method was used to measure and correct the phase difference between the two signals, restoring their orthogonality. After these correction steps, two sinusoidal signals free of DC components were obtained, from which the phase was calculated using the arctangent operation. Although this method effectively mitigates the impact of non-ideal components on phase calculation, it requires substantial computational resources and introduces high latency, primarily due to the division operations required for each signal point during the normalization step. N. Shi adopted a similar signal correction approach [136], as shown in Figure 6c. Building on their work, Y. Wang replaced the FIR filter with a Kalman filter [135], as shown in Figure 6b. By leveraging the sinusoidal characteristics of the input signals, the Kalman filter achieved high-frequency noise suppression, signal normalization, and phase correction at the same time with lower computational complexity. This not only significantly reduced computational resource consumption but also minimized the computational delay to 1.26 us, greatly enhancing the real-time performance of the calculation.

3.1.2. Ellipse Parameter Estimation Methods

In homodyne grating and laser interferometry, ellipse parameter estimation methods demonstrate unique advantages in phase retrieval by modeling the Lissajous figure as an ellipse [51,137,138,139,140,141,142,143,144,195]. These approaches take the two orthogonal signals generated by a homodyne interference system as the horizontal and vertical coordinates of an ellipse and retrieve phase information by fitting geometric parameters such as major axis, minor axis, and rotation angle, as shown in Figure 7. Conventional methods for elliptic parameter estimation primarily include the least squares method [137,140] and the Kalman filtering method [196]. In homodyne grating interference systems, non-ideal factors arising from assembly errors and nonlinearities in optical components—including DC drift, amplitude imbalance, and phase error—cause the signal trajectory to deviate from an ideal circle. Ellipse parameter estimation methods mitigate these errors through least squares fitting or nonlinear optimization algorithms.
Based on Equations (14) and (15), after filtering out high-frequency noise, the phase information of the homodyne signal can be expressed as
Δ ϕ m = L k = a r c t a n I s i n D 2 ( I c o s D 1 ) α + ( I s i n D 2 ) β
where the correction coefficients α and β can be represented as
α = A 2 A 1 c o s ( ϕ 2 )
β = t a n ( ϕ 2 )
The equation of the ellipse is given by
A I c o s 2 + B I c o s I s i n + C I s i n 2 + D I c o s + E I s i n + F = 0
The DC components D 1 and D 2 of the signal can be expressed in terms of the parameters in the ellipse equation as
D 1 = 2 C D B E B 2 4 A C
D 2 = 2 A E B D B 2 4 A C
The correction coefficients α and β can be represented using the parameters of the ellipse equation as
α = 2 A 4 A C B 2
β = B 4 A C B 2
Based on Equations (16)–(23), a relationship can be established between the phase information of the homodyne signal and the ellipse parameters A, B, C, D, E, and F. Thus, by accurately estimating these elliptic parameters, the phase information of the homodyne signal can be derived.
Table 1 shows the comparison between error calibration methods and ellipse parameter estimation methods. Although the error calibration methods can achieve high resolution, their step-by-step calibration consumes substantial hardware computing resources, leading to increased costs. As a result, they are commonly applied in academia. The ellipse parameter estimation methods are more flexible in their implementation: they can utilize Kalman filtering to achieve extremely high precision or employ the least squares method for the pipelined processing of signals. With lower hardware resource consumption, they are widely used in industry.

3.2. Phase Estimation Methods for Heterodyne Signals

Phase calculation for heterodyne interferometers represents one of the key steps towards achieving high-precision displacement measurement, with its essence lying in converting optical frequency shift signals into phase information and realizing high-resolution demodulation. Depending on distinct signal processing principles, existing approaches are primarily categorized into three types: pulse-counting methods [5,197,198,199,200,201,202,203,204,205,206], quadrature phase-locked methods [146,147,148,149,150,207,208], and the Kalman filtering method [151]. In pulse-counting methods, after heterodyne interference signals are converted into sine waves by the PD, a comparator is required to transform them into square waves. The latter two method types, conversely, utilize an analog-to-digital converter (ADC) to acquire sine waves for direct amplitude–phase analysis.
Pulse-counting methods achieve phase calculation through high-frequency pulse filling and time difference measurement, relying on GHz level phase-locked loop (PLL) frequency multiplication technology to enable sub-nanometer resolution and high-speed measurement capabilities. Quadrature phase-locked methods extract phase information via quadrature demodulation and arctangent operation, demonstrating remarkable advantages in picometer-scale resolution. The Spectrum Analysis Method performs phase analysis in the frequency domain, allowing for the easier separation of the signal fundamental frequency from harmonic interferences and exhibiting strong anti-harmonic capabilities. The subsequent discussion systematically addresses the principles, evolutionary, and technological breakthroughs of these three method types.

3.2.1. Pulse-Counting Methods

Pulse-counting methods obtain phase information by using an additional high-frequency clock signal to count the periods of input measurement signals.
These methods convert measurement and reference signals into square waves via a zero-crossing comparator and then input these square waves into counters. By calculating the integer number of signal periods and the number of reference signal periods corresponding to the sub-period segment, the phase difference between the two signals can be accurately determined [197,198].
To improve the resolution of phase measurement, a common approach is to apply PLL-based frequency multiplication to the interference signal, thereby increasing the phase subdivision factor. However, due to the bandwidth limitations of logic devices, the multiplication factor achievable by PLL-based techniques is constrained. To address this issue, Hu proposed a phase measurement method based on PLL frequency multiplication combined with digital delay lines [209]. On the basis of PLL-based multiplication, this method introduces multistage digital delay lines to further enhance the effective phase subdivision factor, thereby improving the resolution of phase and displacement measurements. A series of interferometric signal processing boards have been developed based on this method, achieving an electronic subdivision factor of up to 1024. This enables a maximum measurement resolution of 0.31 nm @optical 2-fold subdivision or 0.15 nm@optical 4-fold subdivision and a maximum measurement speed of 2 m/s@optical 2-fold subdivision or 1 m/s@optical 4-fold subdivision [209,210]. Frank et al. also implemented a heterodyne interferometric phase demodulation system based on this method, achieving a measurement resolution of 0.31 nm and a measurement speed of 2.1 m/s [5].
E. Zhang proposed a novel signal processing method for laser heterodyne interferometry based on rising-edge locking with high-frequency clock signals and digital frequency mixing [170], as shown in Figure 8a. By locking the rising edges of reference and measurement signals with a high-frequency clock and replacing traditional analog mixing with digital frequency mixing, it eliminates multi-edge counting errors and enhances anti-interference capability. Experiments verify that it achieves both high speed, 2.4 m/s, and resolution, 0.008 nm, providing a new solution for high-precision displacement measurement.
T. Yang proposed a novel signal processing method for movement direction identification and phase correction in laser heterodyne interferometry [145], as shown in Figure 8b. The core approach involves establishing four 90° phase intervals based on the reference signal, realizing real-time direction identification and integer fringe counting by detecting the times the rising edge of the measurement signal crosses these intervals. A detailed phase correction method was proposed to solve the fractional phase compensation issue when the initial phase difference is non-zero. The innovation lies in integrating direction identification with phase correction, determining the movement direction through interval division and state changes of the measurement signal, which overcomes measurement errors caused by initial phase differences in traditional methods. Experiments verify the method’s feasibility in bidirectional movement stability tests and nano/micrometer displacement measurements, achieving a measurement accuracy of 6.84 nm. This provides an effective solution for high-precision displacement measurement.
J. Zhou proposed a novel laser interferometric demodulation method based on fringe-to-pulse counting, as shown in Figure 8c, which converts high-frequency orthogonal analog signals into digital signals, obtains displacement increments by comparing four states with a high-frequency clock, and reduces the data rate via subsampling [171]. The innovation lies in eliminating the need for high-speed data acquisition devices with simple hardware and algorithms, enabling real-time high-precision demodulation for both low- and medium-frequency vibration calibrations. Experiments verify that amplitude and phase deviations are within allowable limits.
Pulse-counting methods are widely applied in practice due to their simple algorithm and ease of implementation. They also feature strong algorithm robustness, high sensitivity, and suitability for dynamic measurement tasks. Currently, commercial boards based on pulse-counting methods are available globally. For instance, Agilent has released the N1225A board with 0.15 nm resolution [211]; ZYGO offers the ZMI4104(C) board, also with 0.15 nm resolution [212]. However, inherent limitations exist. First, their accuracy is largely constrained by the precision of zero-crossing conversion, with errors increasing significantly in low-SNR environments. Second, resolution and precision degrade gradually as the frequency of the measured signal increases. Finally, high-frequency signals after multiplication require advanced chip performance.

3.2.2. Quadrature Phase-Locked Methods

Quadrature phase-locked methods extract signal amplitude and phase information through digital signal processing techniques such as mixing and filtering, breaking through the limitation of pulse-counting methods, which rely on periodic counting [213,214,215,216,217].
At present, most mainstream quadrature phase-locked loop techniques employ digital circuits to perform critical operations such as signal mixing and filtering. In comparison with conventional lock-in amplifiers implemented using analog circuits, the digital quadrature phase-locked loop features high precision, free from errors caused by the uncertainty of analog circuits; meanwhile, the algorithm demonstrates better stability, maintaining stable measurement performance over long periods. Additionally, it offers good flexibility, allowing for convenient modification of the algorithm form according to different measurement requirements, thus solving many difficulties faced by analog phase-locked circuits during upgrading [218].
P. K. proposed a lock-in phase measurement method based on an FPGA and a high-speed ADC, which extracts phases of heterodyne interferometric signals using the digital lock-in principle and achieves picometer-level resolution through dynamic window filtering and Doppler frequency shift compensation [146], as shown in Figure 9a. It uses a commercial ADC board with user-defined algorithms to simultaneously track dual-beat frequencies, overcoming bandwidth limitations of traditional phasemeters. Experiments demonstrate a long-term stability of 0.8 pm and reveal polarization effects as the primary error source.
P. Hu proposed a novel phase measurement method based on a digital dual-frequency comb, which forms bandpass filter banks using intrinsic frequency comb signals and narrowband low-pass filters to divide broadband signals from high-speed motion into multiple narrow bands for SNR improvement [147], as shown in Figure 9b. It simultaneously uses two groups of comb signals for phase measurement and output rotation. It solves phase measurement inaccuracies caused by the input signal frequency being the average of two comb frequencies during high-speed motion, enabling high-speed and high-precision phase measurement for heterodyne interferometry.
S. Lu proposed an enhanced lock-in method based on pulse counting for frequency tracking which roughly estimates the measurement signal frequency via pulse counting and generates orthogonal mixing signals to achieve dynamic frequency tracking on an FPGA platform [150], as shown in Figure 9c. The innovation lies in addressing the sensitivity to filter performance and poor generalization capability of traditional quadrature phase-locked methods, achieving resolution errors within tens of picometers and a measurement speed exceeding 1 m/s, thereby significantly enhancing the robustness of the lock-in method.
Taiji used a phasemeter scheme based on the digital phase-locked loop (DPLL), implementing signal mixing, filtering, and frequency tracking via an FPGA and adopting common-mode noise rejection to reduce sampling and frequency jitter noises [148], as shown in Figure 9d. It achieves a sensitivity of 2 π μ r a d / H z , meeting the requirements for 0.01 Hz–1 Hz and 0.1 mHz–1 mHz frequency bands, thus providing key technical support for the Taiji space gravitational wave detection mission.
The phasemeter designed for LISA employs a quadrature phase-locked phase detection scheme based on the DPLL [149], as shown in Figure 9e. To achieve phase locking between the oscillator and the input signal, the scheme utilizes residual tracking error information from two mixed signals to feedback-correct the phase of the digital oscillating signal. Meanwhile, data acquired by the ADC are transmitted to the phase-locked output module to lock the laser for stable acquisition of the required heterodyne frequency. This approach achieves a phase measurement accuracy of 6 π μ r a d / H z in the 1 mHz–1 Hz frequency band.
The advantages of quadrature phase-locked methods are mainly reflected in the following aspects: first, they make full use of signal amplitude information, which can more effectively utilize the information carried by the signal compared with pulse-counting methods, thus improving measurement accuracy; second, they have strong noise resistance, maintaining good measurement performance in noisy environments; third, they can achieve higher precision and resolution to meet the needs of high-precision measurements. However, the selection of the cutoff frequency of the low-pass filter in these methods has an important impact on the dynamic range and output signal noise, which need to be adjusted according to specific measurement scenarios and requirements in practical applications.

3.2.3. Kalman Filtering Method

In heterodyne signal demodulation, quadrature phase-locked methods extract phase information through mixing and filtering, but they exhibit sensitivity to filter performance and tend to suffer from accuracy degradation due to fixed parameters in dynamic or high-noise scenarios. They also struggle to balance real-time performance and accuracy. In contrast, Kalman filtering employs state-space modeling, treating phase, frequency, and amplitude as state variables, and dynamically updates estimations via a prediction-correction mechanism, effectively suppressing random noise and nonlinear interference [219,220,221,222,223]. Its advantages include (1) the adaptive tracking of signal variations without fixed filter parameters; (2) low-latency computation relying only on current sampling data, suitable for real-time scenarios; and (3) enhanced robustness against dynamic frequency drift and environmental noise, making it particularly applicable to high-precision dynamic measurements such as multi-degree-of-freedom grating interferometry.
Y. Ma proposed a heterodyne signal phase solution algorithm based on the first-order extended Kalman filter (EKF), which sets the phase, frequency, and amplitude of the sinusoidal signal as state variables and implements prediction-correction computation via an FPGA [151]. The processing framework of the extended Kalman filter with an FPGA shows in Figure 10. The advantages include (1) achieving a measurement accuracy of 0.03° and a resolution of 0.01°, (2) enabling low-latency computation suitable for high-speed dynamic scenarios, and (3) featuring compact hardware deployment to enhance the real-time performance and noise immunity of multi-degree-of-freedom grating interferometers.
Table 2 shows a comparison of pulse-counting methods, quadrature phase-locked methods, and the Kalman filtering method. Pulse-counting methods exhibit low signal demodulation resolution due to signal zero-crossing errors and clock frequency limitations. However, they do not rely on high-speed ADCs, resulting in lower costs, and thus find applications in both academia and industry. Quadrature phase-locked methods can achieve higher resolution but rely on high-speed ADCs, leading to higher costs. Owing to their excellent performance, they have extensive applications in both academia and industry. The Kalman filtering method can achieve higher resolution than quadrature phase-locked methods. Nevertheless, due to its algorithmic constraints, it cannot implement pipelined signal processing, which limits its signal processing bandwidth. As a result of this limitation, it is generally only applied in academia.

4. Electronic Platforms for Signal Processing

As a core technology in precision displacement metrology, optical interferometry imposes diverse requirements on hardware platforms for its signal processing, encompassing both homodyne and heterodyne signals, with respect to computational capacity, real-time performance, power consumption, and cost. Current mainstream hardware platforms can be categorized into four major types—personal computers (PCs) (Figure 11a,b), microcontrollers/digital signal processors (MCUs/DSPs) (Figure 11c), field-programmable gate arrays (FPGAs) (Figure 11d), and application-specific integrated circuits (ASICs) (Figure 11e)—each suited to full lifecycle scenarios ranging from algorithm verification to industrial mass production. PCs, relying on robust general-purpose computational capabilities and rich software ecosystems, serve as the primary choice for complex algorithm development in laboratory environments. MCUs/DSPs, leveraging their low-cost and low-power advantages, dominate simple signal processing in miniaturized and portable sensors. FPGAs, through hardware logic parallelism, achieve high real-time performance requirements and are well-suited for high-speed feedback control in industrial settings. ASICs, with their ultra-high performance and integration, support large-scale production of high-end grating sensors. These four kinds of platforms complement each other in terms of computational complexity, latency, flexibility, and cost, collectively forming the hardware technical framework for optical interferometric signal processing.

4.1. PCs

PCs represent the most versatile hardware platform for grating interferometric signal processing, with their core value lying in providing a flexible software environment for algorithm development and verification. PCs are built on general-purpose processors based on x86 or ARM architectures, equipped with large-capacity memory and high-speed storage, and support multi-task parallel processing. At the software level, PCs can run professional tools such as MATLAB [225,226,227], Python (paired with NumPy/SciPy toolkits) [228,229,230], and LabVIEW [231,232,233,234,235,236,237,238,239], enabling the rapid implementation of complex algorithms. Their computational capacity is typically measured in GFLOPS (gigaflops, billions of floating-point operations per second), with modern PCs achieving a single-threaded computational capacity of 10 to 100 GFLOPS and exceeding 1 TFLOP (teraflop, one trillion floating-point operations per second) under multi-threaded parallel processing.
PCs are primarily utilized for algorithm prototyping and performance validation in laboratory environments [152,153,154,155,156,157]. For homodyne interferometric signals, phase demodulation requires suppressing DC drift, and PCs can achieve sub-nanometer phase measurement accuracy through complex algorithms such as digital filtering and elliptical fitting to compensate for non-orthogonal errors [233]. For heterodyne interferometric signals, their beat-frequency signals require frequency tracking and phase subdivision [240]. PCs can leverage FFT to analyze beat-frequency stability or employ PLL algorithms to realize high-precision demodulation of dynamic displacement. Furthermore, PCs support joint analysis of multi-signals, providing technical support for system-level error modeling.
The core advantages of PCs lie in their development flexibility and complex task handling capability. Their software ecosystems support rapid algorithm iteration such as adjusting filtering parameters and optimizing fitting models, without requiring hardware redesign; simultaneously, they enable synchronized signal acquisition via DAQ cards, processing, display, and storage, significantly facilitating debugging and analysis. However, the limitations of PCs are also pronounced: their size and power consumption fall short of the requirements for miniaturized devices; the millisecond-level task scheduling latency of general-purpose operating systems cannot accommodate real-time feedback demands in industrial settings, which have microsecond-level latency scenarios in motion control; and the cost of high-performance PCs typically reaches several thousands dollars, making them unsuitable for large-scale deployment.

4.2. MCUs/DSPs

MCUs and DSPs, serving as the core components of embedded systems, exhibit a clear functional division in grating interferometric signal processing: MCUs primarily handle low-complexity control and simple signal processing tasks, while DSPs are dedicated to the efficient implementation of digital signal processing algorithms.
MCUs integrate peripherals such as CPUs, RAM, Flash, ADCs/DACs, etc., with a computing capability of approximately 0.1 to 10 DMIPS (Dhrystone million instructions per second) and power consumption as low as 1 to 100 mW. Their abundant on-chip resources are suitable for implementing logic control and simple signal processing such as low-pass filtering to suppress high-frequency noise and threshold detection for identifying interference fringe zero-crossings. A typical application scenario is low-cost displacement sensors [162,241], where interferometric signals are acquired via on-chip ADC and rough displacement values are output after simple mean filtering.
DSPs are optimized for digital signal processing, featuring built-in hardware multipliers, pipeline architectures, and parallel arithmetic units. They achieve computing capabilities of 100 to 1000 DMIPS with power consumption ranging from 100 mW to 2 W. Their instruction sets are optimized for FFT, convolution, and correlation operations, making them suitable for the real-time processing of high-frequency signals [242]. For example, quadrature demodulation of heterodyne interference that converts MHz level beat-frequency signals into I/Q baseband signals and phase subdivision of homodyne signals that subdivides fringe periods into 1/1000 via arctangent operations can both achieve microsecond-level demodulation through DSP hardware acceleration units.
The core advantages of MCUs/DSPs lie in their low cost and low power consumption [243,244,245]. MCUs are priced at only 0.1 to 10 USD, while DSPs cost approximately 10 to 100 USD, making them suitable for low-cost sensors in mass production. Their power consumption supports battery-powered portable devices. Furthermore, on-chip peripherals such as ADCs and PWMs reduce external circuitry, simplifying system design. However, both exhibit limited computational capabilities: MCUs cannot handle complex algorithms such as high-order filtering and multi-parameter fitting, and DSPs are prone to computational resource saturation under high sampling rates or multi-channel signals. Meanwhile, modifying the code after programming requires recompilation and download, hindering rapid algorithm iteration.

4.3. FPGAs

FPGAs realize signal processing through configurable hardware logic gates, serving as the optimal balance platform for real-time performance and flexibility in grating interferometric signal processing. Based on SRAM or Flash process technologies, their programmable logic units and flip-flops support hardware parallel computing and pipeline design. Mainstream FPGAs such as Xilinx Kintex and Intel Stratix contain tens of thousands to millions of logic elements, integrating DSP slices, high-speed ADC interfaces, and high-speed serial transceivers. They achieve computational latency at the nanosecond level and throughput at the GHz level while supporting dynamic reconfiguration.
The core advantage of FPGAs lies in hardware-level real-time performance, making them particularly suitable for scenarios such as high-speed signal acquisition and preprocessing, as well as real-time phase subdivision [246,247,248,249]. For instance, the beat-frequency signal of heterodyne interference requires sampling at rates above 50 MHz. FPGAs can achieve delay-free sampling via high-speed ADC interfaces and complete preprocessing steps like digital down-conversion and low-pass filtering in hardware pipelines, eliminating the software latency of PC platforms [133]. For phase subdivision in homodyne interference, FPGAs can implement high-precision arctangent operations through customized logic units, subdividing the interference fringe period into 1/1000 to meet the measurement requirements of ultra-precision equipment such as lithography machines [134,135,136]. Furthermore, FPGAs support multi-channel parallel processing, making them suitable for multi-dimensional displacement measurement systems.
FPGAs have excellent real-time performance and flexible configurability. Their hardware parallel architecture enables nanosecond-level latency, meeting the real-time feedback requirements of industrial sites. The dynamic reconfiguration function supports online adjustment of algorithm parameters to adapt to different measurement scenarios. However, FPGAs have a high development threshold, requiring mastery of hardware description languages like Verilog/VHDL, and their design cycles span months [250]. With chip costs ranging from hundreds to thousands of dollars and power consumption from several to tens of watts, FPGAs are more expensive and power-hungry than MCUs/DSPs, limiting their application in low-cost scenarios.

4.4. ASICs

ASICs are custom-designed integrated circuits tailored for specific functions, achieving ultimate optimization in performance and integration for optical interferometric signal processing through hardware-embedded specific processing logic, such as AM-IP4k [224] and GC-NIP for homodyne signals. Based on CMOS technology, ASICs integrate signal acquisition, processing, and output into a single chip, with an area typically at the square-millimeter scale, power consumption as low as the microwatt level, and processing speeds reaching the GHz range.
The core value of ASICs lies in their scale advantages in performance and cost, making them suitable for high-end grating sensors in mass production. Take the displacement measurement module of lithography machines as an example: it needs to simultaneously meet the requirements of sub-nanometer precision, microsecond-level latency, and extremely small size. Furthermore, the mass production cost of ASICs decreases significantly with the increase in yield—when produced in million-unit quantities, the cost per chip can be as low as a few dollars, making them ideal for large-scale deployment of industrial-grade sensors.
ASICs have ultimate performance and ultra-low power consumption. Their customized hardware logic avoids the redundant calculations of general-purpose platforms, enabling processing latency as low as the nanosecond level and power consumption only 1/10 to 1/100 of that of FPGAs. The high integration (full functionality on a single chip) significantly simplifies system design. However, ASICs entail extremely high development costs and extremely long development cycles. Once the design is finalized, modifications are difficult, making ASICs suitable only for scenarios with clear requirements and large production volumes.
Table 3 shows a comparison of different hardware platforms for interferometer signal processing. Among all hardware platforms, PCs exhibit the highest flexibility. Developers can rapidly develop and validate new algorithms on PCs. However, their power consumption, size, and parallel processing capability have hindered large-scale industrial applications, limiting their use to academic research platforms. FPGAs and DSPs strike a balance among flexibility, size, and power consumption. While their flexibility is inferior to that of PCs, they feature a smaller size and lower power consumption, thus finding extensive applications in industry. Furthermore, FPGAs, due to their parallel computing capability, are often employed in multi-channel, high-bandwidth signal processing. ASICs exhibit the lowest flexibility, with long development cycles and high upfront development costs. Nevertheless, they offer low deployment costs, along with minimal size and power consumption, which makes them commonly used in large-scale industrial applications, typically combined with ellipse parameter estimation methods.

5. Conclusions

This paper reviews recent developments in signal processing techniques and platforms for optical interferometric measurement, focusing on both homodyne and heterodyne systems. The goal of signal processing is to enable high-precision, multi-degree-of-freedom, and large-range displacement measurement from the perspective of signal processing.
As one of the leading non-contact methods for precision displacement measurement, optical interferometry offers significant advantages in terms of accuracy, speed, and reliability. It has been widely applied in various industrial sectors, including mechanical engineering, electronics manufacturing, and aerospace. While novel interferometric architectures continue to emerge, increasing attention has been drawn to the importance of signal conditioning techniques for system induced errors. Algorithms that offer high precision, low latency, and minimal computational load have become essential to enabling the deployment of optical interferometers in real-world applications.
Homodyne interferometry employs a single-frequency laser and derives displacement by analyzing phase variations in a DC amplitude-modulated signal. This method features a relatively simple system design but is sensitive to laser stability and environmental noise and is prone to intensity fluctuations and DC drift. Signal processing for homodyne systems primarily focuses on noise suppression, normalization and orthogonalization of the dual sinusoidal signals, and fine phase subdivision. Ellipse parameter estimation methods calibrate orthogonal signals by analyzing their Lissajous figures, from which phase information is extracted. Error calibration methods align the phase of orthogonal signals using known phase relationships.
Heterodyne interferometry utilizes a dual-frequency laser to generate a dynamic beat-frequency signal. Displacement is determined by demodulating phase or frequency variations in this modulated signal. By incorporating frequency-shifting techniques, heterodyne systems achieve effective separation of signal and noise, significantly enhancing the signal-to-noise ratio and robustness against interference. These systems typically offer a broader dynamic range and higher linearity. Signal processing methods for heterodyne signals are more diverse. Pulse-counting methods are computationally simple but are limited by signal quality, frequency stability, and their reliance on periodic counting. Quadrature phase-locked methods, which employ digital phase-locked loop techniques and leverages signal amplitude information, perform better than pulse counting in low-SNR environments. The Kalman filtering approach estimates phase parameters in real time and is well-suited for dynamic applications due to its responsiveness.
As lithography technology advances toward ever-smaller process nodes, the need for the precise, real-time, multi-degree-of-freedom positioning of stages and masks in lithographic systems is becoming increasingly critical. Future development of signal conditioning algorithms will need to address not only the errors arising from optical component imperfections (such as virtual reflections at optical interfaces and non-ideal behaviors of optoelectronic sensors) but also the coupling effects between different degrees of freedom and the influence of these couplings on measurement accuracy. At the same time, reconciling the trade-offs among large-range measurement, high-speed target tracking, and sub-nanometer precision will be a key direction in the evolution of interferometric signal processing techniques. Additionally, the rise of machine learning has opened up new breakthrough directions for interferometric signal processing. Machine learning methods enable more precise analysis and correction of demodulation error patterns while also allowing for the incorporation of information such as ambient temperature and humidity into the model. Furthermore, the increasing emergence of high-computing-ability platforms oriented toward deep learning has made it possible to apply deep learning to interferometric signal processing in real time.

Author Contributions

Conceptualization and investigation, Y.W., F.Z. and X.L.; writing—original draft, Y.W., F.Z. and X.L.; writing—review and editing, Y.W., F.Z., L.L. and X.L.; supervision, X.L.; project administration, X.L.; funding acquisition, X.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China, grant number 62275142; Shenzhen Science and Technology Program, grant number JCYJ20241202125343058; and Shenzhen Stable Supporting Program, grant number WDZC20231124201906001.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available upon request from the corresponding author.

Acknowledgments

The authors would like to thank all our laboratory members who have joined the related projects at Tsinghua University for their contributions to the achievements described in this paper.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Carr, J.; Desmulliez, M.; Weston, N.; McKendrick, D.; Cunningham, G.; McFarland, G.; Meredith, W.; McKee, A.; Langton, C. Miniaturised optical encoder for ultra precision metrology systems. Precis. Eng. 2009, 33, 263–267. [Google Scholar] [CrossRef]
  2. Han, M.; Lei, F.; Shi, W.; Lu, S.; Li, X. Uniaxial MEMS-based 3D reconstruction using pixel refinement. Opt. Express 2022, 31, 536–554. [Google Scholar] [CrossRef] [PubMed]
  3. Lei, F.; Han, M.; Jiang, H.; Wang, X.; Li, X. A phase-angle inspired calibration strategy based on MEMS projector for 3D reconstruction with markedly reduced calibration images and parameters. Opt. Lasers Eng. 2024, 176, 108078. [Google Scholar] [CrossRef]
  4. Zhou, Q.; Qiao, X.; Ni, K.; Li, X.; Wang, X. Depth detection in interactive projection system based on one-shot black-and-white stripe pattern. Opt. Express 2017, 25, 5341–5351. [Google Scholar] [CrossRef]
  5. Demarest, F.C. High-resolution, high-speed, low data age uncertainty, heterodyne displacement measuring interferometer electronics. Meas. Sci. Technol. 1998, 9, 1024. [Google Scholar] [CrossRef]
  6. Yuan, J.; Lyu, B.; Hang, W.; Deng, Q. Review on the progress of ultra-precision machining technologies. Front. Mech. Eng. 2017, 12, 158–180. [Google Scholar] [CrossRef]
  7. Gao, W. Precision Nanometrology: Sensors and Measuring Systems for Nanomanufacturing; Springer: Berlin/Heidelberg, Germany, 2010. [Google Scholar]
  8. Catalucci, S.; Thompson, A.; Piano, S.; Branson III, D.T.; Leach, R. Optical metrology for digital manufacturing: A review. Int. J. Adv. Manuf. Technol. 2022, 120, 4271–4290. [Google Scholar] [CrossRef]
  9. Zhu, C.; Zhuang, Y.; Liu, B.; Huang, J. Review of fiber optic displacement sensors. IEEE Trans. Instrum. Meas. 2022, 71, 1–12. [Google Scholar] [CrossRef]
  10. Sciammarella, C.A. A Review: Optical Methods That Evaluate Displacement. In Advancement of Optical Methods & Digital Image Correlation in Experimental Mechanics, Volume 3: Proceedings of the 2018 Annual Conference on Experimental and Applied Mechanics; Springer: Berlin/Heidelberg, Germany, 2018; pp. 23–52. [Google Scholar]
  11. Zvietcovich, F.; Larin, K.V. Wave-based optical coherence elastography: The 10-year perspective. Prog. Biomed. Eng. 2022, 4, 012007. [Google Scholar] [CrossRef]
  12. Han, Y.; Wu, G.; Feng, D. Vision-based displacement measurement using an unmanned aerial vehicle. Struct. Control Health Monit. 2022, 29, e3025. [Google Scholar] [CrossRef]
  13. Hosseinzadeh Sani, M.; Saghaei, H.; Mehranpour, M.A.; Asgariyan Tabrizi, A. A novel all-optical sensor design based on a tunable resonant nanocavity in photonic crystal microstructure applicable in MEMS accelerometers. Photonic Sens. 2021, 11, 457–471. [Google Scholar] [CrossRef]
  14. Armano, M.; Audley, H.; Baird, J.; Binetruy, P.; Born, M.; Bortoluzzi, D.; Brandt, N.; Castelli, E.; Cavalleri, A.; Cesarini, A.; et al. Sensor noise in LISA pathfinder: In-flight performance of the optical test mass readout. Phys. Rev. Lett. 2021, 126, 131103. [Google Scholar] [CrossRef]
  15. Zuo, C.; Qian, J.; Feng, S.; Yin, W.; Li, Y.; Fan, P.; Han, J.; Qian, K.; Chen, Q. Deep learning in optical metrology: A review. Light. Sci. Appl. 2022, 11, 1–54. [Google Scholar] [CrossRef] [PubMed]
  16. Brown, B.H.; Smallwood, R.H.; Barber, D.C.; Lawford, P.; Hose, D. Medical Physics and Biomedical Engineering; CRC Press: Boca Raton, FL, USA, 2017. [Google Scholar]
  17. Szelitzky, E.; Kuklyte, J.; Mândru, D.; O’Connor, N.E. Low cost angular displacement sensors for biomechanical applications—A review. J. Biomed. Eng. Technol. 2014, 2, 21–28. [Google Scholar]
  18. Masís Solano, M.; Richer, E.; Costantino, S.; Lesk, M.R. Optic nerve head pulsatile displacement in open-angle glaucoma after intraocular pressure reduction measured by optical coherence tomography: A pilot study. Bioengineering 2024, 11, 411. [Google Scholar] [CrossRef]
  19. Knobel, R.G.; Cleland, A.N. Nanometre-scale displacement sensing using a single electron transistor. Nature 2003, 424, 291–293. [Google Scholar] [CrossRef]
  20. Siebert, T.; Hack, E.; Lampeas, G.; Patterson, E.; Splitthof, K. Uncertainty quantification for DIC displacement measurements in industrial environments. Exp. Tech. 2021, 45, 685–694. [Google Scholar] [CrossRef]
  21. Xie, C.; Nguyen, H.; Bui, X.N.; Nguyen, V.T.; Zhou, J. Predicting roof displacement of roadways in underground coal mines using adaptive neuro-fuzzy inference system optimized by various physics-based optimization algorithms. J. Rock Mech. Geotech. Eng. 2021, 13, 1452–1465. [Google Scholar] [CrossRef]
  22. Gao, W.; Haitjema, H.; Fang, F.; Leach, R.; Cheung, C.; Savio, E.; Linares, J.M. On-machine and in-process surface metrology for precision manufacturing. Cirp Ann. 2019, 68, 843–866. [Google Scholar] [CrossRef]
  23. Li, D.; Wang, B.; Tong, Z.; Blunt, L.; Jiang, X. On-machine surface measurement and applications for ultra-precision machining: A state-of-the-art review. Int. J. Adv. Manuf. Technol. 2019, 104, 831–847. [Google Scholar] [CrossRef]
  24. Li, C.; Yan, H.; Qian, X.; Zhu, S.; Zhu, P.; Liao, C.; Tian, H.; Li, X.; Wang, X.; Li, X. A domain adaptation YOLOv5 model for industrial defect inspection. Measurement 2023, 213, 112725. [Google Scholar] [CrossRef]
  25. Sato, R.; Li, K.; Michihata, M.; Takahashi, S.; Gao, W. Advanced sensing and machine learning technologies for intelligent measurement in smart and precision manufacturing. Int. J. Autom. Technol. 2024, 18, 545–580. [Google Scholar] [CrossRef]
  26. Li, C.; Pan, X.; Zhu, P.; Zhu, S.; Liao, C.; Tian, H.; Qian, X.; Li, X.; Wang, X.; Li, X. Style Adaptation module: Enhancing detector robustness to inter-manufacturer variability in surface defect detection. Comput. Ind. 2024, 157, 104084. [Google Scholar] [CrossRef]
  27. Yao, T.F.; Duenner, A.; Cullinan, M. In-line metrology of nanoscale features in semiconductor manufacturing systems. Precis. Eng. 2017, 47, 147–157. [Google Scholar] [CrossRef]
  28. Maitra, V.; Su, Y.; Shi, J. Virtual metrology in semiconductor manufacturing: Current status and future prospects. Expert Syst. Appl. 2024, 249, 123559. [Google Scholar] [CrossRef]
  29. ASML. Twinscan NXE:3600D [EB/OL]. 2022. Available online: https://www.asml.com/en/products/euv-lithography-systems/twinscan-nxe-3600d (accessed on 6 March 2022).
  30. Banks-Sills, L.; Shklovsky, J.; Krylov, S.; Bruck, H.A.; Fourman, V.; Eliasi, R.; Ashkenazi, D. A Methodology for Accurately Measuring Mechanical Properties on the Micro-Scale. Strain 2011, 47, 288–300. [Google Scholar] [CrossRef]
  31. Anandan, N.; George, B. A wide-range capacitive sensor for linear and angular displacement measurement. IEEE Trans. Ind. Electron. 2017, 64, 5728–5737. [Google Scholar] [CrossRef]
  32. Kumar, A.A.; Anandan, N.; George, B.; Mukhopadhyay, S.C. Improved capacitive sensor for combined angular and linear displacement sensing. IEEE Sens. J. 2019, 19, 10253–10261. [Google Scholar] [CrossRef]
  33. Ye, Y.; Zhang, C.; He, C.; Wang, X.; Huang, J.; Deng, J. A review on applications of capacitive displacement sensing for capacitive proximity sensor. IEEE Access 2020, 8, 45325–45342. [Google Scholar] [CrossRef]
  34. Nabavi, M.R.; Nihtianov, S.N. Design strategies for eddy-current displacement sensor systems: Review and recommendations. IEEE Sens. J. 2012, 12, 3346–3355. [Google Scholar] [CrossRef]
  35. Tong, P.; Wu, L.; Yanchen, A.; Su, R.; Wu, Y. A novel eddy current sensor for displacement measurement with high accuracy and long range. IEEE Sens. J. 2023, 23, 12066–12075. [Google Scholar] [CrossRef]
  36. Borovik, S.; Kuteynikova, M.; Sekisov, Y. Reducing the Impact of Influence Factors on the Measurement Results from Single-Coil Eddy Current Sensors. Sensors 2022, 23, 351. [Google Scholar] [CrossRef]
  37. Zhu, H.M.; Wu, Y.Y.; Zheng, W.H.; Huang, Z.W. In-plane and out-of-plane displacement measurement by ultrasonic speckle correlation method (USCM). Arch. Appl. Mech. 2006, 75, 521–526. [Google Scholar] [CrossRef]
  38. Sumi, C. Displacement vector measurement using instantaneous ultrasound signal phase-multidimensional autocorrelation and Doppler methods. IEEE Trans. Ultrason. Ferroelectr. Freq. Control 2008, 55, 24–43. [Google Scholar] [CrossRef]
  39. Berkovic, G.; Shafir, E. Optical methods for distance and displacement measurements. Adv. Opt. Photonics 2012, 4, 441–471. [Google Scholar] [CrossRef]
  40. Fu, X.B.; Liu, B.; Zhang, Y.C. A displacement measurement system based on optical triangulation method. Optoelectron. Lett. 2011, 7, 380–383. [Google Scholar] [CrossRef]
  41. Chen, R.; Li, X.; Wang, X.; Li, J.; Xue, G.; Zhou, Q.; Ni, K. A planar pattern based calibration method for high precision structured laser triangulation measurement. In Proceedings of the Optical Metrology and Inspection for Industrial Applications VI, Hangzhou, China, 20–23 October 2019; SPIE: Cergy, France, 2019; Volume 11189, pp. 212–218. [Google Scholar]
  42. Bai, J.; Li, J.; Wang, X.; Li, X. Research progress of high precision chromatic confocal displacement measurement technology. Laser Optoelectron. Prog 2023, 60, 0312014. [Google Scholar] [CrossRef]
  43. Sato, R.; Li, X.; Fischer, A.; Chen, L.C.; Chen, C.; Shimomura, R.; Gao, W. Signal processing and artificial intelligence for dual-detection confocal probes. Int. J. Precis. Eng. Manuf. 2024, 25, 199–223. [Google Scholar] [CrossRef]
  44. Bai, J.; Li, X.; Wang, X.; Wang, J.; Ni, K.; Zhou, Q. Self-reference dispersion correction for chromatic confocal displacement measurement. Opt. Lasers Eng. 2021, 140, 106540. [Google Scholar] [CrossRef]
  45. Bai, J.; Li, X.; Wang, X.; Zhou, Q.; Ni, K. Chromatic confocal displacement sensor with optimized dispersion probe and modified centroid peak extraction algorithm. Sensors 2019, 19, 3592. [Google Scholar] [CrossRef]
  46. Kurtev, K.I.; Trujillo-Sevilla, J.M.; Rodríguez-Ramos, J.M. Long-Distance Measurements Using a Chromatic Confocal Sensor. Appl. Sci. 2024, 14, 9943. [Google Scholar] [CrossRef]
  47. Bai, J.; Li, X.; Zhou, Q.; Ni, K.; Wang, X. Improved chromatic confocal displacement-sensor based on a spatial-bandpass-filter and an X-shaped fiber-coupler. Opt. Express 2019, 27, 10961–10973. [Google Scholar] [CrossRef]
  48. Bai, J.; Wang, X.; Li, X.; Zhou, Q.; Ni, K. Design and testing of a chromatic dispersion system for displacement application by using a spatial-bandpass-filter. In Proceedings of the Tenth International Symposium on Precision Engineering Measurements and Instrumentation, Kunming, China, 8–10 August 2018; SPIE: Cergy, France, 2019; Volume 11053, pp. 905–911. [Google Scholar] [CrossRef]
  49. Badami, V.; de Groot, P. Displacement measuring interferometry. Handb. Opt. Dimens. Metrol. 2013, 4, 180–199. [Google Scholar]
  50. Haitjema, H. Calibration of displacement laser interferometer systems for industrial metrology. Sensors 2019, 19, 4100. [Google Scholar] [CrossRef]
  51. Požar, T.; Gregorčič, P.; Možina, J. A precise and wide-dynamic-range displacement-measuring homodyne quadrature laser interferometer. Appl. Phys. B 2011, 105, 575–582. [Google Scholar] [CrossRef]
  52. Yan, H.; Duan, H.Z.; Li, L.T.; Liang, Y.R.; Luo, J.; Yeh, H.C. A dual-heterodyne laser interferometer for simultaneous measurement of linear and angular displacements. Rev. Sci. Instruments 2015, 86, 123102. [Google Scholar] [CrossRef] [PubMed]
  53. Huang, G.; Cui, C.; Lei, X.; Li, Q.; Yan, S.; Li, X.; Wang, G. A Review of Optical Interferometry for High-Precision Length Measurement. Micromachines 2024, 16, 6. [Google Scholar] [CrossRef] [PubMed]
  54. Zhang, W.; Wang, Y.; Du, H.; Zeng, Q.; Xiong, X. High-precision displacement measurement model for the grating interferometer system. Opt. Eng. 2020, 59, 045101. [Google Scholar] [CrossRef]
  55. Lee, J.Y.; Jiang, G.A. Displacement measurement using a wavelength-phase-shifting grating interferometer. Opt. Express 2013, 21, 25553–25564. [Google Scholar] [CrossRef]
  56. Lee, J.Y.; Chen, H.Y.; Hsu, C.C.; Wu, C.C. Optical heterodyne grating interferometry for displacement measurement with subnanometric resolution. Sens. Actuators A Phys. 2007, 137, 185–191. [Google Scholar] [CrossRef]
  57. Hsieh, H.L.; Pan, S.W. Development of a grating-based interferometer for six-degree-of-freedom displacement and angle measurements. Opt. Express 2015, 23, 2451–2465. [Google Scholar] [CrossRef] [PubMed]
  58. Guo, D.; Wang, M.; Hao, H. Displacement measurement using a laser feedback grating interferometer. Appl. Opt. 2015, 54, 9320–9325. [Google Scholar] [CrossRef]
  59. Fleming, A.J. A review of nanometer resolution position sensors: Operation and performance. Sens. Actuators A Phys. 2013, 190, 106–126. [Google Scholar] [CrossRef]
  60. Shimizu, Y.; Chen, L.C.; Kim, D.W.; Chen, X.; Li, X.; Matsukuma, H. An insight into optical metrology in manufacturing. Meas. Sci. Technol. 2021, 32, 042003. [Google Scholar] [CrossRef]
  61. Chen, S.; Xue, S.; Xiong, Y.; Peng, X.; Dai, Y. Research progress of ultra-precision measurement of optical surfaces for manufacturing. Laser Optoelectron. Prog. 2023, 60, 0312011. [Google Scholar]
  62. Teimel, A. Technology and applications of grating interferometers in high-precision measurement. Precis. Eng. 1992, 14, 147–154. [Google Scholar] [CrossRef]
  63. Fan, K.C.; Wang, H.Y.; Yang, H.W.; Chen, L.M. Techniques of multi-degree-of-freedom measurement on the linear motion errors of precision machines. Adv. Opt. Technol. 2014, 3, 375–386. [Google Scholar] [CrossRef]
  64. Xu, X.; Tan, Y.; Mu, H.; Li, Y.; Wang, J.; Jin, J. Laser interferometric multi-degree-of-freedom measurement technology in space gravitational-wave detection. Laser Optoelectron. Prog. 2023, 60, 0312006. [Google Scholar]
  65. Bai, Y.; Hu, P.; Lu, Y.; Li, Z.; Zhang, Z.; Tan, J. A six-axis heterodyne interferometer system for the joule balance. IEEE Trans. Instrum. Meas. 2016, 66, 1579–1585. [Google Scholar] [CrossRef]
  66. Cao, F.; Lu, S.; Luo, L.; Wang, S.; Li, X. In-plane and out-of-plane 4-DOF encoder based on grating. In Proceedings of the Second International Conference on Power Electronics and Artificial Intelligence (PEAI 2025), Guangzhou, China, 17–19 January 2025; SPIE: Cergy, France, 2025; Volume 13657, pp. 731–737. [Google Scholar]
  67. Matsukuma, H.; Ishizuka, R.; Furuta, M.; Li, X.; Shimizu, Y.; Gao, W. Reduction in cross-talk errors in a six-degree-of-freedom surface encoder. Nanomanuf. Metrol. 2019, 2, 111–123. [Google Scholar] [CrossRef]
  68. Hori, Y.; Gonda, S.; Bitou, Y.; Watanabe, A.; Nakamura, K. Periodic error evaluation system for linear encoders using a homodyne laser interferometer with 10 picometer uncertainty. Precis. Eng. 2018, 51, 388–392. [Google Scholar] [CrossRef]
  69. Wang, S.; Lu, S.; Cao, F.; Luo, L.; Wang, Y.; Li, X. Absolute grating encoder with high-precision and large measurement range. In Proceedings of the Holography, Diffractive Optics, and Applications XIV, Nantong, China, 12–15 October 2024; SPIE: Cergy, France, 2024; Volume 13240, pp. 182–187. [Google Scholar]
  70. Wang, S.; Zhu, J.; Shi, N.; Luo, L.; Wen, Y.; Li, X. Modeling and test of an absolute four-degree-of-freedom (DOF) grating encoder. In Proceedings of the Optical Metrology and Inspection for Industrial Applications IX, Online, 5–12 December 2022; SPIE: Cergy, France, 2022; Volume 12319, pp. 72–78. [Google Scholar]
  71. Wang, S.; Luo, L.; Li, X. Design and parameter optimization of zero position code considering diffraction based on deep learning generative adversarial networks. Nanomanuf. Metrol. 2024, 7, 2. [Google Scholar] [CrossRef]
  72. Wang, S.; Luo, L.; Zhu, J.; Shi, N.; Li, X. An ultra-precision absolute-type multi-degree-of-freedom grating encoder. Sensors 2022, 22, 9047. [Google Scholar] [CrossRef]
  73. Li, X.; Wang, H.; Ni, K.; Zhou, Q.; Mao, X.; Zeng, L.; Wang, X.; Xiao, X. Two-probe optical encoder for absolute positioning of precision stages by using an improved scale grating. Opt. Express 2016, 24, 21378–21391. [Google Scholar] [CrossRef] [PubMed]
  74. Wang, S.; Luo, L.; Cao, F.; Li, X. Nanoscale Absolute Grating Encoder with Wide Range Directional Discrimination Based on Asymmetric Code and Hybrid Positioning. In Proceedings of the CLEO: Fundamental Science, Charlotte, NC, USA, 5–10 May 2024; Optica Publishing Group: Washington, DC, USA, 2024; p. JTh2A-159. [Google Scholar]
  75. Goßler, S.; Bertolini, A.; Born, M.; Chen, Y.; Dahl, K.; Gering, D.; Gräf, C.; Heinzel, G.; Hild, S.; Kawazoe, F.; et al. The AEI 10 m prototype interferometer. Class. Quantum Gravity 2010, 27, 084023. [Google Scholar] [CrossRef]
  76. Jang, Y.S.; Kim, S.W. Compensation of the refractive index of air in laser interferometer for distance measurement: A review. Int. J. Precis. Eng. Manuf. 2017, 18, 1881–1890. [Google Scholar] [CrossRef]
  77. Ciddor, P.E. Refractive index of air: New equations for the visible and near infrared. Appl. Opt. 1996, 35, 1566–1573. [Google Scholar] [CrossRef]
  78. Wu, J.; Liu, W.; Lu, Y.; Zhang, Y.; Sun, S.; Li, J.; Zhou, Y. Calibration of the geometric error parameters of laser trackers in site environments. IEEE Sens. J. 2023, 23, 7077–7086. [Google Scholar] [CrossRef]
  79. Pengcheng, H.U.; Chang, D.; Tan, J.; Yang, R.; Yang, H.; Haijin, F.U. Displacement measuring grating interferometer: A review. Front. Inf. Technol. Electron. Eng. 2019, 20, 631–654. [Google Scholar] [CrossRef]
  80. Li, X.; Shi, Y.; Wang, P.; Ni, K.; Zhou, Q.; Wang, X. A compact design of optical scheme for a two-probe absolute surface encoders. In Proceedings of the Tenth International Symposium on Precision Engineering Measurements and Instrumentation, Kunming, China, 8–10 August 2018; SPIE: Cergy, France, 2019; Volume 11053, pp. 960–966. [Google Scholar]
  81. Wang, H.; Shi, Y.; Li, X.; Ni, K.; Zhou, Q.; Wang, X. Improvement of absolute positioning of precision stage based on cooperation the zero position pulse signal and incremental displacement signal. In Proceedings of the Journal of Physics: Conference Series, Dalian, China, 26–29 October 2017; IOP Publishing: Bristol, UK, 2018; Volume 986, p. 012002. [Google Scholar]
  82. Liao, B.; Wang, S.; Lin, J.; Dou, Y.; Wang, X.; Li, X. A research on compact short-distance grating interferometer based on ridge prism. In Proceedings of the 2021 International Conference on Optical Instruments and Technology: Optoelectronic Measurement Technology and Systems, Online, 8–10 April 2022; SPIE: Cergy, France, 2022; Volume 12282, pp. 411–418. [Google Scholar]
  83. Gao, W.; Kim, S.W.; Bosse, H.; Haitjema, H.; Chen, Y.; Lu, X.; Knapp, W.; Weckenmann, A.; Estler, W.; Kunzmann, H. Measurement technologies for precision positioning. CIRP Ann. 2015, 64, 773–796. [Google Scholar] [CrossRef]
  84. Li, X.; Yuan, W.; Ni, K.; Zhou, Q.; Wang, X. A two-probe linear encoder by using an arrayed scale grating stitched by multiple separate short gratings. In Proceedings of the Tenth International Symposium on Precision Engineering Measurements and Instrumentation, Kunming, China, 8–10 August 2018; SPIE: Cergy, France, 2019; Volume 11053, pp. 207–212. [Google Scholar]
  85. Ni, K.; Wang, H.; Li, X.; Wang, X.; Xiao, X.; Zhou, Q. Measurement uncertainty evaluation of the three degree of freedom surface encoder. In Proceedings of the Optical Metrology and Inspection for Industrial Applications IV, Beijing, China, 12–14 October 2016; SPIE: Cergy, France, 2016; Volume 10023, pp. 187–192. [Google Scholar]
  86. Luo, L.; Cao, F.; Wang, S.; Li, X. Design and demonstration of a large range precision grating sensor for simultaneous measurement of out-of-plane motions. Measurement 2025, 249, 116799. [Google Scholar] [CrossRef]
  87. Shi, Y.; Zhou, Q.; Li, X.; Ni, K.; Wang, X. Design and testing of a linear encoder capable of measuring absolute distance. Sens. Actuators A Phys. 2020, 308, 111935. [Google Scholar] [CrossRef]
  88. Shi, Y.; Ni, K.; Li, X.; Zhou, Q.; Wang, X. Highly accurate, absolute optical encoder using a hybrid-positioning method. Opt. Lett. 2019, 44, 5258–5261. [Google Scholar] [CrossRef]
  89. Li, X.; Zhou, Q.; Zhu, X.; Lu, H.; Yang, L.; Ma, D.; Sun, J.; Ni, K.; Wang, X. Holographic fabrication of an arrayed one-axis scale grating for a two-probe optical linear encoder. Opt. Express 2017, 25, 16028–16039. [Google Scholar] [CrossRef]
  90. Zhong, Z.; Li, J.; Lu, T.; Li, X. High dynamic wavefront stability control for high-uniformity periodic microstructure fabrication. Precis. Eng. 2025, 93, 216–223. [Google Scholar] [CrossRef]
  91. Shao, C.; Li, X. Technologies for Fabricating Large-Size Diffraction Gratings. Sensors 2025, 25, 1990. [Google Scholar] [CrossRef]
  92. Xue, G.; Lin, L.; Zhai, Q.; Zeng, C.; Wang, X.; Li, X. Development of dielectric-film-based polarization modulation scheme for patterning highly uniform 2d array structures with periodic tunability. Opt. Lasers Eng. 2023, 167, 107627. [Google Scholar] [CrossRef]
  93. Gao, X.; Li, J.; Zhong, Z.; Li, X. Global alignment reference strategy for laser interference lithography pattern arrays. Microsyst. Nanoeng. 2025, 11, 41. [Google Scholar] [CrossRef]
  94. Li, X.; Gao, W.; Shimizu, Y.; Ito, S. A two-axis Lloyd’s mirror interferometer for fabrication of two-dimensional diffraction gratings. CIRP Ann. 2014, 63, 461–464. [Google Scholar] [CrossRef]
  95. Li, X.; Lu, H.; Zhou, Q.; Wu, G.; Ni, K.; Wang, X. An orthogonal type two-axis Lloyd’s mirror for holographic fabrication of two-dimensional planar scale gratings with large area. Appl. Sci. 2018, 8, 2283. [Google Scholar] [CrossRef]
  96. Xue, G.; Lu, H.; Li, X.; Zhou, Q.; Wu, G.; Wang, X.; Zhai, Q.; Ni, K. Patterning nanoscale crossed grating with high uniformity by using two-axis Lloyd’s mirrors based interference lithography. Opt. Express 2020, 28, 2179–2191. [Google Scholar] [CrossRef]
  97. Xue, G.; Zhai, Q.; Lu, H.; Zhou, Q.; Ni, K.; Lin, L.; Wang, X.; Li, X. Polarized holographic lithography system for high-uniformity microscale patterning with periodic tunability. Microsyst. Nanoeng. 2021, 7, 31. [Google Scholar] [CrossRef]
  98. Wang, G.; Xue, G.; Zhai, Q.; Zhu, J.; Yu, K.; Huang, G.; Wang, M.; Zhong, A.; Zhu, L.; Yan, S.; et al. Planar diffractive grating for magneto-optical trap application: Fabrication and testing. Appl. Opt. 2021, 60, 9358–9364. [Google Scholar] [CrossRef]
  99. Kimura, A.; Gao, W.; Kim, W.; Hosono, K.; Shimizu, Y.; Shi, L.; Zeng, L. A sub-nanometric three-axis surface encoder with short-period planar gratings for stage motion measurement. Precis. Eng. 2012, 36, 576–585. [Google Scholar] [CrossRef]
  100. Wang, S.; Liao, B.; Shi, N.; Li, X. A compact and high-precision three-degree-of-freedom grating encoder based on a quadrangular frustum pyramid prism. Sensors 2023, 23, 4022. [Google Scholar] [CrossRef]
  101. Li, X.; Zhu, X.; Zhou, Q.; Wang, H.; Ni, K. Low-cost lithography for fabrication of one-dimensional diffraction gratings by using laser diodes. In Proceedings of the 2015 International Conference on Optical Instruments and Technology: Micro/Nano Photonics and Fabrication, Beijing, China, 17–19 May 2015; SPIE: Cergy, France, 2015; Volume 9624, pp. 51–56. [Google Scholar]
  102. Li, X.; Shimizu, Y.; Ito, S.; Gao, W. Fabrication of scale gratings for surface encoders by using laser interference lithography with 405 nm laser diodes. Int. J. Precis. Eng. Manuf. 2013, 14, 1979–1988. [Google Scholar] [CrossRef]
  103. Li, J.; Tan, S.H.; Li, X. Efficient fabrication and optimization of large-area metallic diffraction gratings. In Proceedings of the International Conference on Advanced Imaging, Detection, and Signal Processing (AIDSP 2024), Harbin, China, 27–29 December 2024; SPIE: Cergy, France, 2025; Volume 13693, pp. 14–19. [Google Scholar]
  104. Hu, H.; Qiu, X.; Wang, J.; Ju, A.; Zhang, Y. Subdivision and direction recognition of λ/16 of orthogonal fringes for nanometric measurement. Appl. Opt. 2009, 48, 6479–6484. [Google Scholar] [CrossRef] [PubMed]
  105. Ma, Y.; Li, J.; Zhao, P.; Li, X. Design of solving circuit for 6-DOF heterodyne grating interferometry system. In Proceedings of the Fourth International Conference on Mechatronics and Intelligent Control (MIC 2024), Nanchang, China, 27–29 December 2024; SPIE: Cergy, France, 2025; Volume 13643, pp. 472–478. [Google Scholar]
  106. Luo, L.; Cao, F.; Wang, S.; Li, X. Coupling error analysis and compensation algorithm for three-degree-of-freedom grating autocollimator. In Proceedings of the International Conference on Advanced Imaging, Detection, and Signal Processing (AIDSP 2024), Harbin, China, 27–29 December 2024; SPIE: Cergy, France, 2025; Volume 13693, pp. 231–238. [Google Scholar]
  107. Li, X.; Gao, W.; Muto, H.; Shimizu, Y.; Ito, S.; Dian, S. A six-degree-of-freedom surface encoder for precision positioning of a planar motion stage. Precis. Eng. 2013, 37, 771–781. [Google Scholar] [CrossRef]
  108. Yu, K.; Zhu, J.; Yuan, W.; Zhou, Q.; Xue, G.; Wu, G.; Wang, X.; Li, X. Two-channel six degrees of freedom grating-encoder for precision-positioning of sub-components in synthetic-aperture optics. Opt. Express 2021, 29, 21113–21128. [Google Scholar] [CrossRef]
  109. Li, X.; Shi, Y.; Xiao, X.; Zhou, Q.; Wu, G.; Lu, H.; Ni, K. Design and testing of a compact optical prism module for multi-degree-of-freedom grating interferometry application. Appl. Sci. 2018, 8, 2495. [Google Scholar] [CrossRef]
  110. Shimizu, Y.; Ito, T.; Li, X.; Kim, W.; Gao, W. Design and testing of a four-probe optical sensor head for three-axis surface encoder with a mosaic scale grating. Meas. Sci. Technol. 2014, 25, 094002. [Google Scholar] [CrossRef]
  111. Adany, P.; Allen, C.; Hui, R. Chirped lidar using simplified homodyne detection. J. Light. Technol. 2009, 27, 3351–3357. [Google Scholar] [CrossRef]
  112. Lin, X.; Su, X.; Wang, J.; Sun, Y.; Hu, P. Laser Interferometer Technology and Instruments for Sub-Nanometer and Picometer Displacement Measurements. Laser Optoelectron. Prog. 2023, 60, 312016. [Google Scholar]
  113. Watchi, J.; Cooper, S.; Ding, B.; Mow-Lowry, C.M.; Collette, C. Contributed Review: A review of compact interferometers. Rev. Sci. Instruments 2018, 89, 121501. [Google Scholar] [CrossRef]
  114. Köning, R.; Wimmer, G.; Witkovskỳ, V. The statistical uncertainty of the Heydemann correction: A practical limit of optical quadrature homodyne interferometry. Meas. Sci. Technol. 2015, 26, 084004. [Google Scholar] [CrossRef]
  115. Vishnyakov, G.; Minaev, V.; Shumsky, E. Homodyne quadrature displacement interferometer. Experimental results. Opt. Spectrosc. 2022, 130, 327–335. [Google Scholar] [CrossRef]
  116. Li, X.; Shimizu, Y.; Ito, T.; Cai, Y.; Ito, S.; Gao, W. Measurement of six-degree-of-freedom planar motions by using a multiprobe surface encoder. Opt. Eng. 2014, 53, 122405. [Google Scholar] [CrossRef]
  117. Xu, X.; Gu, D.; Gao, S.; Sun, L.; Lu, X.; Wang, K.; Bai, J. Back to Michelson interferometer: A precise inspection system for industrial intricate structures defect detection. Meas. Sci. Technol. 2023, 35, 035026. [Google Scholar] [CrossRef]
  118. Kim, J.A.; Kim, J.W.; Kang, C.S.; Eom, T.B.; Ahn, J. A digital signal processing module for real-time compensation of nonlinearity in a homodyne interferometer using a field-programmable gate array. Meas. Sci. Technol. 2008, 20, 017003. [Google Scholar] [CrossRef]
  119. Li, X.H.; Shimizu, Y.; Muto, H.; Ito, S.; Gao, W. Design of a three-axis surface encoder with a blue-ray laser diode. Key Eng. Mater. 2012, 523, 913–918. [Google Scholar] [CrossRef]
  120. Shaddock, D.A. Digitally enhanced heterodyne interferometry. Opt. Lett. 2007, 32, 3355–3357. [Google Scholar] [CrossRef] [PubMed]
  121. Zhu, J.; Wang, G.; Xue, G.; Zhou, Q.; Li, X. Heterodyne three-degree-of-freedom grating interferometer for ultra-precision positioning of lithography machine. In Proceedings of the 2021 International Conference on Optical Instruments and Technology: Optoelectronic Measurement Technology and Systems, Online, 8–10 April 2022; SPIE: Cergy, France, 2022; Volume 12282, pp. 21–32. [Google Scholar]
  122. Zhao, P.; Gao, L.; Cui, C.; Ma, Y.; Yang, M.; Liu, L.; Li, X. A grating interferometer for improving optical subdivision in Z-axis measurement. In Proceedings of the Optical Metrology and Inspection for Industrial Applications XI, Nantong, China, 12–15 October 2024; SPIE: Cergy, France, 2024; Volume 13241, pp. 238–246. [Google Scholar]
  123. Dong Nguyen, T.; Higuchi, M.; Tung Vu, T.; Wei, D.; Aketagawa, M. 10-pm-order mechanical displacement measurements using heterodyne interferometry. Appl. Opt. 2020, 59, 8478–8485. [Google Scholar] [CrossRef] [PubMed]
  124. Dong, N.T.; Tai, N.T.; Hoang, D.V.; Mai, N.T.P.; Tung, V.T.; Thang, V.T. Sub-nanometer displacement measurement using heterodyne interferometer and down-beat frequency technique. In Proceedings of the Regional Conference in Mechanical Manufacturing Engineering, Hanoi, Vietnam, 10–12 June 2022; Springer: Singapore, 2021; pp. 1170–1176. [Google Scholar]
  125. Cui, C.; Yang, M.; Liu, L.; Zhao, P.; Gao, L.; Li, X. A quasi-common-path heterodyne grating interferometer for six-degree-of-freedom measurement. In Proceedings of the Optical Metrology and Inspection for Industrial Applications XI, Nantong, China, 12–15 October 2024; SPIE: Cergy, France, 2024; Volume 13241, pp. 115–125. [Google Scholar]
  126. Liu, L.; Cui, C.; Yang, M.; Zhao, P.; Li, X. Blazed grating collimation strategy for a high-precision heterodyne grating interferometer. In Proceedings of the International Conference on Advanced Imaging, Detection, and Signal Processing (AIDSP 2024), Harbin, China, 27–29 December 2024; SPIE: Cergy, France, 2025; Volume 13693, pp. 81–87. [Google Scholar]
  127. Kunzmann, H.; Pfeifer, T.; Flügge, J. Scales vs. laser interferometers performance and comparison of two measuring systems. CIRP Ann. 1993, 42, 753–767. [Google Scholar] [CrossRef]
  128. Deng, J.; Yan, X.; Wei, C.; Lu, Y.; Li, M.; Xiang, X.; Jia, W.; Zhou, C. Eightfold optical encoder with high-density grating. Appl. Opt. 2018, 57, 2366–2375. [Google Scholar] [CrossRef] [PubMed]
  129. Eom, T.; Choi, T.; Lee, K.; Choi, H.; Lee, S. A simple method for thecompensation of the nonlinearity in the heterodyne interferometer. Meas. Sci. Technol. 2002, 13, 222. [Google Scholar] [CrossRef]
  130. Schmitz, T.; Chu, D.; Houck, L. First-order periodic error correction: Validation for constant and non-constant velocities with variable error magnitudes. Meas. Sci. Technol. 2006, 17, 3195. [Google Scholar] [CrossRef]
  131. Zhu, J.; Yu, K.; Xue, G.; Shi, N.; Zhou, Q.; Wang, X.; Li, X. A simplified two-phase differential decoding algorithm for high precision grating encoder. In Proceedings of the Optical Metrology and Inspection for Industrial Applications VIII, Nantong, China, 10–20 October 2021; SPIE: Cergy, France, 2021; Volume 11899, pp. 49–58. [Google Scholar]
  132. Zhu, J.; Yu, K.; Xue, G.; Zhou, Q.; Wang, X.; Li, X. An improved signal filtering strategy based on EMD algorithm for ultrahigh precision grating encoder. In Proceedings of the Real-time Photonic Measurements, Data Management, and Processing VI, Nantong, China, 10–20 October 2021; SPIE: Cergy, France, 2021; Volume 11902, pp. 140–149. [Google Scholar]
  133. Han, Y.; Ni, K.; Li, X.; Wu, G.; Yu, K.; Zhou, Q.; Wang, X. An fpga platform for next-generation grating encoders. Sensors 2020, 20, 2266. [Google Scholar] [CrossRef]
  134. Shi, N.; Wang, S.; Xue, G.; Liu, M.; Han, Y.; Zhu, J.; Ni, K.; Wang, X.; Li, X. A real-time processing system for dual-channel six-degree-of-freedom grating ruler based on FPGA. In Proceedings of the Optical Design and Testing XI, Nantong, China, 10–20 October 2021; SPIE: Cergy, France, 2021; Volume 11895, pp. 250–260. [Google Scholar]
  135. Wang, Y.; Shi, N.; Li, L.; Ni, K.; Li, X. Instantaneous grating signal subdivision system with non-linear Kalman filters. In Proceedings of the Optical Design and Testing XII, Online, 5–12 December 2022; SPIE: Cergy, France, 2022; Volume 12315, pp. 90–96. [Google Scholar]
  136. Shi, N.; Zhu, J.; Wang, S.; Han, Y.; Yu, K.; Li, X.; Wang, X. FPGA displacement calculation platform and test of two-phase grating interference encoder. In Proceedings of the 2021 International Conference on Optical Instruments and Technology: Optoelectronic Measurement Technology and Systems, Online, 8–10 April 2022; SPIE: Cergy, France, 2022; Volume 12282, pp. 372–383. [Google Scholar]
  137. Heydemann, P.L. Determination and correction of quadrature fringe measurement errors in interferometers. Appl. Opt. 1981, 20, 3382–3384. [Google Scholar] [CrossRef]
  138. Zhang, G.; Xu, L.; Ge, Q.; Wu, X.; Yu, B. Ameliorated PGC demodulation scheme using Taubin least squares fitting of ellipse in fiber optic interferometric sensors. Opt. Fiber Technol. 2023, 80, 103374. [Google Scholar] [CrossRef]
  139. Li, J.; Zhang, Q. Normalized and orthogonal phase retrieval algorithm based on lissajous ellipse fitting. Opt. Lasers Eng. 2022, 151, 106917. [Google Scholar] [CrossRef]
  140. Wu, C.m.; Su, C.s.; Peng, G.S. Correction of nonlinearity in one-frequency optical interferometry. Meas. Sci. Technol. 1996, 7, 520. [Google Scholar] [CrossRef]
  141. Požar, T.; Možina, J. Enhanced ellipse fitting in a two-detector homodyne quadrature laser interferometer. Meas. Sci. Technol. 2011, 22, 085301. [Google Scholar] [CrossRef]
  142. Köning, R.; Wimmer, G.; Witkovskỳ, V. Ellipse fitting by nonlinear constraints to demodulate quadrature homodyne interferometer signals and to determine the statistical uncertainty of the interferometric phase. Meas. Sci. Technol. 2014, 25, 115001. [Google Scholar] [CrossRef]
  143. Fitzgibbon, A.; Pilu, M.; Fisher, R.B. Direct least square fitting of ellipses. IEEE Trans. Pattern Anal. Mach. Intell. 1999, 21, 476–480. [Google Scholar] [CrossRef]
  144. Long, C.; Hu, Q.; Zhao, M.; Li, D.; Ouyang, Z.; Yan, D.M. A triple-stage robust ellipse fitting algorithm based on outlier removal. IEEE Trans. Instrum. Meas. 2023, 72, 1–14. [Google Scholar] [CrossRef]
  145. Yang, T.; Yan, L.; Chen, B.; Liu, Y.; Tian, Q. Signal processing method of phase correction for laser heterodyne interferometry. Opt. Lasers Eng. 2014, 57, 93–100. [Google Scholar] [CrossRef]
  146. Köchert, P.; Flügge, J.; Weichert, C.; Köning, R.; Manske, E. Phase measurement of various commercial heterodyne He–Ne-laser interferometers with stability in the picometer regime. Meas. Sci. Technol. 2012, 23, 074005. [Google Scholar] [CrossRef]
  147. Hu, P.; Wang, J.; Lin, X.; Xing, X.; Fu, H.; Tan, J. Phase measurement method based on digital dual frequency comb for high-precision high-speed heterodyne interferometry. IEEE Sens. J. 2023, 23, 9707–9715. [Google Scholar] [CrossRef]
  148. Lu, S.; Wang, Y.; Li, X. Robust phase measurement of high-speed signals using frequency tracking lock-in method. In Proceedings of the Holography, Diffractive Optics, and Applications XIV, Nantong, China, 12–15 October 2024; SPIE: Cergy, France, 2024; Volume 13240, pp. 188–197. [Google Scholar]
  149. Shaddock, D.; Ware, B.; Halverson, P.; Spero, R.; Klipstein, B. Overview of the LISA Phasemeter. In Proceedings of the AIP Conference Proceedings, Greenbelt, MD, USA, 19–23 June 2006; American Institute of Physics: College Park, MD, USA, 2006; Volume 873, pp. 654–660. [Google Scholar]
  150. Liu, H.; Luo, Z.; Jin, G. The development of phasemeter for Taiji space gravitational wave detection. Microgravity Sci. Technol. 2018, 30, 775–781. [Google Scholar] [CrossRef]
  151. Ma, Y.; Wang, Y.; Lu, S.; Wang, S.; Li, X. High-accuracy and real-time phase solution algorithm based on extended Kalman filter for multi-degree-of-freedom grating interferometer. In Proceedings of the Optoelectronic Devices and Integration XIII, Nantong, China, 12–15 October 2024; SPIE: Cergy, France, 2024; Volume 13236, pp. 99–108. [Google Scholar]
  152. Wang, W.; Lin, J.; Jin, S.; Bai, Y.; Liu, L.; Liu, Z.; Sun, Y.; Li, W. A Real-time Determination Method for Homodyne Grating Interferometer Signal Nonlinear Error based on Duty Cycle Measurement of States. IEEE Trans. Instrum. Meas. 2024, 73, 7004208. [Google Scholar] [CrossRef]
  153. Yuan, B.; Yan, H.M.; Cao, X.Q. A new subdivision method for grating-based displacement sensor using imaging array. Opt. Lasers Eng. 2009, 47, 90–95. [Google Scholar] [CrossRef]
  154. Niu, R.; Chen, G.; Li, R.; Jia, Z.; Feng, Z. A high subdivision and light fluctuation-insensitive approach for grating displacement sensors. Opt. Lasers Eng. 2025, 184, 108545. [Google Scholar] [CrossRef]
  155. Hua-hui, Y.; Wei-li, F.; Fu, L. Subdivision Design of Grating Moire Fringes and Simulation by Simulink. J. Astronaut. Metrol. Meas. 2016, 36, 29. [Google Scholar]
  156. Ye, W.; Zhang, M.; Zhu, Y.; Wang, L.; Hu, J.; Li, X.; Hu, C. Ultraprecision real-time displacements calculation algorithm for the grating interferometer system. Sensors 2019, 19, 2409. [Google Scholar] [CrossRef]
  157. Jan, C.M.; Liu, C.S.; Lin, C.Y. Development of low-cost heterodyne interferometer with virtual electronic phasemeter. Meas. Control 2022, 55, 229–238. [Google Scholar] [CrossRef]
  158. Bo, L.; Jinyan, L. Research on signal subdivision of grating sensor. In Proceedings of the Proceedings of 2011 6th International Forum on Strategic Technology, Harbin, China, 22–24 August 2011; IEEE: New York, NY, USA, 2011; Volume 2, pp. 1235–1238. [Google Scholar]
  159. Yang, Z.; Zhang, Y. Grating subdivision based on single chip computer. In Proceedings of the Second International Conference on Optics and Communication Technology (ICOCT 2022), Hefei, China, 15–17 July 2022; SPIE: Cergy, France, 2022; Volume 12473, pp. 62–67. [Google Scholar]
  160. Tang, J.; Peng, H.; Yang, P.; Guo, S.; Sun, W.; Jin, L.; Xie, K.; Li, M. Design, Analysis, and Implementation of the Subdivision Interpolation Technique for the Grating Interferometric Micro-Displacement Sensor. Photonics 2025, 12, 64. [Google Scholar] [CrossRef]
  161. Chang, C.P.; Chang, S.C.; Wang, Y.C.; He, P.Y. A Novel Analog Interpolation Method for Heterodyne Laser Interferometer. Micromachines 2023, 14, 696. [Google Scholar] [CrossRef]
  162. Hussain, T.; Amin, S.; Zabit, U.; Bernal, O.D.; Bosch, T. A high performance real-time interferometry sensor system architecture. Microprocess. Microsystems 2019, 64, 23–33. [Google Scholar] [CrossRef]
  163. Straube, G.; Fischer Calderón, J.S.; Ortlepp, I.; Füßl, R.; Manske, E. A heterodyne interferometer with separated beam paths for high-precision displacement and angular measurements. Nanomanuf. Metrol. 2021, 4, 200–207. [Google Scholar] [CrossRef]
  164. Huang, Y.; Che, S.; Zheng, G.; Zhu, W.; Zhou, Y.; Xue, Z. An adaptive filter for subdivision of circular grating signal of angle measurement. IEEE Trans. Instrum. Meas. 2022, 71, 1–10. [Google Scholar] [CrossRef]
  165. Zhu, W.; Xu, Y.; Huang, Y.; Xue, Z.; Ye, S. Subdivision method for nonorthogonal Moiré signals. IEEE Trans. Instrum. Meas. 2020, 70, 1–10. [Google Scholar] [CrossRef]
  166. Hou, H.; Cao, G.; Ding, H.; Li, K.; Wang, A. Phase-locked subdivision method of digital signal based on FPGA. J. Phys. Conf. Ser. Iop Publ. 2022, 2181, 012016. [Google Scholar] [CrossRef]
  167. Li, X.; Xiao, S.; Zhou, Q.; Ni, K.; Wang, X. A real-time distance measurement data processing platform for multi-axis grating interferometry type optical encoders. In Proceedings of the Tenth International Symposium on Precision Engineering Measurements and Instrumentation, Kunming, China, 8–10 August 2018; SPIE: Cergy, France, 2019; Volume 11053, pp. 1028–1035. [Google Scholar]
  168. Deng, W.; Peng, X. A high precision phase extraction method in heterodyne interferometer based on FPGA. J. Phys. Conf. Ser. Iop Publ. 2022, 2387, 012002. [Google Scholar] [CrossRef]
  169. Vera-Salas, L.A.; Moreno-Tapia, S.V.; Garcia-Perez, A.; de Jesus Romero-Troncoso, R.; Osornio-Rios, R.A.; Serroukh, I.; Cabal-Yepez, E. FPGA-based smart sensor for online displacement measurements using a heterodyne interferometer. Sensors 2011, 11, 7710–7723. [Google Scholar] [CrossRef]
  170. Zhang, E.; Chen, B.; Yan, L.; Yang, T.; Hao, Q.; Dong, W.; Li, C. Laser heterodyne interferometric signal processing method based on rising edge locking with high frequency clock signal. Opt. Express 2013, 21, 4638–4652. [Google Scholar] [CrossRef]
  171. Zhou, J.; He, W.; Yu, Y. Fringe-to-pulse counting demodulation of a laser interferometer for primary vibration calibration. Measurement 2023, 221, 113490. [Google Scholar] [CrossRef]
  172. Tucker, D. The history of the homodyne and synchrodyne. J. Br. Inst. Radio Eng. 1954, 14, 143–154. [Google Scholar] [CrossRef]
  173. Schumaker, B.L. Noise in homodyne detection. Opt. Lett. 1984, 9, 189–191. [Google Scholar] [CrossRef] [PubMed]
  174. Jakeman, E.; Oliver, C.; Pike, E. Optical homodyne detection. Adv. Phys. 1975, 24, 349–405. [Google Scholar] [CrossRef]
  175. Kazovsky, L.G. Optical heterodyning versus optical homodyning: A comparison. J. Opt. Commun. 1985, 6, 18–24. [Google Scholar] [CrossRef]
  176. Braunstein, S.L. Homodyne statistics. Phys. Rev. A 1990, 42, 474. [Google Scholar] [CrossRef] [PubMed]
  177. Armen, M.A.; Au, J.K.; Stockton, J.K.; Doherty, A.C.; Mabuchi, H. Adaptive homodyne measurement of optical phase. Phys. Rev. Lett. 2002, 89, 133602. [Google Scholar] [CrossRef]
  178. Glance, B. Performance of homodyne detection of binary PSK optical signals. J. Light. Technol. 2003, 4, 228–235. [Google Scholar] [CrossRef]
  179. Seimetz, M. Phase diversity receivers for homodyne detection of optical DQPSK signals. J. Light. Technol. 2006, 24, 3384–3391. [Google Scholar] [CrossRef]
  180. Puttnam, B.J.; Luís, R.S.; Delgado Mendinueta, J.M.; Sakaguchi, J.; Klaus, W.; Kamio, Y.; Nakamura, M.; Wada, N.; Awaji, Y.; Kanno, A.; et al. Self-homodyne detection in optical communication systems. Photonics 2014, 1, 110–130. [Google Scholar] [CrossRef]
  181. Luo, L.; Gao, L.; Wang, S.; Deng, F.; Wu, Y.; Li, X. An ultra-precision error estimation for a multi-axes grating encoder using quadrant photodetectors. In Proceedings of the Optical Metrology and Inspection for Industrial Applications IX, Online, 5–12 December 2022; SPIE: Cergy, France, 2022; Volume 12319, pp. 57–65. [Google Scholar]
  182. Cao, F.; Wang, S.; Luo, L.; Li, X. Displacement compensation algorithm for high-precision absolute six-degree-of-freedom grating encoder. In Proceedings of the Optical Metrology and Inspection for Industrial Applications XI, Nantong, China, 12–15 October 2024; SPIE: Cergy, France, 2024; Volume 13241, pp. 73–81. [Google Scholar]
  183. He, Y.; Orr, B.J. Optical heterodyne signal generation and detection in cavity ringdown spectroscopy based on a rapidly swept cavity. Chem. Phys. Lett. 2001, 335, 215–220. [Google Scholar] [CrossRef]
  184. Takenaka, T.; Tanaka, K.; Fukumitsu, O. Signal-to-noise ratio in optical heterodyne detection for Gaussian fields. Appl. Opt. 1978, 17, 3466–3471. [Google Scholar] [CrossRef]
  185. Huang, C.C. Optical heterodyne profilometer. Opt. Eng. 1984, 23, 365–370. [Google Scholar] [CrossRef]
  186. Abbas, G.; Chan, V.; Yee, T. A dual-detector optical heterodyne receiver for local oscillator noise suppression. J. Light. Technol. 1985, 3, 1110–1122. [Google Scholar] [CrossRef]
  187. Maznev, A.; Nelson, K.; Rogers, J. Optical heterodyne detection of laser-induced gratings. Opt. Lett. 1998, 23, 1319–1321. [Google Scholar] [CrossRef]
  188. Fried, D.L. Optical heterodyne detection of an atmospherically distorted signal wave front. Proc. IEEE 2005, 55, 57–77. [Google Scholar] [CrossRef]
  189. Cui, C.; Gao, L.; Zhao, P.; Yang, M.; Liu, L.; Ma, Y.; Huang, G.; Wang, S.; Luo, L.; Li, X. Towards multi-dimensional atomic-level measurement: Integrated heterodyne grating interferometer with zero dead-zone. Light. Adv. Manuf. 2025, 6, 1–14. [Google Scholar] [CrossRef]
  190. Zhu, J.; Wang, G.; Wang, S.; Li, X. A reflective-type heterodyne grating interferometer for three-degree-of-freedom subnanometer measurement. IEEE Trans. Instrum. Meas. 2022, 71, 1–9. [Google Scholar] [CrossRef]
  191. Yang, M.; Cui, C.; Liu, L.; Zhao, P.; Gao, L.; Li, X. An unfolding-based model for analysis of rotation and translation coupling in grating interferometers. In Proceedings of the Advanced Sensor Systems and Applications XIV, Nantong, China, 12–15 October 2024; SPIE: Cergy, France, 2024; Volume 13243, pp. 56–65. [Google Scholar]
  192. Liu, L.; Cui, C.; Yang, M.; Zhao, P.; Gao, L.; Li, X. An ultra-precision and real-time algorithm for 6-DOF displacements based on four grating interferometers. In Proceedings of the Advanced Sensor Systems and Applications XIV, Nantong, China, 12–15 October 2024; SPIE: Cergy, France, 2024; Volume 13243, pp. 71–81. [Google Scholar]
  193. Gao, L.; Zhu, J.; Huang, G.; Wang, G.; Wang, S.; Ni, K.; Wang, X.; Li, X. A polarization adjustment module for frequency aliasing of heterodyne grating interferometer. In Proceedings of the Optical Metrology and Inspection for Industrial Applications IX, Online, 5–12 December 2022; SPIE: Cergy, France, 2022; Volume 12319, pp. 362–369. [Google Scholar]
  194. Yang, Y.; Wang, G.; Zhu, J.; Wang, S.; Wu, Y.; Li, X. Method and system for phase measurement of the heterodyne interference system. In Proceedings of the 2021 International Conference on Optical Instruments and Technology: Optoelectronic Measurement Technology and Systems, Online, 8–10 April 2022; SPIE: Cergy, France, 2022; Volume 12282, pp. 431–442. [Google Scholar]
  195. Lu, Z.; Liu, B.; Zhang, K.; Lin, H.; Zhang, Y. A method for measuring the inclination of forgings based on an improved optimization algorithm for fitting ellipses. IEEE Trans. Instrum. Meas. 2022, 72, 1–11. [Google Scholar] [CrossRef]
  196. Wang, C.; Burnham-Fay, E.D.; Ellis, J.D. Real-time FPGA-based Kalman Filter for Constant and Non-constant Velocity Periodic Error Correction. Precis. Eng. 2016, 48, 133–143. [Google Scholar] [CrossRef]
  197. Kalisz, J. Review of methods for time interval measurements with picosecond resolution. Metrologia 2003, 41, 17–32. [Google Scholar] [CrossRef]
  198. Kim, M.S.; Kim, S.W. Two-way frequency-conversion phase measurement for high-speed and high-resolution heterodyne interferometry. Meas. Sci. Technol. 2004, 15, 2341. [Google Scholar] [CrossRef]
  199. Szplet, R.; Kalisz, J. Interpolating time counter with 100 ps resolution on a single FPGA device. Instrum. Meas. IEEE Trans. 2000, 49, 879–883. [Google Scholar] [CrossRef]
  200. Keranen, P.; Kostamovaara, J. A Wide Range, 4.2 ps(rms) Precision CMOS TDC With Cyclic Interpolators Based on Switched-Frequency Ring Oscillators. IEEE Trans. Circuits Syst. I Regul. Pap. 2015, 62, 2795–2805. [Google Scholar] [CrossRef]
  201. Koscielnik, D.; Szyduczynski, J.; Rzepka, D.; Andrysiewicz, W.; Miskowicz, M. Optimized design of successive approximation time-to-digital converter with single set of delay lines. In Proceedings of the 2016 Second International Conference on Event-based Control, Communication, and Signal Processing (EBCCSP), Krakow, Poland, 13–15 June 2016. [Google Scholar]
  202. Takayama, M.; Dosho, S.; Takeda, N.; Miyahara, M.; Matsuzawa, A. A time-domain architecture and design method of high speed A-to-D converters with standard cells. IEICE Trans. Electron. 2013, 6, 813–819. [Google Scholar] [CrossRef]
  203. Mantyniemi, A.; Rahkonen, T.; Kostamovaara, J. A CMOS Time-to-Digital Converter (TDC) Based On a Cyclic Time Domain Successive Approximation Interpolation Method. IEEE J. Solid-State Circuits 2009, 44, 3067–3078. [Google Scholar] [CrossRef]
  204. Mantyniemi, A.; Kostamovaara, J. Time-to-digital converter (TDC) based on startable ring oscillators and successive approximation. In Proceedings of the 2014 NORCHIP, Tampere, Finland, 27–28 October 2014. [Google Scholar]
  205. Henzler, S.; Koeppe, S.; Kamp, W.; Mulatz, H.; Schmitt-Landsiedel, D. 90nm 4.7ps-Resolution 0.7-LSB Single-Shot Precision and 19pJ-per-Shot Local Passive Interpolation Time-to-Digital Converter with On-Chip Characterization. In Proceedings of the IEEE International Solid-state Circuits Conference, San Francisco, CA, USA, 3–7 February 2008; pp. 548–635. [Google Scholar]
  206. Wu, J.; Shi, Z. The 10-ps wave union TDC: Improving FPGA TDC resolution beyond its cell delay. In Proceedings of the IEEE Nuclear Science Symposium Conference Record, Dresden, Germany, 19–25 October 2008; pp. 3440–3446. [Google Scholar]
  207. Li, G.; Zhang, C.; Liu, N.; Liu, C. A Real-Time Data Processing Method Based on Orthogonal Phase-Locked Amplification and Its Application in Frequency Domain Semi-Airborne Electromagnetic Detection. IEEJ Trans. Electr. Electron. Eng. 2020, 15, 1160–1168. [Google Scholar] [CrossRef]
  208. Jarrett, J.W.; Liu, X.; Nealey, P.F.; Vaia, R.A.; Cerullo, G.; Knappenberger, K.L. Communication: SHG-detected circular dichroism imaging using orthogonal phase-locked laser pulses. J. Chem. Phys. 2015, 142, 151101. [Google Scholar] [CrossRef]
  209. Hu, P.; Tan, J.; Yang, H.; Fu, H.; Wang, Q. Recent developments in heterodyne laser interferometry at Harbin Institute of Technology. In Proceedings of the Eighth International Symposium on Precision Engineering Measurement and Instrumentation, Chengdu, China, 8–11 August 2012; SPIE: Cergy, France, 2013; Volume 8759, pp. 898–906. [Google Scholar]
  210. Yang, H.; Fu, H.; Hu, P.; Yang, R.; Xu, X.; Liang, Y.; Di, C.; Tan, J. Ultra-precision and high-speed laser interferometric displacement measurement technology and instrument. Laser Optoelectron. Prog 2022, 59, 0922018. [Google Scholar]
  211. Keysight Technologies. N1225A Four-Channel High-Resolution Laser Axis Board [EB/OL]. 2025. Available online: https://www.keysight.com.cn/cn/zh/product/N1225A/four-channel-high-resolution-laser-axis-board.html (accessed on 27 June 2025).
  212. Zygo Corporation. Measurement Electronics [EB/OL]. 2025. Available online: https://www.zygo.com/products/nano-position-sensors/displacement-measuring-interferometers/measurement-electronics (accessed on 27 June 2025).
  213. Losic, N.A. A quadrature-based phase-locked loop. In Proceedings of the IECON’03. 29th Annual Conference of the IEEE Industrial Electronics Society (IEEE Cat. No. 03CH37468), Roanoke, VA, USA, 2–6 November 2003; IEEE: New York, NY, USA, 2003; Volume 3, pp. 2957–2962. [Google Scholar]
  214. Golestan, S.; Guerrero, J.M.; Abusorrah, A.; Al-Hindawi, M.M.; Al-Turki, Y. An adaptive quadrature signal generation-based single-phase phase-locked loop for grid-connected applications. IEEE Trans. Ind. Electron. 2016, 64, 2848–2854. [Google Scholar] [CrossRef]
  215. Xu, C.; Tang, X.; Liu, G. A Novel Method for Frequency Estimation of AMB Rotor Based on Multiresonance Controller-Quadrature Phase-Locked Loop. IEEE Trans. Ind. Electron. 2024, 72, 1849–1859. [Google Scholar] [CrossRef]
  216. Hsieh, G.C.; Hung, J.C. Phase-locked loop techniques. A survey. IEEE Trans. Ind. Electron. 1996, 43, 609–615. [Google Scholar] [CrossRef]
  217. Pin, G.; Chen, B.; Fedele, G.; Parisini, T. Robust frequency-adaptive quadrature phase-locked-loops with Lyapunov-certified global stability. IEEE Trans. Control Syst. Technol. 2022, 31, 467–474. [Google Scholar] [CrossRef]
  218. Guo, Z.K.; Li, Y.G.; Yu, B.C.; Zhou, S.C.; Meng, Q.Y.; Lu, X.X.; Huang, Y.F.; Liu, G.P.; Lu, J. Research progress of lock-in amplifiers. Acta Phys. Sin. 2023, 72, 249901. [Google Scholar] [CrossRef]
  219. Welch, G.; Bishop, G. An Introduction to the Kalman Filter; University of North Carolina at Chapel Hill: Chapel Hill, NC, USA, 1995. [Google Scholar]
  220. Simon, D. Kalman filtering. Embed. Syst. Program. 2001, 14, 72–79. [Google Scholar]
  221. Khodarahmi, M.; Maihami, V. A review on Kalman filter models. Arch. Comput. Methods Eng. 2023, 30, 727–747. [Google Scholar] [CrossRef]
  222. Grewal, M.S. Kalman filtering. In International Encyclopedia of Statistical Science; Springer: Berlin/Heidelberg, Germany, 2011; pp. 705–708. [Google Scholar]
  223. Li, Q.; Li, R.; Ji, K.; Dai, W. Kalman filter and its application. In Proceedings of the 2015 8th International Conference on Intelligent Networks and Intelligent Systems (ICINIS), Tianjin, China, 1–3 November 2015; IEEE: New York, NY, USA, 2015; pp. 74–77. [Google Scholar]
  224. Mikrosensoranwendung Chemnitz GmbH. AM-IP4k Datasheet. 2020. Available online: https://amac-chemnitz.de/images/interpolation/AM-IP4k/50400-DB-1-3-E-IP4k_AMAC.pdf (accessed on 20 June 2025).
  225. Köning, R.; Wimmer, G.; Witkovský, V. EllipseFit4HC: A MATLAB algorithm for demodulation and uncertainty evaluation of the quadrature interferometer signals. In Advanced Mathematical And Computational Tools In Metrology And Testing X; World Scientific: London, UK, 2015; pp. 211–218. [Google Scholar]
  226. Yue, C.; Li, J.; Sun, J.; Zhu, R.; Hou, X.; Zhang, X.; Liu, L.; Chen, W. Homodyne coherent optical receiver for intersatellite communication. Appl. Opt. 2018, 57, 7915–7923. [Google Scholar] [CrossRef]
  227. Bartáková, Z.; Bálek, R. Use of the application of heterodyne laser interferometer in power ultrasonics. Ultrasonics 2006, 44, e1567–e1570. [Google Scholar] [CrossRef]
  228. Stöckling, C.A.; Pöschel, W.; Schulz, S. Optical Simulation and Design of Spatial Heterodyne Spectrometers for Remote Sensing Applications. In Proceedings of the EPJ Web of Conferences, Online, 20 August 2020; EDP Sciences: Les Ulis, France, 2020; Volume 238, p. 12018. [Google Scholar]
  229. Wang, R.; Zhang, Y.; Lu, M.; Wu, Y.; Wang, Y.; Li, W.; Zhang, H.; Li, D.; Wei, F. PyQt5 Coding and Optimization for Heterodyne Detected Vibrational Sum Frequency Generation Spectroscopy. Authorea Prepr. 2025, 139, 103827. [Google Scholar] [CrossRef]
  230. Sahoo, A.; Mathew, J.; Battisti, A.; Tucker, B. Compact Spatial Heterodyne Spectrographs for Future Space-Based Observations: Instrument Modeling and Applications. Sensors 2024, 24, 4709. [Google Scholar] [CrossRef] [PubMed]
  231. He, C.; Hang, L.; Wu, B. Application of homodyne demodulation system in fiber optic sensors using phase generated carrier based on LabVIEW in pipeline leakage detection. In Proceedings of the 2nd International Symposium on Advanced Optical Manufacturing and Testing Technologies: Optical Test and Measurement Technology and Equipment, Xi’an, China, 2–5 November 2005; SPIE: Cergy, France, 2006; Volume 6150, pp. 488–493. [Google Scholar]
  232. Holzer, M.; Mohr, F. Aspects of signal processing in laser vibrometry and their embedded-realisation on FPGA with NI-LabVIEW. In Proceedings of the 2007 17th International Conference Radioelektronika, Brno, Czech Republic, 24–25 April 2007; IEEE: New York, NY, USA, 2007; pp. 1–5. [Google Scholar]
  233. Carneiro, A.C.; Lopez-Barbero, A.P.; Silva, V.N.; Kitano, C.; Galeti, J.H.; dos Santos, A.B. High accuracy homodyne interferometric method for wide dynamic range applications. IEEE Sens. J. 2019, 20, 4221–4228. [Google Scholar] [CrossRef]
  234. Huang, J.B.; Li, Y.; Gu, H.C.; Li, R.Z.; Tan, B.; Yang, G. Homodyne coherent demodulation method for fiber laser hydrophone using a 3x3 coupler with gains adjusted real-time for symmetry. In Proceedings of the International Symposium on Photoelectronic Detection and Imaging 2011: Laser Sensing and Imaging; and Biological and Medical Applications of Photonics Sensing and Imaging, Beijing, China, 24–26 May 2011; SPIE: Cergy, France, 2011; Volume 8192, pp. 391–398. [Google Scholar]
  235. Patel, K.; Nagora, U.; Joshi, H.C.; Pathak, S.; Jadeja, K.A.; Patel, K.; Tanna, R.L. LabVIEW-FPGA-based real-time data acquisition system for ADITYA-U heterodyne interferometry. IEEE Trans. Plasma Sci. 2021, 49, 1891–1897. [Google Scholar] [CrossRef]
  236. Martinussen, H.; Aksnes, A.; Engan, H.E. Heterodyne interferometry for high sensitivity absolute amplitude vibrational measurements. In Proceedings of the Interferometry XIII: Techniques and Analysis, San Diego, CA, USA, 13–17 August 2006; SPIE: Cergy, France, 2006; Volume 6292, pp. 316–326. [Google Scholar]
  237. Wu, Y.; Qin, F.; Li, Y.; Ding, Z.; Xu, R. Designing an automatic frequency stabilization system for an external cavity diode laser using a data acquisition card in the LabVIEW platform. Appl. Sci. 2023, 14, 308. [Google Scholar] [CrossRef]
  238. Hue, J.; Moulin, V.; Oxaran, L.; Berg, V.; Rat, V.; Abdoul-Carime, K.; Dupont, B.; Martin, S.; Sermet, F.; Stanchina, S.; et al. Heterodyne detection for optical FMCW LiDAR: Tests, analysis, and experimental comparisons between photodetectors and configurations. In Proceedings of the Photonic Instrumentation Engineering XI, San Francisco, CA, USA, 27 January–1 February 2024; SPIE: Cergy, France, 2024; Volume 12893, pp. 36–51. [Google Scholar]
  239. Sun, C.; He, X.; Xu, R.; Lu, S.; Pan, X.; Bai, J. Optimization of signal-to-noise ratio of laser heterodyne radiometer. Microw. Opt. Technol. Lett. 2024, 66, e33857. [Google Scholar] [CrossRef]
  240. Nguyen, T.D.; Dinh, G.N. Stability and Resolution of a Conventional Displacement Measuring Heterodyne Interferometer Using a Single Phase-Locked Loop. Exp. Mech. 2023, 63, 1015–1032. [Google Scholar] [CrossRef]
  241. Cavedo, F.; Pesatori, A.; Norgia, M.; di Milano, P.; Solari, G.E. Laser rangefinder for steel pipes characterization. In Proceedings of the 2015 IEEE International Instrumentation and Measurement Technology Conference (I2MTC) Proceedings, Pisa, Italy, 11–14 May 2015; IEEE: New York, NY, USA, 2015; pp. 1387–1390. [Google Scholar]
  242. Norgia, M.; Melchionni, D.; Magnani, A.; Pesatori, A. High-speed self-mixing laser distance sensor. In Proceedings of the AIP Conference Proceedings, Ancona, Italy, 25–27 June 2014; American Institute of Physics: College Park, MD, USA, 2014; Volume 1600, pp. 422–425. [Google Scholar]
  243. Ming, X.; Shuang, W.; Jie, H.; Ruicheng, Q.; Man, Z. Development of Agricultural Implement Visual Navigation Terminal Based on DSP and MCU. Nongye Jixie Xuebao/Transactions Chin. Soc. Agric. Mach. 2015, 46, 21–26. [Google Scholar]
  244. Jyrkkä, K.; Silven, O.; Ali-Yrkkö, O.; Heidari, R.; Berg, H. Component-based development of DSP software for mobile communication terminals. Microprocess. Microsyst. 2002, 26, 463–474. [Google Scholar] [CrossRef]
  245. Feng, Z.; Qiu, M. Design and Implementation of Self-Validating Pneumatic Actuator Hardware System Based on DSP and MCU. Int. J. Hybrid Inf. Technol. 2014, 7, 101–114. [Google Scholar] [CrossRef]
  246. Joost, R.; Salomon, R. Advantages of FPGA-based multiprocessor systems in industrial applications. In Proceedings of the 31st Annual Conference of IEEE Industrial Electronics Society, Raleigh, NC, USA, 6–10 November 2005; IEEE: New York, NY, USA, 2005; pp. 6–12. [Google Scholar]
  247. Rodríguez-Andina, J.J.; Valdes-Pena, M.D.; Moure, M.J. Advanced features and industrial applications of FPGAs—A review. IEEE Trans. Ind. Informatics 2015, 11, 853–864. [Google Scholar] [CrossRef]
  248. Farooq, U.; Marrakchi, Z.; Mehrez, H.; Farooq, U.; Marrakchi, Z.; Mehrez, H. FPGA architectures: An overview. In Tree-Based Heterogeneous FPGA Architectures: Application Specific Exploration and Optimization; Springer: Berlin/Heidelberg, Germany, 2012; pp. 7–48. [Google Scholar]
  249. Yang, H.; Zhang, J.; Sun, J.; Yu, L. Review of advanced FPGA architectures and technologies. J. Electron. 2014, 31, 371–393. [Google Scholar] [CrossRef]
  250. Gandhare, S.; Karthikeyan, B. Survey on FPGA architecture and recent applications. In Proceedings of the 2019 International Conference on Vision Towards Emerging Trends in Communication and Networking (ViTECoN), Vellore, India, 30–31 March 2019; IEEE: New York, NY, USA, 2019; pp. 1–4. [Google Scholar]
Figure 1. Overall writing framework of this manuscript [133,145,146,147,148,149,150,157,162,170,171].
Figure 1. Overall writing framework of this manuscript [133,145,146,147,148,149,150,157,162,170,171].
Sensors 25 05013 g001
Figure 2. Homodyne grating interferometer and laser interferometer in a Michelson-type structure. (a) Homodyne grating interferometer. (b) Homodyne laser interferometer.
Figure 2. Homodyne grating interferometer and laser interferometer in a Michelson-type structure. (a) Homodyne grating interferometer. (b) Homodyne laser interferometer.
Sensors 25 05013 g002
Figure 3. Heterodyne grating interferometer and laser interferometer in a Michelson-type structure. (a) Heterodyne grating interferometer. (b) Heterodyne laser interferometer.
Figure 3. Heterodyne grating interferometer and laser interferometer in a Michelson-type structure. (a) Heterodyne grating interferometer. (b) Heterodyne laser interferometer.
Sensors 25 05013 g003
Figure 4. Comparison between homodyne signal and heterodyne signal. (a) Ideal homodyne signal. (b) Ideal heterodyne signal.
Figure 4. Comparison between homodyne signal and heterodyne signal. (a) Ideal homodyne signal. (b) Ideal heterodyne signal.
Sensors 25 05013 g004
Figure 5. Homodyne signal with non-ideal components.
Figure 5. Homodyne signal with non-ideal components.
Sensors 25 05013 g005
Figure 6. Examples of error calibration methods. (a) Method based on FIR filter and phase correction [133]. (b) Method based on Kalman filter [136]. (c) Method based on phase synchronization [135].
Figure 6. Examples of error calibration methods. (a) Method based on FIR filter and phase correction [133]. (b) Method based on Kalman filter [136]. (c) Method based on phase synchronization [135].
Sensors 25 05013 g006
Figure 7. Lissajous ellipse for ellipse parameter estimation.
Figure 7. Lissajous ellipse for ellipse parameter estimation.
Sensors 25 05013 g007
Figure 8. Examples of pulse-counting methods. (a) Method based on rising-edge locking with high-frequency clock signals and digital frequency mixing [170]. (b) Method for movement direction identification and phase correction [145]. (c) Method based on fringe-to-pulse counting [171].
Figure 8. Examples of pulse-counting methods. (a) Method based on rising-edge locking with high-frequency clock signals and digital frequency mixing [170]. (b) Method for movement direction identification and phase correction [145]. (c) Method based on fringe-to-pulse counting [171].
Sensors 25 05013 g008
Figure 9. Examples of quadrature phase-locked methods. (a) Method using digital lock-in phase principle based on FPGA and high-speed ADC [146]. (b) Method based on digital dual-frequency comb [147]. (c) Enhanced lock-in method based on pulse counting for frequency tracking [150]. (d) Method with phasemeter scheme based on the DPLL [148]. (e) Method with quadrature phase-locked phase detection scheme based on the DPLL [149].
Figure 9. Examples of quadrature phase-locked methods. (a) Method using digital lock-in phase principle based on FPGA and high-speed ADC [146]. (b) Method based on digital dual-frequency comb [147]. (c) Enhanced lock-in method based on pulse counting for frequency tracking [150]. (d) Method with phasemeter scheme based on the DPLL [148]. (e) Method with quadrature phase-locked phase detection scheme based on the DPLL [149].
Sensors 25 05013 g009
Figure 10. Calculation process of Kalman filter [151].
Figure 10. Calculation process of Kalman filter [151].
Sensors 25 05013 g010
Figure 11. Hardware platforms for optic interference signal processing. (a) PC platform [157]. (b) Labview interface on PC. (c) MCU/DSP platform [162]. (d) FPGA platform [133]. (e) ASIC platform [224].
Figure 11. Hardware platforms for optic interference signal processing. (a) PC platform [157]. (b) Labview interface on PC. (c) MCU/DSP platform [162]. (d) FPGA platform [133]. (e) ASIC platform [224].
Sensors 25 05013 g011
Table 1. Comparison of signal processing methods for homodyne signals.
Table 1. Comparison of signal processing methods for homodyne signals.
MethodResolutionBandwidthCostApplication
Error Calibration MethodsHighHighHighAcademia
Ellipse Parameter Estimation MethodsHighHighLowIndustry/Academia
Table 2. Comparison of signal processing methods for heterodyne signals.
Table 2. Comparison of signal processing methods for heterodyne signals.
MethodResolutionBandwidthCostApplication
Pulse-Counting MethodsMediumHighLowIndustry/Academia
Quadrature Phase-Locked MethodsHighHighMediumIndustry/Academia
Kalman Filtering MethodHighMediumMediumAcademia
Table 3. Hardware platform comparison. Dev. stands for development and Dep. for deployment.
Table 3. Hardware platform comparison. Dev. stands for development and Dep. for deployment.
PlatformPowerVolumeDev. CostDev. TimeDep. Cost
PCsHighHighLowShortMedium
MCUs/DSPsLowLowLowShortLow
FPGAsMediumLowLowMediumMedium
ASICsLowLowHighLongLow
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, Y.; Zhao, F.; Luo, L.; Li, X. A Review on Recent Advances in Signal Processing in Interferometry. Sensors 2025, 25, 5013. https://doi.org/10.3390/s25165013

AMA Style

Wang Y, Zhao F, Luo L, Li X. A Review on Recent Advances in Signal Processing in Interferometry. Sensors. 2025; 25(16):5013. https://doi.org/10.3390/s25165013

Chicago/Turabian Style

Wang, Yifeng, Fangyuan Zhao, Linbin Luo, and Xinghui Li. 2025. "A Review on Recent Advances in Signal Processing in Interferometry" Sensors 25, no. 16: 5013. https://doi.org/10.3390/s25165013

APA Style

Wang, Y., Zhao, F., Luo, L., & Li, X. (2025). A Review on Recent Advances in Signal Processing in Interferometry. Sensors, 25(16), 5013. https://doi.org/10.3390/s25165013

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