Next Article in Journal
Development and Test of a Novel High-Precision Inchworm Piezoelectric Motor
Previous Article in Journal
Preparation of Hydrophobic Glass Surfaces by Femtosecond Laser
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances in Laser Linewidth Measurement Techniques: A Comprehensive Review

1
Center for Advanced Laser Technology, School of Electronics and Information Engineering, Hebei University of Technology, Tianjin 300401, China
2
Hebei Key Laboratory of Advanced Laser Technology and Equipment, Tianjin 300401, China
3
Innovation and Research Institute, Hebei University of Technology in Shijiazhuang, Shijiazhuang 050299, China
4
Collaborative Innovation Center for Diamond Laser Technology and Applications, Tianjin 300401, China
*
Authors to whom correspondence should be addressed.
Micromachines 2025, 16(9), 990; https://doi.org/10.3390/mi16090990 (registering DOI)
Submission received: 2 July 2025 / Revised: 5 August 2025 / Accepted: 19 August 2025 / Published: 29 August 2025

Abstract

As a key parameter that defines the spectral characteristics of lasers, the precise measurement of laser linewidth is crucial for a wide range of advanced applications. This review systematically summarizes recent advances in laser linewidth measurement techniques, covering methods applicable from GHz-level broad linewidths to sub-Hz ultranarrow regimes. We begin by presenting representative applications of lasers with varying linewidth requirements, followed by the physical definition of linewidth and a discussion of the fundamental principles underlying its measurement. For broader linewidth regimes, we review two established techniques: direct spectral measurement using high-resolution spectrometers and Fabry–Pérot interferometer-based analysis. In the context of narrow-linewidth lasers, particular emphasis is placed on the optical beating method. A detailed comparison is provided between two dominant approaches: power spectral density (PSD) analysis of the beat signal and phase-noise-based linewidth evaluation. For each technique, we discuss the working principles, experimental configurations, achievable resolution, and limitations, along with comparative assessments of their advantages and drawbacks. Additionally, we critically examine recent innovations in ultra-high-precision linewidth metrology. This review aims to serve as a comprehensive technical reference for the development, characterization, and application of lasers across diverse spectral regimes.

1. Introduction

Laser, an acronym for laser amplification by stimulated emission of radiation, refers to a coherent light source renowned for its exceptional monochromaticity, directionality, spatial coherence, and high brightness. The theoretical foundation of laser technology traces back to Einstein’s seminal work in 1917 [1], which first postulated the existence of stimulated emission. In 1958, Shawlow and Townes [2] made a groundbreaking contribution by proposing the extension of maser principles to the optical regime, articulating the conceptual framework for achieving monochromatic coherent laser amplification through optical resonator configurations. The advent of practical laser systems materialized in 1960 with Maiman’s pioneering demonstration of the first ruby laser [3,4], which generated a high-purity laser at 694.3 nm, heralding the dawn of laser science. Over subsequent decades, laser technology has undergone exponential evolution, becoming indispensable across diverse industrial and scientific domains. Continuous advancements in critical laser parameters including wavelength [5], output power [6], and spectral linewidth [7,8] have driven transformative progress across multiple disciplines, including precision manufacturing [9,10], life sciences [11,12], defense technologies [13,14], information systems [15], and fundamental research [16,17].
Over the past decades, the generation and precise measurement of laser linewidth have remained central themes in laser science, as linewidth is a fundamental parameter that defines a laser’s temporal coherence and spectral purity. The quest for ultra-narrow linewidths has driven significant progress not only in resonator design and oscillator optimization but also in the development of high-performance optical coatings, ultra-stable reference cavities, and advanced linewidth-narrowing techniques. For example, nonlinear optical processes have played a pivotal role. In particular, third-order nonlinear effects, notably stimulated Raman scattering (SRS) [18,19,20,21,22] and stimulated Brillouin scattering (SBS) [23,24,25], have emerged as pivotal research frontiers. SRS facilitates wavelength generation across the transparency window of laser media [26,27,28], significantly advancing the spectral coverage [29] and power scalability [30] of vortex beams. Notably, in 2025, researchers at Macquarie University demonstrated a novel application of SRS in diamond crystals, whereby temporal fluctuations were converted into coherent lattice vibrations that were rapidly dissipated, resulting in a linewidth compression by over four orders of magnitude [31]. Conversely, SBS enables narrow-linewidth laser emission with ultrashort pulse durations and diffraction-limited beam quality [32,33,34,35,36,37,38,39].
In recent years, narrow-linewidth lasers characterized by superior spectral purity, high peak spectral density, ultralong coherence lengths, and exceptionally low phase noise [40,41] have found indispensable applications across a wide range of fields [42,43,44,45], as schematically illustrated in Figure 1. In optical communication [46,47,48], the deployment of narrow-linewidth lasers has enabled systems to surpass traditional radio-frequency (RF) communication in terms of directionality and carrier frequency, resulting in significantly higher data transmission rates, expanded bandwidth capacity, and enhanced communication security [49,50,51]. Laser communication has further facilitated the implementation of inter-satellite links [52], with prominent satellite constellations such as “Kuiper” [53], “Starlink” [54], and “Xingyun” [55] adopting laser-based transmission as a backbone carrier technology. This paradigm shift signifies the transition of space networks from the radio-frequency era to the laser-dominated era. In coherent optical communication systems [56,57,58,59,60,61,62,63], which utilize the phase coherence of lasers for high-speed and long-distance data transmission, linewidth narrowing is essential for minimizing phase noise at the receiver, thereby reducing the bit error rate (BER) [64] and enhancing overall system performance [65,66,67,68,69,70]. Experimental studies have shown that BER in spaceborne coherent systems is highly sensitive to laser linewidth fluctuations; broader linewidths induce increased phase noise, degrade signal-to-noise ratio (SNR), and result in higher BER. For instance, four-phase shift keying (4-PSK) coherent detection systems necessitate laser linewidths below 100 kHz [71,72,73,74]. In 40 Gbps high-speed optical communication systems, optimal reception performance for advanced modulation formats, including eight-quadrature amplitude modulation (8-QAM), 16-QAM, and 64-QAM, require laser linewidths not exceeding 1.2 kHz [75]. Consequently, lasers with linewidths below 8 kHz are essential to meet stringent BER requirements for high-quality communications [76]. Furthermore, the characterization of laser linewidth demands enhanced precision, as carrier linewidth limitations directly impact the frequency-locking capability of optical phase-locked loops [77,78]. In lidar systems [79,80,81,82,83,84,85,86,87,88,89,90,91,92], particularly coherent lidar, the detection range is inherently limited by the coherence length of the laser source. Employing narrow-linewidth lasers can dramatically extend coherence length, thereby improving measurement range and enabling deployment in large-scale and high-resolution applications. Linewidth narrowing also enhances SNR, which is crucial for precise detection. For instance, lasers with linewidths below 100 Hz [93] are employed in automotive frequency-modulated continuous-wave (FMCW) lidar systems, where wavelength-swept detection enables high-resolution, long-range sensing. In spectroscopy [94,95,96,97,98,99,100], narrow-linewidth lasers with high brightness and tunability allow precise alignment with specific absorption lines, significantly enhancing spectral resolution. This capability supports detailed characterization of atomic and molecular transitions, interaction dynamics, and physical constants, thereby driving progress in high-sensitivity spectroscopic techniques [101,102,103]. For space-based gravitational wave detection [104,105,106,107,108], the stability of laser linewidth governs measurement precision of minute distance variations induced by gravitational waves [109,110]. In the field of atomic clocks [111,112,113,114,115,116,117,118], precise linewidth characterization supports the realization of enhanced temporal and frequency standards, which hold critical importance for satellite navigation and high-speed communication systems. Additionally, narrow-linewidth lasers serve as high-spectral-purity sources in distributed fiber optic sensing [119,120,121,122], where improved linewidth characterization enhances both sensitivity and spatial resolution. These lasers also support more accurate remote sensing [123], enable high-performance quantum sensors [124], and contribute to improved inertial navigation systems [125].
The growing demand for high-performance narrow-linewidth lasers is driven by both fundamental scientific research and the rapid expansion of advanced application frontiers [126,127,128]. Key developments such as external cavity design [129,130,131,132,133,134,135,136,137,138,139], optical phase-locked loop (OPLL) technologies [140,141,142], the Pound–Drever–Hall (PDH) technique [143], self-injection locking (SIL) methods [144,145,146], and distributed feedback (DFB) structures [147,148,149,150,151,152,153,154,155,156,157] have propelled laser linewidths into the kHz regime, with cutting-edge systems achieving Hz-level [158,159] and sub-Hz [160,161] spectral resolutions. Concurrently, precision characterization of ultranarrow linewidth has emerged as a critical challenge [162,163,164,165,166,167,168]. Notably, no single measurement technique is universally applicable across all linewidth regimes. Moreover, the development of linewidth measurement technologies has lagged behind the rapid progress in narrow-linewidth laser sources, creating a critical bottleneck for further advancement. As a result, there is an urgent need for convenient, accurate, and scalable linewidth characterization techniques that can support the continued innovation and optimization of high-performance lasers.
Although a variety of linewidth measurement methods have been proposed in the literature, a systematic and comparative evaluation of their principles, performance limits, and practical constraints remains lacking. This review addresses this gap by summarizing and analyzing existing laser linewidth measurement techniques, organized by applicable linewidth regime from broad to ultranarrow.

2. Basic Concept of Linewidth

2.1. What Is Linewidth?

Laser radiation is distinguished from conventional light sources by its exceptional coherence properties, encompassing both spatial and temporal coherence. Spatial coherence quantifies the phase correlation across transverse wavefront dimensions during beam propagation, serving as the prerequisite for achieving collimated beam characteristics. Temporal coherence describes the phase relationship of the beam at different time points along the propagation direction, which is proportional to the monochromaticity of the laser. Temporal coherence can be characterized by linewidth.
The operational dynamics of laser systems are inherently influenced by dual noise mechanisms: quantum noise stemming from spontaneous emission [2,169,170] and classical noise perturbations induced by environmental factors including mechanical vibrations and thermal instabilities [171,172,173]. This means that in addition to the intrinsic spectral width of the laser with a certain width [174], the actual output spectral line is further broadened, which is the root cause of the existence of the laser linewidth. Quantitatively, linewidth is defined as the full width at half maximum (FWHM) of the optical power spectrum. This refers to the spectral width measured at 50% of the peak intensity, as schematically illustrated in Figure 2.
In 1955, Gordon et al. [174] established a theoretical framework for ammonia molecular beam-driven microwave amplifiers. Through a classical power-balance analysis, they derived the fundamental linewidth expression for microwave amplification systems, as follows:
υ t k T ( Δ ω ) 2 P
where kT represents the thermal bath’s spectral power density maintaining equilibrium in the microwave amplifier prior to beam activation, Δω denotes the molecular emission bandwidth, and P corresponds to the emitted power. Both Δω and υt contribute to the FWHM.
Schawlow and Townes proposed the adaptation of Equation (1) to optical regimes [2,175]. The Schawlow–Townes formula can be obtained by simply replacing kT with , which corresponds to the spectral power density of one photon in each mode, as follows:
υ t = h υ ( Δ ω ) 2 P
This expression forms one of the cornerstones of modern laser physics, highlighting the intrinsic relationship between spontaneous emission and spectral purity.
To account for the dynamics of practical laser systems, Haken [176], Lax [177], and Scully [124] independently proposed a generalized linewidth expression for four-level lasers operating above threshold, now known as the Haken–Lax–Scully formula:
υ t = h υ ( Δ ω min ) 2 2 P
where Δωmin refers to the smaller of the natural emission bandwidth and the cavity bandwidth. Experimental measurements by Manes [178] confirmed the general form of this equation, which also accommodates the effects of inhomogeneous broadening.
In essence, laser linewidth is fundamentally determined by the spontaneous emission rate of the gain medium and the structural properties of the resonant cavity [159]. As long as the gain medium is based on the amplification of stimulated emission laser, the lifetime of the upper energy level is limited, and spontaneous emission cannot be avoided. The vibration of the vibration source causes low-frequency noise, and the cavity length disturbance that this causes leads to optical frequency drift, thereby broadening the spectral linewidth. Consequently, linewidth serves as a direct proxy for phase noise evaluation, where narrower linewidth correlate with enhanced frequency stability and suppressed phase fluctuations [179]. Given the multifaceted origins of linewidth variations, ranging from dominant quantum effects to subtle technical noise sources, the development of refined metrological techniques has become imperative for validating linewidth control strategies.

2.2. How to Measure Linewidth?

The advancement of laser linewidth measurement techniques has closely paralleled the progress in linewidth compression technologies. As illustrated in Figure 3, a variety of methods have been developed to meet the demands across different linewidth regimes.
For GHz-level linewidths, the spectroscopic method, a relatively mature commercial technique, utilizes high-resolution optical spectrum analyzers (OSAs) to directly determine the FWHM of the laser emission spectrum [180,181,182,183,184,185]. In the MHz regime, Fabry–Pérot (F-P) interferometry provides another effective approach for linewidth characterization [186,187]. However, as laser linewidths continue to narrow, these methods face increasing limitations. The spectroscopic approach is fundamentally constrained by the resolution limits of OSAs, while F-P interferometry becomes increasingly sensitive to cavity length fluctuations and noise floor interference, ultimately rendering both methods inadequate for the precise measurement of ultra-narrow linewidths [188,189,190]. The ongoing refinement of laser frequency stabilization [44,191] and mode-selection techniques [192] has driven significant advancements in narrow-linewidth laser metrology. The laser linewidth can be obtained not only directly by the laser power spectrum, but also indirectly by the measurement of phase noise. Due to the relative ease of acquiring the power spectrum and its intuitive representation of linewidth characteristics, optical beating methods have gained widespread adoption. These techniques are currently capable of resolving linewidths on the order of kilohertz.
In this review, we focus on four representative methods for directly measuring linewidth based on the signal power spectrum: two-beam interferometry, the Brillouin Stokes optical beating method, delayed self-homodyne, and delayed self-heterodyne techniques. While phase-noise-based linewidth measurement offers higher precision, it often involves more complex mathematical modeling and algorithmic processing. Finally, we summarize and provide perspectives on the future development of advanced linewidth measurement techniques, highlighting the need for greater accuracy, robustness, and ease of implementation in emerging applications.

3. Spectroscopic Method

Laser linewidth can be determined by analyzing the optical spectrum of the emitted light. To achieve this, the spectral information of the laser must first be acquired. Optical gratings exploit their periodic structure to produce constructive interference at specific angles, thereby spatially dispersing different wavelengths. Figure 4a illustrates the operating principle of a blazed grating. Two incident light rays, originating from points A and B, are diffracted by the grating and produce rays AA′ and BB′. In this schematic, d denotes the distance between adjacent grooves (i.e., the grating period), which is also the spacing between the corresponding diffracted rays. The angles α and β represent the incident and diffraction angles, respectively. The relationship governing spectral dispersion is described by the grating equation, as follows:
d ( sin α + sin β ) = m λ
where m indicates the diffraction order (m = 0, ±1, ±2, ……), and λ is the wavelength of the incident laser.
As shown in Figure 4b, a typical grating spectrometer fundamentally consists of an incident slit (S1), collimating mirror (M1), blazed grating (G), focusing mirror (M2), and exit slit (S2). Polychromatic light entering through S1 is collimated by M1 and projected onto the grating, where wavelength dispersion occurs to generate parallel beams at distinct diffraction angles. M2 focuses a specific wavelength onto S2, and the photodetector (PD) behind S2 records the output signal intensity corresponding to different grating rotation angles. By rotating the grating using a precision rotation stage, full-spectrum scanning is achieved, yielding the spectral intensity distribution as illustrated in Figure 4d. The FWHM of the resulting spectral peak represents the linewidth of the measured laser.
The resolution of a grating spectrometer is primarily determined by three instrumental parameters: the effective focal length, the dispersion characteristics of the grating, and the widths of the entrance and exit slits. While the effective focal length and grating dispersion are intrinsic to a given spectrometer design and cannot be modified, the slit widths are adjustable parameters that must be optimized based on experimental requirements.

3.1. Effect of Incident Slit Width on Spectral Linewidth

For the theoretical analysis of the effect of incident slit width on the measured spectral line width, it is assumed that the incident beam is an ideal monochromatic wave, and the width of the exit slit is neglected. As shown in Figure 4b, when the incident slit width of the spectrometer is Δx1, each point on the slit can be regarded as a luminous point, and the angular width of the incident angle of the beam incident on the grating is Δα. Since the slit is located on the focal plane of the cylindrical mirror, the value of Δα can be expressed as follows:
Δ α = Δ x 1 f F L
By differentiating both sides of Equation (4) with respect to the grating incident angle and substituting Equation (5) into the result, the measured spectral line width caused by the slit width can be expressed as follows:
Δ λ = d Δ x 1 m f F L cos α
where fFL denotes the focal length of the cylinder lens.
From Equation (6), it can be observed that the measured spectral line width at this point is proportional to the incident slit width and inversely proportional to fFL.

3.2. Effect of Exit Slit Width on Spectral Line Width

As shown in Figure 4c, when the incident slit width of the spectrometer is Δ x 2 , the corresponding beam emitted from the grating and passing through the exit slit has an angular width of Δβ. By analogy, the measured spectral line width Δλ caused by the exit slit width Δ x 2 can be expressed as follows:
Δ λ = d Δ x 2 m f F L cos β
It can thus be seen that the spectral line width is proportional to the exit slit width and inversely proportional to fFL.

3.3. Effect of Slit Width on Spectrometer Resolution

The resolution of a spectrometer is a parameter that indicates its ability to separate two spectral lines with extremely close wavelengths. It is defined as follows:
R = λ Δ λ s
where λ is the wavelength of the light wave, and Δλs is the resolving limit of the spectrometer. As discussed in Section 3.1 and Section 3.2, Δλs is proportional to the slit width; thus, the resolution R is inversely proportional to both the exit slit width and the incident slit width [193].
In general, reducing the widths of the incident and exit slits is beneficial for improving the resolution of the spectrometer. However, narrowing the incident and exit slits weakens the signal intensity, and excessively small slit widths may render the signal undetectable by photodetectors. Therefore, when using a spectrometer to measure laser linewidth, the slit widths should be minimized as much as possible under the premise that the optical signal can be detected by the photodetector, to improve the measurement accuracy of the laser linewidth.
Early studies adopted multi-spectrometer collaborative measurement to overcome single-device limitations. H. Liu et al. [194] theoretically analyzed the principle of combined two-spectrometer laser linewidth measurement, derived the dispersion rate formula, and enhanced optical path dispersion via their joint use to increase spectral separation. Experimental results showed this method improved He-Ne laser FWHM measurement accuracy by at least five orders of magnitude over single spectrometers, verifying multi-device error complementation potential and offering a cost-effective scheme for resource-limited labs, though performance remains constrained by hardware’s intrinsic physical properties.
Recent advancements have witnessed the convergence of optical encoding and deep learning algorithms. Liu et al. [195] pioneered a dual-phase framework where optical devices encode incident spectral information and reconstruction algorithms decode it, proposing an adaptive deep learning algorithm to mitigate distortions induced by filter array imperfections.
This method, grounded in the principle of grating dispersion, enables laser linewidth determination through spectral feature analysis while simultaneously acquiring wavelength and intensity parameters, thus supporting real-time observation of spectral dynamics. The spectrometer’s free-space optical interface facilitates direct beam collimation and injection, significantly improving operational efficiency [196]. State-of-the-art diffraction grating spectrometers now achieve remarkable performance metrics, including 5 pm wavelength resolution, 5 pm wavelength accuracy, 65 dB dynamic range, and 80 dB stray light suppression ratio. These advancements enable unprecedented precision in spectral characterization, particularly in applications demanding ultra-narrow linewidth analysis, thereby unlocking new potential for high-resolution spectroscopic investigations [197].
Nevertheless, despite this operational simplicity, different spectrometers have been designed to cover a specific wavelength range. Therefore, lasers with wavelengths beyond this range cannot be measured as they cannot be effectively separated or detected. Furthermore, due to limitations such as the resolution being constrained by the slit width, the difficulty in precisely controlling incident and diffraction angles, and potential overlap of spectral lines from different orders, spectrometers can only measure laser linewidths at the GHz order of magnitude or above.

4. F-P Interferometry

F-P interferometry is a laser linewidth measurement technique based on multi-beam interference principles [198,199]. Two configurations are commonly employed: the F-P etalon (with constant cavity length) and the F-P scanner (with tunable cavity length) [200].
The F-P etalon functions as an optical resonator comprising two high-reflectivity parallel surfaces (glass or quartz plates) separated by a medium to form a resonant cavity of length [201,202]. Typical implementations include air-gap etalons and solid fused-silica etalons [203]. The F-P scanner enhances interferometric sensitivity by integrating auxiliary optics such as focusing lenses and polarizers, with cavity designs adopting confocal or quasi-confocal configurations for improved beam alignment [204]. While both variants effectively measure continuous-wave laser linewidth, the F-P etalon suffers from limited sampling rates for low-repetition-rate pulsed lasers, whereas the F-P scanner configuration eliminates pulse frequency constraints, making it ideal for pulsed laser characterization [205].
As illustrated in Figure 5a, monochromatic light incident at angle θ undergoes multiple reflections and transmissions at the coated surfaces, generating parallel reflected/transmitted beams. When the cavity satisfies the resonance condition (L is the integer multiple of half wavelength), standing waves induce constructive or destructive interference. Collimating the transmitted beams via a convex lens produces a concentric ring pattern on the focal plane S’, as shown in Figure 5b. The interferometric ring pattern is then converted into spectral data through coordinate transformation within the interference domain, followed by Gaussian fitting to extract the interferometric linewidth.
For the F-P interferometer, its spectral resolution is typically characterized by the parameter finesse. Higher finesse implies that the interferometer can generate extremely sharp resonance peaks, enabling it to more readily distinguish closely spaced transmission peaks; thus, high finesse corresponds to high spectral resolution. In practice, the total finesse Ft of an F-P interferometer system is influenced by multiple factors and can be expressed as follows [207]:
1 F t = 1 F + 1 F Q + 1 F i
where F denotes the reflectivity-limited finesse of the mirror, FQ represents the quality-limited finesse originating from the mirror surface, and Fi accounts for the illumination-limited finesse induced by the mirror’s illumination conditions (including beam alignment and beam diameter).
Most F-P interferometers are designed with carefully engineered reflective coatings such that, over the entire operating wavelength range and under proper illumination conditions, the reflectivity-limited finesse F dominates the total system finesse Ft (where a higher reflectivity increases F, thereby enhancing the spectral resolution). Regarding the other two finesse components, FQ characterizes the symmetric broadening of spectral lines induced by microscopic irregularities on the mirror surface. These irregularities introduce random optical path differences across different regions of the mirror, leading to blurred broadening of the interference spectrum due to phase incoherence. In contrast, Fi degrades the resolution as the beam diameter increases or the input beam undergoes misalignment. When the finesse is limited by Fi, the measured spectral line shape exhibits asymmetry.
The performance breakthrough of high-finesse fiber Fabry–Pérot scanning interferometers (FFPSI) originates from the synergistic optimization of finesse and spectral resolution. Since the FFPSI employs a built-in single-mode fiber cavity, it is free from diffraction loss or imperfect spatial mode matching that are inherent in bulk devices. Therefore, the FFPSI can be extended to longer cavity lengths, enabling direct high-resolution measurement of narrow laser linewidths. Early advancements involved constructing high-precision FFPSI systems with 10–40 cm cavity lengths, achieving a resolution of 500 kHz and finesse exceeding 500, representing an order-of-magnitude improvement over conventional designs [208]. While this progress solidified FFPSI’s dominance in static spectral analysis, its dynamic measurement capability remained constrained by scanning rate and cavity length limitations. Subsequent efforts targeting dynamic applications focused on optimizing long-cavity FFPSI scanning rates. Kevin Hsu et al. [209] improved the resolution of FFPSI by increasing the cavity length and reducing the scanning rate. They tuned the FFPSI by adjusting the resonant cavity length via a piezoelectric transducer platform, achieving a resolution of approximately 8 kHz for the long-cavity FFPSI at a scanning rate of about 3 ms per free spectral range. Although the measurement device was sensitive to environmental interferences such as ambient temperature drifts, acoustic waves, and mechanical vibrations, thus requiring appropriate shielding, this achievement provides a valuable reference for developing long-cavity FFPSI into an effective, low-cost, and lightweight instrument that can be directly applied to high-resolution spectroscopic research. The extension of FFPSI applications to pulsed laser characterization introduced new challenges in response dynamics. Xue et al. [210] systematically investigated the differential response mechanisms to continuous-wave and pulsed lasers, demonstrating FFPSI’s capability to measure pulse laser linewidths to 100 MHz. However, conventional interferometry exhibited critical limitations in low-repetition-rate short-pulse (LSL) measurements due to convolutional artifacts and stability-dependent errors. Addressing these constraints, Hun Xuanning et al. [206,211] employed a technique combining a Fabry–Pérot etalon with a complementary metal oxide semiconductor (CMOS) beam profiler. They accurately localized the center of the interference pattern via the Hough transform algorithm, followed by the application of a pixel rotation algorithm to convert the two-dimensional interference pattern into one-dimensional spectral information. This significantly increased the number of effective data points, improved spectral resolution, and reduced the error induced by pixel size to below 1 MHz. Subsequently, they mitigated the error induced by transmission spectrum width (TSW) through deconvolution processing, thereby achieving accurate measurement of the linewidth of short pulses at low repetition rates. This methodology transformed hardware limitations into algorithmic optimization opportunities, drastically reducing measurement time and eliminating stringent laser stability requirements.
For the F-P interferometer, increasing the reflection coefficient enhances F, enabling the interferometer to generate sharper resonance peaks and thereby improving spectral resolution. However, this enhancement is constrained by the GHz-scale free spectral range (FSR) of the F-P interferometer, which limits its minimum resolvable frequency difference to the MHz regime. Conversely, lower reflectivity reduces F, broadening the resonance peaks and degrading resolution. A critical balance must be maintained between spectral resolving power and net light-gathering efficiency. A pragmatic compromise involves enlarging the mirror aperture until spectral resolution diminishes by 70% [207]. These reasons together lead to the fact that the F-P interferometer can only measure the laser linewidth of MHz magnitude.

5. Optical Beating Method

Due to the limitation of resolution, the spectrometer can only measure the linewidth of GHz magnitude, while the linewidth limit detected by F-P interferometry is in the order of MHz [212]. For laser linewidth in the kHz regime and below, the optical beating method emerges as the dominant approach, where precise acquisition and analysis of beat frequency signals are critical. Two optical beams with distinct frequencies and wavelenghs (λ1 and λ2) are coherently combined and incident on a PD. The PD’s nonlinear response generates a beat signal at the frequency difference, expressed as follows:
f m = λ 1 λ 2 λ 1 2
where fm presents the difference of the laser frequencies.
The acquired beat signal necessitates sophisticated numerical processing, as systematically outlined in Figure 6, which delineates two distinct analytical workflows and their interrelationships. The data acquisition of beat frequency in time domain is the entry point to obtain the center of linewidth. Route A employs the Wiener–Khinchin theorem [213], wherein autocorrelation of the temporal beat signal is computed prior to Fourier transformation, yielding the laser’s PSD from which the linewidth is directly extracted. In contrast, Route B leverages β-separation theory [214], deriving the linewidth through the intrinsic relationship between phase noise, frequency noise, and linewidth. Since the signal power spectrum contains more intuitive linewidth information and is relatively easy to obtain, most of the linewidth measurement experiments are based on Route A.

5.1. Laser Linewidth Measurement Based on the Signal Power Spectrum (Direct Measurement)

The method for measuring laser linewidth based on the signal power spectrum is characterized by its simplicity and efficiency, and has been widely applied in practical operations. It corresponds to Route A in Figure 7. When two incoherent lasers with Lorentzian line shapes are used, their beat frequency signal retains a Lorentzian line shape, and the PSD of the beat frequency signal can be expressed as follows:
S ( υ ) = Δ υ 2 π [ ( υ f m ) 2 + ( Δ υ 2 ) 2 ]
Δ υ = υ t + υ r
where Δυ is the linewidth of the beat frequency power spectrum, υt is the linewidth of the tested laser, υr is the linewidth of the reference laser, and fm is the frequency difference between the two lasers mentioned above.
Figure 7. Principle of optical beating method [215] (where S1(υ) is the PSD of the tested laser, and S2(υ) is that of the reference laser).
Figure 7. Principle of optical beating method [215] (where S1(υ) is the PSD of the tested laser, and S2(υ) is that of the reference laser).
Micromachines 16 00990 g007
Linewidth measurement based on the optical beating method typically involves two cases, as shown in Figure 7. When the reference laser and the tested laser have identical linewidth, the linewidth of the beat frequency signal becomes twice that of the tested laser (Case 1). This method is suitable for measuring arbitrary linewidth, but its drawback lies in the complexity of acquiring, measuring, and calibrating the reference laser. When the linewidth of the reference laser is significantly smaller than that of the tested laser and can be neglected, the linewidth of the beat frequency signal approximates that of the tested laser (Case 2). This approach can measure the linewidth of most lasers but struggles with precise measurements of narrow linewidth.
Depending on the specific characteristics of the tested laser, either of these two schemes can be flexibly selected to achieve optimal measurement results. This section highlights four measurement methods based on the two cases, including two-beam interferometry, the Brillouin Stokes optical beating method, the delayed self-homodyne method, and the delayed self-heterodyne method.
Two-beam interferometry, based on Case 1, involves coupling two optical signals with closely matched wavelengths after they pass through an isolator. The intermediate-frequency electrical signal generated from the coupling is then detected by a PD. Finally, an electrical spectrum analyzer (ESA) is used to process the detected signal, enabling the calculation of the laser linewidth.
Unlike two-beam interferometry, the Brillouin Stokes optical beating method (based on Case 2) eliminates the need for an additional reference laser. It uses narrow-linewidth second-order Stokes light generated in a fiber resonator as the reference for beating with the tested laser. A resonant tracking circuit (RTC) and piezoelectric ceramic (PZT) are used to control the fiber length, ensuring the pump light resonates in the fiber resonator.
To avoid the stringent environmental requirements of the two-beam interferometry and overcome the linewidth measurement range limitation of the Brillouin Stokes beating method, the delayed self-homodyne method has emerged, featuring a simple structure, wide measurement range, and low optical transmission loss. Based on Case 1, this method processes the tested laser using an unbalanced Mach–Zehnder interferometer (UMZI) [216].
To steer clear of the rigorous environmental demands of two-beam interferometry and get around low-frequency interference in the delayed self-homodyne method, Okoshi et al. [217] proposed the delayed self-heterodyne method, achieving a resolution of 50 kHz. Subsequently, P. Gallion [218] conducted a detailed derivation of its basic principles. Based on Case 1, this method uses a long delay fiber to create a time delay much longer than the coherence time of the original beam, rendering the two beams incoherent during beating. The spectrum of the beat signal is typically fitted with a Lorentzian function, facilitating linewidth parameter measurement [219]. Furthermore, the system incorporates an acousto-optic modulator (AOM) into the system, shifting the laser signal’s center frequency to a high-frequency region to avoid interference from environmental noise near the zero frequency.

5.2. Linewidth Measurement Based on Phase Noise (Indirect Measurement)

In 1986, L.E. Richter et al. [220] theoretically analyzed the laser linewidth measurement principle of this method, proposing that the beat signal exhibits a perfect Lorentzian line shape only when the delay fiber length exceeds six times the coherence length of the laser under test. This results in a 1590 km fiber being required to measure a 100 Hz linewidth [221], and even for 10 kHz linewidths, the delay fiber would still be tens of kilometers long [222]. However, long fibers exacerbate optical power loss and amplify the effects of factors such as Rayleigh scattering [223,224], spectral drift [225], and 1/f noise, causing additional spectral broadening [7,226] and measurement errors [227]. Additionally, high-power lasers in long fibers are prone to SBS, resulting in the conversion of pump energy into Stokes wave and acoustic wave energy, which exacerbates transmission loss and hinders signal detection by the PD. To address these limitations, researchers have proposed various improved schemes for the delayed self-heterodyne method. Table 1 presents the measurement devices, their descriptions and main features, and some references.
Table 1. Four direct measurement methods [228,229,230,231,232,233,234,235,236,237,238,239,240,241,242,243,244,245,246,247,248,249,250,251,252,253,254,255,256].
Table 1. Four direct measurement methods [228,229,230,231,232,233,234,235,236,237,238,239,240,241,242,243,244,245,246,247,248,249,250,251,252,253,254,255,256].
Measuring EquipmentDescriptionCaseMain HallmarksReference
Two-beam interferometryMicromachines 16 00990 i001This method determines the laser linewidth by interfering the test laser with a stabilized reference laser, precisely tuning their frequency offset, detecting the beat signal with a photodetector, and analyzing it using an electrical spectrum analyzer.1High-resolution and high-sensitivity

Strict requirements for reference light sources
0.6 Hz [228]
40 mHz [229]
100 Hz [230]
Brillouin stokes optical beating methodMicromachines 16 00990 i002This technology measures the laser linewidth by using second-order Stokes waves in a fiber resonator as intrinsic references, generating same-direction second-order Stokes via pump excitation, coupling for beat frequency, analyzing FWHM of the spectrum, and controlling resonator length with resonant tracking and piezoelectric ceramics.2

High-resolution
Simple system
(only one laser needed)

Sensitive to environment
Beat signal frequency range limitation due to large Brillouin shift
4.2 kHz [231]
100 kHz [232]
300 Hz [233]
5.5 kHz [234]
860 Hz [235]
3.84 kHz [236]
Delayed self-homodyne methodMicromachines 16 00990 i003This technology uses an unbalanced Mach-Zehnder interferometer architecture to split the tested laser into a time-delayed signal via a fiber delay line and a reference signal, generating an optical beat signal through coherent interference, converting it into an electrical signal, and extracting the laser linewidth by analyzing the PSD of the electrical signal.1Avoiding dependence on the reference laser

Zero frequency interference
18 MHz [237]
460 kHz [238]
Delayed self-heterodyne methodMicromachines 16 00990 i004Mach-Zehnder interferometer-based delayed self-heterodyne methodThe tested laser is split at Coupler1. One arm incorporates a delay fiber, to induce controlled decoherence, while the other arm routes light through an AOM generating a frequency shift. The obtained delayed beam and frequency-shifted beam are coupled at Coupler2 and detected by PD. Finally, ESA is used to collect data and display the detected signal.1Avoiding interference from zero frequency

Long fiber length can introduce 1/f noise
120 kHz [239]
2.58 kHz [240]
Micromachines 16 00990 i005Michelson interferometer-based delayed self-heterodyne methodThe system splits the tested laser via a coupler, introducing Faraday Rotator Mirror (FRM) to stabilize polarization and reduce noise. After passing through delay fiber, the laser is reflected by FRM, redoing the optical path, then interferes at the coupler for 2 times the fiber-length delay. Finally, PD detects and ESA acquires data.1FRM is introduced, maintaining the stability of polarization state and reducing the noise caused by random polarization state drift.
Long fiber length can introduce 1/f noise.
[241]
Micromachines 16 00990 i006Gain compensation loop delay self-heterodyne method
EDFA: erbium-doped optical fiber amplifier
This method uses a fiber ring’s multipass transmission to amplify delay time, reducing required delay fiber length. Resolving beams with different circulation cycles from the fiber ring enables acquiring theoretically infinite photocurrent spectral lines via ESA, with its architecture multiplying temporal delay between beams.1Can measure wide range of laser bands
Sensitive to the environmental noise
Unable to eliminate the impact of 1/f noise
680 Hz [242]
Micromachines 16 00990 i007High-coherence envelope self-coherence detection
VOA: Variable Optical Attenuator
This approach introduces two variable optical attenuators into the modified MZI-based delayed self-heterodyne method system. By establishing a quantitative mapping relationship between the contrast difference between the second spectral peak and the second trough (CDSPST) and the laser linewidth, the linewidth is calculated.1suppressing the spectral broadening induced by 1/f noise in the Gaussian linewidth of laser output150 Hz [243,244]
98 Hz [245]
609 Hz [246]
Micromachines 16 00990 i008Dual-parameter acquisition (DPA) method
SMF: Single-Mode Fiber
This method transforms conventional Lorent-zian fitting based on incoherent interference into dynamic modeling utilizing partially co-herent interference. By extracting the power difference between adjacent extrema in the first-order sidelobe and the frequency deviation between the central frequency and the zeroth-order minimum, the linewidth is measured.1No precise delay fiber length data needed458 Hz [247]
Micromachines 16 00990 i009Linewidth measurement based on short-fiber delayed self-heterodyneIn order to completely eliminate the influence of 1/f noise from the root, a linewidth measurement method using short optical fibers for delayed autoheterodyne has emerged in recent years. When the delay fiber is short, the broadening of the measured power spectrum can be effectively suppressed.1eliminate the influence of 1/f noise151 Hz [248]
944 Hz [249,250]
8 kHz [251]
2.53 kHz [252]
6.1 kHz [253,254]
1.753 kHz [255]
100 Hz [256]
The limitations of heterodyne detection methods primarily manifest in two aspects: beat signal fidelity and systemic noise contamination. First, during beat signal generation, spectral broadening models exhibit systematic deviations from actual linewidths due to compounded effects including fiber nonlinearities, harmonic distortion in photoelectric conversion, and quantization noise aliasing. Second, the frequency-domain characteristics of laser phase noise fundamentally arise from the superposition of spontaneous emission-induced 1/f noise and white noise, which dominantly governs lineshape broadening.
Two-beam interferometry, reliant on interference between independent laser sources, theoretically enable direct linewidth characterization through heterodyne signatures. However, frequency drift and relative intensity noise in reference lasers introduce uncontrollable errors. The Brillouin Stokes optical beating method exploits stimulated Brillouin scattering nonlinearities to map the frequency difference between pump and Stokes waves into measurable electrical signals. Nevertheless, pump power fluctuations and phonon relaxation time uncertainties exacerbate low-frequency noise. Delayed self-homodyne/heterodyne methods circumvent external reference interference through extended fiber delay lines in split-path configurations, yet respectively suffer from zero-frequency ambiguity and 1/f noise amplification.
To solve these predicaments, Elliott et al. were among the first to provide the theoretical relationship between laser line shape and frequency-noise PSD. Then the researchers put forward the β-separation theory which showed the β-separation line divides the frequency noise spectrum into two distinct regions [257], as illustrated in Figure 8. The first region above the β-separation line (Sυ(f) > 8ln(2)f/π2) corresponds to the low-frequency modulation domain, which is dominated by 1/f noise, fundamentally determining the intrinsic linewidth of the measured laser. The other region below the β-separation line (Sυ(f) < 8ln(2)f/π2) represents the high-frequency modulation domain characterized by white noise, resulting in a Lorentzian lineshape [258]. In this regime, the laser linewidth is exclusively governed by the spectral density of frequency noise [259].
Figure 8. Frequency noise distribution and β-separation line [260].
Figure 8. Frequency noise distribution and β-separation line [260].
Micromachines 16 00990 g008
In fact, the noise components in the first region with spectral density higher than its Fourier frequency (Sυ(f) > f) generate Gaussian autocorrelation functions. The Fourier transform of the product of these autocorrelation functions yields the laser line shape. And this line shape is a Gaussian function, whose variance is the sum of the contributions of all high-modulation index noise components. Therefore, we can obtain a good approximation of the laser linewidth through the following simple expression [261]:
F W H M = ( 8 ln ( 2 ) A ) 1 / 2
A = 1 T 0 H ( S υ ( f ) 8 ln ( 2 ) f / π 2 ) S υ ( f ) d f
where H(f) denotes the unit step function, A represents Sυ(f) above the β-separation line in the frequency noise spectrum, and T0 signifies the measurement duration.
Overall, the β-separation theory provides a theoretical basis for linewidth measurement at complete Fourier frequencies. Based on the β-separation theory, researchers have proposed various measurement methods based on phase noise. This section systematically examines four indirect approaches corresponding to Route B in Figure 6: the linewidth measurement method based on cross-correlation and the β algorithm, the frequency discrimination method, the optical coherent reception method based on an interferometer, and the optical coherent reception method based on delayed self-homodyne/self-heterodyne techniques.
The linewidth measurement method based on cross-correlation and the β algorithm integrates mature radio-frequency cross-correlation frequency noise characterization [262] with β full-frequency demodulation; the former suppresses system noise via iterative operations, the latter enables full-band linewidth calculation through frequency noise integration [263,264,265]. Specifically, the laser under test first coherently beats with two reference sources; cross-correlation of the two beat signals eliminates noise to obtain the phase-noise autocorrelation function. Fourier transform yields the phase-noise power spectrum, and subsequent data conversion with β algorithm processing derives the laser linewidth.
The frequency discrimination method measures frequency noise by converting laser frequency fluctuations into intensity variations via a discriminator [266]. Its system includes a quadrature servo control to dynamically adjust laser frequency, maintaining the bias point in quadrature for maximum sensitivity and stabilizing frequency by compensating drift through feedback. This method enables comprehensive noise analysis [267,268] and avoids low-frequency noise by locking to the quadrature point [269] or shifting to the radio frequency domain via self-heterodyne detection [270]. However, PD conversion introduces extra noise; its accuracy depends on the discriminator, and the laser must operate in a narrow frequency range, limiting achievement of the ultra-low noise floor required for millihertz-linewidth laser characterization.
Optical coherent reception based on an interferometer determines the linewidth by analyzing the relationship between the phase noise of the interferometer and the frequency noise power spectrum of the laser [259]. Its distinctive feature lies in the adoption of a phase-difference interferometer system with a 3×3 coupler to measure differential phase information. The three output ports generate split beams with equal amplitude and 120° phase differences. After the signals are collected by an analog card, the phase noise PSD is obtained by analyzing the deviation of frequency-phase characteristics from the linear fitting curve and demodulating the phase information. The laser linewidth is then calculated using the β algorithm.
The optical coherent reception method based on delayed self-homodyne/self- heterodyne realizes the effective demodulation of signal light and local oscillator light by introducing an incoherent-to-coherent receiver (ICR) [69,271,272,273]. The laser under test and the delayed laser, after passing through a polarization controller, enter the ICR with a 90° mixer and are mixed with the local oscillator light, generating in-phase/quadrature (I/Q) signals. These I/Q signals are detected by a balanced photodetector (BPD), converted into electrical signals, and then amplified by a transimpedance amplifier (TIA). The amplified signals are sent to a digital signal processor (DSP) to extract phase noise information, and finally, the linewidth of the laser under testing is calculated. Table 2 presents the measurement devices, their descriptions and main features and some references.
Table 2. Four indirect measurement methods Based on Phase Noise [159,160,265,274,275,276,277].
Table 2. Four indirect measurement methods Based on Phase Noise [159,160,265,274,275,276,277].
Measuring EquipmentDescriptionMain HallmarksReference
Cross-correlation method and β algorithmMicromachines 16 00990 i010ADC:Analog-to-Digital ConverterThe innovation of this methodology lies in the integration of the radio-frequency-domain cross-correlation frequency noise characterization technique with the β full-frequency-domain demodulation algorithm. The cross-correlation method iteratively suppresses system noise through multiple computational operations, while the β-algorithm achieves full-bandwidth linewidth resolution via frequency noise integration.Avoiding linear fitting

Suitable for any noise

Complex calculation
26.9 kHz [159]
Frequency discrimination methodMicromachines 16 00990 i011Polarization Controller: PolConFrequency discrimination is to characterize frequency noise by changing laser frequency fluctuations into intensity variations via a frequency discriminator, where the resulting intensity changes are monitored to derive frequency deviations. A quadrature servo control adjusts laser frequency dynamically via discriminator feedback, maintaining quadrature operation for optimal sensitivity and compensating drift for frequency stabilization.Avoiding interference from zero frequency

Frequency range is limited

Precise control is necessary
0.7 Hz [160]
269 Hz [274]
45 Hz [275]
Optical coherent reception methodMicromachines 16 00990 i012Based on an 120° interferometerThis method determines laser linewidth by establishing the quantitative relationship between interferometric phase noise and the laser’s frequency noise PSD. After signal collection via an analog card, the phase noise PSD is obtained by analyzing frequency-phase characteristics and linear fitting curve deviation from linear scanning, demodulating phase information, and calculating the laser linewidth using the β algorithm.Measure instantaneous phase change directly

Complex structure

High cost
4.36 kHz [265]
(minimum integrated linewidth)
3.58 kHz [265]
(minimum Lorentzian linewidth)
Optical coherent reception methodMicromachines 16 00990 i013Based on delayed self-homodyne/self-heterodyneThe tested laser is split via a coupler into two beams. One propagates directly, while the other undergoes a time delay. Both beams are then routed through a PolCon before entering the incorporating an ICR for generating I/Q signals. The amplified signals are fed into a DSP system to extract phase noise characteristics, ultimately computing the linewidth.Measure instantaneous phase change directly

Complex structure

High cost
50 kHz [276]
20 kHz ~ 2 MHz [277]

6. Linewidth Measurement Method Based on Electronic Information Processing

In the field of laser linewidth characterization, two conventional methodologies relying on beat-note power spectral analysis and phase noise evaluation have undergone continuous refinement and innovation, giving rise to diverse novel measurement approaches. These advancements not only enhance measurement precision and efficiency but also broaden applicability, enabling more accurate assessment of narrow-linewidth laser characteristics.

6.1. Frequency Comb-Based Method

This technique employs heterodyne beating between the emission frequency of the tested laser and the harmonic components of a frequency comb’s repetition rate. A voltage-controlled oscillator (VCO) tracking the beat frequency converts its frequency fluctuations into voltage signals, which are subsequently analyzed through feedback-controlled spectral measurements to derive laser linewidth.
M. Ravaro et al. [278] reported a method for measuring the intrinsic linewidth of terahertz quantum cascade lasers (QCLs) using near-infrared frequency combs. The QCL and frequency comb beams are collinearly focused onto a ZnTe crystal. A tandem configuration of waveplates and a polarizing beam splitter behind the crystal forms an ultrafast near-infrared electro-optic amplitude modulator driven by the terahertz electric field. The balanced detection output is frequency-shifted to match the operational frequency of the VCO via an RF synthesizer. For demodulation, a “tracking oscillator” technique is implemented, where the VCO converts beat frequency fluctuations into voltage signals. A fast Fourier transform analyzer measures the PSD of the output voltage, from which the frequency noise spectral density of the QCL is derived using the calibrated VCO sensitivity. At an output power of 2 mW, this system achieved an intrinsic linewidth measurement of 230 Hz.
This method is based on generating a difference frequency signal between the QCL frequency and the repetition frequency of the near-infrared optical frequency comb [279,280]. It enables the measurement of the frequency noise spectral density of terahertz QCLs at any frequency demonstrated to date. However, it is limited by the constraint that the emission frequency must fall within the comb spectrum bandwidth.

6.2. Optical Feedback Interferometry-Based Method

Compared to conventional interferometric methods, laser self-mixing techniques offer significant advantages by eliminating external detectors and simplifying optical alignment processes [281].
The tested laser is collimated via a lens and focused onto a planar mirror mounted on a piezoelectric actuator, typically driven by sinusoidal voltage modulation. A VOA ensures optimal feedback levels. Light reflected from the mirror interferes with the intracavity optical field, generating self-mixing signals, which are detected by a PD integrated at the laser rear facet, amplified, and digitized using a digital oscilloscope. Linewidth estimation is achieved through statistical analysis of interference fringe period histograms, combined with repeated root-mean-square phase noise measurements at varying target distances. M.C. Caldiroli et al. [282] implemented this methodology with an adjustable pinhole for dynamic feedback regulation, successfully characterizing a mid-infrared QCL with a measured linewidth of 280 kHz.
Compared with delayed self-homodyne and self-heterodyne methods, this method avoids extra external detectors and simplifies optical processing. Yet it has drawbacks: inability to distinguish individual noise sources (only fringe-period-averaged integrated linewidth); optical feedback-induced residual narrowing leading to narrower measured linewidths; thermal effects reducing the laser’s tuning coefficient (declining above 1 kHz [283]), with smaller coefficients reversing high-frequency fluctuations and causing underestimated linewidths.

6.3. Power Area Method (PAM)

The β-separation line method faces limitations in determining linewidth when confronted with complex frequency noise PSD profiles. PAM effectively resolves this challenge by providing estimation errors below 7% for both white noise and flicker frequency noise across nearly all measurement durations, while retaining the simplicity and intuitiveness of the β-separation line approach. Notably, PAM is applicable not only for determining laser linewidth but also for characterizing delayed self-heterodyne beat-note spectral linewidth. Zhou et al. [284] successfully employed this method to estimate linewidth in both delayed self-heterodyne and delayed self-heterodyne beat-note signals.
This method and theory may be applicable and useful in the applications of narrow-linewidth lasers, such as coherent optical communication, high-resolution spectroscopy, and optical frequency combs.
These three types of methods differ from traditional laser linewidth measurement approaches that rely on optical processing. Instead, they optimize laser linewidth measurement methods by leveraging modern electronic information processing techniques (e.g., detection via voltage signals) and algorithm optimization. Furthermore, these methods are complementary in their technical routes; the frequency comb method focuses on improving precision, optical feedback interferometry emphasizes system simplification, and the power area method is dedicated to algorithm optimization.

7. Comparative Analysis of Laser Linewidth Measurement Techniques

Figure 9 illustrates the measurement ranges of reported laser linewidth characterization methods. Selection of an appropriate methodology requires comprehensive consideration of laser performance parameters and available experimental infrastructure to ensure optimal measurement outcomes.

8. Conclusions and Perspectives

In summary, laser linewidth, an essential parameter for evaluating spectral purity and coherence, plays a pivotal role in advancing photonic technologies across diverse domains, including optical communications, quantum computing, and precision sensing. Existing linewidth measurement techniques can be broadly categorized into two regimes: wide-linewidth methods and narrow-linewidth methods. For wide-linewidth lasers, GHz-range spectroscopic methods and MHz-range F-P interferometry offer relatively mature solutions, with numerous commercial instruments readily available. For narrow-linewidth lasers, direct and indirect methods have been developed. Direct approaches rely on signal power spectral analysis and include techniques such as two-beam interferometry, Brillouin Stokes optical beating, delayed self-homodyne, and delayed self-heterodyne methods. Indirect approaches, which rely on phase noise analysis, include the linewidth measurement technique based on cross-correlation and the β-separation line algorithm, the frequency discrimination method, interferometer-based optical coherent reception, and optical coherent reception schemes employing delayed self-homodyne or self-heterodyne techniques.
At present, the convergence of photonic integration, quantum technologies, and artificial intelligence is accelerating the advancement of laser linewidth metrology toward sub-Hz precision, superior noise immunity, and intelligent real-time operation. While substantial progress has been achieved, the continued development of accurate, robust, and low-noise measurement techniques remains critical to support the evolving demands of next-generation ultra-narrow-linewidth laser systems. In particular, the rapid progress of artificial intelligence is propelling linewidth measurement systems toward greater autonomy, adaptability, and interpretability, enabling advanced data analytics, real-time system calibration, and predictive diagnostics. These innovations are expected not only to support the advancement of fundamental scientific instruments such as atomic clocks and gravitational wave detectors, but also to transform laser applications ranging from ultra-fine precision to ultra-wide fields of view, and from short-range to ultra-long-distance scenarios.

Funding

This work was supported by the National Key Research and Development Program of China (No. 2024YFE0206000), National Natural Science Foundation of China (Nos. 62375076 and 61927815), Natural Science Foundation of Tianjin City (No. 22JCYBJC01100), Natural Science Foundation of Hebei Province (No. F2023202063), Natural Science Research Program for Higher Education Institutions in Hebei Province (No. JCZX2025003), and Science and Technology Cooperation Special Project of Shijiazhuang (No. SJZZXC24006).

Data Availability Statement

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Einstein, A. On the Quantum Mechanics of Radiation. Phys. Z. 1917, 18, 121–128. [Google Scholar]
  2. Schawlow, A.L.; Townes, C.H. Infrared and Optical Masers. Phys. Rev. 1958, 112, 1940–1949. [Google Scholar] [CrossRef]
  3. Maiman, T.H. Stimulated Optical Radiation in Ruby. Nature 1960, 187, 493–494. [Google Scholar] [CrossRef]
  4. Maiman, T.H. Optical and Microwave-Optical Experiments in Ruby. Phys. Rev. Lett. 1960, 4, 564–566. [Google Scholar] [CrossRef]
  5. Zeng, X.; Cui, S.; Jiang, H.; Ruan, B.; Cheng, X.; Zhou, J.; Lin, Z.; Yang, X.; Chen, W.; Feng, Y. Single-Frequency Upconverted Laser Generation by Phase Summation. High Power Laser Sci. Eng. 2023, 11, e18. [Google Scholar] [CrossRef]
  6. Legero, T.; Matei, D.G.; Haefner, S.; Grebing, C.; Weyrich, R.; Riehle, F.; Sterr, U.; Zhang, W.; Robinson, J.; Sonderhouse, L.; et al. 1.5 μm Lasers with Sub 10 mHz Linewidth. In Proceedings of the 2017 Conference on Lasers and Electro-Optics (CLEO), San Jose, CA, USA, 14–19 May 2017. [Google Scholar]
  7. Mercer, L.B. 1/f Frequency Noise Effects on Self-Heterodyne Linewidth Measurements. J. Light. Technol. 1991, 9, 485–493. [Google Scholar] [CrossRef]
  8. Pixley, N.C.; Correll, T.L.; Pappas, D.; Matveev, O.I.; Smith, B.W.; Winefordner, J.D. Tunable Resonance Fluorescence Monochromator with Sub-Doppler Spectral Resolution. Opt. Lett. 2001, 26, 1946–1948. [Google Scholar] [CrossRef]
  9. Račiukaitis, G. Ultra-Short Pulse Lasers for Microfabrication: A Review. IEEE J. Sel. Top. Quantum Electron. 2021, 27, 1100112. [Google Scholar] [CrossRef]
  10. Wei-Lian, S. The Development of Laser Processing Technology. Laser Infrared 2006, 36, 755–758. [Google Scholar]
  11. Murray, K.K.; Seneviratne, C.A.; Ghorai, S. High Resolution Laser Mass Spectrometry Bioimaging. Methods 2016, 104, 118–126. [Google Scholar] [CrossRef]
  12. Najda, S.P.; Perlin, P.; Leszczyński, M.; Slight, T.J.; Meredith, W.; Schemmann, M.; Moseley, H.; Woods, J.A.; Valentine, R.; Kalra, S. A Multi-Wavelength (Uv to Visible) Laser System for Early Detection of Oral Cancer. In Proceedings of the Imaging, Manipulation, and Analysis of Biomolecules, Cells, and Tissues XIII, San Francisco, CA, USA, 9–11 February 2015; SPIE: Bellingham, WA, USA, 2015; Volume 9328, pp. 32–37. [Google Scholar]
  13. Kalisky, Y.; Kalisky, O. The Status of High-Power Lasers and Their Applications in the Battlefield. Opt. Eng. 2010, 49, 91003. [Google Scholar] [CrossRef]
  14. Wang, R.-F.; Zhang, Y.-P.; Xu, Z.-Y. Present Situation and Developing Trend of Application of Laser Technique to Military. Infrared Laser Eng. 2007, S1, 308–311. [Google Scholar]
  15. Predehl, K.; Grosche, G.; Raupach, S.M.F.; Droste, S.; Terra, O.; Alnis, J.; Legero, T.; Haensch, T.W.; Udem, T.H.; Holzwarth, R.; et al. A 920-Kilometer Optical Fiber Link for Frequency Metrology at the 19th Decimal Place. Science 2012, 336, 441–444. [Google Scholar] [CrossRef] [PubMed]
  16. Uchida, A.; Amano, K.; Inoue, M.; Hirano, K.; Naito, S.; Someya, H.; Oowada, I.; Kurashige, T.; Shiki, M.; Yoshimori, S.; et al. Fast Physical Random Bit Generation with Chaotic Semiconductor Lasers. Nat. Photonics 2008, 2, 728–732. [Google Scholar] [CrossRef]
  17. Marcu, A.; Stafe, M.; Barbuta, M.; Ungureanu, R.; Serbanescu, M.; Calin, B.; Puscas, N. Photon Energy Transfer on Titanium Targets for Laser Thrusters. High Power Laser Sci. Eng. 2022, 10, e27. [Google Scholar] [CrossRef]
  18. Duan, Y.; Sun, Y.; Zhu, H.; Mao, T.; Zhang, L.; Chen, X. YVO4 Cascaded Raman Laser for Five-Visible-Wavelength Switchable Emission. Opt. Lett. 2020, 45, 2564–2567. [Google Scholar] [CrossRef] [PubMed]
  19. Chen, H.; Bai, Z.; Chen, J.; Li, X.; Zhu, Z.-H.; Wang, Y.; Omatsu, T.; Mildren, R.P.; Lu, Z. Diamond Raman Vortex Lasers. ACS Photonics 2024, 12, 864–869. [Google Scholar] [CrossRef]
  20. Liu, K.; Wang, J.; Chauhan, N.; Harrington, M.W.; Nelson, K.D.; Blumenthal, D.J. Integrated Photonic Molecule Brillouin Laser with a High-Power Sub-100-mHz Fundamental Linewidth. Opt. Lett. 2023, 49, 45–48. [Google Scholar] [CrossRef]
  21. Jin, D.; Bai, Z.; Zhao, Z.; Chen, Y.; Fan, W.; Wang, Y.; Mildren, R.P.; Lü, Z. Linewidth Narrowing in Free-Space-Running Diamond Brillouin Lasers. High Power Laser Sci. Eng. 2023, 11, e47. [Google Scholar] [CrossRef]
  22. Alouini, M.; Danion, G.; Vallet, M. Self-Linewidth-Narrowing Photonic Oscillator. Opt. Express 2025, 33, 1021–1033. [Google Scholar] [CrossRef]
  23. Li, L.; Zhao, Z.; Yu, Q.; Sheng, L.; Qi, Y.; Ding, J.; Yan, B.; Wang, Y.; Lu, Z.; Bai, Z. Multistage Cyclic Filtering System Utilizing a Single-Longitudinal-Mode Brillouin Fiber Laser with a Fiber Ring and Reflective Saturable Absorber. Opt. Commun. 2025, 577, 131434. [Google Scholar] [CrossRef]
  24. Bao, X.; Chen, L. Recent Progress in Distributed Fiber Optic Sensors. Sensors 2012, 12, 8601–8639. [Google Scholar] [CrossRef] [PubMed]
  25. Shaashoua, R.; Kasuker, L.; Kishner, M.; Levy, T.; Rotblat, B.; Ben-Zvi, A.; Bilenca, A. Brillouin Gain Microscopy. Nat. Photonics 2024, 18, 836–841. [Google Scholar] [CrossRef]
  26. Chen, H.; Bai, Z.; Cai, Y.; Yang, X.; Ding, J.; Qi, Y.; Yan, B.; Li, Y.; Wang, Y.; Lu, Z. Order Controllable Enhanced Stimulated Brillouin Scattering Utilizing Cascaded Diamond Raman Conversion. Appl. Phys. Lett. 2023, 122, 092202. [Google Scholar] [CrossRef]
  27. Nishigata, Y.; Sasaki, S.; Miyamoto, K.; Omatsu, T. Cascaded Vector Vortex Mode Generation from a Solid-State Raman Laser. Appl. Opt. 2021, 60, 10638–10642. [Google Scholar] [CrossRef]
  28. Pask, H.M. The Design and Operation of Solid-State Raman Lasers. Prog. Quantum Electron. 2003, 27, 3–56. [Google Scholar] [CrossRef]
  29. Xuan, C.; Zhou, Y.; Yang, X.; Ma, Y.; Rao, A.S.; Omatsu, T.; Bai, Z.; Wan, Y.; Wen, J.; Yusufu, T. Generation of High-order Laguerre-gaussian Modes from an Optical Vortex Pumped Diamond Raman Laser. Laser Photonics Rev. 2024, 18, 2400081. [Google Scholar] [CrossRef]
  30. Antipov, S.; Sabella, A.; Williams, R.J.; Kitzler, O.; Spence, D.J.; Mildren, R.P. 1.2 kW Quasi-Steady-State Diamond Raman Laser Pumped by an M2 = 15 Beam. Opt. Lett. 2019, 44, 2506–2509. [Google Scholar] [CrossRef]
  31. Pahlavani, R.L.; Spence, D.J.; Sharp, A.O.; Mildren, R.P. Linewidth Narrowing in Raman Lasers. APL Photonics 2025, 10, 076107. [Google Scholar] [CrossRef]
  32. Liu, Z.; Fan, R.; Jin, D.; Luo, T.; Li, S.; Li, N.; Li, S.; Wang, Y.; Lu, Z. Quarter Acoustic Period Pulse Compression Using Stimulated Brillouin Scattering in PF-5060. Opt. Express 2022, 30, 12586–12595. [Google Scholar] [CrossRef]
  33. Zhu, L.; Bai, Z.; Chen, Y.; Jin, D.; Fan, R.; Qi, Y.; Ding, J.; Yan, B.; Wang, Y.; Lu, Z. The Effect of Pump Beam Focusing Characteristics on Stimulated Brillouin Scattering. Opt. Commun. 2022, 515, 128205. [Google Scholar] [CrossRef]
  34. Gyger, F.; Liu, J.; Yang, F.; He, J.; Raja, A.S.; Wang, R.N.; Bhave, S.A.; Kippenberg, T.J.; Thévenaz, L. Observation of Stimulated Brillouin Scattering in Silicon Nitride Integrated Waveguides. Phys. Rev. Lett. 2020, 124, 13902. [Google Scholar] [CrossRef]
  35. Tao, Y.; Jiang, M.; Liu, L.; Li, C.; Zhou, P.; Jiang, Z. More than 20 W, High Signal-to-Noise Ratio Single-Frequency All-Polarization-Maintaining Hybrid Brillouin/Ytterbium Fiber Laser. J. Light. Technol. 2022, 41, 678–683. [Google Scholar] [CrossRef]
  36. Deroh, M.; Lucas, E.; Hammani, K.; Millot, G.; Kibler, B. Stabilized Single-Frequency Sub-kHz Linewidth Brillouin Fiber Laser Cavity Operating at 1 μm. Appl. Opt. 2023, 62, 8109–8114. [Google Scholar] [CrossRef] [PubMed]
  37. Cai, Y.; Gao, F.; Chen, H.; Yang, X.; Bai, Z.; Qi, Y.; Wang, Y.; Lu, Z.; Ding, J. Continuous-Wave Diamond Laser with a Tunable Wavelength in Orange–Red Wavelength Band. Opt. Commun. 2023, 528, 128985. [Google Scholar] [CrossRef]
  38. Sun, Y.; Yang, X.; Li, M.; Zeng, X.; Jiang, H.; Feng, Y. Diamond Guide Star Laser Pulsed at a Larmor Frequency. Opt. Express 2024, 32, 46345–46352. [Google Scholar] [CrossRef]
  39. Zhang, W.; Kittlaus, E.; Savchenkov, A.; Iltchenko, V.; Yi, L.; Papp, S.B.; Matsko, A. Monolithic Optical Resonator for Ultrastable Laser and Photonic Millimeter-Wave Synthesis. Commun. Phys. 2024, 7, 177. [Google Scholar] [CrossRef]
  40. Matsui, Y.; Eriksson, U.; Wesstrom, J.-O.; Liu, Y.; Hammerfeldt, S.; Hassler, M.; Stoltz, B.; Carlsson, N.; Siraj, S.; Goobar, E.; et al. Narrow Linewidth Tunable Semiconductor Laser. In Proceedings of the 2016 Compound Semiconductor Week (CSW) Includes 28th International Conference on Indium Phosphide & Related Materials (IPRM) & 43rd International Symposium on Compound Semiconductors (ISCS), Toyama, Japan, 26–30 June 2016. [Google Scholar]
  41. Wang, Q.; Guo, J.; Chen, W.; Liu, J.; Zhu, N. Widely Tunable Distributed Feedback Semiconductor Lasers with Constant Power and Narrow Linewidth. Chin. J. Lasers 2017, 44, 0101004. [Google Scholar] [CrossRef]
  42. Glebov, A.L.; Leisher, P.O.; Wicht, A.; Bawamia, A.; Krüger, M.; Kürbis, C.; Schiemangk, M.; Smol, R.; Peters, A.; Tränkle, G. Narrow Linewidth Diode Laser Modules for Quantum Optical Sensor Applications in the Field and in Space. Soc. Photo-Opt. Instrum. Eng. (SPIE) Conf. Ser. 2017, 10085, 100850F. [Google Scholar]
  43. Ishii, H.; Fujiwara, N.; Watanabe, K.; Kanazawa, S.; Itoh, M.; Takenouchi, H.; Miyamoto, Y.; Kasai, K.; Nakazawa, M. Narrow Linewidth Tunable DFB Laser Array Integrated with Optical Feedback Planar Lightwave Circuit (PLC). IEEE J. Sel. Top. Quantum Electron. 2017, 23, 1501007. [Google Scholar] [CrossRef]
  44. Argence, B.; Chanteau, B.; Lopez, O.; Nicolodi, D.; Abgrall, M.; Chardonnet, C.; Daussy, C.; Darquie, B.; Le Coq, Y.; Amy-Klein, A. Quantum Cascade Laser Frequency Stabilization at the Sub-Hz Level. Nat. Photonics 2015, 9, 456–460. [Google Scholar] [CrossRef]
  45. Deng, X.; Liu, J.; Zang, Q.; Jiao, D.; Gao, J.; Zhang, X.; Wang, D.; Dong, R.; Liu, T. Coherent Phase Transfer via Fiber Using Heterodyne Optical Phase Locking as Optical Amplification. Appl. Opt. 2018, 57, 9620–9624. [Google Scholar] [CrossRef] [PubMed]
  46. Wu, T.-C.; Chi, Y.-C.; Wang, H.-Y.; Tsai, C.-T.; Lin, G.-R. Blue Laser Diode Enables Underwater Communication at 12.4 Gbps. Sci. Rep. 2017, 7, 40480. [Google Scholar] [CrossRef] [PubMed]
  47. Giuliano, G.; Laycock, L.; Rowe, D.; Kelly, A.E. Solar Rejection in Laser Based Underwater Communication Systems. Opt. Express 2017, 25, 33066–33077. [Google Scholar] [CrossRef]
  48. Agrell, E.; Karlsson, M.; Chraplyvy, A.R.; Richardson, D.J.; Krummrich, P.M.; Winzer, P.; Roberts, K.; Fischer, J.K.; Savory, S.J.; Eggleton, B.J. Roadmap of Optical Communications. J. Opt. 2016, 18, 63002. [Google Scholar] [CrossRef]
  49. Biswas, A.; Kovalik, J.M.; Srinivasan, M.; Shaw, M.; Piazzolla, S.; Wright, M.W.; Farr, W.H. Deep Space Laser Communications. In Proceedings of the Free-Space Laser Communication and Atmospheric Propagation XXVIII, San Francisco, CA, USA, 13–18 February 2016; SPIE: Bellingham, WA, USA, 2016; Volume 9739, pp. 209–223. [Google Scholar]
  50. Khalighi, M.A.; Uysal, M. Survey on Free Space Optical Communication: A Communication Theory Perspective. IEEE Commun. Surv. Tuts. 2014, 16, 2231–2258. [Google Scholar] [CrossRef]
  51. Mansour, A.; Mesleh, R.; Abaza, M. New Challenges in Wireless and Free Space Optical Communications. Opt. Lasers Eng. 2017, 89, 95–108. [Google Scholar] [CrossRef]
  52. Liu, L. Laser Communications in Space I Optical Link and Terminal Technology. Chin. J. Lasers 2007, 34, 3. [Google Scholar]
  53. Liang, J.; Chaudhry, A.U.; Yanikomeroglu, H. Phasing Parameter Analysis for Satellite Collision Avoidance in Starlink and Kuiper Constellations. In Proceedings of the 2021 IEEE 4th 5G World Forum (5GWF), Montreal, QC, Canada, 13–15 October 2021; IEEE: New York, NY, USA, 2021; pp. 493–498. [Google Scholar]
  54. Chaudhry, A.U.; Yanikomeroglu, H. Laser Intersatellite Links in a Starlink Constellation: A Classification and Analysis. IEEE Veh. Technol. Mag. 2021, 16, 48–56. [Google Scholar] [CrossRef]
  55. Zhai, H.; Zhang, Z.; Zhang, H.; Wang, B.; Zhao, Y.; Zhang, J. Design and Implementation of the Hardware Platform of Satellite Optical Switching Node. In Proceedings of the 2021 19th International Conference on Optical Communications and Networks (ICOCN), Qufu, China, 23–27 August 2021; IEEE: New York, NY, USA, 2021; pp. 1–3. [Google Scholar]
  56. Guan, H.; Novack, A.; Galfsky, T.; Ma, Y.; Fathololoumi, S.; Horth, A.; Huynh, T.N.; Roman, J.; Shi, R.; Caverley, M.; et al. Widely-Tunable, Narrow-Linewidth III-V/Silicon Hybrid External-Cavity Laser for Coherent Communication. Opt. Express 2018, 26, 7920–7933. [Google Scholar] [CrossRef]
  57. Zhou, K.; Zhao, Q.; Huang, X.; Yang, C.; Li, C.; Zhou, E.; Xu, X.; Wong, K.K.Y.; Cheng, H.; Gan, J.; et al. kHz-Order Linewidth Controllable 1550 nm Single-Frequency Fiber Laser for Coherent Optical Communication. Opt. Express 2017, 25, 19752–19759. [Google Scholar] [CrossRef]
  58. Li, R.; Lin, B.; Liu, Y.; Dong, M.; Zhao, S. A Survey on Laser Space Network: Terminals, Links, and Architectures. IEEE Access 2022, 10, 34815–34834. [Google Scholar] [CrossRef]
  59. Chen, Q.; Lu, Q.; Guo, W. Theory and Simulation of Multi-Channel Interference (MCI) Widely Tunable Lasers. Opt. Express 2015, 23, 18040–18051. [Google Scholar] [CrossRef] [PubMed]
  60. Jiang, C.; Chen, Q.; Wang, K.; Lu, Q.; Lu, M.; Guo, W. Narrow-Linewidth Thermally Tuned Multi-Channel Interference Laser Integrated with a SOA and Spot Size Converter. Opt. Express 2021, 29, 13246–13255. [Google Scholar] [CrossRef]
  61. Chen, Q.; Jiang, C.; Wang, K.; Zhang, M.; Ma, X.; Liu, Y.; Lu, Q.; Guo, W. Narrow-Linewidth Thermally Tuned Multi-Channel Interference Widely Tunable Semiconductor Laser with Thermal Tuning Power below 50 mW. Photonics Res. 2020, 8, 671–676. [Google Scholar] [CrossRef]
  62. Otuya, D.O.; Kasai, K.; Hirooka, T.; Yoshida, M.; Nakazawa, M. 1.92 Tbit/s, 64 QAM Coherent Nyquist Pulse Transmission over 150 km with a Spectral Efficiency of 7.5 Bit/s/Hz. In Proceedings of the OFC 2014, San Francisco, CA, USA, 9–13 March 2014; IEEE: New York, NY, USA, 2014; pp. 1–3. [Google Scholar]
  63. Schuh, K.; Buchali, F.; Idler, W.; Eriksson, T.A.; Schmalen, L.; Templ, W.; Altenhain, L.; Dümler, U.; Schmid, R.; Möller, M. Single Carrier 1.2 Tbit/s Transmission over 300 Km with PM-64 QAM at 100 Gbaud. In Proceedings of the Optical Fiber Communication Conference, Los Angeles, CA, USA, 19–23 March 2017; Optica Publishing Group: Washington, DC, USA, 2017; p. Th5B.5. [Google Scholar]
  64. Kazovsky, L. Performance Analysis and Laser Linewidth Requirements for Optical PSK Heterodyne Communications Systems. J. Light. Technol. 1986, 4, 415–425. [Google Scholar] [CrossRef]
  65. Zhang, H.; Li, Z.; Kavanagh, N.; Zhao, J.; Ye, N.; Chen, Y.; Wheeler, N.V.; Wooler, J.P.; Hayes, J.R.; Sandoghchi, S.R.; et al. 81 Gb/s WDM Transmission at 2μm over 1.15 km of Low-Loss Hollow Core Photonic Bandgap Fiber. In Proceedings of the 2014 the European Conference on Optical Communication (ECOC), Cannes, France, 21–25 September 2014. [Google Scholar]
  66. Gwyn, S.; Watson, S.; Slight, T.; Knapp, M.; Viola, S.; Ivanov, P.; Zhang, W.; Yadav, A.; Rafailov, E.; Haji, M.; et al. Dynamic Device Characteristics and Linewidth Measurement of InGaNGaN Laser Diodes. IEEE Photonics J. 2021, 13, 1500510. [Google Scholar] [CrossRef]
  67. Ip, E.; Kahn, J.M. Addendum to “Feedforward Carrier Recovery for Coherent Optical Communications”. J. Light. Technol. 2009, 27, 2552–2553. [Google Scholar] [CrossRef]
  68. Pfau, T.; Hoffmann, S.; Noe, R. Hardware-Efficient Coherent Digital Receiver Concept with Feedforward Carrier Recovery for M-QAM Constellations. J. Light. Technol. 2009, 27, 989–999. [Google Scholar] [CrossRef]
  69. Kikuchi, K. Fundamentals of Coherent Optical Fiber Communications. J. Light. Technol. 2016, 34, 157–179. [Google Scholar] [CrossRef]
  70. Nakazawa, M.; Okamoto, S.; Omiya, T.; Kasai, K.; Yoshida, M. 256 QAM (64 Gbit/s) Coherent Optical Transmission over 160 km with an Optical Bandwidth of 5.4 GHz. In Proceedings of the Optical Fiber Communication Conference, San Diego, CA, USA, 21–25 March 2010; Optica Publishing Group: Washington, DC, USA, 2010; p. OMJ5. [Google Scholar]
  71. Ip, E.; Kahn, J. Carrier Synchronization for 3-and 4-Bit-per-Symbol Optical Transmission. J. Light. Technol. 2005, 23, 4110–4124. [Google Scholar] [CrossRef]
  72. Barry, J.R.; Kahn, J.M. Carrier Synchronization for Homodyne and Heterodyne Detection of Optical Quadriphase-Shift Keying. J. Light. Technol. 1992, 10, 1939–1951. [Google Scholar] [CrossRef]
  73. Al-Dabbagh, M.D.; Smith, J.; Kleine-Ostmann, T.; Naftaly, M.; Fatadin, I. Characterization of Photonic-Assisted Free-Space Sub-THz Data Transmission. In Proceedings of the 2023 48th International Conference on Infrared, Millimeter, and Terahertz Waves (IRMMW-THz), Montreal, QC, Canada, 17–22 September 2023; IEEE: New York, NY, USA, 2023. [Google Scholar]
  74. Chen, Z.; Wang, K.; Chen, Q.; Jiang, C.; Wei, Y.; Liu, H.; Lu, M.; Lu, Q.; Guo, W. Nano-ITLA Based on Thermo-Optically Tuned Multi-Channel Interference Widely Tunable Laser. J. Light. Technol. 2023, 41, 5405–5411. [Google Scholar] [CrossRef]
  75. Seimetz, M. Osa Laser Linewidth Limitations for Optical Systems with High-Order Modulation Employing Feed Forward Digital Carrier Phase Estimation. In Proceedings of the Optical Fiber Communication Conference, San Diego, CA, USA, 24–28 February 2008; pp. 2470–2472. [Google Scholar]
  76. Li, M.; Guo, Y.; Wang, X.; Fu, W.; Zhang, Y.; Wang, Y. Researching Pointing Error Effect on Laser Linewidth Tolerance in Space Coherent Optical Communication Systems. Opt. Express 2022, 30, 5769–5787. [Google Scholar] [CrossRef] [PubMed]
  77. Lowery, A.J. Amplified-Spontaneous Noise Limit of Optical OFDM Lightwave Systems. Opt. Express 2008, 16, 860–865. [Google Scholar] [CrossRef] [PubMed]
  78. Ferrero, V.; Camatel, S. Optical Phase Locking Techniques: An Overview and a Novel Method Based on Single Side Sub-Carrier Modulation. Opt. Express 2008, 16, 818–828. [Google Scholar] [CrossRef]
  79. Li, Y.; Lin, X.; Yang, Y.; Xia, Y.; Xiong, J.; Song, S.; Liu, L.; Chen, Z.; Cheng, X.; Li, F. Temperature Characteristics at Altitudes of 5–80 km with a Self-Calibrated Rayleigh–Rotational Raman Lidar: A Summer Case Study. J. Quant. Spectrosc. Radiat. Transf. 2017, 188, 94–102. [Google Scholar] [CrossRef]
  80. Xia, Y.; Wang, Z.L.; Cheng, X.W.; Yang, G.T.; Du, L.F.; Wang, J.H.; Yang, Y.; Li, Y.J.; Xiong, J.; Li, F.Q.; et al. All-Solid-State Narrowband Sodium Lidar System and Preliminary Result. Chin. J. Lasers 2015, 42, S113003. [Google Scholar] [CrossRef]
  81. Yang, F.; Ye, Q.; Pan, Z.; Chen, D.; Cai, H.; Qu, R.; Yang, Z.; Zhang, Q. 100-mW Linear Polarization Single-Frequency All-Fiber Seed Laser for Coherent Doppler Lidar Application. Opt. Commun. 2012, 285, 149–152. [Google Scholar] [CrossRef]
  82. Hostetler, C.A.; Behrenfeld, M.J.; Hu, Y.; Hair, J.W.; Schulien, J.A. Spaceborne Lidar in the Study of Marine Systems. In Annual Review of Marine Science; Carlson, C., Giovannoni, S., Eds.; Annual Review: San Mateo, CA, USA, 2018; Volume 10, pp. 121–147. ISBN 1941-1405. [Google Scholar]
  83. Edner, H.; Sunesson, A.; Svanberg, S. NO Plume Mapping by Laser-Radar Techniques. Opt. Lett. 1988, 13, 704–706. [Google Scholar] [CrossRef]
  84. Carlson, C.G.; Dragic, P.D.; Price, R.K.; Coleman, J.J.; Swenson, G.R. A Narrow-Linewidth, Yb Fiber-Amplifier-Based Upper Atmospheric Doppler Temperature Lidar. IEEE J. Sel. Top. Quantum Electron. 2009, 15, 451–461. [Google Scholar] [CrossRef]
  85. Geng, J.; Spiegelberg, C.; Jiang, S. Narrow Linewidth Fiber Laser for 100-km Optical Frequency Domain Reflectometry. Photonics Technol. Lett. IEEE 2005, 17, 1827–1829. [Google Scholar] [CrossRef]
  86. Hu, X.; Yan, Z.A.; Guo, S.Y.; Cheng, Y.Q.; Gong, J.C. Sodium Fluorescence Doppler Lidar to Measure Atmospheric Temperature in the Mesopause Region. Chin. Sci. Bull. 2011, 56, 417–423. [Google Scholar] [CrossRef]
  87. Li, T.; Fang, X.; Liu, W.; Gu, S.-Y.; Dou, X. Narrowband Sodium Lidar for the Measurements of Mesopause Region Temperature and Wind. Appl. Opt. 2012, 51, 5401–5411. [Google Scholar] [CrossRef]
  88. Chen, T.-H.; Huang, C.-Y.; Shia, T.K.; Wun, S.-J.; Hsu, C.-H.; Ku, K.-N.; Lee, C.-S.; Lin, C.-Y.; Chang, P.-C.; Wang, C.-C. A Frequency Digital Pre-Distortion Compensation Method for FMCW LiDAR System. In Proceedings of the Optical Fiber Communication Conference, San Diego, CA, USA, 8–12 March 2020; Optica Publishing Group: Washington, DC, USA, 2020; p. Th2A.23. [Google Scholar]
  89. Sun, J.; Hosseini, E.S.; Yaacobi, A.; Cole, D.B.; Leake, G.; Coolbaugh, D.; Watts, M.R. Two-Dimensional Apodized Silicon Photonic Phased Arrays. Opt. Lett. 2014, 39, 367–370. [Google Scholar] [CrossRef] [PubMed]
  90. Suh, M.-G.; Vahala, K.J. Soliton Microcomb Range Measurement. Science 2018, 359, 884–887. [Google Scholar] [CrossRef] [PubMed]
  91. Wang, Y.; Hua, Z.; Shi, J.; Dai, Z.; Wang, J.; Shao, L.; Tan, Y. Laser Feedback Frequency-Modulated Continuous-Wave LiDAR and 3-D Imaging. IEEE Trans. Instrum. Meas. 2023, 72, 9. [Google Scholar] [CrossRef]
  92. Liu, C.; Guo, Y.; Zhou, Y.; Li, X.; Lu, L.; Li, Y.; Bao, W.; Chen, J.; Zhou, L. Fast-tuning and Narrow-linewidth Hybrid Laser for FMCW Ranging. Laser Photonics Rev. 2025, 19, 2401338. [Google Scholar] [CrossRef]
  93. Morton, P.A.; Xiang, C.; Khurgin, J.B.; Morton, C.D.; Tran, M.; Peters, J.; Guo, J.; Morton, M.J.; Bowers, J.E. Integrated Coherent Tunable Laser (ICTL) with Ultra-Wideband Wavelength Tuning and Sub-100 Hz Lorentzian Linewidth. J. Light. Technol. 2022, 40, 1802–1809. [Google Scholar] [CrossRef]
  94. Li, B.; Luo, S.; Yu, A.; Gao, J.; Sun, P.; Wang, X.; Zuo, D. Sensitive Raman Gas Analysis Using a 500 mW External Cavity Diode Laser at 410 nm. Laser Phys. Lett. 2017, 14, 95701. [Google Scholar] [CrossRef]
  95. Al-Taiy, H.; Wenzel, N.; PreußLer, S.; Klinger, J.; Schneider, T. Ultra-Narrow Linewidth, Stable and Tunable Laser Source for Optical Communication Systems and Spectroscopy. Opt. Lett. 2014, 39, 5826–5829. [Google Scholar] [CrossRef]
  96. Jadbabaie, A.; Pilgram, N.H.; Kłos, J.; Kotochigova, S.; Hutzler, N.R. Enhanced Molecular Yield from a Cryogenic Buffer Gas Beam Source via Excited State Chemistry. New J. Phys. 2020, 22, 22002. [Google Scholar] [CrossRef]
  97. Mitra, D.; Vilas, N.B.; Hallas, C.; Anderegg, L.; Augenbraun, B.L.; Baum, L.; Miller, C.; Raval, S.; Doyle, J.M. Direct Laser Cooling of a Symmetric Top Molecule. Science 2020, 369, 1366–1369. [Google Scholar] [CrossRef]
  98. Anderegg, L.; Cheuk, L.W.; Bao, Y.; Burchesky, S.; Ketterle, W.; Ni, K.-K.; Doyle, J.M. An Optical Tweezer Array of Ultracold Molecules. Science 2019, 365, 1156–1158. [Google Scholar] [CrossRef] [PubMed]
  99. Feng, K.; Cui, J.; Dang, H.; Zhao, S.; Wu, W.; Tan, J. Investigation and Development of a High Spectral Resolution Coherent Optical Spectrum Analysis System. Opt. Express 2016, 24, 25389–25402. [Google Scholar] [CrossRef] [PubMed]
  100. Dong, Y.; Jiang, T.; Teng, L.; Zhang, H.; Chen, L.; Bao, X.; Lu, Z. Sub-MHz Ultrahigh-Resolution Optical Spectrometry Based on Brillouin Dynamic Gratings. Opt. Lett. 2014, 39, 2967–2970. [Google Scholar] [CrossRef] [PubMed]
  101. Verma, M.; Jayich, A.M.; Vutha, A.C. Electron Electric Dipole Moment Searches Using Clock Transitions in Ultracold Molecules. Phys. Rev. Lett. 2020, 125, 153201. [Google Scholar] [CrossRef]
  102. Yu, P.; Hutzler, N.R. Probing Fundamental Symmetries of Deformed Nuclei in Symmetric Top Molecules. Phys. Rev. Lett. 2021, 126, 023003. [Google Scholar] [CrossRef]
  103. National Academies of Sciences, Engineering, and Medicine; Division on Engineering and Physical Sciences; Board on Physics and Astronomy; Committee on Decadal Assessment and Outlook Report on Atomic, Molecular, and Optical Science. Manipulating Quantum Systems: An Assessment of Atomic, Molecular, and Optical Physics in the United States; National Academies Press: Washington, DC, USA, 2020; ISBN 0-309-49951-8. [Google Scholar]
  104. Wang, J.; Qi, K.Q.; Wang, S.X.; Gao, R.; Li, P.; Yang, R.; Liu, H.; Luo, Z. Advance and prospect in the study of laser interferometry technology for space gravitational wave detection. Sci. Sin-Phys Mech. Astron. 2024, 54, 109–127. (In Chinese) [Google Scholar] [CrossRef]
  105. Abbott, B.P.; Abbott, R.; Abbott, T.D.; Abernathy, M.R.; Acernese, F.; Ackley, K.; Adams, C.; Adams, T.; Addesso, P.; Adhikari, R.X.; et al. Observation of Gravitational Waves from a Binary Black Hole Merger. Phys. Rev. Lett. 2016, 116, 061102. [Google Scholar] [CrossRef]
  106. Abbott, B.P.; Abbott, R.; Abbott, T.D.; Abraham, S.; Acernese, F.; Ackley, K.; Adams, C.; Adhikari, R.X.; Adya, V.B.; Affeldt, C. Binary Black Hole Population Properties Inferred from the First and Second Observing Runs of Advanced LIGO and Advanced Virgo. Astrophys. J. Lett. 2019, 882, L24. [Google Scholar] [CrossRef]
  107. Abbott, B.P.; Abbott, R.; Abbott, T.; Abraham, S.; Acernese, F.; Ackley, K.; Adams, C.; Adhikari, R.X.; Adya, V.B.; Affeldt, C. GWTC-1: A Gravitational-Wave Transient Catalog of Compact Binary Mergers Observed by LIGO and Virgo during the First and Second Observing Runs. Phys. Rev. X 2019, 9, 31040. [Google Scholar] [CrossRef]
  108. Kapasi, D.P.; Eichholz, J.; McRae, T.; Ward, R.L.; Slagmolen, B.J.J.; Legge, S.; Hardman, K.S.; Altin, P.A.; McClelland, D.E. Tunable Narrow-Linewidth Laser at 2 μm Wavelength for Gravitational Wave Detector Research. Opt. Express 2020, 28, 3280–3288. [Google Scholar] [CrossRef] [PubMed]
  109. Harry, G.M.; LIGO Sci Collaboration. Advanced LIGO: The next Generation of Gravitational Wave Detectors. Class. Quantum Gravity 2010, 27, 084006. [Google Scholar] [CrossRef]
  110. Kwee, P.; Bogan, C.; Danzmann, K.; Frede, M.; Kim, H.; King, P.; Poeld, J.; Puncken, O.; Savage, R.L.; Seifert, F.; et al. Stabilized High-Power Laser System for the Gravitational Wave Detector Advanced LIGO. Opt. Express 2012, 20, 10617–10634. [Google Scholar] [CrossRef]
  111. Nicholson, T.L.; Campbell, S.L.; Hutson, R.B.; Marti, G.E.; Bloom, B.J.; McNally, R.L.; Zhang, W.; Barrett, M.D.; Safronova, M.S.; Strouse, G.F.; et al. Systematic Evaluation of an Atomic Clock at 2 × 10−18 Total Uncertainty. Nat. Commun. 2015, 6, 6896. [Google Scholar] [CrossRef]
  112. Ludlow, A.D.; Boyd, M.M.; Ye, J.; Peik, E.; Schmidt, P.O. Optical Atomic Clocks. Rev. Mod. Phys. 2015, 87, 637–701. [Google Scholar] [CrossRef]
  113. Bloom, B.J.; Nicholson, T.L.; Williams, J.R.; Campbell, S.L.; Bishof, M.; Zhang, X.; Zhang, W.; Bromley, S.L.; Ye, J. An Optical Lattice Clock with Accuracy and Stability at the 10−18 Level. Nature 2014, 506, 71–75. [Google Scholar] [CrossRef]
  114. Ushijima, I.; Takamoto, M.; Das, M.; Ohkubo, T.; Katori, H. Cryogenic Optical Lattice Clocks. Nat. Photonics 2015, 9, 185–189. [Google Scholar] [CrossRef]
  115. Huntemann, N.; Sanner, C.; Lipphardt, B.; Tamm, C.; Peik, E. Single-Ion Atomic Clock with 3 × 10−18 Systematic Uncertainty. Phys. Rev. Lett. 2016, 116, 63001. [Google Scholar] [CrossRef]
  116. Takamoto, M.; Takano, T.; Katori, H. Frequency Comparison of Optical Lattice Clocks beyond the Dick Limit. Nat. Photonics 2011, 5, 288–292. [Google Scholar] [CrossRef]
  117. Young, A.W.; Eckner, W.J.; Milner, W.R.; Kedar, D.; Norcia, M.A.; Oelker, E.; Schine, N.; Ye, J.; Kaufman, A.M. Half-Minute-Scale Atomic Coherence and High Relative Stability in a Tweezer Clock. Nature 2020, 588, 408–413. [Google Scholar] [CrossRef]
  118. Newman, Z.L.; Maurice, V.; Drake, T.; Stone, J.R.; Briles, T.C.; Spencer, D.T.; Fredrick, C.; Li, Q.; Westly, D.; Ilic, B.R. Architecture for the Photonic Integration of an Optical Atomic Clock. Optica 2019, 6, 680–685. [Google Scholar] [CrossRef]
  119. Zhu, T.; He, Q.; Xiao, X.; Bao, X. Modulated Pulses Based Distributed Vibration Sensing with High Frequency Response and Spatial Resolution. Opt. Express 2013, 21, 2953–2963. [Google Scholar] [CrossRef] [PubMed]
  120. Juarez, J.C.; Taylor, H.F. Field Test of a Distributed Fiber-Optic Intrusion Sensor System for Long Perimeters. Appl. Opt. 2007, 46, 1968–1971. [Google Scholar] [CrossRef] [PubMed]
  121. Liu, T.; Du, Y.; Ding, Z.; Liu, K.; Zhou, Y.; Jiang, J. 40-km OFDR-Based Distributed Disturbance Optical Fiber Sensor. IEEE Photon. Technol. Lett. 2015, 28, 771–774. [Google Scholar] [CrossRef]
  122. Wang, Z.N.; Zeng, J.J.; Li, J.; Fan, M.Q.; Wu, H.; Peng, F.; Zhang, L.; Zhou, Y.; Rao, Y.J. Ultra-Long Phase-Sensitive OTDR with Hybrid Distributed Amplification. Opt. Lett. 2014, 39, 5866–5869. [Google Scholar] [CrossRef]
  123. Harris, M.; Pearson, G.; Vaughan, J.; Letalick, D.; Karlsson, C. The Role of Laser Coherence Length in Continuous-Wave Coherent Laser Radar. J. Mod. Opt. 1998, 45, 1567–1581. [Google Scholar] [CrossRef]
  124. Scully, M.O.; Lamb, W.E. Quantum Theory of an Optical Maser. I. General Theory. Phys. Rev. 1967, 159, 208–226. [Google Scholar] [CrossRef]
  125. Krawinkel, T. Improved GNSS Navigation with Chip-Scale Atomic Clocks; Verlag der Bayerischen Akademie der Wissenschaften: Munich, Germany, 2018; ISBN 3-7696-5235-5. [Google Scholar]
  126. Henderson, S.W.; Suni, P.J.; Hale, C.P.; Hannon, S.M.; Magee, J.R.; Bruns, D.L.; Yuen, E.H. Coherent Laser Radar at 2 mu m Using Solid-State Lasers. IEEE Trans. Geosci. Remote Sens. 1993, 31, 4–15. [Google Scholar] [CrossRef]
  127. Koch, G.J.; Beyon, J.Y.; Barnes, B.W.; Petros, M.; Yu, J.; Amzajerdian, F.; Kavaya, M.J.; Singh, U.N. High-Energy 2 μm Doppler Lidar for Wind Measurements. Opt. Eng. 2007, 46, 116201. [Google Scholar] [CrossRef]
  128. Koch, G.J.; Beyon, J.Y.; Petzar, P.J.; Petros, M.; Yu, J.; Trieu, B.C.; Kavaya, M.J.; Singh, U.N.; Modlin, E.A.; Barnes, B.W.; et al. Field Testing of a High-Energy 2-μm Doppler Lidar. J. Appl. Remote Sens. 2010, 4, 043512. [Google Scholar] [CrossRef]
  129. Fan, Y.; Oldenbeuving, R.M.; Roeloffzen, C.G.; Hoekman, M.; Geskus, D.; Heideman, R.G.; Boller, K.-J. 290 Hz Intrinsic Linewidth from an Integrated Optical Chip-Based Widely Tunable InP-Si3N4 Hybrid Laser. In Proceedings of the 2017 Conference on Lasers and Electro-Optics (CLEO), San Jose, CA, USA, 14–19 May 2017; IEEE: New York, NY, USA, 2017; pp. 1–2. [Google Scholar]
  130. Sun, G.W.; Wei, F.; Zhang, L. Low-Noise External Cavity Semiconductor Lasers Based on Polarization-Maintaining Fiber Bragg Gratings. Chin. J. Laser 2018, 45, 0601004. [Google Scholar]
  131. Fan, Y.; Epping, J.P.; Oldenbeuving, R.M.; Roeloffzen, C.G.; Hoekman, M.; Dekker, R.; Heideman, R.G.; van der Slot, P.J.; Boller, K.-J. Optically Integrated InP-Si3N4 Hybrid Laser. IEEE Photonics J. 2016, 8, 1505111. [Google Scholar] [CrossRef]
  132. Ding, D.; Lv, X.; Chen, X.; Wang, F.; Zhang, J.; Che, K. Tunable High-Power Blue External Cavity Semiconductor Laser. Opt. Laser Technol. 2017, 94, 1–5. [Google Scholar] [CrossRef]
  133. Yu, L.; Lu, D.; Pan, B.; Zhang, L.; Guo, L.; Li, Z.; Zhao, L.J. Widely Tunable Narrow-Linewidth Lasers Using Self-Injection DBR Lasers. IEEE Photonics Technol. Lett. 2014, 27, 50–53. [Google Scholar] [CrossRef]
  134. Shin, D.K.; Henson, B.M.; Khakimov, R.I.; Ross, J.A.; Dedman, C.J.; Hodgman, S.S.; Baldwin, K.G.H.; Truscott, A.G. Widely Tunable, Narrow Linewidth External-Cavity Gain Chip Laser for Spectroscopy between 1.0-1.1 μm. Opt. Express 2016, 24, 27403–27414. [Google Scholar] [CrossRef]
  135. Luvsandamdin, E.; Spieberger, S.; Schiemangk, M.; Sahm, A.; Mura, G.; Wicht, A.; Peters, A.; Erbert, G.; Trankle, G. Development of Narrow Linewidth, Micro-Integrated Extended Cavity Diode Lasers for Quantum Optics Experiments in Space. Appl. Phys. B 2013, 111, 255–260. [Google Scholar] [CrossRef]
  136. Aoyama, K.; Yoshioka, R.; Yokota, N.; Kobayashi, W.; Yasaka, H. Experimental Demonstration of Linewidth Reduction of Laser Diode by Compact Coherent Optical Negative Feedback System. Appl. Phys. Express 2014, 7, 122701. [Google Scholar] [CrossRef]
  137. Guo, Y.; Li, X.; Xu, W.; Liu, C.; Jin, M.; Lu, L.; Xie, J.; Stroganov, A.; Chen, J.; Zhou, L. A Hybrid-Integrated External Cavity Laser with Ultra-Wide Wavelength Tuning Range and High Side-Mode Suppression. In Proceedings of the 2022 Optical Fiber Communications Conference and Exhibition (OFC), San Diego, CA, USA, 6–10 March 2022; IEEE: New York, NY, USA, 2022; pp. 1–3. [Google Scholar]
  138. Daiber, A. Narrow-Linewidth Tunable External Cavity Laser for Coherent Communication. In Proceedings of the 2014 IEEE Photonics Conference, San Diego, CA, USA, 12–16 October 2014; IEEE: New York, NY, USA, 2014; pp. 447–448. [Google Scholar]
  139. Guo, Y.; Li, X.; Jin, M.; Lu, L.; Xie, J.; Chen, J.; Zhou, L. Hybrid Integrated External Cavity Laser with a 172-nm Tuning Range. APL Photonics 2022, 7, 066101. [Google Scholar] [CrossRef]
  140. Tang, L.; Yang, S.; Chen, H.; Chen, M. Hybrid Integrated Low Noise Optical Phase-Locked Loop Based on Self-Injection Locked Semiconductor Laser. J. Light. Technol. 2021, 40, 2033–2039. [Google Scholar] [CrossRef]
  141. Qin, J.; Zhang, L.; Xie, W.; Cheng, R.; Liu, Z.; Wei, W.; Dong, Y. Ultra-Long Range Optical Frequency Domain Reflectometry Using a Coherence-Enhanced Highly Linear Frequency-Swept Fiber Laser Source. Opt. Express 2019, 27, 19359–19368. [Google Scholar] [CrossRef] [PubMed]
  142. Binaie, A.; Ahasan, S.; Krishnaswamy, H. A Spurless and Wideband Continuous-Time Electro-Optical Phase Locked Loop (CT-EOPLL) for High Performance LiDAR. IEEE Open J. Solid-State Circuits Soc. 2021, 1, 235–246. [Google Scholar] [CrossRef]
  143. Idjadi, M.H.; Aflatouni, F. Integrated Pound−Drever−Hall Laser Stabilization System in Silicon. Nat. Commun. 2017, 8, 1209. [Google Scholar] [CrossRef] [PubMed]
  144. Lihachev, G.; Riemensberger, J.; Weng, W.; Liu, J.; Tian, H.; Siddharth, A.; Snigirev, V.; Shadymov, V.; Voloshin, A.; Wang, R.N. Low-Noise Frequency-Agile Photonic Integrated Lasers for Coherent Ranging. Nat. Commun. 2022, 13, 3522. [Google Scholar] [CrossRef]
  145. Far Brusatori, M.; Duplat, D.N.; Degli-Eredi, I.; Nielsen, L.; Tønning, P.L.; Castera, P.; Volet, N.; Heck, M.J. Ultralow-Linewidth Ring Laser Using Hybrid Integration and Generic Foundry Platforms. Opt. Lett. 2022, 47, 2686–2689. [Google Scholar] [CrossRef]
  146. Snigirev, V.; Riedhauser, A.; Lihachev, G.; Churaev, M.; Riemensberger, J.; Wang, R.N.; Siddharth, A.; Huang, G.; Möhl, C.; Popoff, Y. Ultrafast Tunable Lasers Using Lithium Niobate Integrated Photonics. Nature 2023, 615, 411–417. [Google Scholar] [CrossRef]
  147. Liu, H.; Chen, J.; Chen, X.; Jiang, D.; Huang, J. Fabrication of Distributed Feedback Bragg Fiber Laser on Regular Er-Doped Fiber. Chin. J. Lasers 2006, 33, 873–876. [Google Scholar]
  148. Hill, K.O.; Bilodeau, F.; Malo, B.; Kitagawa, T.; Theriault, S.; Johnson, D.C.; Albert, J.; Takiguchi, K. Chirped In-Fiber Bragg Gratings for Compensation of Optical-Fiber Dispersion. Opt. Lett. 1994, 19, 1314–1316. [Google Scholar] [CrossRef]
  149. Hill, K.O.; Malo, B.; Bilodeau, F.; Johnson, D.C.; Albert, J. Bragg Gratings Fabricated in Monomode Photosensitive Optical Fiber by UV Exposure through a Phase Mask. Appl. Phys. Lett. 1993, 62, 1035–1037. [Google Scholar] [CrossRef]
  150. Song, Z.; Qi, H.; Peng, G.; Li, S.; Guo, J.; Wang, C. Research on Fabrication Technology of Fiber Bragg Grating by Phase Mask Moving. J. Optoelectron. 2014, 25, 1294–1297. [Google Scholar]
  151. Song, Z.; Wang, W.; Qi, H.; Guo, J.; Ni, J.; Wang, C. Packaging Technology of Distributed Feedback Fiber Laser. Acta Photonica Sin. 2016, 45, 0814005. [Google Scholar] [CrossRef]
  152. Vivek, K.; Rajesh, R.; Sreehari, C.V.; Kumar, S.S.; Shajahan, K.; Praveen, T.V.; Santhanakrishnan, T.; Moosad, K.P.B. A New Approach of Large Diameter Polymer-Coated Fiber Laser Hydrophone. J. Light. Technol. 2017, 35, 4097–4104. [Google Scholar] [CrossRef]
  153. Launay, F.-X.; Bouffaron, R.; Lardat, R.; Roux, G.; Doisy, M.; Bergogne, C. Acoustic Antenna Based on Fiber Laser Hydrophones; LopezHiguera, J., Jones, J., LopezAmo, M., Santos, J., Eds.; SPIE: Bellingham, WA, USA, 2014; Volume 9157. [Google Scholar]
  154. Ishii, H.; Kasaya, K.; Oohashi, H. Narrow Spectral Linewidth Operation (<160 kHz) in Widely Tunable Distributed Feedback Laser Array. Electron. Lett. 2010, 46, 714–715. [Google Scholar] [CrossRef]
  155. Kasukawa, A.; Mukaihara, T. High Power, Narrow Linewidth Tunable Lasers. In Proceedings of the 2014 OptoElectronics and Communication Conference and Australian Conference on Optical Fibre Technology, Melbourne, Australia, 6–10 July 2014; IEEE: New York, NY, USA, 2014; pp. 965–966. [Google Scholar]
  156. Panyaev, I.S.; Itrin, P.A.; Korobko, D.A.; Fotiadi, A.A. Sub-100-Hz DFB Laser Injection-Locked to PM Fiber Ring Cavity. Light. Technol. IEEE/OSA J. Light. Technol. 2024, 42, 2928–2937. [Google Scholar] [CrossRef]
  157. Yang, D.; Yan, F.; Feng, T. Narrow-Linewidth and Low RIN Tm/Ho Co-Doped Fiber Laser Based on Self-Injection Locking. Opt. Express 2024, 32, 12172–12180. [Google Scholar] [CrossRef] [PubMed]
  158. Ma, W.; Xiong, B.; Sun, C.; Ke, X.; Hao, Z.; Wang, L.; Wang, J.; Han, Y.; Li, H.; Luo, Y. Laser Frequency Noise Characterization by Self-Heterodyne with Both Long and Short Delay. Appl. Opt. 2019, 58, 3555–3563. [Google Scholar] [CrossRef]
  159. Li, Y.; Fu, Z.; Zhu, L.; Fang, J.; Zhu, H.; Zhong, J.; Xu, P.; Chen, X.; Wang, J.; Zhan, M. Laser Frequency Noise Measurement Using an Envelope-Ratio Method Based on a Delayed Self-Heterodyne Interferometer. Opt. Commun. 2019, 435, 244–250. [Google Scholar] [CrossRef]
  160. Gundavarapu, S.; Brodnik, G.M.; Puckett, M.; Huffman, T.; Bose, D.; Behunin, R.; Wu, J.; Qiu, T.; Pinho, C.; Chauhan, N.; et al. Sub-Hertz Fundamental Linewidth Photonic Integrated Brillouin Laser. Nat. Photonics 2019, 13, 60–67. [Google Scholar] [CrossRef]
  161. Day, T.; Gustafson, E.K. Sub-Hertz Relative Frequency Stabilization of Two-Diode Laser-Pumped Nd:YAG Lasers Locked to a Fabry-Perot Interferometer. IEEE J. Quantum Electron. 1992, 28, 1106–1117. [Google Scholar] [CrossRef]
  162. Sane, S.S.; Bennetts, S.; Debs, J.E.; Kuhn, C.C.N.; McDonald, G.D.; Altin, P.A.; Close, J.D.; Robins, N.P. 11 W Narrow Linewidth Laser Source at 780nm for Laser Cooling and Manipulation of Rubidium. Opt. Express 2012, 20, 8915–8919. [Google Scholar] [CrossRef] [PubMed]
  163. Liu, C.; Qi, Y.; Ding, Y.; Zhou, J.; Dong, J.; Wei, Y.; Lou, Q. All-Fiber, High Power Single-Frequency Linearly Polarized Ytterbium-Doped Fiber Amplifier. Chin. Opt. Lett. 2011, 9, 031402. [Google Scholar] [CrossRef]
  164. Zhu, X.; Liu, J.; Bi, D.; Zhou, J.; Diao, W.; Chen, W. Development of All-Solid Coherent Doppler Wind Lidar. Chin. Opt. Lett. 2012, 10, 012801. [Google Scholar] [CrossRef]
  165. Jiang, M.; Xu, H.; Zhou, P.; Zhao, G.; Gu, X. All-Fiber, Narrow Linewidth and Linearly Polarized Fiber Laser in a Single-Mode-Multimode-Single-Mode Cavity. Appl. Opt. 2016, 55, 6121–6124. [Google Scholar] [CrossRef]
  166. Wright, M.W.; Yao, H.; Marciante, J.R. Resonant Pumping of Er-Doped Fiber Amplifiers for Improved Laser Efficiency in Free-Space Optical Communications. NASA/JPL Interplanet. Netw. (IPN) Prog. Rep. 2012, 42–189, 1–20. [Google Scholar]
  167. Psaltis, D. Coherent Optical Information Systems. Science 2002, 298, 1359–1363. [Google Scholar] [CrossRef]
  168. Ricciardi, I.; De Tommasi, E.; Maddaloni, P.; Mosca, S.; Rocco, A.; Zondy, J.-J.; De Rosa, M.; De Natale, P. A Narrow-Linewidth, Frequency-Stabilized OPO for Sub-Doppler Molecular Spectroscopy around 3 μm. In Proceedings of the Nonlinear Optics and Applications VI, Brussels, Belgium, 16–18 April 2012; SPIE: Bellingham, WA, USA, 2012; Volume 8434, pp. 389–397. [Google Scholar]
  169. Li, C.; Xu, S.; Feng, Z.; Xiao, Y.; Mo, S.; Yang, C.; Zhang, W.; Chen, D.; Yang, Z. The ASE Noise of a Yb3+-Doped Phosphate Fiber Single-Frequency Laser at 1083 nm. Laser Phys. Lett. 2014, 11, 025104. [Google Scholar] [CrossRef]
  170. Li, C.; Xu, S.; Yang, C.; Wei, X.; Yang, Z. Frequency Noise of High-Gain Phosphate Fiber Single-Frequency Laser. Laser Phys. 2013, 23, 045107. [Google Scholar] [CrossRef]
  171. Ronnekleiv, E. Frequency and Intensity Noise of Single Frequency Fiber Bragg Grating Lasers. Opt. Fiber Technol. 2001, 7, 206–235. [Google Scholar] [CrossRef]
  172. Foster, S.; Tikhomirov, A.; Milnes, M. Fundamental Thermal Noise in Distributed Feedback Fiber Lasers. IEEE J. Quantum Electron. 2007, 43, 378–384. [Google Scholar] [CrossRef]
  173. Ma, L.; Hu, Z.; Liang, X.; Meng, Z.; Hu, Y. Relaxation Oscillation in Er3+-Doped and Yb3+/Er3+ Co-Doped Fiber Grating Lasers. Appl. Opt. 2010, 49, 1979–1985. [Google Scholar] [CrossRef]
  174. Gordon, J.P.; Zeiger, H.J.; Townes, C.H. The Maser—New Type of Microwave Amplifier, Frequency Standard, and Spectrometer. Phys. Rev. 1955, 99, 1264–1274. [Google Scholar] [CrossRef]
  175. Henry, C. Theory of the Linewidth of Semiconductor Lasers. IEEE J. Quantum Electron. 1982, 18, 259–264. [Google Scholar] [CrossRef]
  176. Haken, H. Theory of Intensity and Phase Fluctuations of a Homogeneously Broadened Laser. Z. Für Phys. 1966, 190, 327–356. [Google Scholar] [CrossRef]
  177. Lax, M. Quantum Noise v: Phase Noise in a Homogeneously Broadened Maser. Phys. Quantum Electron. 1966, 65, 735. [Google Scholar]
  178. Manes, K.R.; Siegman, A.E. Observation of Quantum Phase Fluctuations in Infrared Gas Lasers. Phys. Rev. A 1971, 4, 373–386. [Google Scholar] [CrossRef]
  179. Zhu, T.; Wei, D.; Shi, L.; Huang, L.; Li, J.; Xu, M. Research Progress in Narrow Linewidth Laser Technology (Invited). Laser Optoelectron. Prog. 2024, 61, 0114003. [Google Scholar] [CrossRef]
  180. Hao, S.; Jian-Wen, H.; Xiang, X.; Tao, L.I. Design of Digital Frequency Multiplier for near Infrared and Visible Light Fourier Transform Spectrometer. Infrared 2014, 35, 29–33. [Google Scholar]
  181. Hongbo, W. Research on In-Orbit Spectral Calibration and Polarization Correction Technology of Visible/near-Infrared Push Broom Imaging Spectrometer. Ph.D. Thesis, Graduate School of Chinese Academy of Sciences (Shanghai Institute of Technical Physics), Shanghai, China, 2016. [Google Scholar]
  182. Yin, Y. Research on High Precision CCD Circuit System of Portable Raman Spectrometer. Master’s Thesis, Zhejiang University, Hangzhou, China, 2014. [Google Scholar]
  183. Zheng, Y.; Han, Z.; Li, Y.; Li, F.; Wang, H.; Zhu, R. 3.1 kW 1050 nm Narrow Linewidth Pumping-Sharing Oscillator-Amplifier with an Optical Signal-to-Noise Ratio of 45.5 dB. Opt. Express 2022, 30, 12670–12683. [Google Scholar] [CrossRef]
  184. Wang, Y.; Sun, Y.; Peng, W.; Feng, Y.; Wang, J.; Ma, Y.; Gao, Q.; Zhu, R.; Tang, C. 3.25 kW All-Fiberized and Polarization-Maintained Yb-Doped Amplifier with a 20 GHz Linewidth and near-Diffraction-Limited Beam Quality. Appl. Opt. 2021, 60, 6331–6336. [Google Scholar] [CrossRef]
  185. Wang, Y.; Feng, Y.; Ma, Y.; Chang, Z.; Peng, W.; Sun, Y.; Gao, Q.; Zhu, R.; Tang, C. 2.5 kW Narrow Linewidth Linearly Polarized All-Fiber MOPA with Cascaded Phase-Modulation to Suppress SBS Induced Self-Pulsing. IEEE Photon. J. 2020, 12, 1502815. [Google Scholar] [CrossRef]
  186. Bradberry, G.W.; Vaughan, J.M. Measurement of Narrow Linewidth with a Fabry-Perot Interferometer of Limited Resolution. Opt. Commun. 1977, 20, 307–310. [Google Scholar] [CrossRef]
  187. Takakura, T.; Iga, K.; Tako, T. Linewidth Measurement of a Single Longitudinal Mode AlGaAs Laser with a Fabry-Perot Interferometer. Jpn. J. Appl. Phys. 1980, 19, L725. [Google Scholar] [CrossRef]
  188. Yamamoto, Y.; Mukai, T.; Saito, S. Quantum Phase Noise and Linewidth of a Semiconductor Laser. Electron. Lett. 1981, 17, 327–329. [Google Scholar] [CrossRef]
  189. Fleming, M.; Mooradian, A. Spectral Characteristics of External-Cavity Controlled Semiconductor Lasers. IEEE J. Quantum Electron. 1981, 17, 44–59. [Google Scholar] [CrossRef]
  190. Goodno, G.D.; Book, L.D.; Rothenberg, J.E. Low-Phase-Noise, Single-Frequency, Single-Mode 608 W Thulium Fiber Amplifier. Opt. Lett. 2009, 34, 1204–1206. [Google Scholar] [CrossRef]
  191. Drever, R.W.; Hall, J.L.; Kowalski, F.V.; Hough, J.; Ford, G.M.; Munley, A.J.; Ward, H. Laser Phase and Frequency Stabilization Using an Optical Resonator. Appl. Phys. B 1983, 31, 97–105. [Google Scholar] [CrossRef]
  192. Shaykin, A.A.; Burdonov, K.F.; Khazanov, E.A. A Novel Technique for Longitudinal Mode Selection in Q-Switched Lasers. Laser Phys. Lett. 2015, 12, 125001. [Google Scholar] [CrossRef]
  193. Yang, X.; Li, Z.; Li, H.; Zhou, J.; Zhong, Y. Investigation on the Entrance Slit Width and Exit Slit Width of Grating Spectrometer on the Measured Spectral Line Width. J. Jiaying Univ. (Nat. Sci.) 2008, 26, 38–41. [Google Scholar] [CrossRef]
  194. Liu, H.; Wang, Q.; Tang, Y.; Yan, X. Spectral Linewidth Measurement Enhancement by Two United Spectrometers. Acta Opt. Sin. 2008, 28, 710–714. [Google Scholar] [CrossRef]
  195. Liu, M.; Yang, M.; Zhu, J.; Zhu, H.; Wang, Y.; Ren, Z.; Zhai, Y.; Zhu, H.; Shan, Y.; Qi, H.; et al. High-Sensitivity Computational Miniaturized Terahertz Spectrometer Using a Plasmonic Filter Array and a Modified Multilayer Residual CNN. Nanophotonics 2023, 12, 4375–4385. [Google Scholar] [CrossRef]
  196. Live Lecture Notes: Principles and Selection Guidelines for Different Types of Spectrometers. Available online: https://mp.weixin.qq.com/s/2oauTfHypSLiZei2rlgidQ (accessed on 20 February 2025).
  197. Yokogawa Test & Measurement Corporation. Available online: https://tmi.yokogawa.com/cn/solutions/products/optical-measuring-instruments/optical-spectrum-analyzer/ (accessed on 1 June 2025).
  198. Huang, W.; Xie, J.; Lu, L.; Chen, X.; Ming, H.; Lu, Y.; Wang, A. Spectrum Method of F-P Etalon Spacing High Precision Measurement. Chin. J. Lasers 2003, 30, 739–742. [Google Scholar]
  199. Wang, Y.; Hua, D.; Mao, J. Effect of Propagation Properties of Gaussian Beam on Fabry-Perot Etalon in Lidar. Chin. J. Lasers 2010, 37, 3013–3018. [Google Scholar] [CrossRef]
  200. Gao, W.; Lu, Z.W.; He, W.M.; Dong, Y.K.; Hasi, W.L.J. Characteristics of Amplified Spectrum of a Weak Frequency-Detuned Signal in a Brillouin Amplifier. Laser Part. Beams 2009, 27, 465–470. [Google Scholar] [CrossRef]
  201. Vorobiev, N.; Glebov, L.; Smirnov, V. Single-Frequency-Mode Q-Switched Nd:YAG and Er:Glass Lasers Controlled by Volume Bragg Gratings. Opt. Express 2008, 16, 9199–9204. [Google Scholar] [CrossRef] [PubMed]
  202. Mejia, E.B.; Starodumov, A.N.; Barmenkov, Y.O. Blue and Infrared Up-Conversion in Tm3+-Doped Fluorozirconate Fiber Pumped at 1.06, 1.117, and 1.18 μm. Appl. Phys. Lett. 1999, 74, 1540–1542. [Google Scholar] [CrossRef]
  203. Gurevich, Y.V.; Stürmer, J.; Schwab, C.; Führer, T.; Lamoreaux, S.K.; Quirrenbach, A.; Walther, T. A Laser Locked Fabry-Perot Etalon with 3 cm/s Stability for Spectrograph Calibration. Int. Soc. Opt. Photonics 2014, 9147, 2414–2430. [Google Scholar] [CrossRef]
  204. Thorlabs. Available online: https://www.thorlabs.com (accessed on 1 June 2025).
  205. Ratkoceri, J.; Batagelj, B. Injection-Locked Range and Linewidth Measurements at Different Seed-Laser Linewidths Using a Fabry-Perot Laser-Diode. Opt. Quantum Electron. 2018, 50, 402. [Google Scholar] [CrossRef]
  206. Hun, X.; Bai, Z.; Chen, B.; Wang, J.; Cui, C.; Qi, Y.; Ding, J.; Wang, Y.; Lu, Z. Fabry-Perot Based Short Pulsed Laser Linewidth Measurement with Enhanced Spectral Resolution. Results Phys. 2022, 37, 105510. [Google Scholar] [CrossRef]
  207. Hercher, M. The Spherical Mirror Fabry-Perot Interferometer. Appl. Opt. 1968, 7, 951. [Google Scholar] [CrossRef]
  208. Stone, J.; Marcuse, D. Ultrahigh Finesse Fiber Fabry-Perot Interferometers. J. Light. Technol. 1986, 4, 382–385. [Google Scholar] [CrossRef]
  209. Hsu, K.; Miller, C.M.; Miller, J.W. Speed-of-Light Effects in High-Resolution Long-Cavity Fiber Fabry–Perot Scanning Interferometers. Opt. Lett. 1993, 18, 235–237. [Google Scholar] [CrossRef]
  210. Xue, J.; Chen, W.; Pan, Y.; Shi, J.; Fang, Y.; Xie, H.; Xie, M.; Sun, L.; Su, B. Pulsed Laser Linewidth Measurement Using Fabry–Pérot Scanning Interferometer. Results Phys. 2016, 6, 698–703. [Google Scholar] [CrossRef]
  211. Hun, X.; Bai, Z.X.; Wang, J.; Chen, B.; Cui, C.; Wang, Y.; Lu, Z. Convolution Error Reduction for a Fabry–Pérot-Based Linewidth Measurement: A Theoretical and Experimental Study. Photonics 2022, 9, 1004. [Google Scholar] [CrossRef]
  212. Liu, W.; Hu, C.; Jiang, P.; Wu, B.; Shen, Y. Ultra-High-Resolution Spectrometry Incorporating Two-Dimension Dispersing Spectrometer and Tunable Fabry-Perot Filter. Acta Opt. Sin. 2015, 35, 063007. [Google Scholar]
  213. Cohen, L. Generalization of the Wiener-Khinchin Theorem. IEEE Signal Process. Lett. 1998, 5, 292–294. [Google Scholar] [CrossRef]
  214. Cui, M.; Huang, J.; Yang, X. Review on Methods for Laser Linewidth Measurement. Laser Optoelectron. Prog. 2021, 58, 0900005. [Google Scholar] [CrossRef]
  215. Bai, Z.; Zhao, Z.; Qi, Y.; Ding, J.; Li, S.; Yan, X.; Wang, Y.; Lu, Z. Narrow-Linewidth Laser Linewidth Measurement Technology. Front. Phys. 2021, 9, 768165. [Google Scholar] [CrossRef]
  216. Mach, L. Ueber Einen Interferenzrefraktor. Z. Für Instrumentenkunde 1982, 12, 89–93. [Google Scholar]
  217. Okoshi, T.; Kikuchi, K.; Nakayama, A. Novel Method for High Resolution Measurement of Laser Output Spectrum. Electron. Lett. 1980, 16, 630–631. [Google Scholar] [CrossRef]
  218. Gallion, P.; Debarge, G. Quantum Phase Noise and Field Correlation in Single Frequency Semiconductor Laser Systems. IEEE J. Quantum Electron. 1984, 20, 343–349. [Google Scholar] [CrossRef]
  219. Saito, S.; Yamamoto, Y. Direct Observation of Lorentzian Lineshape of Semiconductor Laser and Linewidth Reduction with External Grating Feedback. Electron. Lett. 1981, 17, 325–327. [Google Scholar] [CrossRef]
  220. Richter, L.; Mandelberg, H.; Kruger, M.; McGrath, P. Linewidth Determination from Self-Heterodyne Measurements with Subcoherence Delay Times. IEEE J. Quantum Electron. 1986, 22, 2070–2074. [Google Scholar] [CrossRef]
  221. Peng, Y. A Novel Scheme for Hundred-Hertz Linewidth Measurements with the Self-Heterodyne Method. Chin. Phys. Lett. 2013, 30, 084208. [Google Scholar] [CrossRef]
  222. Yue, Y.; Qin, B.; Lv, H.; Ouyang, H. Comparison of Several Ultra-Narrow Laser Linewidth Measurement. Opt. Commun. Technol. 2013, 37, 15–17. [Google Scholar]
  223. Okusaga, O.; Cahill, J.; Docherty, A.; Zhou, W.; Menyuk, C.R. Guided Entropy Mode Rayleigh Scattering in Optical Fibers. Opt. Lett. 2012, 37, 683–685. [Google Scholar] [CrossRef]
  224. Moloney, J.V.; Newell, A.C. Nonlinear Optics. Phys. D Nonlinear Phenom. 1990, 44, 1–37. [Google Scholar] [CrossRef]
  225. Camatel, S.; Ferrero, V. Narrow Linewidth CW Laser Phase Noise Characterization Methods for Coherent Transmission System Applications. J. Light. Technol. 2008, 26, 3048–3055. [Google Scholar] [CrossRef]
  226. Murakami, M.; Saito, S. Evolution of Field Spectrum Due to Fiber-Nonlinearity-Induced Phase Noise in in-Line Optical Amplifier Systems. IEEE Photon. Technol. Lett. 1992, 4, 1269–1272. [Google Scholar] [CrossRef]
  227. Olivero, J.J.; Longbothum, R.L. Empirical Fits to the Voigt Line Width: A Brief Review. J. Quant. Spectrosc. Radiat. Transf. 1977, 17, 233–236. [Google Scholar] [CrossRef]
  228. Young, B.C.; Cruz, F.C.; Itano, W.M.; Bergquist, J.C. Visible Lasers with Subhertz Linewidths. Phys. Rev. Lett. 1999, 82, 3799. [Google Scholar] [CrossRef]
  229. Kessler, T.; Hagemann, C.; Grebing, C.; Legero, T.; Sterr, U.; Riehle, F.; Martin, M.J.; Chen, L.; Ye, J. A Sub-40-mHz-Linewidth Laser Based on a Silicon Single-Crystal Optical Cavity. Nat. Photonics 2012, 6, 687–692. [Google Scholar] [CrossRef]
  230. Lee, H.; Suh, M.-G.; Chen, T.; Li, J.; Diddams, S.A.; Vahala, K.J. Spiral Resonators for On-Chip Laser Frequency Stabilization. Nat. Commun. 2013, 4, 2468. [Google Scholar] [CrossRef]
  231. Kueng, A.; Thevenaz, L.; Robert, P. Telenor Laser Linewidth Determination in the Sub-Megahertz Range Using a Brillouin Fibre Laser. In Proceedings of the European Conference on Optical Communication, Oslo, Norway, 19 September 1996; pp. B305–B308. [Google Scholar]
  232. Dong, Y.-K.; Lu, Z.-W.; Lu, Y.-L.; He, W.-M. A New Method of Measuring Ultra-Narrow Laser Line-Width. Harbin Gongye Daxue Xuebao (J. Harbin Inst. Technol.) 2005, 37, 670–673. [Google Scholar]
  233. Sevillano, P.; Subías, J.; Heras, C.; Pelayo, J.; Villuendas, F. Brillouin Induced Self-Heterodyne Technique for Narrow Line Width Measurement. Opt. Express 2010, 18, 15201–15206. [Google Scholar] [CrossRef]
  234. Roy, A.S.; Kumar K., P. Low Jitter Measurement of Laser Linewidth Using Brillouin Induced Self-Heterodyne Method. In Proceedings of the Frontiers in Optics 2017, Washington, DC, USA, 18–21 September 2017. [Google Scholar] [CrossRef]
  235. Xu, Y.; Xiang, D.; Ou, Z.; Lu, P.; Bao, X. Random Fabry–Perot Resonator-Based Sub-kHz Brillouin Fiber Laser to Improve Spectral Resolution in Linewidth Measurement. Opt. Lett. 2015, 40, 1920–1923. [Google Scholar] [CrossRef]
  236. Boschung, J.; Thévenaz, L.; Robert, P.A. High-Accuracy Measurement of the Linewidth of a Brillouin Fibre Ring Laser. Electron. Lett. 1994, 30, 1488–1489. [Google Scholar] [CrossRef]
  237. Iiyama, K.; Hayashi, K. Delayed Self-Homodyne Method Using Solitary Monomode Fibre for Laser Linewidth Measurements. Electron. Lett. 1989, 25, 1589–1590. [Google Scholar] [CrossRef]
  238. Ludvigsen, H.; Tossavainen, M.; Kaivola, M. Laser Linewidth Measurements Using Self-Homodyne Detection with Short Delay. Opt. Commun. 1998, 155, 180–186. [Google Scholar] [CrossRef]
  239. Kojima, K.; Horiguchi, Y.; Koike-Akino, T.; Shimakura, Y.; Enoki, K.; Yagyu, E. Separation of Semiconductor Laser Intrinsic Linewidth and 1/f Noise Using Multiple Fiber Length for Self-Heterodyne Method. In Proceedings of the Optical Fiber Communication Conference 2015, Los Angeles, CA, USA, 22–26 March 2015. [Google Scholar]
  240. Chen, J.; Chen, C.; Guo, Q.; Sun, J.; Zhang, J.; Zhou, Y.; Liu, Z.; Yu, Y.; Qin, L.; Ning, Y. Linear Polarization and Narrow-Linewidth External-Cavity Semiconductor Laser Based on Birefringent Bragg Grating Optical Feedback. Opt. Laser Technol. 2024, 170, 110211. [Google Scholar] [CrossRef]
  241. Canagasabey, A.; Michie, A.; Canning, J.; Holdsworth, J.; Fleming, S.; Wang, H.-C.; Aslund, M.L. A Comparison of Delayed Self-Heterodyne Interference Measurement of Laser Linewidth Using Mach-Zehnder and Michelson Interferometers. Sensors 2011, 11, 9233–9241. [Google Scholar] [CrossRef]
  242. Chen, X.; Han, M.; Zhu, Y.; Dong, B.; Wang, A. Implementation of a Loss-Compensated Recirculating Delayed Self-Heterodyne Interferometer for Ultranarrow Laser Linewidth Measurement. Appl. Opt. 2006, 45, 7712. [Google Scholar] [CrossRef]
  243. Huang, S.; Zhu, T.; Liu, M.; Huang, W. Precise Measurement of Ultra-Narrow Laser Linewidths Using the Strong Coherent Envelope. Sci. Rep. 2017, 7, 41988. [Google Scholar] [CrossRef]
  244. Huang, S.; Zhu, T.; Cao, Z.; Liu, M.; Deng, M.; Liu, J.; Li, X. Laser Linewidth Measurement Based on Amplitude Difference Comparison of Coherent Envelope. IEEE Photonics Technol. Lett. 2016, 28, 759–762. [Google Scholar] [CrossRef]
  245. Bai, Q.; Yan, M.; Xue, B.; Gao, Y.; Wang, D.; Wang, Y.; Zhang, M.; Zhang, H.; Jin, B. The Influence of Laser Linewidth on the Brillouin Shift Frequency Accuracy of BOTDR. Appl. Sci. 2019, 9, 58. [Google Scholar] [CrossRef]
  246. Wang, Y.; Yang, L.; Liao, Z.; Zhang, P.; Dai, S. Ultra-High Signal-to-Noise Ratio and Ultra-Narrow Linewidth Brillouin Fiber Laser for Linewidth Measurement at 2 μm. J. Light. Technol. 2025, 43, 2759–2763. [Google Scholar] [CrossRef]
  247. Wang, Z.; Ke, C.; Zhong, Y.; Xing, C.; Wang, H.; Yang, K.; Cui, S.; Liu, D. Ultra-Narrow-Linewidth Measurement Utilizing Dual-Parameter Acquisition through a Partially Coherent Light Interference. Opt. Express 2020, 28, 8484–8493. [Google Scholar] [CrossRef]
  248. Xue, M.; Zhao, J. Laser Linewidth Measurement Based on Long and Short Delay Fiber Combination. Opt. Express 2021, 29, 27118–27126. [Google Scholar] [CrossRef] [PubMed]
  249. Gao, J.; Jiao, D.; Liu, J.; Deng, X.; Zang, Q.; Zhang, X.; Wang, D.; Zhang, X.; Liu, T. Laser Linewidth Measurement Based on Recirculating Self-Heterodyne Method with Short Fiber. Acta Opt. Sin. 2021, 41, 0712002. [Google Scholar] [CrossRef]
  250. Gao, J.; Jiao, D.; Deng, X.; Liu, J.; Zhang, L.; Zang, Q.; Zhang, X.; Liu, T.; Zhang, S. A Polarization-Insensitive Recirculating Delayed Self-Heterodyne Method for Sub-Kilohertz Laser Linewidth Measurement. Photonics 2021, 8, 137. [Google Scholar] [CrossRef]
  251. Jia, Y.; Ou, P.; Yang, Y.; Zhang, C. Short Fibre Delayed Self-Heterodyne Interferometer for Ultranarrow Laser Linewidth Measurement. J. Beijing Univ. Aeronaut. Astronaut. 2008, 34, 568. [Google Scholar]
  252. He, Y.; Hu, S.; Liang, S.; Li, Y. High-Precision Narrow Laser Linewidth Measurement Based on Coherent Envelope Demodulation. Opt. Fiber Technol. 2019, 50, 200–205. [Google Scholar] [CrossRef]
  253. Zhao, Z.; Bai, Z.; Jin, D.; Qi, Y.; Ding, J.; Yan, B.; Wang, Y.; Lu, Z.; Mildren, R.P. Narrow Laser-Linewidth Measurement Using Short Delay Self-Heterodyne Interferometry. Opt. Express 2022, 30, 30600–30610. [Google Scholar] [CrossRef]
  254. Bai, Z.; Zhao, Z.; Chen, X.; Qi, Y.; Ding, J.; Yan, B.; Wang, Y.; Lu, Z.; Mildren, R.P. A Lorentzian Narrow-Linewidth Demodulation Scheme Based on a Short Fiber Delayed Self-Heterodyne Technique. Appl. Phys. Express 2022, 15, 106502. [Google Scholar] [CrossRef]
  255. Zhang, C.; Huang, L.; Guan, T.; Mao, Y.; Dang, L.; Lan, T.; Shi, L.; Gao, L.; Zhu, T. Laser Coherence Linewidth Measurement Based on Deterioration of Coherent Envelope. Opt. Laser Technol. 2024, 172, 110498. [Google Scholar] [CrossRef]
  256. Wu, L.; Ji, Z.; Ma, W.; Su, D.; Zhao, Y.; Xiao, L.; Jia, S. Narrow Laser Linewidth Measurement with the Optimal Demodulated Lorentzian Spectrum. Appl. Opt. 2024, 63, 1847–1853. [Google Scholar] [CrossRef] [PubMed]
  257. Xu, D.; Lu, B.; Yang, F.; Chen, D.; Cai, H.; Qu, R. Narrow Linewidth Single-Frequency Laser Noise Measurement Based on a 3 × 3 Fiber Coupler. Chin. J. Lasers 2016, 43, 0102004. [Google Scholar]
  258. Stéphan, G.M.; Tam, T.T.; Blin, S.; Besnard, P.; Têtu, M. Laser Line Shape and Spectral Density of Frequency Noise. Phys. Rev. A—At. Mol. Opt. Phys. 2005, 71, 43809. [Google Scholar] [CrossRef]
  259. Von Bandel, N.; Myara, M.; Sellahi, M.; Souici, T.; Dardaillon, R.; Signoret, P. Time-Dependent Laser Linewidth: Beat-Note Digital Acquisition and Numerical Analysis. Opt. Express 2016, 24, 27961–27978. [Google Scholar] [CrossRef]
  260. Bai, Y.; Yan, F.; Feng, T.; Han, W.; Zhang, L.; Cheng, D.; Bai, Z.; Wen, X. Demonstration of Linewidth Measurement Based on Phase Noise Analysis for a Single Frequency Fiber Laser in the 2 μm Band. Laser Phys. 2019, 29, 075102. [Google Scholar] [CrossRef]
  261. Schilt, S.; Bucalovic, N.; Tombez, L.; Dolgovskiy, V.; Schori, C.; Di Domenico, G.; Zaffalon, M.; Thomann, P. Frequency Discriminators for the Characterization of Narrow-Spectrum Heterodyne Beat Signals: Application to the Measurement of a Sub-Hertz Carrier-Envelope-Offset Beat in an Optical Frequency Comb. Rev. Sci. Instrum. 2011, 82, 123116. [Google Scholar] [CrossRef] [PubMed]
  262. Walls, W.F. Cross-Correlation Phase Noise Measurements. In Proceedings of the Frequency Control Symposium, Hershey, PA, USA, 27–29 May 1992. [Google Scholar]
  263. Di Domenico, G.; Schilt, S.; Thomann, P. Simple Approach to the Relation between Laser Frequency Noise and Laser Line Shape. Appl. Opt. 2010, 49, 4801–4807. [Google Scholar] [CrossRef] [PubMed]
  264. Bucalovic, N.; Dolgovskiy, V.; Schori, C.; Thomann, P.; Domenico, G.D.; Schilt, S. Experimental Validation of a Simple Approximation to Determine the Linewidth of a Laser from Its Frequency Noise Spectrum. Appl. Opt. 2012, 51, 4582–4588. [Google Scholar] [CrossRef] [PubMed]
  265. Luo, X.; Chen, C.; Ning, Y.; Zhang, J.; Chen, J.; Zhang, X.; Li, L.; Wu, H.; Zhou, Y.; Qin, L. Single Polarization, Narrow Linewidth Hybrid Laser Based on Selective Polarization Mode Feedback. Opt. Laser Technol. 2022, 154, 108340. [Google Scholar] [CrossRef]
  266. Bava, E.; Galzerano, G.; Svelto, C. Frequency-Noise Sensitivity and Amplitude-Noise Immunity of Discriminators Based on Fringe-Side Fabry-Perot Cavities. IEEE Trans. Ultrason. Ferroelectr. Freq. Control 2002, 49, 1150–1159. [Google Scholar] [CrossRef]
  267. Sorin, W.V.; Chang, K.W. Frequency Domain Analysis of an Optical FM Discriminator. J. Light. Technol. 1992, 10, 787–793. [Google Scholar] [CrossRef]
  268. Kikuchi, K.; Okoshi, T. Measurement of FM Noise, AM Noise, and Field Spectra of 1.3 μm InGaAsP DFB Lasers and Determination of the Linewidth Enhancement Factor. IEEE J. Quantum Electron. 1985, 21, 1814–1818. [Google Scholar] [CrossRef]
  269. Li, J.; Lee, H.; Chen, T.; Vahala, K.J. Characterization of a High Coherence, Brillouin Microcavity Laser on Silicon. Opt. Express 2012, 20, 20170–20180. [Google Scholar] [CrossRef]
  270. Van Exter, M.P.; Kuppens, S.J.M.; Woerdman, J.P. Excess Phase Noise in Self-Heterodyne Detection. IEEE J. Quantum Electron. 1992, 28, 580–584. [Google Scholar] [CrossRef]
  271. Huynh, T.N.; Smyth, F.; Nguyen, L.; Barry, L.P. Effects of Phase Noise of Monolithic Tunable Laser on Coherent Communication Systems. Opt. Express 2012, 20, B244–B249. [Google Scholar] [CrossRef]
  272. Huynh, T.N.; Nguyen, L.; Barry, L.P. Delayed Self-Heterodyne Phase Noise Measurements with Coherent Phase Modulation Detection. IEEE Photon. Technol. Lett. 2011, 24, 249–251. [Google Scholar] [CrossRef]
  273. Taylor, M.G. Phase Estimation Methods for Optical Coherent Detection Using Digital Signal Processing. J. Light. Technol. 2009, 27, 901–914. [Google Scholar] [CrossRef]
  274. Chauhan, N.; Isichenko, A.; Liu, K.; Wang, J.; Zhao, Q.; Behunin, R.O.; Rakich, P.T.; Jayich, A.M.; Fertig, C.; Hoyt, C.W.; et al. Visible Light Photonic Integrated Brillouin Laser. Nat. Commun. 2021, 12, 4685. [Google Scholar] [CrossRef]
  275. Yang, D.; Li, T.; Yan, F.; Feng, T.; Yu, C.; Qin, Q.; Wang, X.; Guo, H.; Jiang, Y.; Cai, Y.; et al. Demonstration of an Ultra-Wideband Wavelength Self-Adaptive and High-Precision Single-Frequency Laser Linewidth Measurement System. Opt. Express 2024, 32, 40488. [Google Scholar] [CrossRef] [PubMed]
  276. Duthel, T.; Clarici, G.; Fludger, C.R.; Geyer, J.C.; Schulien, C.; Wiese, S. Laser Linewidth Estimation by Means of Coherent Detection. IEEE Photon. Technol. Lett. 2009, 21, 1568–1570. [Google Scholar] [CrossRef]
  277. Sutili, T.; Figueiredo, R.C.; Conforti, E. Laser Linewidth and Phase Noise Evaluation Using Heterodyne Offline Signal Processing. J. Light. Technol. 2016, 34, 4933–4940. [Google Scholar] [CrossRef]
  278. Ravaro, M.; Barbieri, S.; Santarelli, G.; Jagtap, V.; Linfield, E.H. Measurement of the Intrinsic Linewidth of Terahertz Quantum Cascade Lasers Using a Near-Infrared Frequency Comb. Opt. Express 2013, 20, 25654–25661. [Google Scholar] [CrossRef]
  279. Ravaro, M.; Manquest, C.; Sirtori, C.; Barbieri, S.; Santarelli, G.; Blary, K.; Lampin, J.-F.; Khanna, S.P.; Linfield, E.H. Phase-Locking of a 2.5 THz Quantum Cascade Laser to a Frequency Comb Using a GaAs Photomixer. Opt. Lett. 2011, 36, 3969–3971. [Google Scholar] [CrossRef]
  280. Barbieri, S.; Gellie, P.; Santarelli, G.; Ding, L.; Maineult, W.; Sirtori, C.; Colombelli, R.; Beere, H.; Ritchie, D. Phase-Locking of a 2.7-THz Quantum Cascade Laser to a Mode-Locked Erbium-Doped Fibre Laser. Nat. Photonics 2010, 4, 636–640. [Google Scholar] [CrossRef]
  281. Giuliani, G.; Norgia, M. Laser Diode Linewidth Measurement by Means of Self-Mixing Interferometry. IEEE Photonics Technol. Lett. 2000, 2, 1028–1030. [Google Scholar] [CrossRef]
  282. Cardilli, M.C.; Dabbicco, M.; Mezzapesa, F.P.; Scamarcio, G. Linewidth Measurement of Mid Infrared Quantum Cascade Laser by Optical Feedback Interferometry. Appl. Phys. Lett. 2016, 108, 6165–6206. [Google Scholar] [CrossRef]
  283. Tombez, L.; Francesco, J.D.; Schilt, S.; Domenico, G.D.; Hofstetter, D. Frequency Noise of Free-Running 4.6 μm Distributed Feedback Quantum Cascade Lasers near Room Temperature. Opt. Lett. 2011, 36, 3109–3111. [Google Scholar] [CrossRef]
  284. Zhou, Q.; Qin, J.; Xie, W.; Liu, Z.; Tong, Y.; Dong, Y.; Hu, W. Power-Area Method to Precisely Estimate Laser Linewidth from Its Frequency-Noise Spectrum. Appl. Opt. 2015, 54, 8282–8289. [Google Scholar] [CrossRef]
Figure 1. Technological applications of lasers across different linewidth regimes. (Note: linewidth requirements depend on factors such as pulse duration, power level, and application context).
Figure 1. Technological applications of lasers across different linewidth regimes. (Note: linewidth requirements depend on factors such as pulse duration, power level, and application context).
Micromachines 16 00990 g001
Figure 2. Definition of Laser Linewidth Based on FWHM.
Figure 2. Definition of Laser Linewidth Based on FWHM.
Micromachines 16 00990 g002
Figure 3. Overview of laser linewidth measurement techniques across different linewidth regimes.
Figure 3. Overview of laser linewidth measurement techniques across different linewidth regimes.
Micromachines 16 00990 g003
Figure 4. Schematic and components of a grating spectrometer. (a) Working principle of a blazed grating. (b) Incident slit with defined width. (c) Exit slit with defined width. (d) Measured laser spectrum with linewidth determined by FWHM.
Figure 4. Schematic and components of a grating spectrometer. (a) Working principle of a blazed grating. (b) Incident slit with defined width. (c) Exit slit with defined width. (d) Measured laser spectrum with linewidth determined by FWHM.
Micromachines 16 00990 g004
Figure 5. (a) Optical path diagram of the F-P interferometer. (b) Recorded false-color image of an interference pattern generated using the F-P etalon [206]. L: cavity length of F-P interferometry.
Figure 5. (a) Optical path diagram of the F-P interferometer. (b) Recorded false-color image of an interference pattern generated using the F-P etalon [206]. L: cavity length of F-P interferometry.
Micromachines 16 00990 g005
Figure 6. Relationship among beat frequency signals processed by different methods.
Figure 6. Relationship among beat frequency signals processed by different methods.
Micromachines 16 00990 g006
Figure 9. Measurement Ranges of Laser Linewidth Characterization Methods.
Figure 9. Measurement Ranges of Laser Linewidth Characterization Methods.
Micromachines 16 00990 g009
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

Liu, Z.; Zheng, H.; Li, C.; Qi, Z.; Zhang, C.; Li, T.; Bai, Z. Advances in Laser Linewidth Measurement Techniques: A Comprehensive Review. Micromachines 2025, 16, 990. https://doi.org/10.3390/mi16090990

AMA Style

Liu Z, Zheng H, Li C, Qi Z, Zhang C, Li T, Bai Z. Advances in Laser Linewidth Measurement Techniques: A Comprehensive Review. Micromachines. 2025; 16(9):990. https://doi.org/10.3390/mi16090990

Chicago/Turabian Style

Liu, Zhongtian, Hao Zheng, Chunwei Li, Zunhan Qi, Cunwei Zhang, Tie Li, and Zhenxu Bai. 2025. "Advances in Laser Linewidth Measurement Techniques: A Comprehensive Review" Micromachines 16, no. 9: 990. https://doi.org/10.3390/mi16090990

APA Style

Liu, Z., Zheng, H., Li, C., Qi, Z., Zhang, C., Li, T., & Bai, Z. (2025). Advances in Laser Linewidth Measurement Techniques: A Comprehensive Review. Micromachines, 16(9), 990. https://doi.org/10.3390/mi16090990

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

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop