Next Article in Journal
Properties of Diamond-like Coatings in Tribological Systems Lubricated with Ionic Liquid
Previous Article in Journal
RETRACTED: Cui et al. Effects of Cr2O3 Content on Microstructure and Mechanical Properties of Al2O3 Matrix Composites. Coatings 2021, 11, 234
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Photon Detector Technology for Laser Ranging: A Review of Recent Developments

1
College of Information Technology, Jilin Normal University, Siping 136000, China
2
College of Electrical Engineering, Weihai Innovation Research Institute, Qingdao University, Qingdao 266000, China
3
State Grid Siping Electric Power Supply Company, Siping 136000, China
*
Author to whom correspondence should be addressed.
Coatings 2025, 15(7), 798; https://doi.org/10.3390/coatings15070798
Submission received: 26 May 2025 / Revised: 3 July 2025 / Accepted: 7 July 2025 / Published: 8 July 2025

Abstract

Laser ranging technology holds a key position in the military, aerospace, and industrial fields due to its high precision and non-contact measurement characteristics. As a core component, the performance of the photon detector directly determines the ranging accuracy and range. This paper systematically reviews the technological development of photonic detectors for laser ranging, with a focus on analyzing the working principles and performance differences of traditional photodiodes [PN (P-N junction photodiode), PIN (P-intrinsic-N photodiode), and APD (avalanche photodiode)] (such as the high-frequency response characteristics of PIN and the internal gain mechanism of APD), as well as their applications in short- and medium-range scenarios. Additionally, this paper discusses the unique advantages of special structures such as transmitting junction-type and Schottky-type detectors in applications like ultraviolet light detection. This article focuses on photon counting technology, reviewing the technological evolution of photomultiplier tubes (PMTs), single-photon avalanche diodes (SPADs), and superconducting nanowire single-photon detectors (SNSPDs). PMT achieves single-photon detection based on the external photoelectric effect but is limited by volume and anti-interference capability. SPAD achieves sub-decimeter accuracy in 100 km lidars through Geiger mode avalanche doubling, but it faces challenges in dark counting and temperature control. SNSPD, relying on the characteristics of superconducting materials, achieves a detection efficiency of 95% and a dark count rate of less than 1 cps in the 1550 nm band. It has been successfully applied in cutting-edge fields such as 3000 km satellite ranging (with an accuracy of 8 mm) and has broken through the near-infrared bottleneck. This study compares the differences among various detectors in core indicators such as ranging error and spectral response, and looks forward to the future technical paths aimed at improving the resolution of photon numbers and expanding the full-spectrum detection capabilities. It points out that the new generation of detectors represented by SNSPD, through material and process innovations, is promoting laser ranging to leap towards longer distances, higher precision, and wider spectral bands. It has significant application potential in fields such as space debris monitoring.

1. Introduction

A photon detector is one of the core components of a laser ranging system. Since its advent in the 1960s, laser ranging technology has been widely applied in military, aerospace, industrial manufacturing, and other fields, leveraging its unique advantages of high-precision measurement and non-contact operation [1]. The basic principle of laser ranging is to calculate the target distance by measuring the time difference between the laser pulse from transmission to reception; the photon detector, as the core component of the receiving end, is responsible for converting the weak optical signal into an electrical signal, and its performance directly determines the accuracy and range of the ranging system [2]. This paper introduces the working principles of all kinds of detectors and then points out the development directions of detection units in the laser ranging field.
The first photon detector was created in 1917 by T.W. Coase using thallium sulfide (Tl2S) material [3]. In the early days of the development of photon detectors, their military application value was quickly explored. Photon detectors were gradually applied to laser ranging in the late 20th century [4]. At that time, people carried out research on time-dependent photon counting methods, making full use of the correlation characteristics of target photons, and effectively extracting the target echo photons mixed in the noise. This lay the foundation for the application of photon detectors in laser ranging.
Compared with traditional optoelectronic applications such as optical communication and biomedical imaging, the detectors in the field of laser ranging exhibit three unique technical characteristics: Firstly, the extreme requirement for time synchronization accuracy. Laser ranging relies on the measurement of the flight time of nanosecond-level laser pulses (for example, in the case of the Earth–Moon ranging, a distance of 400,000 km corresponds to a round-trip time of approximately 2.5 s, and a detector time jitter of less than 50 ps is required to achieve sub-meter accuracy [5]), which demands that the response delay fluctuation of the detector be controlled within the picosecond range [6], while optical communication detectors focus more on linear responses at GHz-level modulation rates, and the time accuracy requirements of the two differ by 3–4 orders of magnitude. Secondly, the ability to capture single-photon-level weak light signals [7]. In a 3000 km satellite ranging scenario, the laser echo power can be reduced to below 10−15 W (corresponding to less than 10 photons per second), requiring the detector to achieve a single-photon detection efficiency of >90% in the 1550 nm band, while biological fluorescence imaging detectors, although requiring low dark counts, typically operate at microwatt-level light intensities, with the detection efficiency requirement being only 30%–50%. Thirdly, a strong design for resisting background light interference [8]. Laser ranging needs to identify nanosecond-level laser pulses in sunlight (with background light intensity reaching 105 lux), so the detector must have a dynamic range of 108:1 or more. This unique requirement drives the development of laser ranging detectors towards the “high time resolution—single-photon sensitivity—resistance to strong background light” technical direction, and also makes emerging technologies such as SNSPD the core support for long-distance high-precision ranging.
Subsequently, with the continuous development and advancement of technology, the application of photodetectors in the field of laser ranging has been continuously expanded and deepened. For instance, in 2017, the large-aperture superconducting array single-photon detector developed by Nanjing University was applied to space debris detection [9]; in 2019, Nanjing University collaborated with Sun Yat-sen University and other institutions to use a high-time-precision superconducting array single-photon detector for 400,000 km Earth–Moon laser ranging [10]; in 2022, the near-infrared single-photon detector independently developed by the Changchun Satellite Observation Station of the National Astronomical Observatories of the Chinese Academy of Sciences successfully achieved satellite laser ranging [11].
As the detection unit of the laser ranging system, photodetectors operate based on the photoelectric effect. The photoelectric effect can be either internal photoelectric effect or external photoelectric effect. In the internal photoelectric effect, the electrons in the material absorb the energy of photons and transition from a lower energy level to a higher energy level, thereby generating photogenerated charge carriers (electrons and holes), changing the electrical properties of the material. This effect does not involve electrons escaping from the material surface but causes changes within the material, mainly including the photoconductive effect and photoelectric effect. The external photoelectric effect occurs when electrons on the surface of a metal or semiconductor material absorb the energy of photons and escape from the surface. Based on the internal photoelectric effect, there are photovoltaic detectors (photodiodes), avalanche diodes, etc., while those based on the external photoelectric effect include phototubes and photomultiplier tubes [12]. As shown in Figure 1, when the P-N junction structure is excited by light, electrons absorb photons and gain energy exceeding the band gap width. They then transition from the valence band across the band gap to the conduction band, generating a photogenerated electron–hole pair, which is a key physical process for carrier generation in photoelectric detection.
Due to different ranging targets, ranging methods, and ranging accuracy, detectors with different working modes should be selected. In this paper, various kinds of photodetectors are summarized and analyzed, including traditional laser ranging technology and detection systems, such as photodiodes, PIN photodiodes, and APD avalanche diodes. The paper also discusses the emerging photon counting laser ranging technology and its related single-photon detectors, including photomultiplier tubes, SPADs (single-photon avalanche diodes), and superconducting single-photon detectors.

2. Photodiodes

A photodiode is a kind of semiconductor photoelectric device based on the internal photoelectric effect. As shown in Figure 2, it is usually composed of a PN junction, and the P and N regions are, respectively, different types of semiconductor materials. A depletion layer is formed between the P and N regions, and this region is very sensitive to the response of light. Photodiodes have different packaging forms (common patch type, in-line type, etc.), to adapt to different application scenarios. Common types of photodiodes are presented in Table 1.

2.1. PIN-Type Photodiodes

A PIN photodiode is a semiconductor photoelectric device based on internal photoelectric effect [13]. It is composed of a P-type semiconductor, an intrinsic semiconductor (I layer), and an N-type semiconductor. Among them, the intrinsic semiconductor layer is located between the p-type and n-type semiconductors, and the thickness is usually between a few microns and tens of microns. There are electrodes in the P and N zones, respectively, which are used to connect the external circuit [14]. Planar structure or mesa structure is usually used for packaging to improve the stability and reliability of the device.
When light strikes the PIN photodiode, the energy of the photon is absorbed by the semiconductor material. If the energy of the photon is greater than the bandgap energy of the semiconductor, it will excite the electrons in the valence band to transition to the conduction band, producing electron–hole pairs. Under the action of the built-in electric field of the PN junction, the electrons drift to the N region and the holes drift to the P region, thus forming a photogenerated current. Due to the existence of the intrinsic semiconductor layer, the width of the depletion layer is increased, and the response speed and linearity of the device are improved. Therefore, it has the advantages of fast response speed, high sensitivity, good linearity, small size, light weight, and low power consumption.
In 2016, Yu Zhou et al. designed and tested a laser rangefinder based on the pulse method [15]. (The principles of the pulse method and the phase method are shown in Figure 3 and Figure 4, respectively.) In this study, a PIN diode is selected as the receiving end of the system, and its ranging limit range is 10 m. Several groups of close-range measurement errors are tested and analyzed. In the same year, Ye Daohuan et al. designed a high-precision laser dynamic testing system based on the working characteristics and technical index requirements of the Chang’e-5 laser rangefinder [15]. Combined with the characteristics of laser pulse signal and the technical requirements of the system, a PIN photodiode becomes the first choice. The experimental results show that the dynamic system meets the technical requirements of measurement accuracy better than 0.15 m and minimum simulation distance less than 15 m, which provides a dynamic measurement method for a pulse ground object rangefinder [16]. In 2017, in order to reduce the cost of the existing laser cloud altimeter, Lv Xueju designed a portable laser cloud altimeter system by using a low-power laser and a PIN diode according to the measurement principles of the pulse method [17]. At the same time, through theoretical analysis and the mathematical modeling of the reflected signal of the photodetector, the sensitivity problem faced by the PIN photodiode in the application scenario of this portable laser cloud height measuring system was solved.
PIN photodiodes are more suitable for short-distance high-precision ranging scenarios, such as small-part size measurement, object position detection, and low-power laser ranging systems. This is because their structure consists of a P-type semiconductor, an intrinsic layer, and an N-type semiconductor. The existence of the intrinsic layer not only widens the depletion layer width, enabling it to have good linear response and anti-noise ability in short-distance scenarios, but also results in the device not having an internal gain mechanism like APD. Due to the lack of internal gain, when applied to medium-to-long-distance laser ranging, the detection ability of PIN photodiodes for weak light signals significantly decreases—the photogenerated carriers generated by incident light cannot achieve current amplification through the internal structure of the device, resulting in low sensitivity; at the same time, although the intrinsic layer improves the response linearity in short-distance scenarios, it is relatively slower in charge separation when facing weak signals attenuated in medium-to-long-distance transmission, resulting in a slower response speed. Moreover, the lack of gain characteristics makes the device more susceptible to environmental noise interference in medium-to-long-distance scenarios, because the signal-to-noise ratio of weak effective signals is lower compared to noise, further limiting its applicability in medium-to-long-distance ranging.

2.2. Avalanche Diodes

Photodiodes mainly rely on the motion of photogenerated carriers under an electric field to form photocurrent, and there is no internal signal amplification mechanism. However, the avalanche diode adds a high electric field region to the structure. When the photogenerated charge carriers enter the region, they will accelerate under the action of the electric field [12] and generate more charge carriers through impact ionization, so as to realize the gain amplification of the signal. This internal gain mechanism enables the avalanche diode to detect much weaker light signals, and the detection ability of weak light signals is much higher than that of photodiodes. For example, in optical fiber communication, the optical signal becomes very weak after long-distance transmission, and the avalanche diode can better receive and detect such a signal.
In 2000, Pellegrini et al. carried out an experiment with Si-APD, using a Q-switched laser with a repetition rate of 25 MHz (wavelength: 850 nm; pulse width: 10 ps; pulse energy: 10 pJ) to measure the range of non-cooperative targets, and successfully achieved a detection range of 50 m [18]. And a depth resolution of 3 mm was obtained with 10 returned photons [15]. In 2014, Liu Haidi et al. studied and adjusted the phase method laser ranging technology when they studied high-precision short-range measurement [19]. In order to meet the experimental requirements, the Harbin Institute of Technology research team designed a photodetector based on APD by itself, and compared it with the ordinary PIN photodiode. The detector has outstanding advantages in terms of cut-off frequency size and response time. In 2016, Degnan et al. carried out related research work and successfully developed a single-photon laser ranging system. The system uses a pulse laser with a frequency of 60 kHz and a pulse width of sub-nanosecond level [20]. Under the airborne platform, the Gm-APD array is used to receive the reflected echo of the target, so as to realize the measurement accuracy of sub-decimeter level. In 2017, Peng Wei et al. used phase-difference infrared ranging and counting in close-range laser ranging, and selected APD as a photodetector [21]. Phase-based laser ranging calculates the target range by measuring the phase difference between the emitted laser and the reflected laser. As shown in Figure 4, the transmitter sends out the modulated laser signal, which is reflected by the target, and the reflected signal is received by the receiver. By comparing the phase difference between the transmitted signal and the received signal, the round-trip time of the laser can be calculated, and then the target distance can be obtained. The accuracy of the phase method depends on the modulation frequency and the sensitivity of the phase detection, and it is generally suitable for high-precision ranging at short and medium distances.
The InGaAs/InP avalanche photodiode is a kind of high-sensitivity detector based on III-V semiconductor materials, which is suitable for optical signal detection in the near-infrared band. As shown in Figure 5, its structure consists of multiple layers of semiconductor materials, including an InP substrate, an InGaAs absorption layer, and an InP doubling layer. The InGaAs layer is responsible for absorbing photons and generating electron–hole pairs, while the InP layer amplifies the signal by achieving avalanche multiplication of charge carriers through regions of high electric field. This structural design enables InGaAs/InP APDs to have high quantum efficiency and low noise characteristics in the near-infrared band, which is widely used in the field of fiber communication and laser ranging.
APD is widely used in short-range precision ranging and medium- and near-range laser ranging because of its internal gain structure, and its performance is better than that of the PIN photodiode. However, the avalanche diode operating in linear mode has low gain, high noise, and is greatly affected by temperature. Therefore, complex peripheral control circuits and temperature compensation circuits are required to ensure the sensitivity of APD, which increases the cost and the volume of the laser rangefinder. With the continuous improvement of distance measurement requirements, research on new ranging methods and emerging single-photon detectors has seen an upsurge.

3. Single-Photon Detectors

SPD plays an important role in the fields of quantum communication, space detection, and national defense construction [22,23]. SPD is a highly sensitive photodetector capable of detecting single photons [24]. It is usually based on specific physical effects, such as the photoelectric effect and the avalanche multiplication effect, to convert the energy of a single photon into a measurable electrical signal. A single-photon detector is mainly composed of a photosensitive material, an electron multiplication structure, and a signal processing circuit. Photosensitive materials absorb photons and create electron–hole pairs, and electron-multiplying structures amplify these initial electron–hole pairs, producing an electrical signal strong enough to be detected by signal processing circuits.
Single-photon detectors have extremely high sensitivity and are able to detect extremely weak optical signals or even the presence of single photons. The response speed is fast, and it is able to react to the arrival of photons in a short time. This is important for real-time monitoring and high-speed data acquisition. Being sensitive to only a single photon, the noise generated when no photons are incident is very low. This helps to improve the signal-to-noise ratio and obtain more accurate measurement results. It can detect different wavelengths of light and has a wide spectral response range.
Laser ranging mainly includes time of flight (TOF), interferometry, and triangulation [25]. In recent years, with the development of single-photon detectors, laser ranging has derived a new ranging method, namely, the single-photon laser ranging method. In the single-photon laser ranging system, the photon detector can respond to the photon, so as to achieve longer distance measurement [26]. In laser ranging, especially long-distance ranging, the reflected light signal is very weak [27]. The high sensitivity of single-photon detectors can effectively detect these weak signals and improve the accuracy and range of ranging [28]. For example, in the LIDAR system of driverless cars, single-photon detectors can help the vehicle to perceive the distance information of the surrounding environment more accurately.

3.1. Photomultiplier Single-Photon Detectors

PMT is a kind of vacuum electronic device based on the external photoelectric effect and the secondary electron emission effect, which is often used as a single-photon detector [29]. The photomultiplier tube is mainly composed of a photocathode, an electron multiplication system, and an anode [29] (Table 2). When photons are incident on the photocathode, electrons are released from the photocathode material [30]. These electrons are accelerated by the electric field and hit the first multiplication pole in the electron multiplication system, generating more secondary electrons. This process is repeated in multiple multiplication poles, allowing the number of electrons to increase exponentially. Eventually, a large number of electrons reach the anode, forming a measurable electrical pulse signal.
PMT has the advantages of high sensitivity, fast response speed, wide dynamic range, and good stability. It has a high detection efficiency for a single photon and can detect extremely weak light signals [31]. The response time is short and can respond to photons in nanoseconds, making it suitable for high-speed measurement and time-resolved applications. It can measure from a single photon to a relatively strong light signal with a wide dynamic range [32]. Under proper working conditions, the performance is stable and reliable, and is not easily affected by environmental factors.
In the early years, photomultiplier tubes were widely used in Earth satellite ranging. In 2011, Wang Hanji et al. from the Harbin Institute of Technology used photomultiplier tubes to complete the experiment of laser ranging based on time-dependent single-photon counting technology [33]. However, the current laser rangefinder is developing in the direction of miniaturization, lightweight, and integration. The photomultiplier single-photon detector suffers from inherent drawbacks—bulky volume, poor anti-interference capability (prone to magnetic field interference), complex internal vacuum chamber structure requiring manual fabrication, high cost, narrow dynamic response range, and the need for multiple sets of operating voltages—rendering it incompatible with new-generation laser rangefinders. Currently, the development of new lasers focuses on the realization of large range under the premise of ensuring as small a size as possible, so the solid-state single-photon detector has found more and more applications.

3.2. Single-Photon Avalanche Diodes

SPAD is a kind of highly sensitive light detector, capable of detecting the presence of a single photon [34,35]. When a photon is incident on a photosensitive region of the SPAD, the photon is absorbed and creates an electron–hole pair [36]. Under the action of the reverse bias voltage, this initial carrier causes the avalanche multiplication effect in the high electric field region. As the avalanche process progresses, the number of carriers increases exponentially, producing a large detectable current pulse. By detecting this current pulse, it is possible to determine whether a single photon has been detected.
SPAD has the advantages of extremely high sensitivity, fast response, low noise, and easy miniaturization and integration, as well as the ability to detect extremely weak light signals, even the presence of a single photon [37,38]. This gives it a unique advantage in areas that require highly sensitive detection, such as quantum communication, high-altitude detection, and biomedical imaging. The response speed is extremely fast, and it can respond to photons at the picosecond level. This is important for applications that require real-time measurement and high-speed data acquisition. The dark count rate of SPAD is a key parameter reflecting its intrinsic noise performance, which is typically lower under optimized operating conditions (e.g., reduced temperature or bias voltage), leading to minimal noise generation in the absence of incident photons. This helps to improve the signal-to-noise ratio of the signal and obtain more accurate measurement results. SPAD can be manufactured using a semiconductor process and is small, lightweight, and easy to integrate into a variety of optical systems.
Figure 6 shows the typical structure of an avalanche diode. The left figure represents an unprotected ring structure. An avalanche diode is a semiconductor optoelectronic device based on the internal photoelectric effect. Its structure design endows it with unique photoelectric properties and signal amplification capabilities. Its structure composition includes a semiconductor material substrate, a PN junction, and a high electric field region. The semiconductor material substrate serves as the main support of the avalanche diode, providing the basic physical environment for the various performance of the device; the PN junction is the core part of the avalanche diode, formed by the combination of a P-type semiconductor and an N-type semiconductor. At the PN junction, due to the difference in carrier concentration, an intrinsic electric field is formed. The high electric field region is formed near the PN junction through specific structural designs (such as the distribution of doping concentration). When photogenerated carriers (electron–hole pairs) enter this region, they will accelerate their movement under the strong electric field effect.
Single-photon detectors are widely used in long-distance laser ranging. In 2009, McCarthy et al. carried out a specific research effort in which they added a scanning time-of-flight system to the detection system. In this detection, Si-SPAD (with an active area diameter of 180 μm and jitter of about 400 ps) was used, and a time-dependent single-photon counting technique was used to detect the operation [39]. Through the above settings and application of technology, the detection of low-feature targets beyond 325 m was successfully performed in a daytime environment, and the detection effect of centimeter-level depth resolution was achieved. In 2016, Luo Yuan et al. used an avalanche diode single-photon detector of Geiger mode in the 100 km long-range LIDAR ranging system, mainly for the measurement of small targets at a long distance [40]. In 2018, DuB, PangC et al. demonstrated a high-speed long-distance photon counting and ranging system, and completed a 165 km measurement, in which APD was selected for detection [41].
A single-photon avalanche diode has the advantages of high quantum efficiency, low power consumption, large operating spectrum range, small size, and low operating voltage [42]. However, because it works in Geiger mode, there is a dead time after each signal acquisition, and it cannot terminate the avalanche itself, so a peripheral quenching circuit is required to suppress the avalanche. At the same time, SPAD works near the breakdown voltage; at this time, the dark current and other noise is very large, and the breakdown voltage is extremely sensitive to temperature, so a temperature control circuit is needed to maintain a constant temperature [41]. Due to the limitation of intrinsic physical properties of the material, it has a high dark count and a relatively low detection rate, which reduces the detection probability of effective echo photons. Therefore, SPAD, a common detector with single-photon sensitivity, has been unable to meet the application requirements of single-photon detection in long-distance near-infrared bands. In this context, the emerging superconducting nanowire single-photon detector has become an important technology for the detection unit of long-distance near-infrared laser ranging [43].

3.3. Superconducting Nanowire Single-Photon Detectors

In the field of single-photon detection, the development of technology has always faced continuous challenges related to material limitations and performance trade-offs. Early Si-APD and PMT were limited by their spectral response windows, being mainly capable of effective single-photon detection within the visible light band (400–760 nanometers). This limitation stemmed from the bandgap energy of silicon, which limited its ability to absorb photons longer than 1100 nanometers. For example, Si-APD operating in the Geiger mode had a dark count rate exceeding 100 cps in the near-infrared (NIR) region, while PMT had inherent limitations such as a bulky vacuum tube structure and susceptibility to magnetic field interference, making it unsuitable for compact or high-mobility applications [44].
The subsequent development of InGaAs/InP SPD met the demand for near-infrared (1000–1650 nanometers) detection, but introduced new performance differences. Although these devices utilized the smaller bandgap of InGaAs to expand the spectral response, the detection efficiency in the 1550 nanometer wavelength band was typically lower than 50%, far below the 70%–80% efficiency achievable in the visible light range. Additionally, InGaAs/InP SPD was affected by temperature-dependent breakdown voltage, requiring complex thermoelectric cooling systems to maintain stability, which reduced its feasibility in practical applications.
One of the key limitations faced by these detectors is the “dead time” phenomenon, which occurs during the recovery phase after a single-photon event. In the Geiger-mode photomultiplier tube, the avalanche multiplication process triggered by the photon cannot terminate on its own; instead, it requires an external suppression circuit to suppress the current. This suppression and subsequent charging of the detector capacitance form a time window ranging from tens of nanoseconds to several hundred nanoseconds, during which the device cannot respond to subsequent photons. For example, a standard Si-APD may have a dead time of approximately 50 nanoseconds, which limits its maximum count rate to about 20 MHz and may lead to missed detections in high-photon-flux conditions.
The emergence of superconducting single-photon detector technology marked a significant technological breakthrough, but its development was hindered by operational conditions. Early SNSPD performed poorly in low count rates (below 10 MHz), large time jitter (over 100 picoseconds), and strict requirements for low-temperature environments (below 4 K), typically requiring liquid helium cooling. These factors significantly hindered its widespread adoption in practical applications [45].
In this context, SNSPD emerged as a transformative solution, leveraging the quantum properties of superconducting materials. These detectors are composed of niobium nitride (NbN) nanowires, which maintain a temperature below the critical temperature (typically 2.1–4 Kelvin), taking advantage of the zero-resistance state of superconductors. When absorbing a single photon, the local superconducting state of the nanowire is disrupted, breaking Cooper pairs and generating normal state electron–hole pairs, which leads to a transient resistance change. This detectable resistance fluctuation within picoseconds enables ultrafast response times and unprecedented performance metrics, such as a detection efficiency of 95% at 1550 nanometer wavelength, a dark count rate of less than 1 per second, and a time jitter of less than 50 picoseconds, making superconducting nanodevices the benchmark for advanced single-photon detection in challenging environments such as satellite laser ranging and deep space communication [46].
Compared with other types of detectors, the core differences in SNSPD lie in the following aspects: Firstly, its detection efficiency is much higher than that of traditional detectors. In the 1550 nm band, it can reach over 95%, while the detection efficiency of SPAD in the near-infrared band is usually lower than 50%. Secondly, the dark count rate is extremely low, less than 1 cps (counts per second), which is much lower than that of APD and SPAD, which range from tens to hundreds of cps. Thirdly, its spectral response range covers from visible light to mid-infrared, breaking through the bottleneck of Si-based detectors in the near-infrared. Fourthly, the time jitter is extremely small (less than 50 ps), which is better than the nanosecond-level jitter of PMT. In addition, SNSPD needs to work in an extremely low-temperature environment (requiring cooling to 2.1 K–4 K), while PMT, SPAD, etc., can operate at room temperature or under simple temperature control. This characteristic poses challenges for the cooling system when it comes to applications in scenarios with strict requirements for volume and power consumption, such as aerospace and quantum communication. However, precisely because of this, its performance demonstrates an irreplaceable advantage in long-distance laser ranging (such as 3000 km satellite ranging, with an accuracy of 8 mm).
SNSPD exhibits extremely high detection efficiency, approaching or even reaching 100%, which provides a significant advantage in applications requiring high-precision detection. This makes it a great advantage in applications where high-precision detection is required. In the absence of incident photons, SNSPD has a very low dark count rate. This helps improve the signal-to-noise ratio of the signal and obtain more accurate measurement results. The response speed is fast, and it can respond to photons at the picosecond level. This is important for applications that require real-time measurement and high-speed data acquisition. Moreover, SNSPD has a wide spectrum response range, which can detect photons in a wide spectrum range from visible light to infrared light. Due to the properties of superconducting materials, the noise generated by SNSPD is very low, helping to improve the sensitivity and accuracy of detection.
Table 3 presents the performance parameters of superconducting nanowire single-photon detectors (SNSPDs) developed by different institutions. The meanings and performance of each parameter are as follows: Wavelength (Wavelength/nm) refers to the wavelength of the light signal that the detector operates on, with 1550 nm and 1064 nm being commonly used wavelengths in laser ranging, covering the near-infrared spectrum. Detection efficiency (Detection efficiency/%) reflects the detector’s ability to capture photons. For example, the SNSPD developed by SIMIT achieves 95% in the 1550 nm band, approaching the theoretical limit, meaning that very few photons will be missed. Dark count rate (Dark count rate/cps) characterizes the frequency of false triggering when no photons are incident. The dark count rates of NIST and USTC devices are lower than 1 cps, indicating extremely low noise levels and suitability for weak signal detection. Repetition rate (Repetition rate/MHz) represents the maximum number of counts that the detector can process per second. The JPL device has a repetition rate of 500 MHz, suitable for high-speed photon flow scenarios. Time jitter (Time jitter/ps) measures the stability of the response time. The MIT device has a jitter of 30 ps, enabling sub-millimeter-level ranging accuracy. Temperature (Temperature/K) and refrigeration method (Refrigeration method) indicate the requirements of the working environment. For example, NJU uses a 1.8 K hybrid refrigeration system, and a low-temperature environment is crucial for maintaining superconducting properties.
The abbreviations in the table correspond to the following institutions: NIST (National Institute of Standards and Technology of the United States), MIT (Massachusetts Institute of Technology), NICT (Japan Institute of Information and Communications Technology), SIMIT (Shanghai Institute of Microsystem and Information Technology of the Chinese Academy of Sciences), USTC (University of Science and Technology of China), NJU (Nanjing University), and JPL (Jet Propulsion Laboratory of the United States). The research conducted by these institutions represents the international cutting-edge level of SNSPD technology.
The superconducting nanowire single-photon detector has the characteristics of high sensitivity, high time accuracy, fast detection speed, and low dark count [52]. The energy gap in superconductors is generally three orders of magnitude smaller than that in semiconductor detectors, and is more suitable for near-infrared light detection from the perspective of quantum efficiency. It has important application prospects in laser ranging and other fields. In 2009, Lamb et al. [53] from Heriot-Watt University in the UK carried out relevant research and practice. They used SNSPD (which at the time had a jitter of about 70 ps) to detect ground-to-ground targets at a wavelength of 1550 nm. Finally, a ground-to-ground resolution of 1 cm and a depth resolution of 4 mm were successfully achieved at a distance of 330 m [53]. In 2016, Zhang Sen et al. applied it to ultra-long-distance ground object ranging, completed a 126 km field ranging experiment with a visibility of 80 km, and calculated that the range limit was 280 km [54]. In the same year, LiH, ChenS, YouL, and others designed and completed the satellite laser ranging based on SNSPD, which was about 3000 km away from the ground station of Sheshan Observatory, with an accuracy of 8 mm [55]. In 2017, Zhujiang et al. used a superconducting nanowire single-photon detector to detect fog echo signals ranging from 42.3 km to 63.5 km [56]. In 2020, Lixing You from Yunnan Observatory of the Chinese Academy of Sciences carried out laser ranging experiments. In this process, array detection technologies such as array superconducting nanowire single-photon detector and multi-channel event timekeeper were applied to achieve the accurate detection of small targets with an orbital altitude of 1000 km and a radar cross-section area of 0.045 m2 [57].
According to the experimental results of the above team, SNSPD has excellent performance in detection efficiency, dark count, detection speed, time jitter, response band, etc. SNSPD is the single-photon detector with the best comprehensive performance at present, and is widely used in the field of long-distance laser ranging [58]. However, the production of SNSPD is difficult, and the entire nanowire width must be kept consistent in the production of nanowire; otherwise, the unique sensitivity will be lost. Compared with traditional small-size SNSPD devices, the complexity in the fabrication process of large-array and large-format SNSPD (Superconducting Nanowire Single-Photon Detector) devices increases significantly., which leads to the decrease in device performance consistency and yield, restricting their further development and application. If this problem can be solved, the performance of laser ranging systems is expected to be further improved. Despite the difficulties in preparation, SNSPD has achieved a breakthrough in all the key parameters of the current single-photon detector, and as a revolutionary single-photon detection technology, it will be the future development direction of the laser ranging detection field.

4. Summary and Prospects

4.1. Summary

Photon detectors are core to laser ranging systems, with their performance directly determining ranging accuracy and range. This paper reviews traditional photodiodes (PN, PIN, and APD) and emerging single-photon detectors (PMT, SPAD, and SNSPD). PN photodiodes have low dark current but slow response, suitable for short-range applications like color sensing. PIN photodiodes, with an intrinsic layer, offer fast response and high sensitivity, achieving sub-meter accuracy within 10 m. APDs use avalanche gain for medium-range precision, such as millimeter-level accuracy at 50 m, but require complex temperature control.
For single-photon detection, PMTs achieve sensitivity via the external photoelectric effect but are limited by size and magnetic interference. SPADs in Geiger mode enable sub-decimeter accuracy at 100 km but face dark counting and temperature stability issues. SNSPDs, based on superconducting materials, show exceptional performance, namely, 95% detection efficiency at 1550 nm, dark count rates below 1 cps, and 8 mm accuracy in 3000 km satellite ranging, breaking through the near-infrared bottleneck.

4.2. Future Prospects

Building on current advancements, the future development of photon detectors for laser ranging will focus on several interrelated directions.
Future research will prioritize material innovation to expand spectral coverage. NbN-based nanowire alloys like NbN-TaN aim to extend detection to the mid-infrared (3–5 μm) while maintaining ultra-low dark counts, supporting applications such as atmospheric CO2 laser ranging. Hybrid structures combining perovskite semiconductors with superconducting nanowires may enable seamless detection from visible to infrared wavelengths, crucial for multi-band environmental monitoring.
System miniaturization is another key frontier. Integrating micro-pulse tube refrigeration technologies could scale down cryogenic systems to fit within 3U CubeSats, enabling global space debris tracking with 8 mm accuracy at 10,000 km ranges. Exploring topological insulators like Bi2Se3 for room-temperature single-photon detection may revolutionize tactical laser rangefinders, eliminating the need for bulky cooling systems.
Advanced signal processing will enhance detection performance. Deep neural networks can suppress thermal noise to reduce effective dark counts below 0.1 cps, while quantum tomography algorithms may enable photon number resolution—extracting amplitude information from SNSPD signals to improve atmospheric composition analysis. Combining SNSPDs with entangled photons could achieve sub-millimeter accuracy in interstellar laser ranging for deep-space probe navigation.
Cross-disciplinary applications will also expand. SNSPD’s high efficiency makes it ideal for quantum radar systems, overcoming classical limits in low-signal scenarios like stealth aircraft detection. Adaptation for single-photon fluorescence imaging could enable sub-cellular resolution in live tissue with picosecond time gating. These advancements will drive laser ranging toward longer distances, higher precision, and broader spectral coverage, unlocking new possibilities in space exploration, autonomous transportation, and quantum science.

Author Contributions

Conceptualization, Z.L. and X.J.; methodology, X.J.; validation, C.Y., X.J., and K.W.; formal analysis, K.W.; investigation, C.Y.; resources, X.J.; writing—original draft preparation, Z.L.; writing—review and editing, K.W.; supervision, K.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Youth Innovation Technology Project of Higher School in Shandong Province (No. 2022KJ139).

Conflicts of Interest

Author Changfu Yuan was employed by the company State Grid Siping Electric Power Supply Company. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Lloyd, D.R. What was measured in Millikan’s study of the photoelectric effect? Am. J. Phys. 2015, 83, 765–772. [Google Scholar] [CrossRef]
  2. Liu, Y. The application and optimization of phase method in laser distance measurement. In Journal of Physics: Conference Series; IOP Publishing: Bristol, UK, 2021; p. 022002. [Google Scholar]
  3. Rogalski, A. History of infrared detectors. Opto-Electron. Rev. 2012, 20, 279–308. [Google Scholar] [CrossRef]
  4. Kim, I.; Martins, R.J.; Jang, J.; Badloe, T.; Khadir, S.; Jung, H.-Y.; Kim, H.; Kim, J.; Genevet, P.; Rho, J. Nanophotonics for light detection and ranging technology. Nat. Nanotechnol. 2021, 16, 508–524. [Google Scholar] [CrossRef]
  5. Yin, Y.; Sun, J.; Huang, L.; Jiang, P.; Wang, X.; Ding, C. Moon imaging performance of FAST radio telescope in bistatic configuration with other radars. Remote Sens. 2023, 15, 4045. [Google Scholar] [CrossRef]
  6. Adams, B.W.; Elagin, A.; Frisch, H.; Obaid, R.; Oberla, E.; Vostrikov, A.; Wagner, R.G.; Wang, J.; Wetstein, M. Timing characteristics of large area picosecond photodetectors. In Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment; Elsevier: Amsterdam, The Netherlands, 2015; Volume 795, pp. 1–11. [Google Scholar]
  7. Zhang, W.; You, L.; Li, H.; Huang, J.; Lv, C.; Zhang, L.; Liu, X.; Wu, J.; Wang, Z.; Xie, X. NbN superconducting nanowire single photon detector with efficiency over 90% at 1550 nm wavelength operational at compact cryocooler temperature. Sci. China Phys. Mech. Astron. 2017, 60, 120314. [Google Scholar] [CrossRef]
  8. Van Wolferen, H.A.; Abelmann, L. Laser interference lithography. In Lithography: Principles, Processes and Materials; Nova Publishers: Hauppauge, NY, USA, 2011; pp. 133–148. [Google Scholar]
  9. Zhang, H.; Long, M.; Deng, H.; Cheng, S.; Wu, Z.; Zhang, Z.; Zhang, A.; Sun, J. Developments of space debris laser ranging technology including the applications of picosecond lasers. Appl. Sci. 2021, 11, 10080. [Google Scholar] [CrossRef]
  10. Müller, J.; Murphy, T.W.; Schreiber, U.; Shelus, P.J.; Torre, J.-M.; Williams, J.G.; Boggs, D.H.; Bouquillon, S.; Bourgoin, A.; Hofmann, F. Lunar Laser Ranging: A tool for general relativity, lunar geophysics and Earth science. J. Geod. 2019, 93, 2195–2210. [Google Scholar] [CrossRef]
  11. Leach, J.; Fleming, F.; Paul, D.J.; Tan, C.H.; Ng, J.S.; Henderson, R.K.; Buller, G.S. Single-photon detection for long-range imaging and sensing. Optica 2023, 10, 1124–1141. [Google Scholar]
  12. Skompska, M. Hybrid conjugated polymer/semiconductor photovoltaic cells. Synth. Met. 2010, 160, 1–15. [Google Scholar] [CrossRef]
  13. Ferreira, M.F.; Castro-Camus, E.; Ottaway, D.J.; López-Higuera, J.M.; Feng, X.; Jin, W.; Jeong, Y.; Picqué, N.; Tong, L.; Reinhard, B.M.; et al. Roadmap on optical sensors. J. Opt. 2017, 19, 083001. [Google Scholar] [CrossRef]
  14. Veendrick, H.J. Nanometer CMOS ICs; Springer: Berlin/Heidelberg, Germany, 2017. [Google Scholar]
  15. Ma, H.; Luo, Y.; He, Y.; Pan, S.; Ren, L.; Shang, J. The short-range, high-accuracy compact pulsed laser ranging system. Sensors 2022, 22, 2146. [Google Scholar] [CrossRef] [PubMed]
  16. Hai, Y.; Li, Y.; Ma, X.; Zhao, X.; Zou, Y.; Hou, L.; Wang, D. Optical system design for small size laser ranging. In Proceedings of the 2015 International Conference on Optoelectronics and Microelectronics (ICOM), Changchun, China, 16–18 July 2015; pp. 115–118. [Google Scholar]
  17. Meng, Y.; Chen, W.; Li, W.; Zhao, Y.; Lu, C.; Cai, Y.; Chen, F.; Li, Z. Research on FPGA-based laser measuring instrument for surface flatness. In Proceedings of the 4th Optics Young Scientist Summit (OYSS 2020), Ningbo, China, 4–7 December 2020; pp. 349–357. [Google Scholar]
  18. Pellegrini, S.; Buller, G.S.; Smith, J.M.; Wallace, A.M.; Cova, S. Laser-based distance measurement using picosecond resolutiontime-correlated single-photon counting. Meas. Sci. Technol. 2000, 11, 712. [Google Scholar] [CrossRef]
  19. Petrie, G.; Toth, C.K. Introduction to laser ranging, profiling, and scanning. In Topographic Laser Ranging and Scanning; CRC Press: Boca Raton, FL, USA, 2018; pp. 1–28. [Google Scholar]
  20. Degnan, J.J. Scanning, multibeam, single photon lidars for rapid, large scale, high resolution, topographic and bathymetric mapping. Remote Sens. 2016, 8, 958. [Google Scholar] [CrossRef]
  21. Wehr, A. LiDAR systems and calibration. In Topographic Laser Ranging and Scanning; CRC Press: Boca Raton, FL, USA, 2018; pp. 159–200. [Google Scholar]
  22. Hawkins, A.R.; Schmidt, H. Handbook of Optofluidics; CRC Press: Boca Raton, FL, USA, 2010. [Google Scholar]
  23. Feng, L.-X.; Jiang, M.-S.; Bao, W.-S.; Li, H.-W.; Zhou, C.; Wang, Y. Controlling a sine wave gating single-photon detector by exploiting its filtering loophole. Chin. Phys. B 2018, 27, 080305. [Google Scholar] [CrossRef]
  24. Schaart, D.; Brunner, S. Bismuth Silicate as Detector Material for Top-Pet. Patent OCT-16-069. 2018. Available online: https://research.tudelft.nl/en/publications/bismuth-silicate-as-detector-material-for-top-pet (accessed on 3 July 2025).
  25. Yang, Y.; Wang, H.; Huang, Z.; Cao, Y.; Gui, H. A high resolution laser ranging system based on time-correlated single-photon counting technology. In Proceedings of the International Symposium on Optoelectronic Technology and Application 2014: Laser and Optical Measurement Technology; and Fiber Optic Sensors, Beijing, China, 13–15 May 2014; pp. 116–122. [Google Scholar]
  26. Gerardino, A.; Francardi, M.; Gaggero, A.; Mattioli, F.; Leoni, R.; Balet, L.; Chauvin, N.; Marsili, F.; Fiore, A. Nanophotonic technologies for single-photon devices. Opto-Electron. Rev. 2010, 18, 352–365. [Google Scholar] [CrossRef]
  27. Ren, M.; Gu, X.; Liang, Y.; Kong, W.; Wu, E.; Wu, G.; Zeng, H. Laser ranging at 1550 nm with 1-GHz sine-wave gated InGaAs/InP APD single-photon detector. Opt. Express 2011, 19, 13497–13502. [Google Scholar] [CrossRef]
  28. Luo, W.; Weng, Q.; Long, M.; Wang, P.; Gong, F.; Fang, H.; Luo, M.; Wang, W.; Wang, Z.; Zheng, D.; et al. Room-temperature single-photon detector based on single nanowire. Nano Lett. 2018, 18, 5439–5445. [Google Scholar] [CrossRef]
  29. Wang, W.; Jiang, Y.; Wang, C.; Li, X. Research Progress on Detection of Weak Light Signal by Photomultiplier Tube. In Proceedings of the 2021 International Conference on Advanced Electrical Equipment and Reliable Operation (AEERO), Beijing. China, 15–17 October 2021; pp. 1–5. [Google Scholar]
  30. Pichon, L.; Moignard, B.; Lemasson, Q.; Pacheco, C.; Walter, P. Development of a multi-detector and a systematic imaging system on the AGLAE external beam. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2014, 318, 27–31. [Google Scholar] [CrossRef]
  31. Doroud, K.; Rodriguez, A.; Williams, M.; Zichichi, A.; Zuyeuski, R. Comparative timing measurements of LYSO and LFS to achieve the best time resolution for TOF-PET. In Proceedings of the 2014 IEEE Nuclear Science Symposium and Medical Imaging Conference (NSS/MIC), Seattle, WA, USA, 8–15 November 2014; pp. 1–4. [Google Scholar]
  32. Ahlrichs, A.; Sprenger, B.; Benson, O. Photon counting and timing in quantum optics experiments. In Advanced Photon Counting: Applications, Methods, Instrumentation; Springer: Berlin/Heidelberg, Germany, 2015; pp. 319–341. [Google Scholar]
  33. Villa, F.; Markovic, B.; Bellisai, S.; Bronzi, D.; Tosi, A.; Zappa, F.; Tisa, S.; Durini, D.; Weyers, S.; Paschen, U.; et al. SPAD smart pixel for time-of-flight and time-correlated single-photon counting measurements. IEEE Photonics J. 2012, 4, 795–804. [Google Scholar] [CrossRef]
  34. He, Y.; Yu, C.; Yu, Y.; Bao, J.; Yang, B.; Li, X. Design and fabrication of integrated negative feedback resistor for InGaAs/InP avalanche photodiode. Infrared Phys. Technol. 2024, 142, 105566. [Google Scholar] [CrossRef]
  35. Tripathy, S.; Tyagi, K.; Pratap, P. A comparative study of various superconductors for Superconducting Nanowire Single Photon Detector applications. iScience 2024, 27, 110779. [Google Scholar] [CrossRef] [PubMed]
  36. Mazzillo, M.C.; Sanfilippo, D.N. Geiger-mode avalanche photodiode with high signal-to-noise ratio, and corresponding manufacturing process. U.S. Patent No. 9,105,789, 11 August 2015. [Google Scholar]
  37. Wang, F.-X.; Chen, W.; Li, Y.-P.; He, D.-Y.; Wang, C.; Han, Y.-G.; Wang, S.; Yin, Z.-Q.; Han, Z.-F. Non-Markovian property of afterpulsing effect in single-photon avalanche detector. J. Light. Technol. 2016, 34, 3610–3615. [Google Scholar] [CrossRef]
  38. Wang, X.; Tang, Z. Circular dichroism studies on plasmonic nanostructures. Small 2017, 13, 1601115. [Google Scholar] [CrossRef]
  39. Mccarthy, A.; Collins, R.J.; Krichel, N.J.; Fernandez, V.; Wallace, A.; Buller, G.S. Long-range time-of-flight scanning sensor based on high-speed time-correlated single-photon counting. Appl. Opt. 2009, 48, 6241–6251. [Google Scholar] [CrossRef] [PubMed]
  40. Ullrich, A.; Pfennigbauer, M. Linear LIDAR versus Geiger-mode LIDAR: Impact on data properties and data quality. In Laser Radar Technology and Applications XXI; SPIE: Bellingham, WA, USA, 2016; pp. 29–45. [Google Scholar]
  41. Du, B.; Pang, C.; Wu, D.; Li, Z.; Peng, H.; Tao, Y.; Wu, E.; Wu, G. High-speed photon-counting laser ranging for broad range of distances. Sci. Rep. 2018, 8, 4198. [Google Scholar] [CrossRef]
  42. Guan, Y.; Li, H.; Xue, L.; Yin, R.; Zhang, L.; Wang, H.; Zhu, G.; Kang, L.; Chen, J.; Wu, P. Lidar with superconducting nanowire single-photon detectors: Recent advances and developments. Opt. Lasers Eng. 2022, 156, 107102. [Google Scholar] [CrossRef]
  43. Guo, D.; Qu, Y.; Zhou, X.; Sun, J.; Yin, S.; Lu, J.; Liu, F. Research on Automatic Tracking and Size Estimation Algorithm of “Low, Slow and Small” Targets Based on Gm-APD Single-Photon LIDAR. Drones 2025, 9, 85. [Google Scholar] [CrossRef]
  44. Liang, Y.; Huang, J.; Ren, M.; Feng, B.; Chen, X.; Wu, E.; Wu, G.; Zeng, H. 1550-nm time-of-flight ranging system employing laser with multiple repetition rates for reducing the range ambiguity. Opt. Express 2014, 22, 4662–4670. [Google Scholar] [CrossRef] [PubMed]
  45. Xu, R.; Zheng, F.; Qin, D.; Yan, X.; Zhu, G.; Kang, L.; Zhang, L.; Jia, X.; Tu, X.; Jin, B.; et al. Demonstration of polarization-insensitive superconducting nanowire single-photon detector with Si compensation layer. J. Light. Technol. 2017, 35, 4707–4713. [Google Scholar] [CrossRef]
  46. Takeuchi, S. Photonic quantum information: Science and technology. Proc. Jpn. Acad. Ser. B 2016, 92, 29–43. [Google Scholar] [CrossRef]
  47. Natarajan, C.M.; Tanner, M.G.; Hadfield, R.H. Superconducting nanowire single-photon detectors: Physics and applications. Supercond. Sci. Technol. 2012, 25, 063001. [Google Scholar] [CrossRef]
  48. Hadfield, R.H. Single-photon detectors for optical quantum information applications. Nat. Photonics 2009, 3, 696–705. [Google Scholar] [CrossRef]
  49. Zhang, L.; Kang, L.; Chen, J.; Wu, P. Development of Superconductor Nanowire Single Photon Detector; Research Institute of Superconductor Electronics of Nanjing University: Nanjing, China, 2014. [Google Scholar]
  50. Hu, P.; Li, H.; You, L.; Wang, H.; Xiao, Y.; Huang, J.; Yang, X.; Zhang, W.; Wang, Z.; Xie, X. Detecting single infrared photons toward optimal system detection efficiency. Opt. Express 2020, 28, 36884–36891. [Google Scholar] [CrossRef]
  51. Esmaeil Zadeh, I.; Chang, J.; Los, J.W.; Gyger, S.; Elshaari, A.W.; Steinhauer, S.; Dorenbos, S.N.; Zwiller, V. Superconducting nanowire single-photon detectors: A perspective on evolution, state-of-the-art, future developments, and applications. Appl. Phys. Lett. 2021, 118, 190502. [Google Scholar] [CrossRef]
  52. Gu, M.; Kang, L.; Zhang, L.; Jia, T.; Wan, C.; Xu, R.; Yang, X.; Wu, P.; Witzigmann, B.; Osiński, M.; et al. High detection efficiency and rate superconducting nanowire single-photon detector with a composite optical structure. In Physics and Simulation of Optoelectronic Devices XXIII; SPIE: Bellingham, WA, USA, 2015; pp. 287–294. [Google Scholar]
  53. Lamb, R.A. A review of ultra-short pulse lasers for military remote sensing and rangefinding. Technol. Opt. Countermeas. VI 2009, 7483, 61–75. [Google Scholar]
  54. Zhou, G.; Zhou, X.; Song, Y.; Xie, D.; Wang, L.; Yan, G.; Hu, M.; Liu, B.; Shang, W.; Gong, C.; et al. Design of supercontinuum laser hyperspectral light detection and ranging (LiDAR)(SCLaHS LiDAR). Int. J. Remote Sens. 2021, 42, 3731–3755. [Google Scholar] [CrossRef]
  55. Xue, L.; Li, Z.; Zhang, L.; Zhai, D.; Li, Y.; Zhang, S.; Li, M.; Kang, L.; Chen, J.; Wu, P.; et al. Satellite laser ranging using superconducting nanowire single-photon detectors at 1064 nm wavelength. Opt. Lett. 2016, 41, 3848–3851. [Google Scholar] [CrossRef]
  56. Choquer, M. Surface Acoustic Waves Integrated with Solid-State Single Photon Sources and Superconducting Electronics; University of California: Santa Barbara, CA, USA, 2024. [Google Scholar]
  57. You, L. Superconducting nanowire single-photon detectors for quantum information. Nanophotonics 2020, 9, 2673–2692. [Google Scholar] [CrossRef]
  58. Kang, L.; Chen, J.; Zhang, L.; Wu, P.H. Superconducting nanowires for detecting single photons at telecommunication wavebands. In Proceedings of the 2010 3rd International Nanoelectronics Conference (INEC), Hong Kong, China, 3–8 January 2010; pp. 28–29. [Google Scholar]
Figure 1. Schematic diagram of internal photoelectric effect in PN junction.
Figure 1. Schematic diagram of internal photoelectric effect in PN junction.
Coatings 15 00798 g001
Figure 2. Schematic diagram of PIN-type photodiode structure.
Figure 2. Schematic diagram of PIN-type photodiode structure.
Coatings 15 00798 g002
Figure 3. Schematic diagram of the pulse method.
Figure 3. Schematic diagram of the pulse method.
Coatings 15 00798 g003
Figure 4. Schematic diagram of the phase method.
Figure 4. Schematic diagram of the phase method.
Coatings 15 00798 g004
Figure 5. Structural profile of In0.53Ga0.47As/InP APD.
Figure 5. Structural profile of In0.53Ga0.47As/InP APD.
Coatings 15 00798 g005
Figure 6. Typical structure of avalanche diode (left picture without protective ring).
Figure 6. Typical structure of avalanche diode (left picture without protective ring).
Coatings 15 00798 g006
Table 1. Common photodiodes.
Table 1. Common photodiodes.
SpeciesPeculiaritiesPurpose
PNThe dark current is small, but the response speed is relatively lowColor sensor, phototriode, linear image sensor, spectrophotometer, camera exposure meter, etc.
PINHigh sensitivity, fast response speed, wide bandwidth, but the dark current is relatively large and there is low temperature sensitivityOptical communication, optical fiber, remote control, high voltage rectifiers, RF and DC controlled microwave switches, photodetectors, photovoltaic cells, etc.
APDFast response and internal gain mechanismHigh-speed optical communication, high-speed optical detection, laser rangefinder, positron emission tomography, and other fields
Table 2. Structure of photomultiplier tube.
Table 2. Structure of photomultiplier tube.
ComponentsPurpose
Glass bulbIsolate the air and provide a vacuum working environment for the functional parts of the photomultiplier
Ceramic substrateFixed electron multiplier pole and photocathode, anode, and other parts
PhotocathodeThe photoelectric effect is used to convert photons into electrons
Electron multiplication systemThe photoelectrons are multiplied and amplified by the principle of secondary electron emission
AnodeAccept the multiplied electrons and output them as a current
ResistorThe potential difference between adjacent electron doubling poles ensures that electrons can multiply between electron doubling poles
Table 3. SNSPD R&D organization and performance information.
Table 3. SNSPD R&D organization and performance information.
NameWavelength/nmDetection Efficiency/%Dark Count Rate/cpsRepetition Rate/MHzTime Jitter/psTemperature/KRefrigeration Method
NIST [47]1550931101500.1Dilution refrigeration machine
MIT [48]155057100100301.6Pulse tube chiller
NICT [48]15508410050682.1Stirling refrigerator
SIMIT [7]15509510020662.1Stirling refrigerator
USTC1550920.127450<500.1Liquid helium + pulse tube refrigeration
NJU [49,50]15509550100401.8Hybrid refrigeration system
JPL [51]155096250026.81.5Micro-pulse tube refrigeration machine
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

Li, Z.; Jin, X.; Yuan, C.; Wang, K. Photon Detector Technology for Laser Ranging: A Review of Recent Developments. Coatings 2025, 15, 798. https://doi.org/10.3390/coatings15070798

AMA Style

Li Z, Jin X, Yuan C, Wang K. Photon Detector Technology for Laser Ranging: A Review of Recent Developments. Coatings. 2025; 15(7):798. https://doi.org/10.3390/coatings15070798

Chicago/Turabian Style

Li, Zhihui, Xin Jin, Changfu Yuan, and Kai Wang. 2025. "Photon Detector Technology for Laser Ranging: A Review of Recent Developments" Coatings 15, no. 7: 798. https://doi.org/10.3390/coatings15070798

APA Style

Li, Z., Jin, X., Yuan, C., & Wang, K. (2025). Photon Detector Technology for Laser Ranging: A Review of Recent Developments. Coatings, 15(7), 798. https://doi.org/10.3390/coatings15070798

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

Article Metrics

Back to TopTop