Next Article in Journal
Wearable Flexible Wireless Pressure Sensor Based on Poly(vinyl alcohol)/Carbon Nanotube/MXene Composite for Health Monitoring
Previous Article in Journal
Atomic-Scale Revelation of Voltage-Modulated Electrochemical Corrosion Mechanism in 4H-SiC Substrate
Previous Article in Special Issue
Enhanced UVC Responsivity of Heteroepitaxial α-Ga2O3 Photodetector with Ultra-Thin HfO2 Interlayer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances in Silicon-Based UV Light Detection

1
Department of Semiconductor Engineering, Gachon University, Seongnam-si 13120, Republic of Korea
2
Gachon Bionano Research Institute, Gachon University, Seongnam-si 13120, Republic of Korea
3
Department of Physics and Semiconductor Science, Gachon University, Seongnam-si 13120, Republic of Korea
*
Authors to whom correspondence should be addressed.
Micromachines 2025, 16(10), 1130; https://doi.org/10.3390/mi16101130
Submission received: 28 August 2025 / Revised: 24 September 2025 / Accepted: 26 September 2025 / Published: 30 September 2025
(This article belongs to the Special Issue Photodetectors and Their Applications)

Abstract

Silicon (Si), the cornerstone semiconductor in the micro-electronics industry, can provide a cost-efficient platform with mature technologies for photodetection in visible and near-infrared regions. However, its intrinsic properties, such as a narrow bandgap and the shallow penetration depth of ultraviolet (UV) light into its surface with surface trap states, remain challenges, rendering it unsuitable for effective UV light detection. Various techniques have been reported to circumvent these surface defect-induced difficulties. In addition, wide-bandgap semiconductors that favor UV light absorption in a solar-blind way have been combined with Si for UV light detection in order to retain the device’s compatibility with Si-CMOS processes, though it still faces challenges that need to be overcome. This review starts with concepts of basic parameters of photodetectors and categorizes UV photodetectors according to their detection mechanisms. We also present a review of wide-bandgap semiconductor-based UV light detectors and those based on Si, with a discussion of surface defect minimization. In addition, we review the hybrid structure of the two kinds, i.e., wide-bandgap semiconductors and Si, and discuss their properties that produce synergistic effects. Lastly, we provide conclusions and outlooks for the possible development of next-generation UV light detectors based on Si.

1. Introduction

Ultraviolet (UV) radiation occupies the invisible band of the electromagnetic spectrum, with wavelengths ranging from 100 to 400 nm. UV radiation is further classified into four categories, namely, UV-A (315–400 nm), UV-B (280–315 nm), UV-C (200–280 nm), and Vacuum UV (VUV) (l00–200 nm). The sun is one of the sources of UV light, though substantial amounts of its light are screened out via the absorption of atmospheric gases [1]. The full spectrum of electromagnetic waves is depicted in Figure 1.
UV light has found applications ranging from water purification [2,3,4,5] and photolithography [6,7,8,9,10,11] to optical communication [12,13], due to its high photon energies and lower degree of diffraction. Such a variety of applications of UV light requires its accurate and precise detection, generating continued interest in the development of highly efficient UV detectors. The numerous UV detection systems that have been developed can be categorized into photoelectric effect-based photomultiplier tubes [14], photoconductive detectors [15,16], photovoltaic detectors [17,18], field-effect transistors (FETs) [19,20,21], pyroelectric detectors [22,23], and thermoelectric detectors [24,25]. UV detectors usually employ wide-bandgap semiconductors that generate photogenerated electrons and holes only when the incident photon energy is equal to or greater than their bandgap energies. These wide-bandgap semiconductors include inorganic compound semiconductors [26,27,28,29,30], organic materials [26,27,28,29,30,31,32,33], and perovskite materials [34,35]. The wide bandgap energy required for UV photon absorption can also be realized by reducing the size of photon-absorbing material into nanoscales, such as quantum wells [36], quantum wires [37,38], and quantum dots [39,40,41].
Silicon (Si), widely used in semiconductors for detecting visible light, has a bandgap energy of about 1.12 eV at room temperature [42,43] and a relatively long carrier diffusion length of hundreds of micrometers [44,45]. However, its narrow bandgap energy allows it to absorb photons mostly in the visible and IR regions, leading to solar sensitivity. Moreover, the UV penetration depth into Si’s surface is known to be around 10 nm, resulting in photogenerated carriers near the surface that cannot diffuse into the bulk, but are annihilated through carrier recombination via surface defects and traps [46]. Furthermore, an unwanted oxidized layer of SiO2 on Si’s surface can absorb or scatter UV photons, with the consequence of rolling back UV light absorption by Si [47,48].
Despite the abovementioned drawbacks of Si for UV light detection, it possesses a considerably large absorption coefficient of over 106 cm−1 in the UV range. Figure 2 presents the absorption coefficient of Si as a function of wavelength, ranging from the 200 nm (UV) to 1500 (IR) region. Its strong absorption in the UV region makes it inherently fit for UV detection. In addition, its compatibility with the pre-matured technologies of complementary metal oxide semiconductors (CMOSs) favors mass production of Si-based UV detectors at a low cost due to their lower power consumption and their easy integration into a system (e.g., system on a chip) [49,50]. These aspects, together with other encouraging characteristics of Si, including the long lifetime of photogenerated carriers and its abundance in nature, have tempted researchers to develop Si-based UV detectors. This review article is structured as follows. Section 2 presents the definition of several parameters relevant to the performance of photodetectors that were not restricted to those for UV detection. Section 3 provides a classification of UV photodetectors according to detection mechanisms. Section 4 briefly reviews photodetectors based on wide-bandgap semiconductors, along with their performance. Section 5 covers Si-based photodetectors, along with their properties and progress made to overcome the relevant challenges. The hybrid structure of wide-bandgap semiconductors and Si is also presented with a relevant discussion. Lastly, we provide conclusions and outlooks for the future direction of Si-based UV light detectors.

2. Parameters Relevant to Photodetector Performance

A photodetector, in general, is characterized by a set of parameters that determine its performance in detecting incident light power. These parameters include the dark current (Id), total current (It), photocurrent (Ip), applied voltage bias, responsivity (R), external quantum efficiency (EQE), spectral response (SR), response time (tr), signal to noise ratio (SNR), noise equivalent power (NEP), specific detectivity (D), and linearity and dynamic range. In this section, we define the abovementioned parameters as follows:
i.
Dark Current (Id)
Dark current (Id) is the current that flows through the photodetector circuit at a given voltage bias when no light is incident on the detector active area. It arises primarily due to thermally excited carriers through mid-gap trap states, including those near the surface or the interface of a light-absorbing semiconductor/medium. Id restricts the photodetector sensitivity, thus being considered one of the most significant parameters for detecting light with low optical power [52,53].
ii.
Photocurrent (Iph)
Photocurrent (Iph) is the current obtained by subtracting dark current from the current measured when light is incident on a photodetector area. Iph should have no dependence on ambient temperature but be directly in proportion to the optical power of light incident on the active photodetector area. When subtracting dark current for estimating Iph, light absorption-induced thermal energy dissipation that may change the level of dark current must be considered.
iii.
Responsivity (R)
Responsivity (R) is the ratio of the Iph with respect to the power of incident light ( P i n ), thus having a unit of Ampere/Watt. It manifests how effectively a photodetector converts optical power to electrical current. Particularly, in photodetectors with no internal gain such as in photodiodes, R is directly proportional to internal quantum efficiency η , representing a probability that a photo-generated pair of charge carriers are collected per unit photon absorbed, as given by
η = ( I p h / e ) / ( P a / ω ) = ( ω / e ) ( R / A ) ,
where e is the electronic charge, P a is the absorbed light power, ω is the frequency of absorbed light, is the Plank constant, and P a = A   P i n ( A < 1 is the absorption efficiency). In general, η depends on the materials and structure of a photodetector.
Typically, R < 1 , and it becomes larger at a longer wavelength since there is a greater number of photons present in 1 W optical power at longer wavelengths. Meanwhile, R can exceed unity in some cases, including a photodiode with a very large η value (e.g., close to unity) at a longer wavelength, for instance, near infrared. In addition, internal gain available from multiple charge carriers generated from absorbed single photon can account for R > 1 , such as in photoconductive detectors and avalanche photodiodes.
iv.
External Quantum Efficiency (EQE)
External quantum efficiency (EQE) describes how effectively incident photons are converted to photogenerated carriers in a photodetector, taking into account the optical losses of incident light due to light reflection/scattering from the surface of the detector’s active medium and light transmission through it. This allows EQE to be given by the following:
E Q E = η A .
Similarly η , EQE can exceed unity (100%) in cases of internal gain due to the contribution of multiple carriers to current per single photon absorbed.
v.
Voltage Bias
The voltage bias is the external voltage applied across the electrode of a photodetector to separate photogenerated electrons and holes in typical photoconductive detectors or even accelerate them into hot carriers with huge kinetic energy, such as in an avalanche photodiode. Figure 3a,b show the R versus UV wavelength for various bias voltages, while Figure 3c,d show the EQD versus UV wavelengths for various bias voltages in a photodetector. As the bias voltage rises from zero, R and EQE increase until a given bias voltage. This is simply due to the fact that enhanced fields increase the drift current but suppress the radiative recombination rate. However, when increasing the bias voltage beyond the given point, defect state-mediated non-radiative recombination begins to play a significant role due to a widened depletion region in photodiodes such as Shockley–Read–Hall recombination [54,55] and trap-assisted tunneling [56,57,58].
vi.
Response Time
Photodetector response time refers to the time it takes for photocurrent to rise or fall when light is turned on or off, being inversely proportional to the device bandwidth. Response time of the order of magnitude of ns or ps can support high-speed optical signal processing. Many factors affect the response time, such as traverse time of photogenerated carriers across the device electrodes, carrier mobility, carrier diffusion time, carrier lifetime, the device resistance-capacitance product, and carrier trapping/non-radiative recombination [60,61,62].
vii.
Noise EquivalentOptical Power (NEP)
Noise equivalent optical power (NEP) refers to the incident optical power that produces a signal-to-noise ratio of unity, and scales with the square root of the electrical noise power. The electrical noise is comprised mainly of shot noise, dark current noise, and Nyquist noise—thus, NEP scales with the square root of the device bandwidth Δ f (Hz).
viii.
Specific Detectivity (D*)
Detectivity (D) is defined by the reciprocal of NEP, representing the photodetector sensitivity. However, NEP scales with Δ f and A , where A is the photodetector active area to which dark current is approximately proportional. Therefore, NEP can be normalized by ( Δ f ) A  into a parameter independent of the active area size and bandwidth. Then, the specific detectivity (D*) is defined by the reciprocal of the normalized NEP so that D* = A Δ f / N E P . It is the figure of merit expressed in the unit of Jones ( c m H z / W ), which is used for estimating how low optical power of light can be detected regardless of the size of the photodetector active area and its bandwidth. In cases where dark current noise dominates electrical noise, such as in the detection of low light power and infrared light detection, D* can be approximated as follows:
D *   R A / 2 e I d .

3. Classification of Photodetectors by Detection Mechanism

The development of UV photodetectors has evolved into a diverse family of devices, each customized for a particular spectral region and sensitivity. In this section, we categorize them according to light-detection mechanisms, though this classification may equally apply to photodetectors for visible and infrared light wavelengths. This classification may help us to estimate which types of photodetectors would be favorable for UV light detection in terms of suppression of dark current, optimization of responsivity, and internal gain harvest through carrier multiplication (see Table 1 and Table 2).

3.1. Photomultiplier Tube (PMT) for Photodetection

Electromagnetic waves in a wide spectrum ranging from deep UV to near infrared can be detected using the photoelectric effect, whereby electrons are ejected from the material surface when UV photons are incident on it. Such photoelectrons are generated when a photon energy exceeds the work function/electron affinity of the material, which acts as a photocathode, as shown in Figure 4. A voltage as high as hundreds of volts is applied between a photocathode and a dynode, subsequently accelerating photoelectrons toward it. They eventually hit a dynode, generating secondary electrons via impact ionization, e.g., inelastic collision whereby bound electrons in a dynode are knocked out into the vacuum (secondary electrons). The secondary electrons are then accelerated and hit the next dynode, leading to cascading generation of hot electrons, which are finally collected by the anode of the detector. The cascading multiplication of photoelectrons in this kind of device, called a photomultiplier tube (PMT), accounts for electrical signal gain, which could grow exponentially to be enormous enough to detect the single-photon level. Despite this ultrahigh sensitivity, PMTs face challenges in their miniaturization and robustness. PMT finds various applications in fluorescence, nuclear/high-energy physics, and medical imaging [63,64,65]. For the detection of deep UV and daylight UV, alkali metal compounds such as CsI and Ce2Te can be used for a photocathode of PMT [66,67,68]. Moreover, bi-alkali metal compounds such as K2CsSb can be used for a photocathode of PMT that detects near-UV and blue light [69,70].

3.2. Photodiodes and Photovoltaic Devices for Light Detection

In general, a photodiode relies on a heterostructure that creates diode characteristics of electrical current, such as p-n/p-i-n structures of heterostructure semiconductors or a metal–semiconductor structure. The junction potential at the interface between p- and n-type semiconductors in a p-n structure (Figure 5A) can be inherently formed by majority carrier diffusion, which is induced by the Fermi level difference. Insertion of an intrinsic semiconductor between p- and n-type semiconductors results in an enhanced depletion region, increasing the photon absorption region, as shown in Figure 5B. The junction potential can also be enhanced by an external voltage bias connected in a reverse way, boosting effective separation of photogenerated electron–hole pairs [72,73,74,75,76]. The reverse bias reduces not only the carrier transit time across the electrodes but also device capacitance due to a widened depletion region, resulting in faster response. Removal of an external voltage from the device circuit leads to the photovoltaic mode, which favors lowering the dark current but suffers from a slow response compared to photodiodes [77,78].
In both photodiode and photovoltaic devices, a region depleted of carriers (depletion region) at the interface can absorb UV light only if its bandgap energy is equal to or greater than the photon energies. Wide-bandgap materials can lend themselves to both the absorption of photon energies greater than 3.1 eV (corresponding to 200 nm) and solar blind response, such as GaN [79,80], AlGaN [81,82], SiC [83], and diamond [84,85].
Figure 5. Energy structures of (A) p-n junction semiconductors and (B) p-i-n junction semiconductors. Reproduced from [86] (Exploration, John Wiley & Sons 2022). Licensed under CC BY 4.0.
Figure 5. Energy structures of (A) p-n junction semiconductors and (B) p-i-n junction semiconductors. Reproduced from [86] (Exploration, John Wiley & Sons 2022). Licensed under CC BY 4.0.
Micromachines 16 01130 g005

3.3. Photoconductive Detectors for Light Detection

In general, the conductivity of a channel through which current flows can increase when photogenerated carriers begin to be added to it, leading to a current increase in proportion to the light intensity incident on the active area of the device. Photodetectors employing this principle for light detection are called photoconductive detectors or photo-resistive detectors (because resistivity also changes as its reciprocal due to the same reason, i.e., photogenerated carriers) [15,87,88,89]. Photoconductive light detectors take advantage of the photo-induced changes in conductivity in a device. For instance, the abovementioned photodiodes use conductivity increase due to the creation of photogenerated carriers in the depletion region, making them fall within this category (the photoconductive mode) under a reverse bias across heterojunctions such as p-n/p-i-n structures. This external bias voltage would drive photogenerated carriers for photocurrent that increases with light intensity. Another example is a current flow channel of a single semiconductor across which an external bias is applied. When the channel semiconductor absorbs photons, photogenerated carriers increase conductivity, producing a current increase.
Very often, they have non-negligible internal gain for current, for which multiple carriers are collected at electrodes under a device neutrality, given a single pair of photogenerated electrons and holes, particularly when a photogenerated carrier’s lifetime is longer than a carrier transit time [90,91,92]. This eventually increases R and Id while decreasing the device bandwidth. As shown in Figure 6, a semiconductor material is connected between two metal electrodes under an external bias voltage. As photogenerated carriers are generated upon incident light absorption, external bias drives those carriers for current. For UV light detection, wide-bandgap semiconductors can be adopted between device electrodes as a medium where UV photon absorption takes place. This medium also plays a channel through which current flows, meaning its carrier mobility needs to be high for high device bandwidth. The media used for UV photon absorption in photoconductive detectors include ZnO [93], SiC [94], GaN [79], AlGaN [95], gallium oxide (Ga2O3) [96], Graphene [87], and Si [97].

3.4. Field-Effect Transistor-Based Photodetectors

Field-effect transistors (FETs) have been used for photodetection via channel conductivity modulation by photogenerated carriers [99]. As shown in Figure 7, its structure typically comprises source, drain, gate, photon-absorbing media (which could also be the same as the channel itself), an oxide dielectric layer (e.g., SiO2 or Al2O3), a channel, and a substrate. Like the role of a gate bias voltage that changes channel conductivity, incident photons can act as a photogate [100]: one type of photogenerated carriers can be trapped in an absorber (in cases of small dimensions such as a quantum dot), near the absorber–channel interface, or channel–dielectric interface. Then, trapped carriers can produce local electric fields as extra gate fields, therefore strongly affecting the gate voltage threshold in FET performance. This eventually enables great change in conductivity to be achieved. Photogenerated carriers themselves in a channel can also contribute to photoconductivity change. FET-based photodetectors take advantage of ultrahigh responsivity, dark current suppression, low power consumption, and compatibility with CMOS processes for integrated chips [101], while facing the challenge of limited bandwidth and degradation issues, particularly in hybrid structures, such as with perovskite and quantum dots [102,103,104].

3.5. Avalanche Photodiodes for Light Detection

An avalanche photodiode (APD) utilizes an avalanche effect for which a reverse bias voltage is adjusted to be close to or beyond the breakdown voltage, i.e., tens to hundreds of volts. This high electric field that adds up to the inherent junction field accelerates photogenerated carriers in the junction (carrier depletion) of p-n/p-i-n structure semiconductors, kicking them into (unthermalized) hot carriers with enormous kinetic energies. They collide with lattice atoms, knocking their bound electrons into a conduction band (creating additional free electrons) while generating holes in a valence band. Those additional carriers are again accelerated and undergo similar procedures, giving rise to carrier multiplication in a cascading process, as shown in Figure 8. This avalanche effect leads to an ultrahigh gain of about 10–100 for current, with a consequence of ultrahigh R, making the device suited for low-light detection. However, the random characteristics of impact ionization by electron–atom collision produce excess noise as the trade-off for very high gain. Avalanche photodiodes can be operated either in a linear mode or Geiger mode, depending on the external bias magnitude (smaller or greater than a breakdown voltage). SiC is one of the widely used semiconductors as an absorbing medium for UV light APD due to its wide bandgap energy (3.2 eV) and high breakdown voltage (100–400 V for a thin junction, while more than 1 kV for a thick junction) [106,107]. AlGaN and GaN can also be used as a UV light-absorbing medium with additional benefits due to their good solar blindness [108,109]. Si can also be used as an absorbing medium due to its high absorption coefficient at the UV light spectral wavelengths, as mentioned above. However, the very short penetration depth of UV light into Si hinders it from being used as a deep-UV light-absorbing medium due to its surface trap-mediated recombination of photogenerated carriers. Instead, Si-based APD has found application in APD for near-UV light detection [110,111].

4. Wide-Bandgap Semiconductors and Metal–Semiconductor–Metal Structure for UV Light Detectors

Photodetectors for UV light detection use wide-bandgap semiconductors as light-absorbing media, such as SiC, ZnO, TiO2, GaN, AlGaN, Ga2O3, AlN, diamond, perovskite oxides (SrTiO3, BaTiO3), and their hybrid combinations. These materials benefit from the large bandgap energies of 3.2–6.2 eV, solar blindness, and high absorption coefficients at UV wavelengths. Some of them also exhibit the large excitonic binding energies, such as ZnO ( ~ 60 meV) [112], AlGaN ( ~ 25–80 meV) [113,114], AlN ( ~ 71–80 meV) [115,116], and diamond ( ~ 70–80 meV) [117,118], enabling the detection of specific UV wavelengths. In addition, among those materials, reasonably high mobility of carriers can also be found in SiC ( ~ 800–1000 cm 2 /V · s for electrons, ~ 100 cm 2 /V ·; s for holes) [119,120], GaN ( ~ 1000 cm 2 /V · s for electrons, ~ 200 cm 2 /V · s for holes) [121,122], and diamond ( ~ 2200 cm 2 /V · s for electrons, ~ 1600 cm 2 /V · s for holes) [123,124,125], allowing for high-bandwidth detectors.
Wide-bandgap semiconductors are often utilized in “metal-semiconductor-metal (MSM) photodiodes with interdigitated electrodes. They have gained popularity for UV light detection due to their high bandwidth, low dark current, easy fabrication (standard microfabrication techniques), and monolithic integrability with other electronic devices. Interdigitated electrodes of gold or platinum can be patterned like interlaced fingers on a semiconductor (light-absorbing medium), as shown in Figure 9. The metal–semiconductor interface in the MSM structure forms back-to-back Schottky barriers whose height can be modulated by externally applied bias voltage. Interdigitated electrode structures enable the electric fields between electrodes to be homogeneous while the carrier transit time between them would be short, being beneficial for high bandwidth (fast response). An interdigitated electrode structure can also be optimized for dark current suppression and capacitance minimization. As the semiconductors for UV light absorption, GaN [126], AlGaN [127], ZnO [128], SiC [129], and diamond are used [130]. There could be carrier-trapped states at the metal–semiconductor interface affecting Id and the effective bias voltage due to additional local static fields from trapped carriers.
The characteristics of UV light detectors with various wide-bandgap semiconductors are briefly summarized in Table 1. However, it is worth mentioning that wide-bandgap semiconductors still face difficulties in cost-effective production of their crystals [33,132,133], industrial scalability of their crystalline wafer size with high quality [26], and monolithic integration with CMOS-based optoelectronic components [134].
Table 1. Wide-bandgap semiconductor-based materials used for UV light detectors.
Table 1. Wide-bandgap semiconductor-based materials used for UV light detectors.
Active MaterialBandgap Energy (eV)UV Wavelength (nm)Photodetector TypeResponsivity (A/W)Bias Voltage (Volt)Ref.
AlN~6.2193photoconductive0.3920[135]
Boron Nitride (BN)~6200photoconductive0.09520[136]
Diamond~5.48222photoconductive22.650[137]
α-Ga2O3~5.1–5.3230photoconductive2.7110[138]
€-Ga2O3~4.9240photoconductive52.7720[139]
WO3/AlGaN/GaN~3.2240Photo-FET1.67 × 1040.5[140]
AlGaN~3.4–6.2250photoconductive~1065[141]
Si/SiC~1.12/3.26260Photo-FET4.63 × 1055[142]
TiO2~3.2260photoconductive1995[143]
AlGaN/GaN~3.4–6.2/3.4265Photo-FET3.6  ×  107−8.2[144]
Graphene/SiC~0/3.26275photoconductive5.425 × 1031[145]
SnO2~3.6322photoconductive1.353 × 10310[146]
ZnO~3.37325photoconductive2.6 × 1048[147]
GaN3.4340photoconductive~1.3 × 1041[148]
CsPbCl3 nanowires~3.03360Photovoltaic0.3980[149]
SiC3.26360n+/p/n Photo-FET2.02 × 10437[150]
GaSe/GaN~3.4362p-n heterojunction1.38 × 105−4[151]
W18O49/TiO2~3.2365Photoconductive1.6 × 1041[152]
AlGaN/GaN~3.4–6.2/3.4365Photo-FET1  ×  106−8.2[144]

5. Silicon-Based UV Detectors: Structural Advancements and Challenges

The use of silicon for UV photodetectors can offer many advantages. One of them is the availability of mature CMOS-compatible technologies for passivation, metallization, photolithography, and doping. It can also offer cost-effective fabrication of high-quality detectors, large-scale wafer-based production of detector arrays, and integrability with other electronic components. Despite its bandgap energy of 1.12 eV (1100 nm) and indirect bandgap nature, it exhibits large absorption coefficients at UV wavelengths (Figure 2), as mentioned above. However, Si absorbs visible photons in addition to UV photons but shows very low R, particularly in the deep UV regions (near 300 nm), as shown in Figure 10. The low sensitivity of Si to UV light is due to shallow penetration depth (a few nanometers) below its surface and significant surface trap-induced non-radiative recombination of photogenerated carriers. Various surface engineering techniques and sensitization strategies could be applied to improve the UV light absorption and reduce the surface-trap-induced non-radiative recombination, finally enhancing the UV detection performance of Si.

5.1. Surface Passivation

The shallow penetration depth of UV light into Si leads to the generation of photogenerated carriers, mostly near its surface. The Si surface contains defect states, such as dangling bonds, that are unwantedly created during processes of etching and polishing. These tend to play the surface-trap state as non-radiative recombination centers. The surface-induced effects for non-radiative decay of carriers can be reduced by depositing a thin dielectric layer on the Si surface, such as SiO2, Al2O3, and Si3N4, subject to its transparency to a given UV wavelength. These passivation effects come into play in various ways. First, such a deposited layer can make a chemical saturation of the surface dangling bonds, which neutralizes the recombination centers (chemical passivation), such as found in the deposition of thermally grown SiO2, whose oxygen (O) bonds with unsaturated Si atoms [154,155]. Second, the dielectric layers can possess fixed charges, such as negative fixed charges in an atomically deposited layer of Al2O3 on Si [156,157]. These fixed charges generate surface-normal electric fields and then create a space charge layer near the interface, which is depleted of carriers. This eventually reduces electron–hole recombination rates, called electric field passivation. The abovementioned Si passivation technologies with SiO2 and Al2O3 are widely reported for Si-based solar cells. Si passivation for UV light detection has yet to be used to develop a method that achieves a sufficiently high EQE and requires an understanding of its working principle, including the relevant carrier dynamics.
A surface passivation approach suggested by Nizdak et al. [158] involved delta-doped surface passivation using molecular beam epitaxy (MBE). For such passivation, a single atomic layer of boron was deposited on a Si substrate, followed by epitaxially overgrown Si that creates a stable dipole and effectively passivates the interface traps. This process enabled suppression of surface recombination and eliminated the charge-induced instabilities of Si. Further deposition of multiple delta-doped layers was performed to control the band bending and directly extend the depletion region to the Si surface for efficient carrier collection. This led the device to achieve a stable high EQE of 80% at a 206 nm UV wavelength, which corresponds to a responsivity of 0.1329 (A/W).

5.2. Ion Implantation

Unwanted carrier recombination near the surface can also be reduced by implanting external ions or distributing impurities near the surface region of pure Si wafers in the form of vacancies and interstitials, benefitting the successful collection of UV-photogenerated carriers [159]. For instance Kuroda et al. [160] achieved approximately 100% internal QE with the estimated R of 0.2016 (A/W) under 250 nm UV illumination using a Si photodiode fabricated with low-energy ion implantation in the following way: the p-type Si wafer was initially annealed at high temperature both to flatten its surface at an atomic level and to reduce the surface trap states at an interface between Si and SiO2 layer; the oxidized layer already existed or generated during annealing process. Figure 11a presents a cross-sectional view of a p-type Si wafer with an atomically flat surface achieved by standard RCA cleaning followed by annealing of the Si wafer. Figure 11b shows a conventional Si surface with a rough surface (cross-sectional view). Subsequently, low-energy arsenic ion implantation introduced donor impurities, creating an n-type region near the surface so that a charged spacer layer forms due to positively ionized donors (As+). Therefore, the induced electric fields hindered the carrier recombination and improved the internal QE of the photodetector. The device showed stable performance under prolonged (1000 min) UV illumination.
Low-energy boron implantation was conducted on the plasma-etched surface of black silicon (b-Si) [161] to generate the 1.5 μ m deep p-type region near the surface, as shown in Figure 12. Near the surface, the ionized acceptors (implemented boron ions) built up surface-normal electrical fields together with ionized donors of the n-type Si, generating a surface-near region depleted of carriers. This eventually reduces the carrier recombination rate. Carrier recombination rate near the surface is further reduced by atomically layered deposition of a 50 nm Al2O3 film, as mentioned above. This structure yields R of 0.15 (A/W) at 200 nm wavelength and 0.3 (A/W) at 400 nm. A similar device designed by Garin et al. [162] achieved EQE of 132% with a corresponding R value of 0.213 (A/W).

5.3. Quantum Cutting

The principle of quantum cutting is straightforward yet profound. A fluorescent material, immobilized on the surface of Si in a photodetector, absorbs high-energy UV photons and emits two or more photons with lower energy through a cascade of energy transitions of downconversion [163]. In such a scenario, the emitted photon energy should fall within the detection range of the Si photodetector. The downconversion by photon splitting, called “quantum cutting”, is theoretically expected to achieve a high EQE exceeding 100% since a single UV photon can lead to the generation of multiple electron-hole pairs in the underlying Si substrate.
The implication of quantum cutting in UV detecting devices relies on the development of efficient down-conversion materials. Lanthanide-doped phosphors [163,164], Silicates [165], and perovskite nano-crystals and quantum dots [166,167] have recently garnered attention as downconversion materials. Lanthanides have the unique energy levels of a ladder pattern for 4f electrons, which are shielded by 5s/5p orbitals, making them less vulnerable to host lattice vibration. This leads to long lifetimes of excited states and enhanced radiative decay. Lanthanides can serve both as sensitizers that absorb UV photons and as activators that are excited and emit photons of lower energies (near-infrared wavelengths). Quantum cutting materials are directly applicable using conventional coating methods or can be designed with luminescent concentrators (i.e., waveguide) to collect emission light effectively upon UV absorption in the 200 nm to 400 nm range [153,168].
For Si-based detectors, Shao et al. spin-coated the compound materials, i.e., La3+, Yb3+ co-doped CsPbClBr2 quantum dots, on Si and achieved an EQE of 55.6% and an R of 0.131 (A/W) at the UV wavelength of 240 nm [169]. The La3+ ions reportedly suppress the non-radiative decay by reducing the defect, whereas the Yb3+ helps the UV photon down-conversion to near-infrared photons, as in the case of CsPbCl3. An R of 0.14 A/W at approximately 300 nm UV is achievable [170]. An R of 0.006 A/W by Dy3+-doped CsPbCl2Br1 [171] and an R of 0.009 (A/W) for Eu3+-doped CsPbCl2Br1 [172] quantum dots on Si under UV 320 nm radiation have also been reported.
Despite the encouraging results mentioned above, the practical implementation of quantum cutting in Si-based photodetectors remains challenging. The optimization of dopant concentration and achieving high efficiency in energy transfer between sensitizers and activators is not straightforward. For quantum dots, their long-term stability and the environmental factors also remain key concerns. The development of more robust quantum cutting materials in the future could pave the way for next-generation and highly sensitive Si-based photodetectors.

5.4. Surface Etching and Plasmonic Surface Nano-Structuring

The UV light sensitivity of Si can be enhanced through surface etching, which produces micro- and nano-scaled texturing, such as pores, cones, and pyramids. Microscale roughness of the Si surface can reduce the surface reflections, thus making absorption more likely to take place. Nanoscale roughness may create quantum confinement effects in Si, with a consequence of widening effective bandgap energies and corresponding enhancement of UV photon absorption [173]. Furthermore, reflected photons from one micro- and nano-scaled structure in a random direction can be absorbed by neighboring micro- and nano-scaled structures, resulting in strong light absorption [174]. However, photogenerated carriers due to such photons absorbed in shallow regions of locally textured structures suffer carrier recombination due to defects before carrier collection at electrodes [175].
A UV photodetector based on surface-etched n-type Si has been reported by Asad et al. [176]. They performed electrochemical etching of the Si surface, where electrical current modulated the etching ratio, under UV light illumination (here, UV light for etching purposes). Then, the etching-generated nano-porous Si was treated with Nd:YAG nanosecond pulsed laser for structural improvement and recrystallization. The photodetector utilized interdigitated Pt electrodes in the MSM structure mentioned above, and the R for detecting 380 nm UV light was estimated as 2.01 (A/W).
Gold (Au) nanoparticles deposited over the porous surface could further increase R due to surface plasmon resonance (SPR) effects. The collective oscillation of conduction electrons on Au nanoparticles can occur coherently with incident light of a specific wavelength [177,178,179]. SPR leads to strong absorption and elastic scattering of light, generating enhanced local fields around the particles [180,181]. By carefully controlling the size and shape of plasmonic nanoparticles, the UV absorption can be enhanced [182]. Ismail et al. [183] recently reported the surface etching followed by the thermal oxidation of Si. Au nanoparticles were immobilized on the etched surface, enabling R of 0.205 (A/W) to be achieved for detecting UV light at 365 nm.
Interestingly, silicon alone can simultaneously act as both a plasmonic resonator and an active absorbing medium, as demonstrated in a recent experiment by Tanaka et al. [184]. They used electron-beam lithography followed by reactive ion etching to carve periodic corrugations with a pitch of 210 nm on an n-type Si substrate, as shown in Figure 13. The corrugated structure of Si enabled the SPR effects to be activated without using noble metals at deep UV wavelengths since Si had the negative values of the relative permittivity at those wavelengths, such as 266 nm. SPR effects produced enhanced local fields near the surface, leading to improved absorption and, consequently, enhanced efficiency in the photogeneration of carriers. The device showed R of 0.18 A/W and 0.3 A/W at 266 nm at bias voltages of 7 V and 10 V, respectively. This demonstration of SPR-induced amplification of the E field with a nanoscale grating offers new routes for CMOS-compatible Si UV photodetectors.

5.5. Wide-Bandgap Semiconductor Integrated Si

Wide-bandgap semiconductors that are suited not only for UV light absorption but also for solar-blind applications can be deposited over the Si surface to construct hybrid heterostructures. Here, wide-bandgap semiconductors serve as a medium where UV photon absorption and subsequent carrier generation take place, whereas the Si, the underlying layer, acts as a carrier capture and transport layer. This hybrid structure can provide the advantages of better selectivity for UV light absorption and effective passivation of Si, which may reduce the carrier recombination rate. Basically, this hybrid structure still possesses advantages of using wide-bandgap materials such as the suppression of dark current due to fewer number of carriers excited at room temperature, availability of high applied bias due to high breakdown voltage, resistance against radiation damage, and chemical/thermal/mechanical robustness in addition to the aforementioned solar blindness and high efficiency of UV photon absorption [134,185,186,187,188,189].
However, it is worthwhile to note that this hybrid structure faces several challenges, including lattice mismatch-induced dislocation of crystals [190,191,192], thermal expansion mismatch [193,194,195], hybrid interface-induced defects [196,197], and the requirement of a (wide-bandgap) crystal growing temperature too high to be compatible with Si-based CMOS processes [198,199]. The crystal dislocations in such a hybrid structure lead to high etch pit density, causing non-radiative recombination centers to be generated for high dark current and degrading photodetectors over time. Thermal expansion mismatch can also develop unwanted strains during crystal growing at high temperatures or annealing, disabling efficient wafer-scale fabrication [200]. To circumvent the aforementioned challenges related to lattice and thermal mismatch, a suite of sophisticated defect and strain engineering strategies is being applied by researchers. One method is strain management through buffer layer engineering on the Si surface, which serves as a foundation for wide-bandgap materials [201,202,203,204]. The buffer layer, which is usually graded, absorbs the stress and manages the mismatch during the growth of wide-bandgap materials by bending and trapping the dislocations. Consequently, the dislocation density in the active area decreases, which leads to improved performance of the device. Another method is interfacial chemical engineering, which involves the creation of a spotless atomic interface by removing the native contamination from the Si surface, such as SiO2, through chemical reactions and replacing it with the desired oxides of the same family of the active material considered for deposition [205,206]. This oxide layer with intermediate lattice parameters acts as a template for the subsequent deposition of wide-bandgap materials.
In addition, the inherent surface properties found in wide-bandgap semiconductors alone, such as surface defects and Fermi energy pinning, may need to be considered when they are brought into contact with Si to avoid substantial surface leakage current due to additional passivation [207,208,209,210,211,212,213,214]. Table 2 displays the performance of UV photodetectors based on Si, including wide-bandgap semiconductor-integrated Si, operated at various UV wavelengths. In some cases, the EQE, which exceeds 100%, indicates photoconductive gain more than unity, such as in avalanche multiplication.
Table 2. Performance of Si-based UV light detectors at various UV wavelengths.
Table 2. Performance of Si-based UV light detectors at various UV wavelengths.
MaterialDevice TypeWavelength nmEQE %Responsivity (A/W)Refs.
Boron implanted n-type black Sip-i-n diode2001000.15[161]
n-type black Sip-i-n diode2001320.213[162]
p-type Sip-n diode206800.13[158]
β-Ga2O3/p-type SiliconMSM Schottky diode230378.50.702[206]
La3+, Yb3+ co-doped CsPbClBr2/Si p-n diode24055.60.13[169]
p-type Si n+–p–n diode2501000.2016[160]
SiC/SiMSM Schottky diode25089.30.18[215]
β-Ga2O3/p-type SiMSM Schottky diode25019,25038.8[216]
β-Ga2O3/AiN/p-type SiMSM Schottky diode250588011.84[216]
SiC/p-type SiMSM Schottky diode/MOSFET2602.2 × 1084.63 × 105[129]
Corrugated n-type Siphotoconductor2661400.3[184]
AlGaN/Sip-i-n diode27454.30.12[201]
La3+, Yb3+ co-doped CsPbCl3/Siphotodiode30057.90.14[170]
Eu3+-doped CsPbCl2Br1/Siphoto diode3203.410.009[172]
Dy3+-doped CsPbCl2Br1/Siphoto diode3202.30.006[171]
GaN/SiAsymmetric MSM diode34021,20058[217]
ZnO/Siheterojunction photoconductor35024,10068[218]
Porous n-type SiMSM diode3806582.01[176]
Au NPs/porous p-type SiMSM diode36569.690.205[183]
Boron-implanted n-type black Sip-i-n diode4001000.3[161]

5.6. Comparative Analysis

The establishment of Si-based stable and standard photodetectors requires careful consideration of the balance between their performance and industrial viability. Among the abovementioned techniques, the surface passivation and external ion implantation techniques are industrially scalable with long-term stability and are compatible with CMOS technology despite the modest gains in responsivity in the UV region. Considering an enhanced performance, wide-bandgap materials integrated with Si offer high responsivities with inherently robust stability. Although their deposition methods are also well-established, their cross-contaminations pose a threat to full-scale CMOS compatibility, as the CMOS foundries require extreme purity. The thermal and lattice mismatches between the wide-bandgap materials and Si could also lead to performance degradation over time, reducing the overall effectiveness of the photodetectors. UV photodetectors based on the quantum cutting mechanism, despite being a novel method of photon down-conversion to measure UV light intensity, also face severe limitations in industrial-scale production and CMOS incompatibility issues due to contamination. The poor stability of the fluorescent materials is still evolving.
Surface etching of Si also boosts its sensitivity in the UV region. However, it calls into question the large-scale uniformity. Large-scale etching has its own scaling issues that introduce non-uniformity in etching across the area, which leads to non-reproducible performances of the final devices. The plasmonic nano-structuring to create a corrugated Si surface through electron beam lithography could be proven useful in full-scale integration into CMOS technology, but with a large amount of energy required for the lithography device. Further research is needed to improve the responsivity to even higher levels.

6. Conclusions

Ultraviolet photodetectors based on wide-bandgap semiconductors have typically been reported due to their many advantages, such as high responsivity, solar blindness, chemical/thermal/mechanical robustness, and good tolerance against radiation damage. However, they are expensive to produce in high crystal quality and incompatible with Si-based CMOS processes, all while quite often having inherently rich surface defect states. Alternatively, the use of Si as an active medium ensures Si-based CMOS compatibility, allowing for the easy integration of Si-based UV detectors with other electronic devices into a chip. The significant absorption coefficients of Si at UV wavelengths can also support the use of Si as an active medium. However, the penetration depth of UV light into the Si surface is notably limited, rendering surface-near photogenerated carriers to encounter the surface trap states. This eventually hinders the effective transport of photogenerated carriers to electrodes. Recently reported advancements in surface passivation, ion implantation, quantum cutting, surface nano-texturing, and hybrid integration with wide-bandgap materials have improved Si photodetectors’ responsivities up to a considerable level. In particular, Si integration with wide-bandgap semiconductors can be regarded as viable for near-future commercialization given the fact that the aforementioned surface treatment/engineering can also be applied to the hybrid structure of Si-wide-bandgap semiconductor to maximize the expected benefits mentioned above.

Author Contributions

Conceptualization, A.K., S.H. and H.J.; investigation, A.K. and H.J.; resources, H.J.; data curation, A.K. and H.J.; writing—original draft, A.K. and H.J.; writing—review and editing, S.H. and H.J.; funding acquisition, S.H. and H.J. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Gachon University research fund of 2025 (GCU-202500560001) and also supported by the National Research Foundation of Korea (NRF) grant funded by the Korean Government (MSIT) (No. RS-2023-00279149).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Blaustein, A.R.; Searle, C. Ultraviolet Radiation. In Encyclopedia of Biodiversity, 2nd ed.; Levin, S.A., Ed.; Academic Press: Waltham, MA, USA, 2013; pp. 296–303. [Google Scholar]
  2. Wood, J.; Monge, M.; Seto, E.P.; Ratliff, K.; Ford, B.; Aslett, D.; Abdel-Hady, A.; Mendez Sandoval, L. Sterilization of Stainless-Steel Surfaces Using Ultraviolet Radiation Produced by Light-Emitting Diodes. Astrobiology 2025, 25, 550–562. [Google Scholar] [CrossRef]
  3. Belay, E.; Abafoge, H.; Ahmed, F.; Alemu, B.; Sisay, S. Design and Development of Automated Ultraviolet (UV-C) Surface Sterilizer and Disinfection Device Using a User-Centered Design Approach. F1000Research 2025, 14, 512. [Google Scholar] [CrossRef]
  4. Sankurantripati, S.; Duchaine, F. Indoor air quality control for airborne diseases: A review on portable UV air purifiers. Fluids 2024, 9, 281. [Google Scholar] [CrossRef]
  5. Haque, S.N.; Bhuyan, M.M.; Jeong, J.-H. Radiation-induced hydrogel for water treatment. Gels 2024, 10, 375. [Google Scholar] [CrossRef] [PubMed]
  6. Yu, Z.-w.; Zheng, M.; Fan, H.-y.; Liang, X.-h.; Tang, Y.-l. Ultraviolet (UV) radiation: A double-edged sword in cancer development and therapy. Mol. Biomed. 2024, 5, 49. [Google Scholar] [CrossRef]
  7. Mittal, A.; Kumar, M.; Gopishankar, N.; Kumar, P.; Verma, A.K. Quantification of narrow band UVB radiation doses in phototherapy using diacetylene based film dosimeters. Sci. Rep. 2021, 11, 684. [Google Scholar] [CrossRef]
  8. Onatayo, D.A.; Srinivasan, R.S.; Shah, B. Ultraviolet radiation transmission in building’s fenestration: Part ii, exploring digital imaging, uv photography, image processing, and computer vision techniques. Buildings 2023, 13, 1922. [Google Scholar] [CrossRef]
  9. Mojeski, J.A.; Almashali, M.; Jowdy, P.; Fitzgerald, M.E.; Brady, K.L.; Zeitouni, N.C.; Colegio, O.R.; Paragh, G. Ultraviolet imaging in dermatology. Photodiagnosis Photodyn. Ther. 2020, 30, 101743. [Google Scholar] [CrossRef]
  10. Harada, K.; Horinouchi, R.; Murakami, M.; Isii, M.; Kamashita, Y.; Shimotahira, N.; Suehiro, F.; Nishi, Y.; Murata, H.; Nishimura, M. The disinfectant effects of portable ultraviolet light devices and their application to dentures. Photodiagnosis Photodyn. Ther. 2025, 51, 104434. [Google Scholar] [CrossRef]
  11. Casini, B.; Scarpaci, M.; Chiovelli, F.; Leonetti, S.; Costa, A.L.; Baroni, M.; Petrillo, M.; Cavallo, F. Antimicrobial efficacy of an experimental UV-C robot in controlled conditions and in a real hospital scenario. J. Hosp. Infect. 2025, 156, 72–77. [Google Scholar] [CrossRef]
  12. Kaushal, H.; Kaddoum, G. Applications of Lasers for Tactical Military Operations. IEEE Access 2017, 5, 20736–20753. [Google Scholar] [CrossRef]
  13. Wang, S.; Li, Z.; Xu, Z.; Yang, Z.; Gong, C.; Xu, Z.; Long, F.; Tang, J.; Chi, N.; Shen, C. A Large FoV UV LED Array Transmitter Enabling Optical Wireless Communication Over 1.1 km. IEEE Photonics Technol. Lett. 2025, 37, 563–566. [Google Scholar] [CrossRef]
  14. Gu, K.; Wu, K.; Zhang, Z.; Ohsawa, T.; Huang, J.; Koide, Y.; Toda, M.; Liao, M. Synergistic Effect of Surface States and Deep Defects for Ultrahigh Gain Deep-Ultraviolet Photodetector with Low-Voltage Operation. Adv. Funct. Mater. 2025, 35, 2420238. [Google Scholar] [CrossRef]
  15. Hou, X.; Liu, Y.; Bai, S.; Yu, S.; Huang, H.; Yang, K.; Li, C.; Peng, Z.; Zhao, X.; Zhou, X.; et al. Pyroelectric Photoconductive Diode for Highly Sensitive and Fast DUV Detection. Adv. Mater. 2024, 36, 2314249. [Google Scholar] [CrossRef]
  16. Wu, L.; Zhang, M.; Shi, X.; Hu, P.; Yu, H.; Teng, F. The Influence of Surface Treatment on the Performance of Photoconductive Detectors Based on Silicon Wafer. IEEE Sens. J. 2024, 24, 20476–20484. [Google Scholar] [CrossRef]
  17. Luo, X.; Zhang, Y.; Liu, L.; Berbille, A.; Wang, K.; Han, G.; Zhu, L.; Wang, Z.L. A self-powered Ag/β-Ga2O3 photodetector with broadband response from 200 to 980 nm based on the photovoltaic and pyro-phototronic effects. J. Mater. Sci. Technol. 2025, 206, 125–134. [Google Scholar] [CrossRef]
  18. Yu, P.; You, S.; Liu, X.; Zhu, Z.-K.; Zeng, Y.; Luo, J. Self-Powered Broadband Photodetection Ranging from X-ray to UV–Vis Light in a Polar Perovskite Induced by Bulk Photovoltaic Effect. J. Phys. Chem. Lett. 2024, 15, 11767–11772. [Google Scholar] [CrossRef] [PubMed]
  19. Khurana, M.; Upasana; Saxena, M.; Gupta, M. TCAD based investigation of junctionless phototransistor for UVC radiation detection. Opt. Laser Technol. 2024, 172, 110486. [Google Scholar] [CrossRef]
  20. Wang, C.; Zilong, W.; Yichao, W.; Cichao, Y.; Qianyi, Z.; Xinru, W.; Yu, G.; Ji, Z. Low voltage flexible high performance organic field-effect transistor and its application for ultraviolet light detectors. Soft Mater. 2024, 22, 183–191. [Google Scholar] [CrossRef]
  21. Li, Q.; Li, X.; Ding, L.; Li, Y.; Ma, J.; Sang, S.; Minami, T. Organic Photodetectors Achieving UV–Visible Broadband Detection via P3HT-OFET Epitaxial Integration. ACS Appl. Electron. Mater. 2025, 7, 6166–6176. [Google Scholar] [CrossRef]
  22. Liu, Y.; Sun, W.; Sa, Z.; Li, R.; Yang, Z.; Jia, Y.; Sun, X.; Chen, F. Ultrafast and Ultra-Broadband Ferroelectric Photo-Pyroelectric Detectors with Long-Term Stability. Adv. Funct. Mater. 2025, 35, 2502467. [Google Scholar] [CrossRef]
  23. Zhao, X.; Weng, W.; Tang, L.; Xu, H.; Ma, Y.; Rong, H.; Ni, H.; Zhang, J.; Luo, J.; Sun, Z. Broadband Pyro-Phototronic Effect in Lead-Free Double Perovskite Crystal Enables UV-to-NIR and Polarization-Sensitive Detection. Chem. Mater. 2025, 37, 3384–3391. [Google Scholar] [CrossRef]
  24. Kuzanyan, A.A.; Nikoghosyan, V.R.; Kuzanyan, A.S. Ultraviolet thermoelectric single photon detector with high signal-to-noise ratio. Opt. Eng. 2024, 63. [Google Scholar] [CrossRef]
  25. Kuzanyan, A.A.; Nikoghosyan, V.R.; Kuzanyan, A.S. Nanoscale thermoelectric detection pixel for single-photon detection from far ultraviolet to near infrared. Opt. Eng. 2024, 63, 067102. [Google Scholar] [CrossRef]
  26. Yeboah, L.A.; Abdul Malik, A.; Oppong, P.A.; Acheampong, P.S.; Morgan, J.A.; Addo, R.A.A.; Williams Henyo, B.; Taylor, S.T.; Zudor, W.M.; Osei-Amponsah, S. Wide-Bandgap Semiconductors: A Critical Analysis of GaN, SiC, AlGaN, Diamond, and Ga2O3 Synthesis Methods, Challenges, and Prospective Technological Innovations. Intell. Sustain. Manuf. 2025, 2, 10011. [Google Scholar] [CrossRef]
  27. Woods-Robinson, R.; Han, Y.; Zhang, H.; Ablekim, T.; Khan, I.; Persson, K.A.; Zakutayev, A. Wide band gap chalcogenide semiconductors. Chem. Rev. 2020, 120, 4007–4055. [Google Scholar] [CrossRef] [PubMed]
  28. Fang, X.; Bando, Y.; Gautam, U.K.; Ye, C.; Golberg, D. Inorganic semiconductor nanostructures and their field-emission applications. J. Mater. Chem. 2008, 18, 509–522. [Google Scholar] [CrossRef]
  29. Sun, Y.; Rogers, J.A. Inorganic semiconductors for flexible electronics. Adv. Mater. 2007, 19, 1897–1916. [Google Scholar] [CrossRef]
  30. Xie, C.; Lu, X.T.; Tong, X.W.; Zhang, Z.X.; Liang, F.X.; Liang, L.; Luo, L.B.; Wu, Y.C. Recent progress in solar-blind deep-ultraviolet photodetectors based on inorganic ultrawide bandgap semiconductors. Adv. Funct. Mater. 2019, 29, 1806006. [Google Scholar] [CrossRef]
  31. Bronstein, H.; Nielsen, C.B.; Schroeder, B.C.; McCulloch, I. The role of chemical design in the performance of organic semiconductors. Nat. Rev. Chem. 2020, 4, 66–77. [Google Scholar] [CrossRef]
  32. Zhang, J.-L.; Nan, Y.-X.; Li, H.-G.; Qiu, W.-M.; Yang, X.; Wu, G.; Chen, H.-Z.; Wang, M. A new wide bandgap organic semiconductor and its application in organic UV sensors with tunable response wavelength. Sens. Actuators B Chem. 2012, 162, 321–326. [Google Scholar] [CrossRef]
  33. Yoshikawa, A.; Matsunami, H.; Nanishi, Y. Development and applications of wide bandgap semiconductors. In Wide Bandgap Semiconductors: Fundamental Properties and Modern Photonic and Electronic Devices; Springer: Berlin/Heidelberg, Germany, 2007; pp. 1–24. [Google Scholar]
  34. He, C.; Liu, X. The rise of halide perovskite semiconductors. Light Sci. Appl. 2023, 12, 15. [Google Scholar] [CrossRef] [PubMed]
  35. Abiram, G.; Thanihaichelvan, M.; Ravirajan, P.; Velauthapillai, D. Review on Perovskite Semiconductor Field–Effect Transistors and Their Applications. Nanomaterials 2022, 12, 2396. [Google Scholar] [CrossRef]
  36. Fox, M.; Ispasoiu, R. Quantum wells, superlattices, and band-gap engineering. In Springer Handbook of Electronic and Photonic Materials; Springer: Cham, Switzerland, 2017; p. 1. [Google Scholar]
  37. Sun, Y.; Wei, Y.; Li, M.; Zhang, Y.; Li, X.; Fan, L.; Li, Y. Wet chemical synthesis of ultrathin γ-Ga2O3 quantum wires enabling far-UVC photodetection with ultrahigh selectivity and sensitivity. J. Phys. Chem. Lett. 2024, 15, 4301–4310. [Google Scholar] [CrossRef]
  38. Yakimenko, I.I.; Yakimenko, I.P. Electronic properties of semiconductor quantum wires for shallow symmetric and asymmetric confinements. J. Phys. Condens. Matter 2021, 34, 105302. [Google Scholar] [CrossRef]
  39. Kamal, A.; Hong, S.; Ju, H. Carbon Quantum Dots: Synthesis, Characteristics, and Quenching as Biocompatible Fluorescent Probes. Biosensors 2025, 15, 99. [Google Scholar] [CrossRef]
  40. Bera, D.; Qian, L.; Tseng, T.-K.; Holloway, P.H. Quantum dots and their multimodal applications: A review. Materials 2010, 3, 2260–2345. [Google Scholar] [CrossRef]
  41. Reimann, S.M.; Manninen, M. Electronic structure of quantum dots. Rev. Mod. Phys. 2002, 74, 1283. [Google Scholar] [CrossRef]
  42. Fan, Q.; Chai, C.; Wei, Q.; Yang, Y. Two novel silicon phases with direct band gaps. Phys. Chem. Chem. Phys. 2016, 18, 12905–12913. [Google Scholar] [CrossRef]
  43. Oliphant, E.; Mantena, V.; Brod, M.; Snyder, G.J.; Sun, W. Why does silicon have an indirect band gap? Mater. Horiz. 2025, 12, 3073–3083. [Google Scholar] [CrossRef]
  44. Schroder, D.K. Carrier lifetimes in silicon. IEEE Trans. Electron Devices 2002, 44, 160–170. [Google Scholar] [CrossRef]
  45. Bullis, W.M.; Huff, H.R. Interpretation of carrier recombination lifetime and diffusion length measurements in silicon. J. Electrochem. Soc. 1996, 143, 1399. [Google Scholar] [CrossRef]
  46. Niewelt, T.; Steinhauser, B.; Richter, A.; Veith-Wolf, B.; Fell, A.; Hammann, B.; Grant, N.E.; Black, L.; Tan, J.; Youssef, A.; et al. Reassessment of the intrinsic bulk recombination in crystalline silicon. Sol. Energy Mater. Sol. Cells 2022, 235, 111467. [Google Scholar] [CrossRef]
  47. Amaral, S.P.; Grosche, L.C.; Sousa, J.P.S. Synergistic UV protection: Hybrid TiO2@SiO2/UVabs nanoparticles with augmented UV-shielding efficiency for the development of durable protective coatings. Prog. Org. Coat. 2025, 205, 109286. [Google Scholar] [CrossRef]
  48. Saravanan, S.; Dubey, R. Synthesis of SiO2Nanoparticles by Sol-Gel Method and Their Optical and Structural Properties. Sci. Technol. 2020, 23, 105–112. [Google Scholar]
  49. Izhaky, N.; Morse, M.T.; Koehl, S.; Cohen, O.; Rubin, D.; Barkai, A.; Sarid, G.; Cohen, R.; Paniccia, M.J. Development of CMOS-Compatible Integrated Silicon Photonics Devices. IEEE J. Sel. Top. Quantum Electron. 2006, 12, 1688–1698. [Google Scholar] [CrossRef]
  50. Kannan, S.; Chang, R.; Potluri, H.; Zhang, S.; Li, J.; Xu, B.; Hsu, H.C. High Density Integration of Silicon Photonic Chiplets for 51.2T Co-packaged Optics. In Proceedings of the 2024 IEEE 74th Electronic Components and Technology Conference (ECTC), Denver, CO, USA, 28–31 May 2024; pp. 81–84. [Google Scholar]
  51. Miskovsky, N.M.; Cutler, P.H.; Mayer, A.; Weiss, B.L.; Willis, B.; Sullivan, T.E.; Lerner, P.B. Nanoscale Devices for Rectification of High Frequency Radiation from the Infrared through the Visible: A New Approach. J. Nanotechnol. 2012, 2012, 512379. [Google Scholar] [CrossRef]
  52. Simone, G.; Dyson, M.J.; Meskers, S.C.J.; Janssen, R.A.J.; Gelinck, G.H. Organic Photodetectors and their Application in Large Area and Flexible Image Sensors: The Role of Dark Current. Adv. Funct. Mater. 2020, 30, 1904205. [Google Scholar] [CrossRef]
  53. Nodari, D.; Hart, L.J.F.; Sandberg, O.J.; Furlan, F.; Angela, E.; Panidi, J.; Qiao, Z.; McLachlan, M.A.; Barnes, P.R.F.; Durrant, J.R.; et al. Dark Current in Broadband Perovskite–Organic Heterojunction Photodetectors Controlled by Interfacial Energy Band Offset. Adv. Mater. 2024, 36, 2401206. [Google Scholar] [CrossRef]
  54. Alkauskas, A.; Dreyer, C.E.; Lyons, J.L.; Van de Walle, C.G. Role of excited states in Shockley-Read-Hall recombination in wide-band-gap semiconductors. Phys. Rev. B 2016, 93, 201304. [Google Scholar] [CrossRef]
  55. Höhr, T.; Schenk, A.; Fichtner, W. Revised Shockley–Read–Hall lifetimes for quantum transport modeling. J. Appl. Phys. 2004, 95, 4875–4882. [Google Scholar] [CrossRef]
  56. Mandurrino, M.; Goano, M.; Vallone, M.; Bertazzi, F.; Ghione, G.; Verzellesi, G.; Meneghini, M.; Meneghesso, G.; Zanoni, E. Semiclassical simulation of trap-assisted tunneling in GaN-based light-emitting diodes. J. Comput. Electron. 2015, 14, 444–455. [Google Scholar] [CrossRef]
  57. Labed, M.; Kim, K.; Kim, K.H.; Hong, J.; Rim, Y.S. Trap-assisted tunneling in type II Ag2O/β-Ga2O3 self-powered solar blind photodetector. Sens. Actuators A Phys. 2024, 372, 115368. [Google Scholar] [CrossRef]
  58. Fregolent, M.; Piva, F.; Buffolo, M.; De Santi, C.; Cester, A.; Higashiwaki, M.; Meneghesso, G.; Zanoni, E.; Meneghini, M. Advanced defect spectroscopy in wide-bandgap semiconductors: Review and recent results. J. Phys. D Appl. Phys. 2024, 57, 433002. [Google Scholar] [CrossRef]
  59. Zhou, X.; Jiang, D.; Yang, X.; Duan, Y.; Zhang, W.; Zhao, M.; Liang, Q.; Gao, S.; Hou, J.; Zheng, T. Voltage-dependent responsivity of ZnO Schottky UV photodetectors with different electrode spacings. Sens. Actuators A Phys. 2018, 284, 12–16. [Google Scholar] [CrossRef]
  60. Goushcha, A.; Tabbert, B. On response time of semiconductor photodiodes. Opt. Eng. 2017, 56, 097101. [Google Scholar] [CrossRef]
  61. Zeng, Z.; Wang, Y.; Michel, P.; Strauß, F.; Wang, X.; Braun, K.; Scheele, M. Ultrafast Hot Carrier Cooling Enabled van der Waals Photodetectors at Telecom Wavelengths. Nano Lett. 2025, 25, 3497–3504. [Google Scholar] [CrossRef] [PubMed]
  62. Kaur, A.; Sahu, B.P.; Biswas, A.; Dhar, S. p-(001)NiO/n-(0001)ZnO heterojunction based ultraviolet photodetectors with controllable response time. Semicond. Sci. Technol. 2025, 40, 045011. [Google Scholar] [CrossRef]
  63. Hayashi, T. New photomultiplier tubes for medical imaging. IEEE Trans. Nucl. Sci. 1989, 36, 1078–1083. [Google Scholar] [CrossRef]
  64. Kume, H.; Suzuki, S.; Oba, K. Recent Development of Photomultiplier Tubes for Nuclear and Medical Applications. IEEE Trans. Nucl. Sci. 1985, 32, 355–359. [Google Scholar] [CrossRef]
  65. Suzuki, S.; Matsushita, T.; Suzuki, T.; Kimura, S.; Kume, H. New position sensitive photomultiplier tubes for high energy physics and nuclear medical applications. IEEE Trans. Nucl. Sci. 1988, 35, 382–386. [Google Scholar] [CrossRef]
  66. Rai, R.; Singh, B. Optical and structural properties of CsI thin film photocathode. Nucl. Instrum. Methods Phys. Res. Sect. A Accel. Spectrometers Detect. Assoc. Equip. 2015, 785, 70–76. [Google Scholar]
  67. Simons, D.; Fraser, G.; De Korte, P.; Pearson, J.; De Jong, L. UV and XUV quantum detection efficiencies of CsI-coated microchannel plates. Nucl. Instrum. Methods Phys. Res. Sect. A Accel. Spectrometers Detect. Assoc. Equip. 1987, 261, 579–586. [Google Scholar] [CrossRef]
  68. Yusof, Z.; Denchfield, A.; Warren, M.; Cardenas, J.; Samuelson, N.; Spentzouris, L.; Power, J.; Zasadzinski, J. Photocathode quantum efficiency of ultrathin Cs2Te layers on Nb substrates. Phys. Rev. Accel. Beams 2017, 20, 123401. [Google Scholar] [CrossRef]
  69. Moorhead, M.E.; Tanner, N.W. Optical properties of an EMI K2CsSb bialkali photocathode. Nucl. Instrum. Methods Phys. Res. Sect. A Accel. Spectrometers Detect. Assoc. Equip. 1996, 378, 162–170. [Google Scholar] [CrossRef]
  70. Sun, J.; Jin, M.; Wang, X.; Si, S.; Ren, L.; Hou, W.; Qiao, F.; Zhao, M.; Gu, Y.; Huang, G.; et al. Enhanced photoemission capability of bialkali photocathodes for 20-inch photomultiplier tubes. Nucl. Instrum. Methods Phys. Res. Sect. A Accel. Spectrometers Detect. Assoc. Equip. 2020, 971, 164021. [Google Scholar] [CrossRef]
  71. Mansmann, R.; Sipkens, T.A.; Menser, J.; Daun, K.J.; Dreier, T.; Schulz, C. Detector calibration and measurement issues in multi-color time-resolved laser-induced incandescence. Appl. Phys. B 2019, 125, 126. [Google Scholar] [CrossRef]
  72. Bie, Y.Q.; Liao, Z.M.; Zhang, H.Z.; Li, G.R.; Ye, Y.; Zhou, Y.B.; Xu, J.; Qin, Z.X.; Dai, L.; Yu, D.P. Self-powered, ultrafast, visible-blind UV detection and optical logical operation based on ZnO/GaN nanoscale p-n junctions. Adv. Mater. 2011, 23, 649–653. [Google Scholar] [CrossRef]
  73. Monroy, E.; Munoz, E.; Sánchez, F.; Calle, F.; Calleja, E.; Beaumont, B.; Gibart, P.; Muñoz, J.; Cussó, F. High-performance GaN pn junction photodetectors for solar ultraviolet applications. Semicond. Sci. Technol. 1998, 13, 1042. [Google Scholar] [CrossRef]
  74. Shi, Z.; Gao, X.; Xie, M.; Xie, T.; Li, Z.; Zhao, H.; Wang, Y.; Gao, X. Dual-Sided Multiband Ultraviolet Communication System Based on Ga2O3/GaN PN Junction Photodetectors. ACS Appl. Mater. Interfaces 2025, 17, 11325–11333. [Google Scholar] [CrossRef]
  75. Wang, J.; Li, Q.; Mi, W.; Wang, D.; Xu, M.; Xiao, L.; Zhang, X.; Luan, C.; Zhao, J. Fast response self-powered solar-blind UV photodetector based on NiO/Ga2O3 pn junction. Mater. Sci. Semicond. Process. 2025, 186, 109084. [Google Scholar] [CrossRef]
  76. Wei, S.; Chen, B.; Zhu, J.; Wu, X.; Zhang, C.; Zhang, D.W.; Sun, Q.-Q.; Ji, L.; Hu, S. Self-Powered Broad-Spectrum UV–vis–NIR Photodetector Based on p-Si/Zn–Sn–Al–O Heterojunction. ACS Appl. Electron. Mater. 2025, 7, 5069–5079. [Google Scholar] [CrossRef]
  77. Padha, B.; Ahmed, Z.; Dutta, S.; Pandey, A.; Padha, N.; Yadav, I.; Arya, S. Transient response of low-temperature ALD-grown ZnO thin film-based p–i–n UV photodetector. Opt. Mater. 2025, 167, 117264. [Google Scholar] [CrossRef]
  78. Wang, Y.; Li, S.; Cao, J.; Jiang, Y.; Zhang, Y.; Tang, W.; Wu, Z. Improved response speed of β-Ga2O3 solar-blind photodetectors by optimizing illumination and bias. Mater. Des. 2022, 221, 110917. [Google Scholar] [CrossRef]
  79. Zhang, R.; Wang, G.; Zhang, Q.; Wang, S.; Hu, X.; Liu, L.; Lv, S.; Chen, W.; Xu, X.; Zhang, L. Recent progress in GaN-based ultraviolet photodetectors. J. Mater. Chem. C 2025, 13, 10972–10996. [Google Scholar] [CrossRef]
  80. Vashishtha, P.; Verma, A.K.; Walia, S.; Gupta, G. A solar-blind ultraviolet photodetector with self-biasing capability, controlled by surface potential based on GaN hexagonal nano-spikes. Mater. Lett. 2024, 368, 136708. [Google Scholar] [CrossRef]
  81. Maraj, M.; Yaoze, L.; Sun, W. Overview of the structural effects on the performance of AlGaN solar-blind UV detectors. J. Lumin. 2025, 121178. [Google Scholar] [CrossRef]
  82. Fu, Y.; Liu, B.; Zhan, J.; Zheng, F.; Sun, Z. Characterization of high-performance AlGaN-based solar-blind UV photodetectors. Optoelectron. Lett. 2025, 21, 402–406. [Google Scholar] [CrossRef]
  83. De Napoli, M. SiC detectors: A review on the use of silicon carbide as radiation detection material. Front. Phys. 2022, 10, 898833. [Google Scholar] [CrossRef]
  84. Pacheco, E.; Zhou, B.; Aldalbahi, A.; Zhou, A.F.; Feng, P.X. Zero-biased and visible-blind UV photodetectors based on nitrogen-doped ultrananocrystalline diamond nanowires. Ceram. Int. 2022, 48, 3757–3761. [Google Scholar] [CrossRef]
  85. Angelone, M.; Bombarda, F.; Cesaroni, S.; Marinelli, M.; Raso, A.M.; Verona, C.; Verona-Rinati, G. X-Ray and UV Detection Using Synthetic Single Crystal Diamond. Instruments 2025, 9, 9. [Google Scholar] [CrossRef]
  86. Nguyen, T.M.H.; Shin, S.G.; Choi, H.W.; Bark, C.W. Recent advances in self-powered and flexible UVC photodetectors. Exploration 2022, 2, 20210078. [Google Scholar] [CrossRef]
  87. Yin, G.; Bi, K.; Guo, M.; Li, C.; Huang, D.; Liu, H.; Liu, L.; Xiong, W.; Chen, J.; Wang, H. Graphene Ultraviolet Photoconductive Detector Enhanced by Plasmon Local Field with Sea-Urchin-Like Gold Structure. Plasmonics 2025, 20, 1–10. [Google Scholar] [CrossRef]
  88. Cadatal-Raduban, M.; Kato, T.; Horiuchi, Y.; Olejníček, J.; Kohout, M.; Yamanoi, K.; Ono, S. Effect of substrate and thickness on the photoconductivity of nanoparticle titanium dioxide thin film vacuum ultraviolet photoconductive detector. Nanomaterials 2021, 12, 10. [Google Scholar] [CrossRef]
  89. Whitfield, M.D.; McKeag, R.D.; Pang, L.Y.; Chan, S.S.; Jackman, R.B. Thin film diamond UV photodetectors: Photodiodes compared with photoconductive devices for highly selective wavelength response. Diam. Relat. Mater. 1996, 5, 829–834. [Google Scholar] [CrossRef]
  90. Kim, H.; Rothschild, M.; Lee, D.H.; Kim, C.S.; Park, J.; Min, B.-C.; Lee, S. Bias-Switchable Photodetector from Broad-Band to UV-Selective Detection Mode Leveraging Nanolayered Dual-Schottky Junction. ACS Appl. Nano Mater. 2022, 5, 17891–17899. [Google Scholar] [CrossRef]
  91. Basyooni, M.; Kabatas, M.A.; En-Nadir, R.; Rahmani, K.; Eker, Y.R. Positive and Negative Photoconductivity in Ir Nanofilm-Coated MoO3 Bias-Switching Photodetector. Micromachines 2023, 14, 1860. [Google Scholar] [CrossRef]
  92. Xu, Z.-Q.; Deng, H.; Xie, J.; Li, Y.; Zu, X.-T. Ultraviolet photoconductive detector based on Al doped ZnO films prepared by sol–gel method. Appl. Surf. Sci. 2006, 253, 476–479. [Google Scholar] [CrossRef]
  93. Wang, L.W.; Wu, T.Y.; Chu, S.Y. Metal-Free All Transparent Zinc Oxide-Based Ultraviolet Photodetectors With Simple Bilayer Structure. IEEE Sens. J. 2024, 24, 22344–22350. [Google Scholar] [CrossRef]
  94. Huang, M.; Yang, B.; Xu, G.; Lu, J.; Guo, Y.; Weng, L.; Sun, D. Enhancement of high-temperature solar-blind UV photodetector performance on 4H–SiC substrates by deposition of ultrawide bandgap diamond films. Carbon 2025, 237, 120130. [Google Scholar] [CrossRef]
  95. Yang, W. AlGaN UV Photodetectors. In III-V Nitride Semiconductors; CRC Press: Boca Raton, FL, USA, 2022; pp. 675–691. [Google Scholar]
  96. Nie, Y.; Jiao, S.; Yang, S.; Zhao, Y.; Gao, S.; Wang, D.; Yang, X.; Li, Y.; Fu, Z.; Li, A.; et al. Achieving Ultra-Low Dark Current in β-Ga2O3 Photoconductive Photodetectors for Anti-Interference Optical Human–Machine Interaction Systems via Gallium Interstitials Engineering. Small 2025, 21, 2501442. [Google Scholar] [CrossRef]
  97. Hillebrand, M.; Blecher, F.; Sterzel, J.; Böhm, M. An Amorphous Silicon Photoconductor for UV Detection. MRS Proc. 2003, 762, A18.15. [Google Scholar] [CrossRef]
  98. Hou, Y.; Mei, Z.; Du, X. Semiconductor ultraviolet photodetectors based on ZnO and MgxZn1−xO. J. Phys. D Appl. Phys. 2014, 47, 283001. [Google Scholar] [CrossRef]
  99. Dacey, G.C.; Ross, I.M. The Field Effect Transistor. Bell Syst. Tech. J. 1955, 34, 1149–1189. [Google Scholar] [CrossRef]
  100. Shin, J.; Yoo, H. Photogating Effect-Driven Photodetectors and Their Emerging Applications. Nanomaterials 2023, 13, 882. [Google Scholar] [CrossRef]
  101. Sahni, S.; Luo, X.; Liu, J.; Xie, Y.-h.; Yablonovitch, E. Junction field-effect-transistor-based germanium photodetector on silicon-on-insulator. Opt. Lett. 2008, 33, 1138–1140. [Google Scholar] [CrossRef] [PubMed]
  102. Li, Y.; Zhao, Y.; Ruocco, A.; Wang, M.; Li, B.; Akhavan, S. Printed Lithography of Graphene-Perovskite Quantum Dot Hybrid Photodetectors on Paper Substrates. ACS Appl. Mater. Interfaces 2025, 17, 6716–6727. [Google Scholar] [CrossRef] [PubMed]
  103. Fu, J.; Nie, C.; Sun, F.; Li, G.; Wei, X. Photodetectors based on graphene–semiconductor hybrid structures: Recent progress and future outlook. Adv. Devices Instrum. 2023, 4, 0031. [Google Scholar] [CrossRef]
  104. Guo, M.; Li, J.; Zhang, X.; Bai, G.; Chen, X.; Teng, X.; Lou, Z.; Hou, Y.; Teng, F.; Hu, Y. High-Performance CsPbBr3 Quantum Dot/ZTO Heterojunction Phototransistor with Enhanced Stability and Responsivity. J. Phys. Chem. Lett. 2025, 16, 1634–1643. [Google Scholar] [CrossRef] [PubMed]
  105. Xie, C.; Liu, C.-K.; Loi, H.-L.; Yan, F. Perovskite-Based Phototransistors and Hybrid Photodetectors. Adv. Funct. Mater. 2020, 30, 1903907. [Google Scholar] [CrossRef]
  106. Zheng, J.; Wang, X.; Liu, B.; Wu, R.; Zhao, J.; Liu, Y.; Hua, X.; Yang, Y.; Hao, Z.; Guo, A. Advances in Photovoltaic Detectors: Principles, Challenges, and the Role of Avalanche Detectors in High-Sensitivity Detection. In Light-Driven Materials and Devices: Fundamentals and Emerging Applications; IntechOpen Limited: London, UK, 2025. [Google Scholar]
  107. Su, L.; Xu, W.; Zhou, D.; Ren, F.; Chen, D.; Zhang, R.; Zheng, Y.; Lu, H. Avalanche mechanism analysis of 4H-SiC nip and pin avalanche photodiodes working in Geiger mode. Chin. Opt. Lett. 2021, 19, 092501. [Google Scholar] [CrossRef]
  108. Jeong, H.; Garzda, E.; Ji, M.-H.; Cho, M.; Detchprohm, T.; Shen, S.-C.; Otte, A.; Dupuis, R.D. Low-temperature geiger-mode characterization of a gallium nitride pin avalanche photodiode. IEEE J. Quantum Electron. 2023, 59, 1–8. [Google Scholar] [CrossRef]
  109. Gautam, L.; Jaud, A.G.; Lee, J.; Brown, G.J.; Razeghi, M. Geiger-mode operation of AlGaN avalanche photodiodes at 255 nm. IEEE J. Quantum Electron. 2021, 57, 1–6. [Google Scholar] [CrossRef]
  110. Liu, Q.; Xu, L.; Jin, Y.; Zhang, S.; Wang, Y.; Hu, A.; Guo, X. Ultraviolet response in coplanar silicon avalanche photodiodes with CMOS compatibility. Sensors 2022, 22, 3873. [Google Scholar] [CrossRef]
  111. Vinogradov, S.; Popova, E.; Schmailzl, W.; Engelmann, E. Tip Avalanche Photodiode–A New Wide Spectral Range Silicon Photomultiplier. In Radiation Detection Systems; CRC Press: Boca Raton, FL, USA, 2021; pp. 257–288. [Google Scholar]
  112. Özgür, Ü.; Alivov, Y.I.; Liu, C.; Teke, A.; Reshchikov, M.A.; Doğan, S.; Avrutin, V.; Cho, S.-J.; Morkoç. A comprehensive review of ZnO materials and devices. J. Appl. Phys. 2005, 98. [Google Scholar] [CrossRef]
  113. Park, S.-H.; Kim, J.-J.; Kim, H.-M. Exciton binding energy in wurtzite InGaN/GaN quantum wells. J. Korean Phys. Soc. 2004, 45. [Google Scholar]
  114. Cretì, A.; Tobaldi, D.M.; Lomascolo, M.; Tarantini, I.; Esposito, M.; Passaseo, A.; Tasco, V. Exciton effects in low-barrier GaN/AlGaN quantum wells. J. Phys. Chem. C 2022, 126, 14727–14734. [Google Scholar] [CrossRef]
  115. Grunert, M.; Großmann, M.; Runge, E. Predicting exciton binding energies from ground-state properties. Phys. Rev. B 2024, 110, 075204. [Google Scholar] [CrossRef]
  116. Li, J.; Nam, K.B.; Nakarmi, M.L.; Lin, J.Y.; Jiang, H.X.; Carrier, P.; Wei, S.-H. Band structure and fundamental optical transitions in wurtzite AlN. Appl. Phys. Lett. 2003, 83, 5163–5165. [Google Scholar] [CrossRef]
  117. Dean, P.J.; Lightowlers, E.C.; Wight, D.R. Intrinsic and Extrinsic Recombination Radiation from Natural and Synthetic Aluminum-Doped Diamond. Phys. Rev. 1965, 140, A352–A368. [Google Scholar] [CrossRef]
  118. Clark, C.D.; Dean, P.J.; Harris, P.V. Intrinsic edge absorption in diamond. Proc. R. Soc. London. Ser. A Math. Phys. Sci. 1964, 277, 312–329. [Google Scholar] [CrossRef]
  119. Mnatsakanov, T.; Levinshtein, M.; Pomortseva, L.; Yurkov, S. Carrier mobility model for simulation of SiC-based electronic devices. Semicond. Sci. Technol. 2002, 17, 974. [Google Scholar] [CrossRef]
  120. Ayalew, T. SiC Semiconductor Devices Technology, Modeling and Simulation. Ph.D. Thesis, Technische Universität Wien, Vienna, Austria, 2004. [Google Scholar]
  121. Schwierz, F. An electron mobility model for wurtzite GaN. Solid-State Electron. 2005, 49, 889–895. [Google Scholar] [CrossRef]
  122. Mnatsakanov, T.T.; Levinshtein, M.E.; Pomortseva, L.I.; Yurkov, S.N.; Simin, G.S.; Asif Khan, M. Carrier mobility model for GaN. Solid-State Electron. 2003, 47, 111–115. [Google Scholar] [CrossRef]
  123. Liu, J.; Yu, H.; Shao, S.; Tu, J.; Zhu, X.; Yuan, X.; Wei, J.; Chen, L.; Ye, H.; Li, C. Carrier mobility enhancement on the H-terminated diamond surface. Diam. Relat. Mater. 2020, 104, 107750. [Google Scholar] [CrossRef]
  124. Looi, H.J.; Jackman, R.B.; Foord, J.S. High carrier mobility in polycrystalline thin film diamond. Appl. Phys. Lett. 1998, 72, 353–355. [Google Scholar] [CrossRef]
  125. Li, Z.; Chen, L.; Liu, C.; Tai, J.; Zhao, Y.; Yin, H. High carrier mobility of diamond (100) enabled by surface modification using doped hexagonal boron nitride. Appl. Phys. Lett. 2025, 126. [Google Scholar] [CrossRef]
  126. Abutawahina, M.S.; Abbas, A.A.; Hamzah, N.; Ng, S.; Quah, H.; Ahmed, N.M.; Shaveisi, M. Experimental and simulation analyses of bulk GaN-based metal-semiconductor-metal ultraviolet photodetectors. Mater. Sci. Eng. B 2025, 319, 118323. [Google Scholar] [CrossRef]
  127. Choubey, B.; Ghosh, K. Highly Sensitive and Fast Response AlGaN-Based MSM Solar-Blind Photodetector. IEEE Trans. Electron Devices 2025, 72, 4204–4210. [Google Scholar] [CrossRef]
  128. Thalayarathna, D.; Samintha, M.; Attygalle, D.; Amarasinghe, D.S. Capacitive Response Study of ZnO Based Metal-Semiconductor-Metal Ultraviolet Photodetector. In Proceedings of the 2024 Moratuwa Engineering Research Conference (MERCon), Moratuwa, Sri Lanka, 8–10 August 2024; pp. 472–476. [Google Scholar]
  129. Zhang, W.; Chen, Y.; Yang, J.; Feng, S.; Li, B.; Lu, W.; Liu, F.; Wan, J. A Novel Ultraviolet Photodetector With High Responsivity and Low Operating Voltage Based on Hybrid Si/SiC Technology. IEEE Trans. Electron Devices 2025, 72, 2411–2416. [Google Scholar] [CrossRef]
  130. Hu, Y.; Liu, C.; Cao, G.; Zhang, X.; Chen, Y.; Dai, H.; He, X.; Long, H.; Yu, S.; Xu, X. Mpcvd Growth of Single Crystal Diamond (001) Films For Fast Speed Solar-Blind Uv Photodetectors. SSRN, 2024; preprint. [Google Scholar]
  131. Das, N.; Fadakar, F. Application of Metal-Semiconductor-Metal Photodetector in High-Speed Optical Communication Systems. In Advances in Optical Communication; Das, N., Ed.; IntechOpen: Rijeka, Croatia, 2014. [Google Scholar] [CrossRef]
  132. Kim, M.; Seo, J.-H.; Singisetti, U.; Ma, Z. Recent advances in free-standing single crystalline wide band-gap semiconductors and their applications: GaN, SiC, ZnO, β-Ga2O3, and diamond. J. Mater. Chem. C 2017, 5, 8338–8354. [Google Scholar] [CrossRef]
  133. Liu, C.J.; Peng, T.; Wang, B.; Guo, Y.; Lou, Y.; Zhao, N.; Wang, W.; Chen, X.L. Progress in single crystal growth of wide bandgap semiconductor SiC. In Proceedings of Materials Science Forum; Trans Tech Publications Ltd.: Bäch, Switzerland, 2019; pp. 35–45. [Google Scholar]
  134. Yuvaraja, S.; Khandelwal, V.; Tang, X.; Li, X. Wide bandgap semiconductor-based integrated circuits. Chip 2023, 2, 100072. [Google Scholar] [CrossRef]
  135. Zheng, W.; Huang, F.; Zheng, R.; Wu, H. Low-Dimensional Structure Vacuum-Ultraviolet-Sensitive (λ < 200 nm) Photodetector with Fast-Response Speed Based on High-Quality AlN Micro/Nanowire. Adv. Mater. 2015, 27, 3921–3927. [Google Scholar] [CrossRef] [PubMed]
  136. Liu, X.; Wu, T.; Zhao, J.; Zhu, J.; Chen, X.; Yu, H.; Gao, Y.; Zhou, J.; Chen, Z. High-Sensitivity Amorphous Boron Nitride Vacuum Ultraviolet Photodetectors. IEEE Electron Device Lett. 2025, 46, 76–79. [Google Scholar] [CrossRef]
  137. Qiu, M.; Jia, Z.; Yang, M.; Li, M.; Shen, Y.; Liu, C.; Nishimura, K.; Jiang, N.; Wang, B.; Lin, C.-T.; et al. High-performance single crystal diamond pixel photodetector with nanosecond rise time for solar-blind imaging. Diam. Relat. Mater. 2024, 144, 110996. [Google Scholar] [CrossRef]
  138. Seo, M.; Han, C.; Kim, Y.; Son, J.; Heo, J. Large-area α-Ga2O3 Metal–Semiconductor–Metal Photodetectors for Detecting Ultra-low UV-C Light amid Sunlight in Arc Monitoring. IEEE Sens. J. 2025, 25, 34549–34557. [Google Scholar] [CrossRef]
  139. Yang, Y.; Han, D.; Wu, S.; Lin, H.; Zhang, J.; Zhang, W.; Ye, J. High-performance solar-blind ultraviolet photodetector arrays based on two-inch ϵ-Ga2O3 films for imaging applications. J. Phys. D Appl. Phys. 2025, 58, 105108. [Google Scholar] [CrossRef]
  140. Sun, J.; Zhang, S.; Zhan, T.; Liu, Z.; Wang, J.; Yi, X.; Li, J.; Sarro, P.M.; Zhang, G. A high responsivity and controllable recovery ultraviolet detector based on a WO3 gate AlGaN/GaN heterostructure with an integrated micro-heater. J. Mater. Chem. C 2020, 8, 5409–5416. [Google Scholar] [CrossRef]
  141. Yoshikawa, A.; Ushida, S.; Nagase, K.; Iwaya, M.; Takeuchi, T.; Kamiyama, S.; Akasaki, I. High-performance solar-blind Al0.6Ga0.4N/Al0.5Ga0.5N MSM type photodetector. Appl. Phys. Lett. 2017, 111, 191103. [Google Scholar] [CrossRef]
  142. Abdulrahman, A.F.; Abd-Alghafour, N.M.; Almessiere, M.A. A high responsivity, fast response time of ZnO nanorods UV photodetector with annealing time process. Opt. Mater. 2023, 141, 113869. [Google Scholar] [CrossRef]
  143. Xue, H.; Kong, X.; Liu, Z.; Liu, C.; Zhou, J.; Chen, W.; Ruan, S.; Xu, Q. TiO2 based metal-semiconductor-metal ultraviolet photodetectors. Appl. Phys. Lett. 2007, 90, 201118. [Google Scholar] [CrossRef]
  144. Zhang, H.; Liang, F.; Song, K.; Xing, C.; Wang, D.; Yu, H.; Huang, C.; Sun, Y.; Yang, L.; Zhao, X.; et al. Demonstration of AlGaN/GaN-based ultraviolet phototransistor with a record high responsivity over 3.6 × 107 A/W. Appl. Phys. Lett. 2021, 118, 242105. [Google Scholar] [CrossRef]
  145. Wang, X.; Tang, C.; Li, R.; Zhu, X.; Jin, W.; Hao, L.; Sun, R.; Liu, J.; Ma, Y.; Ma, L. High-Responsivity UV Photodetectors Based on High-Mobility Graphene Epitaxially Grown on 4H-SiC. ACS Appl. Nano Mater. 2025, 8, 15799–15807. [Google Scholar] [CrossRef]
  146. Fu, R.; Jiang, X.; Wang, Y.; Xia, D.; Li, B.; Ma, J.; Xu, H.; Shen, A.; Liu, Y. A high responsivity, high detectivity, and high response speed MSM UVB photodetector based on SnO2 microwires. Nanoscale 2023, 15, 7460–7465. [Google Scholar] [CrossRef] [PubMed]
  147. Liu, J.S.; Shan, C.X.; Li, B.H.; Zhang, Z.Z.; Yang, C.L.; Shen, D.Z.; Fan, X.W. High responsivity ultraviolet photodetector realized via a carrier-trapping process. Appl. Phys. Lett. 2010, 97. [Google Scholar] [CrossRef]
  148. Liu, L.; Yang, C.; Patanè, A.; Yu, Z.; Yan, F.; Wang, K.; Lu, H.; Li, J.; Zhao, L. High-detectivity ultraviolet photodetectors based on laterally mesoporous GaN. Nanoscale 2017, 9, 8142–8148. [Google Scholar] [CrossRef]
  149. Wu, X.; Sun, J.; Shao, H.; Zhai, Y.; Li, L.; Chen, W.; Zhu, J.; Dong, B.; Xu, L.; Zhou, D.; et al. Self-powered UV photodetectors based on CsPbCl3 nanowires enabled by the synergistic effect of acetate and lanthanide ion passivation. Chem. Eng. J. 2021, 426, 131310. [Google Scholar] [CrossRef]
  150. Huang, D.; Zhao, X.; Guo, S.; Fu, X.; Cui, P.; Ye, S.; Yu, Z.; He, Y. High Responsivity Analysis of 4H-SiC Phototransistor. IEEE Sens. Lett. 2025, 9, 1–4. [Google Scholar] [CrossRef]
  151. Wang, Y.J.; Chen, J.W.; Hu, T.G.; Huang, Y.Q.; Zhu, W.K.; Li, W.H.; Hu, Y.; Wei, Z.M.; Fan, Z.C.; Zhao, L.X.; et al. High-Performance Ultraviolet Photodetector Based on the Vertical GaSe/GaN Heterojunction. Small 2025, 21, 2407473. [Google Scholar] [CrossRef]
  152. Zheng, L.; Yang, Y.; Bowen, C.R.; Jiang, L.; Shu, Z.; He, Y.; Yang, H.; Xie, Z.; Lu, T.; Hu, F.; et al. A high-performance UV photodetector with superior responsivity enabled by a synergistic photo/thermal enhancement of localized surface plasmon resonance. J. Mater. Chem. C 2023, 11, 6227–6238. [Google Scholar] [CrossRef]
  153. Huang, X.; Han, S.; Huang, W.; Liu, X. Enhancing solar cell efficiency: The search for luminescent materials as spectral converters. Chem. Soc. Rev. 2013, 42, 173–201. [Google Scholar] [CrossRef]
  154. Canfield, L.R.; Kerner, J.; Korde, R. Stability and quantum efficiency performance of silicon photodiode detectors in the far ultraviolet. Appl. Opt. 1989, 28, 3940–3943. [Google Scholar] [CrossRef] [PubMed]
  155. Tosaka, A.; Nishiguchi, T.; Nonaka, H.; Ichimura, S. Low-Temperature Oxidation of Silicon using UV-Light-Excited Ozone. Jpn. J. Appl. Phys. 2005, 44, L1144. [Google Scholar] [CrossRef]
  156. Hoex, B.; Gielis, J.J.H.; van de Sanden, M.C.M.; Kessels, W.M.M. On the c-Si surface passivation mechanism by the negative-charge-dielectric Al2O3. J. Appl. Phys. 2008, 104. [Google Scholar] [CrossRef]
  157. Hoex, B.; Schmidt, J.; Pohl, P.; van de Sanden, M.C.M.; Kessels, W.M.M. Silicon surface passivation by atomic layer deposited Al2O3. J. Appl. Phys. 2008, 104. [Google Scholar] [CrossRef]
  158. Nikzad, S.; Hoenk, M.; Jewell, A.D.; Hennessy, J.J.; Carver, A.G.; Jones, T.J.; Goodsall, T.M.; Hamden, E.T.; Suvarna, P.; Bulmer, J.; et al. Single Photon Counting UV Solar-Blind Detectors Using Silicon and III-Nitride Materials. Sensors 2016, 16, 927. [Google Scholar] [CrossRef]
  159. Current, M.I. Ion implantation of advanced silicon devices: Past, present and future. Mater. Sci. Semicond. Process. 2017, 62, 13–22. [Google Scholar] [CrossRef]
  160. Kuroda, R.; Nakazawa, T.; Hanzawa, K.; Sugawa, S. Highly ultraviolet light sensitive and highly reliable photodiode with atomically flat Si surface. In Proceedings of the International Image Sensor Workshop, Nanae, Japan, 8–11 June 2011; pp. 38–41. [Google Scholar]
  161. Setälä, O.E.; Chen, K.; Pasanen, T.P.; Liu, X.; Radfar, B.; Vähänissi, V.; Savin, H. Boron-Implanted Black Silicon Photodiode with Close-to-Ideal Responsivity from 200 to 1000 nm. ACS Photonics 2023, 10, 1735–1741. [Google Scholar] [CrossRef] [PubMed]
  162. Garin, M.; Heinonen, J.; Werner, L.; Pasanen, T.P.; Vähänissi, V.; Haarahiltunen, A.; Juntunen, M.A.; Savin, H. Black-Silicon Ultraviolet Photodiodes Achieve External Quantum Efficiency above 130. Phys. Rev. Lett. 2020, 125, 117702. [Google Scholar] [CrossRef]
  163. Vergeer, P.; Vlugt, T.; Kox, M.; Den Hertog, M.; Van der Eerden, J.; Meijerink, A. Quantum cutting by cooperative energy transfer in YbxY1−xPO4:Tb3+. Phys. Rev. B—Condens. Matter Mater. Phys. 2005, 71, 014119. [Google Scholar] [CrossRef]
  164. Chen, G.; Yang, C.; Prasad, P.N. Nanophotonics and nanochemistry: Controlling the excitation dynamics for frequency up-and down-conversion in lanthanide-doped nanoparticles. Acc. Chem. Res. 2013, 46, 1474–1486. [Google Scholar] [CrossRef] [PubMed]
  165. Mayavan, A. Comprehensive Review on Downconversion/Downshifting Silicate-Based Phosphors for Solar Cell Applications. ACS Omega 2024, 9, 16880–16892. [Google Scholar] [CrossRef]
  166. Rahman, N.U.; Khan, W.U.; Li, W.; Khan, S.; Khan, J.; Zheng, S.; Su, T.; Zhao, J.; Aldred, M.P.; Chi, Z. Simultaneous enhancement in performance and UV-light stability of organic–inorganic perovskite solar cells using a samarium-based down conversion material. J. Mater. Chem. A 2019, 7, 322–329. [Google Scholar] [CrossRef]
  167. Jiang, L.; Chen, W.; Zheng, J.; Zhu, L.; Mo, L.e.; Li, Z.; Hu, L.; Hayat, T.; Alsaedi, A.; Zhang, C. Enhancing the photovoltaic performance of perovskite solar cells with a down-conversion Eu-complex. ACS Appl. Mater. Interfaces 2017, 9, 26958–26964. [Google Scholar] [CrossRef]
  168. Ding, N.; Xu, W.; Liu, H.; Jing, Y.; Wang, Z.; Ji, Y.; Wu, J.; Shao, L.; Zhu, G.; Dong, B. Highly DUV to NIR-II responsive broadband quantum dots heterojunction photodetectors by integrating quantum cutting luminescent concentrators. Light Sci. Appl. 2024, 13, 289. [Google Scholar] [CrossRef]
  169. Shao, L.; Liu, Q.; Liu, X.; Wu, W.; Liu, M. Boosting UV responsivity of silicon photodetectors through efficient quantum-cutting with La3+, Yb3+ co-doped perovskite quantum dots. J. Lumin. 2024, 272, 120657. [Google Scholar] [CrossRef]
  170. Yang, G.; Zheng, C.; Zhu, Y.; Li, X.; Huang, J.; Xu, X.; Liu, W.; Cui, S.; Pan, G. Efficient quantum cutting of lanthanum and ytterbium ions co-doped perovskite quantum dots towards improving the ultraviolet response of silicon-based photodetectors. J. Alloys Compd. 2022, 921, 166097. [Google Scholar] [CrossRef]
  171. Hao, M.; Cheng, S.; He, Y.; Xiang, W.; Ding, N.; Xu, W.; Ma, C.G.; Liang, X. Dy3+ Doped All-Inorganic Perovskite Nanocrystals Glass toward High-Performance and High-Stability Silicon Photodetectors. Laser Photonics Rev. 2023, 17, 2200748. [Google Scholar] [CrossRef]
  172. Ding, J.; Mu, S.; Xiang, W.; Ding, N.; Xu, W.; Liang, X. Eu3+ doped CsPbCl2Br1 nanocrystals glass for enhanced the ultraviolet response of Si photodetectors. J. Lumin. 2023, 254, 119530. [Google Scholar] [CrossRef]
  173. Ramalingam, G.; Kathirgamanathan, P. Quantum Confinement Effect of 2D Nanomaterials. In Quantum Dots: Fundamental and Applications; BoD—Books on Demand: Norderstedt, Germany, 2020; p. 11. [Google Scholar]
  174. Lv, J.; Zhang, T.; Zhang, P.; Zhao, Y.; Li, S. Review application of nanostructured black silicon. Nanoscale Res. Lett. 2018, 13, 110. [Google Scholar] [CrossRef]
  175. Zhao, Z.; Zhang, Z.; Jing, J.; Gao, R.; Liao, Z.; Zhang, W.; Liu, G.; Wang, Y.; Wang, K.; Xue, C. Black silicon for near-infrared and ultraviolet photodetection: A review. APL Mater. 2023, 11, 021107. [Google Scholar] [CrossRef]
  176. Thahe, A.A.; Dahi, A.; Qaeed, M.A.; Farhat, O.F.; Bakhtiar, H.; Allam, N.K. Engineered etching and laser treatment of porous silicon for enhanced sensitivity and speed of Pt/n-PSi/Pt UV photodetectors. Nanoscale Adv. 2025, 7, 2955–2966. [Google Scholar] [CrossRef] [PubMed]
  177. Link, S.; El-Sayed, M.A. Spectral Properties and Relaxation Dynamics of Surface Plasmon Electronic Oscillations in Gold and Silver Nanodots and Nanorods. J. Phys. Chem. B 1999, 103, 8410–8426. [Google Scholar] [CrossRef]
  178. Petryayeva, E.; Krull, U.J. Localized surface plasmon resonance: Nanostructures, bioassays and biosensing—A review. Anal. Chim. Acta 2011, 706, 8–24. [Google Scholar] [CrossRef]
  179. Seok, J.S.; Ju, H. Plasmonic Optical Biosensors for Detecting C-Reactive Protein: A Review. Micromachines 2020, 11, 895. [Google Scholar] [CrossRef]
  180. Willets, K.A.; Van Duyne, R.P. Localized Surface Plasmon Resonance Spectroscopy and Sensing. Annu. Rev. Phys. Chem. 2007, 58, 267–297. [Google Scholar] [CrossRef] [PubMed]
  181. Kaushal, S.; Nanda, S.S.; Yi, D.K.; Ju, H. Effects of Aspect Ratio Heterogeneity of an Assembly of Gold Nanorod on Localized Surface Plasmon Resonance. J. Phys. Chem. Lett. 2020, 11, 5972–5979. [Google Scholar] [CrossRef]
  182. Mustafa, D.E.; Yang, T.; Xuan, Z.; Chen, S.; Tu, H.; Zhang, A. Surface plasmon coupling effect of gold nanoparticles with different shape and size on conventional surface plasmon resonance signal. Plasmonics 2010, 5, 221–231. [Google Scholar] [CrossRef]
  183. Ismail, R.A.; Alwan, A.M.; Ahmed, A.S. Preparation and characteristics study of nano-porous silicon UV photodetector. Appl. Nanosci. 2017, 7, 9–15. [Google Scholar] [CrossRef]
  184. Tanaka, Y.-i.; Ono, A.; Inami, W.; Kawata, Y. Silicon Plasmonics for Enhanced Responsivity of Silicon Photodetectors in Deep-Ultraviolet Region. Phys. Rev. Lett. 2025, 134, 226901. [Google Scholar] [CrossRef]
  185. Cao, F.; Liu, Y.; Liu, M.; Han, Z.; Xu, X.; Fan, Q.; Sun, B. Wide Bandgap Semiconductors for Ultraviolet Photodetectors: Approaches, Applications, and Prospects. Research 2024, 7, 0385. [Google Scholar] [CrossRef]
  186. Sang, L.; Liao, M.; Sumiya, M. A comprehensive review of semiconductor ultraviolet photodetectors: From thin film to one-dimensional nanostructures. Sensors 2013, 13, 10482–10518. [Google Scholar] [CrossRef]
  187. Landis, G.A.; Loferski, J.J.; Beaulieu, R.; Sekula-Moisé, P.A.; Vernon, S.M.; Spitzer, M.B.; Keavney, C.J. Wide-bandgap epitaxial heterojunction windows for silicon solar cells. IEEE Trans. Electron Devices 1990, 37, 372–381. [Google Scholar] [CrossRef]
  188. Bean, J.C. Silicon-based semiconductor heterostructures: Column IV bandgap engineering. Proc. IEEE 2002, 80, 571–587. [Google Scholar] [CrossRef]
  189. Zhong, C.; Luo, L.; Tan, H.; Geng, K. Band gap optimization of the window layer in silicon heterojunction solar cells. Sol. Energy 2014, 108, 570–575. [Google Scholar] [CrossRef]
  190. Liu, J.; Zhang, J. Nanointerface chemistry: Lattice-mismatch-directed synthesis and application of hybrid nanocrystals. Chem. Rev. 2020, 120, 2123–2170. [Google Scholar] [CrossRef]
  191. Chu, S.; Tsang, W.; Chiu, T.; Macrander, A. Lattice-mismatch-generated dislocation structures and their confinement using superlattices in heteroepitaxial GaAs/InP and InP/GaAs grown by chemical beam epitaxy. J. Appl. Phys. 1989, 66, 520–530. [Google Scholar] [CrossRef]
  192. Wang, Z.; Chen, Z.; Zhang, H.; Zhang, Z.; Wu, H.; Jin, M.; Wu, C.; Yang, D.; Yin, Y. Lattice-mismatch-induced twinning for seeded growth of anisotropic nanostructures. ACS Nano 2015, 9, 3307–3313. [Google Scholar] [CrossRef] [PubMed]
  193. Hudgins, J.L.; Simin, G.S.; Santi, E.; Khan, M.A. An assessment of wide bandgap semiconductors for power devices. IEEE Trans. Power Electron. 2003, 18, 907–914. [Google Scholar] [CrossRef]
  194. Takahashi, K.; Yoshikawa, A.; Sandhu, A. Wide Bandgap Semiconductors; Springer: Berlin/Heidelberg, Germany, 2007. [Google Scholar]
  195. Rossettos, J.; Shen, X. On the axial and interfacial shear stresses due to thermal mismatch in hybrid composite sheets. Compos. Sci. Technol. 1995, 54, 417–422. [Google Scholar] [CrossRef]
  196. Alred, J.M.; Zhang, Z.; Hu, Z.; Yakobson, B.I. Interface-induced warping in hybrid two-dimensional materials. Nano Res. 2015, 8, 2015–2023. [Google Scholar] [CrossRef]
  197. Kopperschmidt, P.; Scholz, R. Interface defects in integrated hybrid semiconductors by wafer bonding. Phys. B Condens. Matter 2001, 308, 1205–1208. [Google Scholar] [CrossRef]
  198. Du, Y.; Jiang, H.; Zhu, B.; Yan, H.; Chai, Y.; Tsoi, C.C.; Zhang, X.; Wang, C. A Universal Bonding Strategy for Achieving CMOS-Compatible Silicon Heterogeneous Integration. Adv. Mater. Technol. 2025, 10, 2402063. [Google Scholar] [CrossRef]
  199. Garrity, E.; Ciobanu, T.; Zakutayev, A.; Stevanovic, V. Emerging ultra-wide band gap semiconductors for future high-frequency electronics. arXiv 2025. [Google Scholar] [CrossRef]
  200. Yang, Y.; Tu, J.; Zhan, G.; Liu, H.; Fan, X.; Chen, X.; Hu, X.; Zhuang, N. Growth Considering Lattice Mismatch and Thermal Expansion and Magneto-Optical Properties of Bismuth-Doped Rare Earth Iron Garnet Crystals. Cryst. Growth Des. 2023, 23, 1598–1602. [Google Scholar] [CrossRef]
  201. Li, J.; Maidebura, Y.; Zhang, Y.; Wu, G.; Su, Y.; Zhuravlev, K.; Wei, X. AlGaN-Based Ultraviolet PIN Photodetector Grown on Silicon Substrates Using SiN Nitridation Process and Step-Graded Buffers. Crystals 2024, 14, 952. [Google Scholar] [CrossRef]
  202. Mukherjee, K.; Selvidge, J.; Hughes, E.; Norman, J.; Shang, C.; Herrick, R.; Bowers, J. Kinetically limited misfit dislocations formed during post-growth cooling in III–V lasers on silicon. J. Phys. D Appl. Phys. 2021, 54, 494001. [Google Scholar] [CrossRef]
  203. Ghosh, S.; Hinz, A.M.; Frentrup, M.; Alam, S.; Wallis, D.J.; Oliver, R.A. Design of step-graded AlGaN buffers for GaN-on-Si heterostructures grown by MOCVD. Semicond. Sci. Technol. 2023, 38, 044001. [Google Scholar] [CrossRef]
  204. Riah, B.; Camus, J.; Ayad, A.; Rammal, M.; Zernadji, R.; Rouag, N.; Djouadi, M.A. Hetero-Epitaxial Growth of AlN Deposited by DC Magnetron Sputtering on Si(111) Using a AlN Buffer Layer. Coatings 2021, 11, 1063. [Google Scholar] [CrossRef]
  205. Vura, S.; Muazzam, U.U.; Kumar, V.; Vanjari, S.C.; Muralidharan, R.; Digbijoy, N.; Nukala, P.; Raghavan, S. Monolithic Epitaxial Integration of β-Ga2O3 with 100 Si for Deep Ultraviolet Photodetectors. ACS Appl. Electron. Mater. 2022, 4, 1619–1625. [Google Scholar] [CrossRef]
  206. Yen, C.-C.; Huang, T.-M.; Chen, P.-W.; Chang, K.-P.; Wu, W.-Y.; Wuu, D.-S. Role of Interfacial Oxide in the Preferred Orientation of Ga2O3 on Si for Deep Ultraviolet Photodetectors. ACS Omega 2021, 6, 29149–29156. [Google Scholar] [CrossRef]
  207. Roccaforte, F.; Giannazzo, F.; Iucolano, F.; Eriksson, J.; Weng, M.; Raineri, V. Surface and interface issues in wide band gap semiconductor electronics. Appl. Surf. Sci. 2010, 256, 5727–5735. [Google Scholar] [CrossRef]
  208. Setera, B.; Christou, A. Challenges of overcoming defects in wide bandgap semiconductor power electronics. Electronics 2021, 11, 10. [Google Scholar] [CrossRef]
  209. Ganose, A.M.; Scanlon, D.O.; Walsh, A.; Hoye, R.L. The defect challenge of wide-bandgap semiconductors for photovoltaics and beyond. Nat. Commun. 2022, 13, 4715. [Google Scholar] [CrossRef]
  210. Vines, L.; Monakhov, E.; Kuznetsov, A. Defects in semiconductors. J. Appl. Phys. 2022, 132, 190401. [Google Scholar] [CrossRef]
  211. Shao, G. Work function and electron affinity of semiconductors: Doping effect and complication due to fermi level pinning. Energy Environ. Mater. 2021, 4, 273–276. [Google Scholar] [CrossRef]
  212. Newman, N.; Spicer, W.; Kendelewicz, T.; Lindau, I. On the Fermi level pinning behavior of metal/III–V semiconductor interfaces. J. Vac. Sci. Technol. B Microelectron. Process. Phenom. 1986, 4, 931–938. [Google Scholar] [CrossRef]
  213. Faschinger, W.; Ferreira, S.; Sitter, H. Doping limitations in wide gap II–VI compounds by Fermi level pinning. J. Cryst. Growth 1995, 151, 267–272. [Google Scholar] [CrossRef]
  214. Liu, X.; Choi, M.S.; Hwang, E.; Yoo, W.J.; Sun, J. Fermi level pinning dependent 2D semiconductor devices: Challenges and prospects. Adv. Mater. 2022, 34, 2108425. [Google Scholar] [CrossRef] [PubMed]
  215. Aldalbahi, A.; Li, E.; Rivera, M.; Velazquez, R.; Altalhi, T.; Peng, X.; Feng, P.X. A new approach for fabrications of SiC based photodetectors. Sci. Rep. 2016, 6, 23457. [Google Scholar] [CrossRef] [PubMed]
  216. Gao, C.; Wang, Y.; Fu, S.; Xia, D.; Han, Y.; Ma, J.; Xu, H.; Li, B.; Shen, A.; Liu, Y. High-Performance Solar-Blind Ultraviolet Photodetectors Based on β-Ga2O3 Thin Films Grown on p-Si(111) Substrates with Improved Material Quality via an AlN Buffer Layer Introduced by Metal–Organic Chemical Vapor Deposition. ACS Appl. Mater. Interfaces 2023, 15, 38612–38622. [Google Scholar] [CrossRef] [PubMed]
  217. Kang, Y.; Li, F.; Yu, H.; Wang, D.; Sun, H. A van der Waals metal contact for asymmetric GaN-based ultraviolet photodetectors. In Proceedings of the 2024 IEEE Photonics Conference (IPC), Rome, Italy, 10–14 November 2024. [Google Scholar]
  218. Tsanakas, M.D.; Jaros, A.; Fleming, Y.; Efthimiadou, M.; Voss, T.; Leturcq, R.; Gardelis, S.; Kandyla, M. Wavelength-Selective, High-Speed, Self-Powered Isotype Heterojunction n+-ZnO/n-Si Photodetector with Engineered and Tunable Spectral Response. Adv. Mater. Technol. 2025, 10, 2401740. [Google Scholar] [CrossRef]
Figure 1. The spectral range of electromagnetic radiation.
Figure 1. The spectral range of electromagnetic radiation.
Micromachines 16 01130 g001
Figure 2. Wavelength-dependent absorption coefficient of silicon. Reproduced from [51] (Journal of Nanotechnology, 2012), licensed under CC BY 4.0.
Figure 2. Wavelength-dependent absorption coefficient of silicon. Reproduced from [51] (Journal of Nanotechnology, 2012), licensed under CC BY 4.0.
Micromachines 16 01130 g002
Figure 3. Responsivity spectra of ZnO thin films based UV photodetector with interdigitated electrodes for various bias voltage. (a,b) are for 3 μ m electrode spacing, while (c,d) are for 8 μ m electrode spacing. Reproduced with permission from [59] (Sensors and Actuators A: Physical, Elsevier 2018).
Figure 3. Responsivity spectra of ZnO thin films based UV photodetector with interdigitated electrodes for various bias voltage. (a,b) are for 3 μ m electrode spacing, while (c,d) are for 8 μ m electrode spacing. Reproduced with permission from [59] (Sensors and Actuators A: Physical, Elsevier 2018).
Micromachines 16 01130 g003
Figure 4. A schematic of a typical PMT, illustrating sequential processes, photoelectron ejection by a photoelectric effect, photoelectron acceleration for acceleration (dashed line), secondary electron emission at a dynode, the following cascading processes for multiple electrons generation/their acceleration, and their collection by an anode. Reproduced with permission from [71] (Applied Physics B: Lasers and Optics 2019).
Figure 4. A schematic of a typical PMT, illustrating sequential processes, photoelectron ejection by a photoelectric effect, photoelectron acceleration for acceleration (dashed line), secondary electron emission at a dynode, the following cascading processes for multiple electrons generation/their acceleration, and their collection by an anode. Reproduced with permission from [71] (Applied Physics B: Lasers and Optics 2019).
Micromachines 16 01130 g004
Figure 6. A schematic of a photoconductive UV light detector. The UV light absorption in a semiconductor (Blue) generates charge carriers. Externally connected bias voltage drives movement of the charge carriers so that they are collected at the ohmic electrodes (Black). For a given mobility of semiconductor, both the channel length L and electric field across the electrodes determine the carrier transit time. Reproduced with permission from [98] (Journal of Physics D: Applied Physics). Copyright 2014.
Figure 6. A schematic of a photoconductive UV light detector. The UV light absorption in a semiconductor (Blue) generates charge carriers. Externally connected bias voltage drives movement of the charge carriers so that they are collected at the ohmic electrodes (Black). For a given mobility of semiconductor, both the channel length L and electric field across the electrodes determine the carrier transit time. Reproduced with permission from [98] (Journal of Physics D: Applied Physics). Copyright 2014.
Micromachines 16 01130 g006
Figure 7. A schematic of two types of FET-based photodetectors. (a) The FET where a channel itself absorbs photons. (b) The FET where light-absorbing layer is deposited on a channel layer for enhanced light absorption. Reproduced with permission from [105] Advanced Functional Materials, copyright 2020.
Figure 7. A schematic of two types of FET-based photodetectors. (a) The FET where a channel itself absorbs photons. (b) The FET where light-absorbing layer is deposited on a channel layer for enhanced light absorption. Reproduced with permission from [105] Advanced Functional Materials, copyright 2020.
Micromachines 16 01130 g007
Figure 8. Avalanche effects in a p-i-n structure under reverse bias. V0 denotes the inherent junction potential when there is no external bias, while V denotes the external bias voltage.
Figure 8. Avalanche effects in a p-i-n structure under reverse bias. V0 denotes the inherent junction potential when there is no external bias, while V denotes the external bias voltage.
Micromachines 16 01130 g008
Figure 9. A schematic of an MSM photodetector with interdigitated metal electrodes. The yellow arrow represents the direction of incident light. Reproduced from [131] (Advances in Optical Communications 2014). Licensed under CC BY 4.0.
Figure 9. A schematic of an MSM photodetector with interdigitated metal electrodes. The yellow arrow represents the direction of incident light. Reproduced from [131] (Advances in Optical Communications 2014). Licensed under CC BY 4.0.
Micromachines 16 01130 g009
Figure 10. Spectral response curve (Responsivity R versus wavelength) of Si used in solar cells under solar radiation. R drops rapidly at above 1100 nm due to the cutoff given by bandgap energy of 1.12 eV. Reproduced with permission from [153] (Chemical Society Reviews 2013).
Figure 10. Spectral response curve (Responsivity R versus wavelength) of Si used in solar cells under solar radiation. R drops rapidly at above 1100 nm due to the cutoff given by bandgap energy of 1.12 eV. Reproduced with permission from [153] (Chemical Society Reviews 2013).
Micromachines 16 01130 g010
Figure 11. A schematic of cross cross-section of thin n+ layer formed on (a) atomically flat Si surface and (b) conventional Si surface. Reproduced from [160] (J-STAGE, ITE Transactions on Media Technology and Applications 2014) (free access).
Figure 11. A schematic of cross cross-section of thin n+ layer formed on (a) atomically flat Si surface and (b) conventional Si surface. Reproduced from [160] (J-STAGE, ITE Transactions on Media Technology and Applications 2014) (free access).
Micromachines 16 01130 g011
Figure 12. Boron-implanted b-Si photodiode. The SiO2 layer serves as masking material. Reproduced from [161] (ACS Photonics 2023) Licensed under CC-BY 4.0.
Figure 12. Boron-implanted b-Si photodiode. The SiO2 layer serves as masking material. Reproduced from [161] (ACS Photonics 2023) Licensed under CC-BY 4.0.
Micromachines 16 01130 g012
Figure 13. SEM images of periodic corrugation on Si surface. (a) Top view and (b) cross-sectional view. Reproduced from [184] (Physical Review Letters 2025); licensed under CC BY 4.0.
Figure 13. SEM images of periodic corrugation on Si surface. (a) Top view and (b) cross-sectional view. Reproduced from [184] (Physical Review Letters 2025); licensed under CC BY 4.0.
Micromachines 16 01130 g013
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

Kamal, A.; Hong, S.; Ju, H. Advances in Silicon-Based UV Light Detection. Micromachines 2025, 16, 1130. https://doi.org/10.3390/mi16101130

AMA Style

Kamal A, Hong S, Ju H. Advances in Silicon-Based UV Light Detection. Micromachines. 2025; 16(10):1130. https://doi.org/10.3390/mi16101130

Chicago/Turabian Style

Kamal, Arif, Seongin Hong, and Heongkyu Ju. 2025. "Advances in Silicon-Based UV Light Detection" Micromachines 16, no. 10: 1130. https://doi.org/10.3390/mi16101130

APA Style

Kamal, A., Hong, S., & Ju, H. (2025). Advances in Silicon-Based UV Light Detection. Micromachines, 16(10), 1130. https://doi.org/10.3390/mi16101130

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