Next Article in Journal
Optimizing Nanosilica-Enhanced Polymer Synthesis for Drilling Fluids via Response Surface Methodology: Enhanced Fluid Performance Analysis
Previous Article in Journal
Assessing the Shear Capacity of Screw Connectors in Composite Columns of Cold-Formed Steel and Concrete Infill
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances in One-Dimensional Metal Sulfide Nanostructure-Based Photodetectors with Different Compositions

School of Materials Science and Engineering, Henan University of Technology, Zhengzhou 450001, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
J. Compos. Sci. 2025, 9(6), 262; https://doi.org/10.3390/jcs9060262
Submission received: 23 April 2025 / Revised: 21 May 2025 / Accepted: 23 May 2025 / Published: 26 May 2025
(This article belongs to the Section Composites Manufacturing and Processing)

Abstract

:
One-dimensional (1D) nanomaterials have attracted considerable attention in the fabrication of nano-scale optoelectronic devices owing to their large specific surface areas, high surface-to-volume ratios, and directional electron transport channels. Compared to 1D metal oxide nanostructures, 1D metal sulfides have emerged as promising candidates for high-efficiency photodetectors due to their abundant surface vacancies and trap states, which facilitate oxygen adsorption and dissociation on their surfaces, thereby suppressing intrinsic carrier recombination while achieving enhanced optoelectronic performance. This review focuses on recent advancements in the performance of photodetectors fabricated using 1D binary metal sulfides as primary photosensitive layers, including nanowires, nanorods, nanotubes, and their heterostructures. Initially, the working principles of photodetectors are outlined, along with the key parameters and device types that influence their performance. Subsequently, the synthesis methods, device fabrication, and photoelectric properties of several extensively studied 1D metal sulfides and their composites, such as ZnS, CdS, SnS, Bi2S3, Sb2S3, WS2, and SnS2, are examined. Additionally, the current research status of 1D nanostructures of MoS2, TiS3, ReS2, and In2S3, which are predominantly utilized as 2D materials, is explored and summarized. For systematic performance evaluation, standardized metrics encompassing responsivity, detectivity, external quantum efficiency, and response speed are comprehensively tabulated in dedicated sub-sections. The review culminates in proposing targeted research trajectories for advancing photodetection systems employing 1D binary metal sulfides.

1. Introduction

Over recent decades, low-dimensional nanomaterials have attracted considerable research interest across diverse domains, including electronics, optoelectronics, bioimaging, and photonic systems, owing to their readily tunable structural, electronic, optical, and chemical characteristics [1,2,3]. These materials are categorized into zero-dimensional (0D), one-dimensional (1D), and two-dimensional (2D) systems based on their structural dimensionality. Precise dimensionality control allows for the systematic modulation of their physical constraints, thereby regulating their dimensional scales and surface-to-volume ratios, while fundamentally determining their distinctive optical, thermal, electronic, and mechanical properties [4,5,6,7,8,9,10]. Among these materials, one-dimensional (1D) nanomaterials exhibit unparalleled advantages for photodetection applications. Nanowires (NWs), nanorods (NRs), and nanotubes (NTs) not only inherit the intrinsic benefits of low-dimensional systems but also amplify light–matter interactions through their anisotropic geometry, enabling directional carrier transport and enhanced light trapping [11,12,13,14,15]. In recent years, these materials have attracted considerable attention within the photonics community, particularly demonstrating critical functionalities in photocatalysis, solar energy conversion, and photodetection technologies. These distinctive attributes establish one-dimensional nanostructures as highly promising architectures for addressing the inherent constraints of traditional photodetection systems [16,17,18].
As critical signal transduction components, photodetectors exhibit extensive applicability in fiber-optic communications, advanced imaging systems, environmental sensing, and space exploration technologies, while demonstrating transformative potential for next-generation optical data storage and binary switching operations within integrated optoelectronic architectures [19,20,21,22,23]. Traditional photodetectors based on Si [24] and InGaAs [25], while technologically mature, are limited by their narrow spectral response ranges (Si: 300–1100 nm; InGaAs: 800–1700 nm), high dark current densities (>10−7 A cm−2), and rigid substrate constraints. These limitations hinder their compatibility with emerging requirements for flexible electronics and broadband detection applications. While metal oxide nanostructures (e.g., ZnO, TiO2, Ga2O3) [26,27,28,29,30,31,32,33,34,35,36] emerged as alternatives by offering morphological controllability, their wide bandgaps (~3.3 eV) and surface vacancy deficiencies fundamentally restrict their light absorption efficiency and carrier dynamics [37,38,39,40]. In comparison, metal sulfides possess bandgap engineering potential through controlled sulfur stoichiometry tuning, enabling wide-spectrum photon absorption across extended wavelength regimes [41]. For example, CuS exhibits superior photoresponsivity (0.41 A W−1) [42] and a microsecond-level response time, significantly outperforming Cu2S (0.22 A W−1, second-level response) [43]; stoichiometric modulation from SnS2 to SnS enables continuous bandgap tuning from 2.9 eV to 1.32 eV, thereby achieving ultraviolet-to-near-infrared spectral response coverage (UV-NIR, ~300–1300 nm) [44]. Additionally, surface sulfur vacancies serve as effective electron trapping sites, which suppress carrier recombination and facilitate efficient carrier separation, consequently reducing dark current densities (<10−10 A cm−2) [45,46,47]. This synergistic interplay—spectral tunability coupled with recombination suppression—establishes metal sulfides as having enhanced optoelectronic capabilities relative to their conventional semiconductor and metal oxide counterparts. Some representative demonstrations of these capabilities include the following: CdS nanowires achieve exceptional UV–visible photoresponsivity (104 A W−1), outperforming ZnO nanowires by two orders of magnitude, while SnS2-based heterostructures demonstrate broad-spectrum detectivity surpassing 1013 Jones, comparable to commercial InGaAs photodetectors. The unique advantage of 1D architectures originates from their inter-dimensional synergy between structural confinement and compositional engineering. Distinct from 2D planar configurations or 0D quantum-confined architectures, the anisotropic morphology of 1D systems concurrently reduces charge migration pathways and extends recombination lifetimes—a dimensional contradiction resolved through directional confinement modulation [48,49,50,51]. For example, 1D SnS2 nanostructures demonstrate a specific detectivity of 1016 Jones and a responsivity of 103 A W−1, surpassing SnS2-based quantum dots (1011 Jones, 0.1 A W−1) and thin films (2.02 × 1010 Jones, 1.95 A W−1) by several orders of magnitude [44,52]. Meanwhile, 1D WS2 photodetectors exhibit a responsivity of 0.777 A W−1, which is 10 times higher than their thin-film counterparts (0.314 A W−1) [53,54].
While current reviews focus primarily on 2D material systems, this work proposes a systematic classification of 1D metal sulfides via sulfur stoichiometry modulation (mono-, di-, and polysulfides), providing a novel analytical paradigm for elucidating structure–property relationships in photodetection physics. This review systematically examines synthesis methodologies for one-dimensional metal sulfides and their engineered composites (including doping strategies, p-n/n-n heterojunction configurations), with representative material systems spanning ZnS, CdS, SnS, WS2, SnS2, MoS2, ReS2, Sb2S3, Bi2S3, In2S3, and TiS3. The enhanced optoelectronic performance of these devices is systematically assessed through key figures of merit—responsivity (R), detectivity (D*), external quantum efficiency (EQE), and temporal response dynamics. The concluding section delineates prospective research directions for 1D metal sulfide photodetectors, addressing fundamental scientific challenges alongside application-specific optimization strategies. This analysis establishes a methodological framework for guiding functional optoelectronic materials development.
This review establishes a pioneering framework elucidating how structural superiority synergistically interacts with sulfur-mediated electronic engineering to achieve unprecedented photodetection performance benchmarks. We perform critical techno-scientific evaluations of cutting-edge advances in quasi-1D transition metal sulfide systems (ZnS, CdS, SnS, WS2, etc.). Departing from conventional fragmented approaches to synthesis protocols and device physics, this investigation establishes systematic correlations between morphological/compositional engineering and device-level performance metrics—responsivity (A W−1), specific detectivity (Jones), external quantum efficiency (%), and temporal response characteristics.

2. Photoelectric Detector Basic Principle and Performance Parameters

2.1. Basic Principles of Photoelectric Detectors

A photodetector is formally defined as an optoelectronic transducer demonstrating photoconductivity modulation within photoactive media under photon irradiation. Its configuration primarily consists of two functional elements: a photon-sensitive semiconductor matrix and metallic charge transport electrodes. The operational dynamics are predominantly dictated by interface potential barrier formation at Schottky junctions and specific device configurations. These mechanisms can include the photoconductivity effect [55,56], the photovoltaic effect [57,58], the photo-thermoelectric effect [59,60], and the pyroelectric effect [61,62], among others. Photon absorption with energy surpassing the detector’s bandgap triggers electronic transitions from valence to conduction bands, generating mobile charge carriers. Under intrinsic built-in or externally applied electric fields, these liberated electrons undergo directional long-range transport, while concomitant hole migration through valence band states enhances bulk conductivity. The resultant electron–hole pair dissociation enables electrode-collectable photocurrent generation, demonstrating intensity-proportional behavior relative to incident photon flux at operational wavelengths. Subsequent sections will systematically evaluate critical performance metrics encompassing photocurrent density, spectral responsivity (R), normalized detectivity (D*), external quantum efficiency (EQE), and temporal response dynamics.

2.2. Basic Performance Parameters of Photodetector Devices

To enable rigorous performance comparison across photodetector architectures, standardized evaluation metrics have been established. This section delineates their formal definitions and elucidates the physical principles governing these fundamental photodetection parameters.

2.2.1. Photocurrent (Iph)

Iph is defined as the ratio of photogenerated electron–hole pairs (e-h+) collected by the electrode at the edge of the semiconductor. In a photodetector, the photocurrent is estimated by the following equation [63,64,65]:
I p h = I l i g h t I d a r k
where Iph denotes the photocurrent, and Ilight and Idark are the currents obtained by the device under light and dark conditions, respectively.

2.2.2. Responsiveness (R)

R denotes the photoelectric conversion efficiency of the photodetector. It is usually calculated using the following formula [63,64,65]:
R = I p h P S = I l i g h t I d a r k P S
Iph is the photogenerated current, which refers to the current of the photodetector device under light (Ilight) minus the current in the dark (Idark), P is the incident light power density, and S is the effective area of light irradiation.

2.2.3. Specific Detection Rate (D*)

D* denotes the sensitivity of the photodetector device for low-signal detection and is defined as the reciprocal of the Noise Equivalent Power (NEP). NEP is the incident optical power at a signal-to-noise ratio of 1, i.e., the noise current i divided by R. D* can be expressed as follows [63,64,65]:
D * = ( S f ) 1 / 2 N E P = R ( S f ) 1 / 2 i
where Δf is the bandwidth. When the dark current of the device is the main contributor to the noise, D* can be simplified as follows:
D * = R ( 2 e I d a r k / S ) 1 / 2

2.2.4. External Quantum Efficiency (EQE)

EQE is the average value of the number of electrons released by a photodetector device for each incident photon, and it is related to the wavelength of the incident light (I). EQE can be calculated by the following formula [63,64,65]:
E Q E = N C N I = h c e λ R
NC and NI denote the number of photogenerated carriers and incident photons, respectively.

2.2.5. Response Time (τ)

The response speed of a photodetector is a characterization of how quickly a detector responds to light, and it is usually measured by the response time. The response time of a detector is usually the time it takes for the output signal to rise from zero to 63.2% of the peak (i.e., 1-1/e) or from the peak to 36.8% of the peak (i.e., 1/e) under the irradiation of pulsed light. It is also possible to take 10% to 90% of the rising edge of the detector’s optical impulse response curve as the rise time (tr), or 90% to 10% of the falling edge as the falling time (tf). In general, there are two main aspects that limit the response speed of a detector: (1) the resistance–capacitance-induced RC delay time; (2) the transit time of photogenerated carriers between electrodes. The response time can be approximated as follows [66]:
τ = τ R C 2 + τ t r 2
where τ R C is the circuit RC delay and τ t r is the carrier time. In addition, for devices with parasitic photoconductance, the capture and release of photogenerated carriers by defects will also seriously affect the response speed of the device.

2.2.6. Photoconductivity Gain (G)

G is defined as the number of carriers passing through the external circuit corresponding to each incident photon in the photodetector device, and can also be calculated from the ratio of the carrier lifetime ( τ l ) to the inter-electrode transition time ( τ 1 ), which gives rise to a gain when the carrier lifetime is greater than the transition time. G can be calculated according to the following equation [63,64,65]:
G = τ l τ 1 = τ l d 2   μ V
where d is the distance between the two electrodes, μ is the mobility of the majority charge carriers, and V is the applied voltage.

2.2.7. Carrier Capture Coefficient (c)

In semiconductor device physics, the carrier capture coefficient (typically denoted as c or σ) is a key parameter that describes the efficiency of defects, traps, or impurity centers in capturing carriers (electrons or holes) in the material. It directly relates to the dynamics of carriers and influences the electrical characteristics of the device, such as response speed, dark current, and noise. The rate equation for carrier capture by defects is given by [67]:
R c a p t u r e = c · N t · n
where R c a p t u r e , N t , n , and c are, respectively, the capture rates, defect state density, free carrier concentration, and capture factor.
When introducing the carrier thermal velocity ( v t h ), the capture coefficient can be expressed as follows:
c = σ · v t h
σ : Capture cross-section, reflecting the “effective capture area” of defects.
v t h : Carrier thermal motion velocity, it is related to temperature ( v t h T ).

2.2.8. Linear Dynamic Range (LDR)

The linear dynamic range (LDR(λ)) of a photodetector is defined as the range across which the photocurrent rises with an increasing optical power of incident light. The detector can detect the incident signal within this range of optical power. The device’s responsiveness should ideally stay consistent as light intensity increases. Typically, the whole dynamic range is calculated from the NEP of a photodetector to the optical power when photocurrent saturates. It is often expressed in decibels (dB). The LDR(λ) can be represented (in dB) as follows [63,64,65]:
L D R λ = 20 l o g 10 P ( λ ) m a x P ( λ ) m i n
where P(λ)max and P(λ)min are denoted as the highest and lowest limits of optical power in a certain wavelength range. Notably, having a suitably high LDR(λ) means being able to retain continuous sensitivity from strong-light to weak-light conditions, which is a necessity for weak-light sensing.

2.3. I-V Characteristics in Relation to Ohmic Contacts and Schottky Barriers

Current–voltage (I-V) characterization serves as a fundamental diagnostic tool for probing charge transport mechanisms in semiconductor architectures. The curve profile constitutes a critical indicator of interface quality and carrier injection dynamics between electrode–semiconductor interfaces. Nonlinear I-V characteristics reveal energy-filtering Schottky barrier signatures, whereas linear behavior demonstrates barrier-free Ohmic contact formation.

2.3.1. Schottky Barrier and Nonlinear I-V Curves

The Schottky barrier formation originates from interfacial energy-level misalignment between metallic work functions and semiconductor electron affinities. This energetic discrepancy induces localized band bending at the junction, generating a rectifying potential barrier that restricts majority carrier transport, thereby manifesting nonlinear charge injection behavior in I-V characteristics [68]. The observed asymmetric nonlinearity in Au/ZnSe/Au devices’ I-V characteristics confirms their Schottky-type interfacial characteristics at their electrode–semiconductor interfaces, as opposed to non-Ohmic charge injection pathways [69]. As shown in Figure 1a, a typical Schottky barrier I-V curve demonstrates that, under forward bias, current increases rapidly with voltage, while under reverse bias, current is significantly suppressed [70].

2.3.2. Ohmic Contact and Linear I-V Curves

Ohmic contacts are characterized by minimal interfacial resistance at metal–semiconductor interfaces, facilitating unrestricted majority carrier injection across the junction while exhibiting linear current–voltage characteristics. This optimized electronic state originates from energy-level alignment between metal work functions and semiconductor Fermi energies under thermodynamic equilibrium conditions [72]. For example, the linear I-V curve in Cr/Au electrode contacts with ITO nanowires confirms high-quality Ohmic contact formation [73]. As illustrated in Figure 1b, a typical Ohmic I-V curve demonstrates efficient current transport even at low voltages, thereby optimizing device performance [71]. Table 1 presents the work functions of different electrode materials and their corresponding I-V curve characteristics.

2.3.3. Comparison of Electrode Contact Configurations

The optimization of contact geometries in photodetectors requires engineering electrode–semiconductor interface configurations to optimize critical performance parameters encompassing responsivity, response speed, external quantum efficiency (EQE), and specific detectivity (D*) [76,77]. As illustrated in Figure 2, three primary contact configurations are typically employed: conventional (planar) contacts, interdigitated contacts, and hybrid contact structures [78]. This strategic selection of contact geometry enables performance enhancement through optimized carrier collection efficiency and reduced parasitic resistance.
Additionally, the work function and interfacial characteristics of different electrode materials (e.g., Au, Cr, ITO) significantly influence the selection of contact geometry [72,74]. Table 2 systematically compiles the performance metrics of ZnSe-based MSM UV photodetectors with varied contact geometries and electrode compositions under standardized operating conditions. The tabulated data reveal that discrete contact architectures incorporating different metallic electrodes demonstrate statistically significant performance discrepancies across key device parameters [78]. The interdigitated Ni/Au electrode configuration achieves a photoresponsivity of 5.4 A W−1, demonstrating a two-fold enhancement over its Cr/Au counterpart (2.23 A W−1). Conversely, conventional Cr/Au contacts exhibit superior performance (1.25 A W−1) compared to standard Ni/Au architectures (0.79 A W−1). This electrode-material-governed polarity inversion demonstrates the pivotal role of contact geometry optimization in precisely modulating photodetector operational parameters, thereby expanding their functional versatility within next-generation optoelectronic architectures.

2.4. Self-Biasing Behavior of Photodetectors

The self-biasing operation of photodetectors manifests as spontaneous photocurrent generation through the efficient separation and transport of photogenerated charge carriers under zero-external-bias conditions (V = 0 V). This phenomenon fundamentally originates from built-in potential gradients within device architectures, predominantly established at heterointerfaces or p-n junction boundaries. Under photon illumination, photogenerated carriers undergo field-driven separation via these intrinsic potential gradients, thereby enabling the photodetectors’ self-powered photodetection functionality.

2.4.1. Formation of Built-In Electric Field

The formation of built-in electric fields originates from carrier diffusion dynamics and space charge redistribution driven by Fermi-level gradients across p-n heterojunctions, culminating in thermodynamic equilibrium through band potential alignment and field stabilization [79,80]. As shown in Figure 3, when p-type Bi2Se3 and n-type Bi2S2 come into contact, electrons diffuse from the n-type (higher Fermi level) to the p-type (lower Fermi level) region, while holes diffuse in the opposite direction due to the Fermi-level mismatch. This process generates a space charge region proximal to the interface: positively charged donor ions are immobilized within the n-region, while negatively charged acceptor ions localize in the p-region. Thermodynamic equilibrium is established upon Fermi-level alignment across the interface, culminating in a stabilized built-in electric field [81,82].

2.4.2. Charge Separation and Transport Mechanism

The charge separation process is critical in photodetectors, with its efficiency directly determining device performance. As illustrated in Figure 4a, when photons irradiate the material surface, energy is transferred to electrons within the material. Excited electrons transit from the valence band to the conduction band, generating electron–hole pairs. Simultaneously, the built-in electric field induces band bending at the heterojunction or p-n junction interface. Electrons diffuse from the higher-energy conduction band material to the lower-energy counterpart, while holes migrate in the opposite direction through the valence band. Finally, a self-powered mode is formed [83,84,85]. When an external electric field is applied (Figure 4b), carriers are ultimately collected by electrodes: electrons accumulate near the ITO electrode (~4.8 eV) and holes enrich around the Ag electrode (~4.26 eV). This spatial separation reduces carrier recombination and enhances charge collection efficiency, enabling high-performance photodetection [86].

3. High-Performance Metal Sulfide-Based Photodetectors

3.1. Material Selection

Metal sulfide compounds constitute an extensive class of inorganic materials that have become a subject of intensive investigation for diverse emerging optoelectronic applications, including photodetectors, solar cells, and light-emitting diodes [87]. These compounds are present in the form of natural minerals such as sphalerite (ZnS), galena (PbS), and pyrite (FeS2), which are inexpensive, abundant, and have different structural types [88]. Metal and sulfur elements, via distinct coordination configurations, yield structurally simple metal sulfides, while these compounds demonstrate remarkable diversity in both architectures and polymorphic crystalline phases, including isotropic/anisotropic arrangements and centrosymmetric versus non-centrosymmetric structural motifs [89]. One-dimensional metal sulfides exhibit significantly higher surface-to-volume ratios than their layered counterparts, facilitating superior optoelectronic performance via increased carrier densities and broad spectral sensitivity extending from ultraviolet (UV) to near-infrared (NIR) regimes. These compounds inherently support the classification of photodetectors into UV, visible, and NIR operational categories according to their intrinsic bandgap energetics. Wide-bandgap systems such as ZnS and SnS2 are employed for UV detection, while visible-responsive compounds including CdS, WS2, Sb2S3, MoS2, and ReS2 operate within the 400–700 nm range. Narrow-bandgap variants like TiS3, SnS, and In2S3 facilitate NIR photodetection. Furthermore, metal sulfides are categorized as monosulfides (ZnS, CdS, SnS), disulfides (SnS2, WS2, MoS2, ReS2), or polysulfides (Sb2S3, In2S3, TiS3), each exhibiting distinct structural and device property variations. Table 3 lists the important structural, optical and electrical properties of various metal sulfide nanostructures, materials whose superior structural, optical, and electrical properties have attracted considerable interest from researchers for high-performance photodetectors [90]. In the next section, we discuss in detail the progress of research on one-dimensional metal sulfide-based photodetectors.

3.2. One-Dimensional Binary Metal Sulfide-Based Photodetectors

This chapter systematically examines contemporary developments in one-dimensional metal sulfide photodetectors, beginning with the fundamental optoelectronic properties and device implementations of ZnS, CdS, and SnS monosulfide nanostructures, along with their heterojunction architectures. The discussion subsequently expands to transition metal disulfides (WS2, SnS2, MoS2, ReS2), analyzing their photodetection performance metrics and unique electronic configurations. Eventually, we highlight the superior optoelectronic conversion characteristics of higher-sulfur-content compounds, including Sb2S3, Bi2S3, and In2S3, establishing an expanded material selection framework and fundamental references for developing advanced photodetection systems. This comprehensive review aims to guide researchers towards a deeper understanding of the direction of one-dimensional metal sulfide-based photodetectors and inspire them to innovate in this field.

3.2.1. Photodetectors Based on a Single Sulfur–Metal Compound

One-dimensional metal monosulfide compounds exhibiting 1:1 metal–sulfur stoichiometric ratios have emerged as frontrunners in photodetector technologies. Notably, ZnS-based ultraviolet (UV) detectors and CdS-based visible-light photodetectors have attracted significant scientific focus, owing to their facile synthetic routes synergized with superior optoelectronic metrics, such as enhanced photosensitivity and ultrafast response dynamics. Although SnS research remains comparatively underdeveloped, with fewer mature fabrication protocols relative to analogous systems, the unique optoelectronic signatures arising from its intrinsic p-type semiconducting nature reveal significant opportunities for engineering advanced p–n heterojunction photodetection platforms.

Zinc Sulfide (ZnS)-Based Photodetectors

As an important constituent of the metal–sulfur group of compounds, ZnS exhibits a bandgap energy of 3.72 eV (sphalerite) or 3.77 eV (fibrillar zincite) [124,125]. Devices prepared with the compound find primary applications in the domains of lasers, photocatalysis, flat panel displays, and UV photodetectors [126,127,128]. This section methodically delineates contemporary scientific progress in one-dimensional ZnS nanostructured photodetectors, encompassing systematic classifications of pristine ZnS-based configurations, doped variants, and heterojunction architectures within photodetection systems.
ZnS nanostructures demonstrate ultrafast UV photoresponse characteristics, attributed to their wide bandgap energy, enhanced electron mobility, and superior electrical conductivity. Building upon these findings, Sue et al. [129] fabricated ZnS NWs on p-type Si substrates via thermochemical vapor deposition (TCVD), subsequently constructing photodetectors with sputter-deposited Pt electrodes. The optimized devices exhibited notable optoelectronic performance under 365 nm UV illumination (38 mW/cm2), achieving a photocurrent gain of 0.572, a responsivity of 2.761 A W−1, and a response speed with rise/recovery times of 3.2/3.6 s. However, intrinsic deep-level defects in ZnS NWs have been identified as critical mediators of substantial carrier scattering phenomena during charge transport dynamics, thereby inducing progressive degradation in their photoluminescence quantum yield. Following this rationale, Dai’s team [130] developed a CVT (chemical vapor transport) strategy to grow single-crystalline ZnS nanowires with reduced defect density on Au-coated sapphire substrates, subsequently constructing nanophotonic devices through the precise integration of individual nanowires with Au interdigitated electrodes. The synthesized nanowires demonstrated superior crystallinity and pronounced photoluminescence characteristics. The systematic evaluation of single-nanowire UV detectors under dark conditions and 325 nm illumination (Figure 5a) revealed a ten-fold photocurrent enhancement compared to dark current levels. Device optimization at 1 V bias with 325 nm excitation (1.25 mW cm−2) yielded exceptional performance metrics, including the maximum switching ratio and sub 100 ms response dynamics (Figure 5b), collectively confirming the viability of defect-engineered ZnS nanowires for high-performance UV photodetection applications.
In comparison to nanowire structures, nanotubes exhibit a greater surface-area-to-volume ratio, which allows for enhanced light absorption and reflection in optoelectronic devices. Consequently, An et al. [131] successfully synthesized single-crystalline ZnS nanotubes (NTs) on Au-catalyzed Si substrates via a vapor-phase deposition process. Complementing this synthesis, the team fabricated metal–semiconductor–metal (MSM) photodetector architectures by integrating silver nanowire (Ag NW) networks as transparent electrodes through microinjection patterning. Under −2 V bias, the optimized devices exhibited a high photocurrent of 1.32 μA and an ultralow dark current of 8.44 pA, yielding an exceptional switching ratio (~1.56 × 105) and rapid response dynamics (rise/decay times: 0.12 s/0.4 s) (Figure 5c,d). At 40 mW cm−2 illumination intensity, these ZnS NT-based devices demonstrated superior optoelectronic performance in the following metrics: responsivity (16.5 A W−1), photoconductive gain (8.92), detectivity (1.41 × 109 Jones), and ultralow Noise Equivalent Power (0.0001 pW Hz−1/2). This enhanced functionality originates from their nanotubular morphology, where confined photon trajectories and interfacial electric fields synergistically boost light–matter interactions and charge separation efficiencies. The research team further developed an innovative self-powered ZnS-NT/Si-NW MSM UV photodetector and systematically evaluated its optoelectronic performance [132]. The device exhibits pronounced photovoltaic functionality under zero-bias conditions, arising from the built-in interfacial electric field at the Ag/ZnS heterointerface. Under optical excitation, photoinduced electron–hole pairs in ZnS experience field-driven directional separation and transport dynamics: electrons are preferentially transported to the Ag nanowire collector electrode via drift–diffusion mechanisms, while holes remain spatially confined within the ZnS matrix, collectively establishing a sustained photogenerated current. This inherent self-driven carrier management mechanism facilitates bias-free photocurrent generation through optimized charge partitioning and transport kinetics. In self-powered mode, the device achieves a dark current of 14.9 pA and a photocurrent of 0.29 μA, yielding a substantial switching ratio of 19,173. Figure 5e depicts five repeat cycles under on/off switching illumination. It can be seen that the device can be alternatively switched between on- and off-state conditions with excellent reproducibility and stability—exhibiting rapid rise/decay times of 0.09 s/0.07 s (ambient) and 0.19 s/0.10 s (vacuum) under pulsed illumination (Figure 5f). Under 297 nm UV irradiation (40 mW cm−2), the detector delivers a responsivity of 2.56 A W−1, a photoconductive gain of 13.6, and a detectivity of 1.67 × 1010 Jones, establishing a foundational framework for designing high-efficiency 1D ZnS-based photodetection systems.
The ordered architecture of nanoarray configurations demonstrates significant advantages over isolated nanostructures through enhanced light-harvesting capabilities derived from expanded active areas. Liang et al. [133] employed metal–organic chemical vapor deposition (MOCVD) to fabricate vertically aligned ZnS nanoarrays (NWAs) on GaAs substrates. Figure 6a schematically depicts the single-nanowire device architecture, wherein Cr/Au electrodes are lithographically defined at both terminals of individual ZnS nanowires on silicon substrates. An analogous fabrication strategy was employed for the nanowire array (NWA) configuration, incorporating conformally deposited Au thin-film electrodes across the NWA surface (Figure 6b). Systematic photoelectronic characterization under 325 nm illumination quantitatively revealed divergent operational parameters: the individual nanowire device exhibited a photocurrent of 1.5 pA under 10 V applied bias (Figure 6c), whereas the NWA architecture generated a 508 pA photocurrent with substantially suppressed dark current at an applied bias of only 1 V (Figure 6d). Spectral selectivity testing at 442 nm confirmed minimal visible-light response, affirming its UV-specific detection capabilities. Cyclic switching stability measurements (Figure 6e,f) were conducted to quantify the temporal response characteristics. The nanowire array (NWA) device demonstrated superior reproducibility, operational stability, and significantly accelerated response speeds (tr = 5 ms, tf = 40 ms) compared to the single-nanowire device (tr, tf < 1.5 s), as evidenced by five consecutive illumination on/off cycles. The optimized NWA photodetector achieved exceptional figures of merit, including 1.87 A W−1 responsivity and 710% external quantum efficiency. This performance enhancement mechanism is mechanistically attributed to two synergistically coupled mechanisms: (1) interwire coupling phenomena that effectively suppress nonradiative recombination pathways; (2) vertically ordered nanoarchitectures acting as resonant photonic microcavities that enable enhanced photon localization through multi-scale light–matter interactions. When integrated with ultralow operational voltage thresholds and wafer-scale manufacturing protocols, these rationally designed nanoarray systems establish technologically promising platforms for implementing next-generation energy-autonomous optoelectronic device systems.
Nanostructured ZnS with atomic-level crystallinity and mitigated compensated defect pairs manifests a highly insulating nature, thereby imposing critical constraints on its optoelectronic device integration capabilities. Strategic doping emerges as an effective approach to modulate these properties, as demonstrated by Saeed et al. [134] through their hydrothermal synthesis of Mn2+-doped ZnS nanorods (NRs). The team fabricated UV photodetectors by integrating silver electrode networks via radio-frequency magnetron sputtering. Doping-induced bandgap narrowing (3.46 eV vs. 3.77 eV for pristine ZnS) extended the spectral response while maintaining visible-blind UV selectivity. As shown in Figure 7a, the device exhibited a 310 nm UV-triggered photoresponse with 1.8 A W−1 responsivity and 719% external quantum efficiency at 0.1 V bias under 0.5 W cm−2 illumination. Figure 7b demonstrates the time-dependent photoresponse characteristics of the UV photodetector. The photocurrent maintained remarkable stability, with an average value of 145.5 μA over hundreds of illumination cycles. Upon UV light termination, the photocurrent rapidly decayed to its low-current state, indicating excellent repeatable stability under operational conditions. The device achieved fast response dynamics (rise/decay times of 16 ms/1.1 ms), with this enhancement attributed to Mn2+-induced surface dangling bonds that accelerated carrier recombination through defect-mediated pathways. Additionally, Jiang et al. [135] synthesized n-type Al-doped ZnS nanowires through thermal co-evaporation, with indium tin oxide (ITO) electrodes deposited via pulsed laser deposition (PLD) to fabricate photodetector devices. As demonstrated in Figure 7c, these devices exhibit wavelength-dependent photocurrent enhancement under 100 mW cm−2 illumination, achieving UV-region detection currents equivalent to visible-range dark current magnitudes. At optimized Al doping (ZnS:Al = 10:1 molar ratio) under 5 V bias with 254 nm illumination (300 mW cm−2), the detectors demonstrated record responsivity (3.1 × 106 A W−1) and photoconductive gain (1.5 × 107). Their enhanced performance stems from two synergistic mechanisms: (1) Surface-band bending, promoting the spatial separation of photogenerated carriers. (2) Aluminum dopants acting as hole-trapping sites, suppressing carrier recombination and extending lifetime (t ≈ 445 s). While exhibiting prolonged response times (tr = 153 s, tf = 445 s) compared to their Mn2+-doped counterparts, attributed to delayed hole release from Al-induced trapping sites, the exceptional UV responsivity and gain characteristics of these detectors compensate for the temporal limitations in their non-switching detection applications.
Surface passivation and heterojunction engineering have been proven to be effective in enhancing photoresponsive performance by mitigating nonradiative trap states. Zhang et al. [136] synthesized ZnS/InP core–shell nanowires (NWs) via chemical vapor deposition (CVD) and fabricated photodetectors by lithographically patterning Cr/Au electrodes on individual NWs aligned on Si/SiO2 substrates. As shown in Figure 7d, when light irradiates a material, it generates photogenerated carriers (electron–hole pairs). The number of these carriers is proportional to the incident optical power. As the optical power increases, the number of generated carriers increases, leading to an enhanced photocurrent. Carrier generation scales linearly with incident power (1.87 μW cm−2), where heterostructured devices exhibit 18.7 times higher photocurrents than pristine ZnS NWs. Under 323 nm UV illumination at 5 V bias, the heterojunction demonstrates 295 A W−1 responsivity and a 10.9 pA photocurrent—two orders of magnitude greater than undoped ZnS NWs (2.0 A W−1, 0.1 pA). The optimized device achieves a photoconductive gain of 1.10 × 103 and a detectivity of 1.65 × 1013 Jones. Figure 7f displays the I-T curves measured under different bias voltages of 2 V, 5 V, and 8 V. The corresponding I-T curves under all the tested biases exhibit excellent reproducibility and stability, further confirming the superior performance of the ZnS/InP NW heterostructure-based device. At 5 V bias, the device demonstrates rapid response characteristics, with a rise time (tr) of 0.75 s and a decay time (tf) of 0.5 s. Complementary to these approaches, Lou et al. [137] fabricated 1D ZnS/CdS nanowire heterostructures via a two-step thermal evaporation process, subsequently patterning Cr/Au bilayer electrodes through lithographic microfabrication. The photodetector demonstrates broadband spectral operation with a photocurrent of 1.2 nA and dark current of 10 fA at 450 nm (213 μW cm−2), achieving an exceptional current switching ratio (~1.2 × 105). Concurrently, the device exhibits ultrahigh detectivity (2.23 × 1014 Jones) and rapid temporal response characteristics (rise/fall times: 5 ms/7 ms), establishing a benchmark for dual-band UV-vis ultrafast detection. Advancing UV optoelectronics, Lee et al. [138] engineered ZnS/ZnO core–shell nanorods through dual-zone horizontal tube furnace synthesis and atomic layer deposition (ALD). This architecture exhibits enhanced near-band-edge (NBE) UV emission by suppressing deep-level (DL) defects through interfacial energy band alignment. The improved performance results from synergistic carrier transport between the ZnS core and ZnO shell, which amplifies UV radiative recombination while passivating surface trap states. This work provides critical insights into the development of high-efficiency UV photodetectors with suppressed visible-wavelength interference.

Cadmium Sulfide (CdS)-Based Photodetectors

As an intrinsic n-type semiconductor, CdS exhibits unique electronic properties characterized by a wide direct bandgap (~2.4 eV), elevated refractive index, reduced work function, and efficient charge carrier transport [139,140,141]. These inherent attributes have enabled CdS nanomaterials to function as critical components in diverse optoelectronic systems, including optical actuators, photonic waveguides, field-effect transistors (FETs), light-emitting diodes, and radiation detectors [142,143,144]. This sub-section systematically reviews recent progress in one-dimensional CdS nanostructure-based photodetection technologies, with a particular focus on comparative performance analyses of undoped, chemically modified, and heterojunction-engineered device architectures.
Hierarchical CdS nanowire (NW) architectures exhibit superior surface-to-volume ratios, photon-harvesting efficiency, and charge transport dynamics compared to their conventional single-crystalline counterparts [145,146]. Li et al. [147] synthesized branched CdS NWs through vapor transport synthesis and fabricated a photodetector device via the lithographic patterning of Cr/Au electrodes onto individual NWs aligned on SiO2/Si substrates. As shown in Figure 8b, the photocurrent exhibits intensity-dependent behavior under 470 nm illumination (0.93 mW cm−2). When light irradiates a material, it generates photogenerated carriers (electron–hole pairs). The number of these carriers is proportional to the incident optical power. As the optical power increases, the number of generated carriers increases, leading to an enhanced photocurrent. At 5 V bias, the device achieves an ultralow dark current (4.7 fA) and a high photocurrent (92.2 pA), yielding a switching ratio of 1.96 × 104 with detectivity reaching 4.27 × 1012 Jones. Figure 8c displays the dynamic photocurrent response of the monolayer CdS nanowire photodetector under a 2 V bias voltage. The photocurrent rapidly increases to a steady state upon illumination and decays promptly after light cessation, revealing the device’s superior response speed and operational stability. Figure 8d presents a magnified view of a single cycle from Figure 8c, with a measured rise time (tr) and decay time (tf) of 0.3 s and 0.4 s, respectively.
To mitigate gate bias effects and suppress dark current generation, Zheng et al. [148] fabricated single CdS nanowire (NW) devices via chemical vapor deposition (CVD), integrating Cr/Au electrodes through metal evaporation and patterning. A 200 nm ferroelectric P(VDF-TrFE) polymer layer was spin-coated onto the NW channel (Figure 9a), inducing carrier depletion via polarization-generated electrostatic fields. Post-depletion, the device exhibited a dark current of 10−12 A and a photocurrent of 1.13 μA under 375 nm UV illumination (18 mW cm−2), yielding an ultrahigh switching ratio (~106). This performance enhancement stemmed from suppressed hot-electron tunneling currents, enabling photon-driven carrier dominance. The optimized device demonstrated exceptional performance in the following metrics: photoconductive gain (8.6 × 105), responsivity (2.6 × 105 A W−1), and detectivity (2.3 × 1016 Jones) at 0.01 mW cm−2 bias. Temporal response analysis (Figure 9c) revealed rapid dynamics, with 12.6 ms rise and 180 ms decay times. Complementarily, Zhao et al. [149] synthesized low-defect CdS nanorods (NRs) on SiO2/Si substrates via CVD, fabricating Ti/Au-electrode photodetectors through lithography. The devices exhibited exceptional blue-light sensitivity (450 nm, 0.5 mW cm−2), with responsivity peaking at 1.23 × 104 A W−1 (Figure 9d). Further performance benchmarks included an external quantum efficiency of 3.5 × 106%, a detectivity of 2.8 × 1011 Jones, and response/recovery times of 0.82/0.84 s.
Self-powered photodetectors, capable of achieving optoelectronic conversion and detection without an external power supply, demonstrate significant potential for energy-efficient applications. Chai et al. [150] synthesized core–shell Sb/CdS nanowires through a two-step chemical vapor deposition (CVD) process, and the device architecture is illustrated in Figure 10a. The single NW-based PDs were prepared by photolithography, thermal evaporation, and stripping processes at both ends of a single NW on a Si/SiO2 substrate in order to construct Cr/Au source/drain electrodes. The device boasts superior photocurrent performance, and Figure 10b presents the increasing photocurrent and built-in open-circuit voltage with rising light intensity at different light intensities for 470 nm illumination. Figure 10c exhibits the time-dependent photoresponse of the device in self-powered mode at a 700 nm wavelength. The photocurrent remains virtually unchanged even after hundreds of switching cycles, demonstrating remarkable stability and reproducibility. It exhibits excellent stability and reproducibility with a substantially enhanced photocurrent (12.85 pA) versus dark current (45.5 fA), achieving a high on/off current ratio of 3.54 × 103. Its better optical response speed is demonstrated by the single optical on/off cycle transient response in Figure 10d, whose rise time and recovery time are less than 0.384 s and 0.312 s, respectively. Also, the device has high optical R, D*, and G values of 93.62 A/W, 2.33 × 1014 Jones, and 2.47 × 102 under illumination at the above wavelength with 0 V bias and a power of 19.12 μW/cm2, respectively, demonstrating excellent detection performance.
Conventional photodetectors relying solely on photovoltaic effects exhibit limited functionality in cryogenic environments due to their significantly degraded response at low temperatures. To address this limitation, Chang et al. [151] engineered a SnS/CdS heterojunction photodetector synergizing pyroelectric (temperature-induced polarization changes generating transient currents [152]) and photovoltaic effects. The device architecture comprised hydrothermally grown CdS nanorod arrays on FTO glass, overlaid with thermally evaporated SnS films and Au electrodes (Figure 11a). Temperature- and power-dependent analyses (Figure 11b) reveal enhanced pyroelectric currents with increasing optical power (0.08 mW cm−2) and decreasing temperature (300 → 130 K), demonstrating the effective substitution of carrier migration (photocurrent) with lattice-vibration-driven currents (pyroelectric) for low-temperature operation. Figure 11c quantifies this transition, showing pyroelectric dominance (Ipyro/Iphoto ratio increasing from 10% to 400%) under cryogenic, low-irradiance conditions. At 650 nm illumination, the device achieves 10.4 mA W−1 responsivity and 3.56 × 1011 Jones detectivity at zero bias, with sub 30 ms response/decay times. Complementing this approach, Ren et al. [86] developed self-powered p-n heterojunction detectors through the hydrothermal deposition of n-CdS nanorods onto p-Si substrates, capped with sputtered ITO/Ag electrodes. The devices exhibit a broadband response (405–1064 nm), with 64.8 mA W−1 responsivity and 1.31 × 1010 Jones detectivity at 2.55 mW cm−2, alongside rapid 190.8/298.4 μs response/recovery dynamics under zero bias. These dual-mechanism architectures establish new paradigms for extreme-environment photodetection systems.
To investigate the influence of heterojunction interfacial structures on the optoelectronic performance of devices, our research team fabricated a photodiode (Figure 12a) by constructing head-to-head PANI/CdS p-n heterojunction nanowire arrays with ITO/Au composite electrodes [153]. The device exhibited excellent rectification characteristics and diode behavior, demonstrating high sensitivity in the visible spectrum, particularly at 420 nm. As shown in Figure 12b, the rectification ratio of the PANI/CdS heterojunction nanowire array reached 34.1 under 5.21 mW cm−2 illumination, representing a three-fold enhancement compared to the dark-state ratio (11). Repeatability and response speed are critical parameters for evaluating a photodetector’s ability to track rapidly varying optical signals. Figure 12c illustrates the switching cycle performance under 420 nm illumination, where the device exhibited a dark current of 0.21 μA and a photocurrent of 1.23 μA, achieving an on/off ratio of 5.8 with high stability and repeatability. The rapid response dynamics arise from the high surface-to-volume ratio of the PANI/CdS heterojunction nanowires, coupled with surface defects and dangling bonds that act as recombination centers, accelerating free carrier recombination and thereby reducing decay time. We also regulated device performance by constructing heterointerfaces with diverse morphologies, including PTCM (poly{3-thiophene carboxylic acid methyl ester})/PbS axially deeply inserted heterojunction nanowire arrays with enlarged interfacial areas, which exhibited superior performance, with a switching ratio of up to 83.5 [154]. To achieve infinite interfacial contact and enhance carrier separation efficiency, we synthesized homogeneous CdS/PPV (poly(p-phenylene vinylene)) hybrid nanowire arrays (NWAs) via porous anodic aluminum oxide (AAO) template-assisted electrochemical co-deposition. Prior to deposition, Au electrodes were sputtered onto one side of the AAO template, followed by ITO glass coating to fabricate novel photodetectors [155]. Compared to pure-CdS nanowire arrays, the CdS/PPV hybrid NWAs demonstrated a significantly enhanced photoresponse. As shown in Figure 12d, under 545 nm illumination at 4.2 mW/cm2, the current increase in the homogeneous device (schematic in Figure 12d inset) far exceeded that of pure-CdS-based devices. The CdS/PPV hybrid NWA device exhibited a stable photocurrent over multiple light on/off cycles, with a dark current of 0.027 μA at a 5 V bias and a stabilized photocurrent of 1.457 μA under illumination, achieving an exceptional switching ratio of 53.4 (Figure 12e), approximately 17 times higher than those of pure-CdS devices (2.8). This improvement primarily stems from PPV acting as a bridge to consume or transfer photogenerated holes and electrons, thereby suppressing electron–hole recombination and enhancing the photoresponse. Under the same illumination, Figure 12f shows that the device’s on/off ratio significantly increased with rising light intensity across a bias range of −5 V to +5 V, confirming the suitability of CdS/PPV hybrid NWAs for 545 nm green-light detection. Based on these findings, we systematically summarized and analyzed recent advances in sulfide-based photodetectors and p-n heterojunction self-powered devices. Additionally, by optimizing the preparation of ZnO nanorod arrays, we established a solid foundation for future studies on ZnO-template-based sulfide photodetectors [63,156].
To extend the spectral detection range of CdS-based devices, Rani et al. [157] synthesized TiO2/CdS nanorod heterostructures via hydrothermal synthesis, fabricating UV photodetectors with spin-coated Ag electrodes. The devices exhibit an exceptional photocurrent-to-dark current ratio of 439.66 (Figure 13a), enabled by ultralow dark currents, which enhance detection sensitivity. Under 365 nm illumination, responsivity and external quantum efficiency (EQE) peak at 2.865 A W−1 and 971.36%, respectively, at 40 μW cm−2 (Figure 13b,c). The anomalous EQE >100% arises from synergistic interfacial phenomena: built-in electric fields in the heterojunction, Type-II band alignment prolonging carrier lifetimes, and suppressed recombination through graded band structures [158]. This performance is further quantified by a Noise Equivalent Power of 3.64 × 10−9 W Hz−1/2 and a detectivity of 9.90 × 1012 Jones. Temporal response profiling (Figure 13d) reveals rapid dynamics, with a 0.99 s rise time and a 0.49 s decay time. These results underscore bandgap-engineered heterostructures as critical enablers of broadband, high-performance photodetection systems.

Tin Sulfide (SnS)-Based Photodetectors

SnS is a bandgap-tunable p-type semiconductor material with indirect and direct bandgap energies of approximately 1.0–1.3 eV and 1.3–1.5 eV, respectively [159,160]. It exhibits well-tuned conductance, high carrier mobility, and wide wavelength absorption. This allows SnS to be employed in a multitude of applications, including field-effect transistors, spin-valley lasers, and photodetectors [161,162,163].
P-type SnS nanowires (NWs) hold critical importance for constructing advanced p-n heterostructures. Zheng et al. [164] fabricated p-type SnS NWs via chemical vapor deposition (CVD), transferring individual NWs onto SiO2/Si substrates. Electron-beam lithography defined the drain/source contacts, followed by the thermal evaporation of Cr/Au electrodes to complete the photodetector assembly. The device exhibits an optical gain of 2.8 × 104 under 830 nm illumination (0.05 mW cm−2), coupled with remarkable responsivity (1.6 × 104 A W−1) and detectivity (2.4 × 1012 Jones). Temporal response profiling reveals ultrafast dynamics with a 1.2 ms rise time and 15.1 ms decay time. The asymmetric response kinetics likely originate from trap-mediated carrier dynamics, where surface states or defect centers delay trapped carrier release, extending excess charge persistence.
In a separate study, Chen et al. [165] synthesized p-type SnS nanowires (NWs) with uniform nano-scale diameters via chemical vapor deposition (CVD), transferring them onto SiO2/Si substrates for device fabrication. Thermally evaporated Cr/Au electrodes were patterned to construct photodetectors, which demonstrated high responsivity (2.6 × 102 A W−1), photoconductive gain (3.9 × 102), and detectivity (1.8 × 1013 Jones) under 830 nm illumination (0.12 mW cm−2, 1 V bias). The devices exhibited ultrafast response dynamics, with rise and decay times of 9.6 ms and 14 ms, respectively. This enhanced performance originated from the NWs’ reduced diameter and exceptional crystallinity, which minimized surface defects and established efficient carrier transport pathways, thereby accelerating generation–recombination kinetics. These results highlight the potential of low-dimensional p-type SnS architectures for next-generation photodetectors requiring high sensitivity, rapid response, and energy-efficient operation.

Comparative Analysis

This section systematically reviews synthesis methodologies and performance characteristics of one-dimensional (1D) monosulfide-based photodetectors. Comparative analysis reveals that nanotube architectures with hollow-core geometries surpass their nanowire/nanorod counterparts in photodetection efficiency, attributable to the enhanced light trapping through their tubular cavity resonance effects. For instance, ZnS nanotube (NT) devices demonstrate superior responsivity (16.5 A W−1 vs. 2.761 A W−1 for nanowires) and photocurrent generation (508 pA). Ferroelectric polymer-polarized CdS nanowire detectors achieve ultralow dark currents (10−12 A) through carrier depletion mechanisms, yielding exceptional detectivity (2.6 × 1016 Jones) exceeding conventional structures by 2–4 orders of magnitude. Strategic doping substantially modulates semiconductor conductivity, exemplified by Al-doped ZnS detectors exhibiting record responsivity (3.1 × 106 A W−1) and photoconductive gain (1.5 × 107). Heterojunction engineering simultaneously reduces interfacial recombination via built-in electric fields while lowering carrier injection barriers, enabling ultrafast response dynamics. The ZnS/CdS nanowire heterostructure exemplifies this dual functionality, achieving broadband UV-vis detection (2.23 × 1014 Jones) through complementary bandgap alignment. Notably, SnS/CdS composite detectors demonstrate cryogenic operability by coupling photovoltaic and pyroelectric effects, attaining 10.4 mA W−1 responsivity and 3.56 × 1011 Jones detectivity under low-temperature conditions. Table 4 quantitatively benchmarks these performance metrics, establishing 1D monosulfide architectures as versatile platforms for extreme-environment photodetection technologies.

3.2.2. Photodetectors Based on Disulfide Metal Compounds

Transition metal disulfides exhibit superior optoelectronic characteristics, stemming from their unique crystallographic and electronic configurations. These materials demonstrate exceptional carrier mobility and ultrafast response dynamics, making them ideal for high-sensitivity photodetection systems—SnS2 for UV-vis spectral ranges and WS2 for visible–NIR regimes. Current disulfide-based photodetector research remains predominantly concentrated on lamellar MoS2 and ReS2 architectures at the two-dimensional (2D) level. However, 2D systems suffer from ambient instability and interfacial degradation under prolonged operational stresses, fundamentally limiting device longevity. In contrast, single-crystalline 1D disulfide nanostructures provide atomically smooth surfaces with minimal defect densities, enhanced surface-to-volume ratios, and intrinsic structural anisotropy along the longitudinal axis, which enables directional carrier transport with reduced scattering losses.

Selenium Sulfide (SnS2)-Based Photodetectors

SnS2, as a homodimeric derivative of SnS, emerges as an environmentally benign n-type semiconductor characterized by chemical stability and an indirect bandgap spanning 2.1–2.4 eV [166,167]. Its strong photoconductive response and superior visible-range absorption coefficients (α > 105 cm−1) have driven its extensive application in photovoltaics, photocatalytic systems, and broadband photodetection technologies [168,169,170].
To explore one-dimensional SnS2 synthesis pathways and to improve SnS2 applications, Zervos et al. [171], by investigating the conversion of SnO2 to SnS2 NWs, found that the tetragonal SnO2 NW structure can be completely converted to a hexagonal SnS2 NW structure at 400 °C. This represents a novel approach to the synthesis of SnS2 NWs, and it paves the way for new avenues of research in the field of one-dimensional SnS2-based photodetectors. Subsequently, Zheng et al. [172] synthesized n-type SnS2 nanowires (NWs) via gas-phase deposition and spin-coated a ferroelectric polymer layer (polyvinylidene fluoride-trifluoroethylene, P(VDF-TrFE)) onto the NWs. Photodetectors were fabricated by depositing Cr/Au electrodes via vacuum thermal evaporation and vapor stripping, creating a localized ferroelectric polarization field (Figure 14a). Under a −17.2 V side-gate voltage, the device exhibited an ultralow dark current (3 × 10−13 A) and an elevated photocurrent (2.74 μA), achieving a switching ratio of 107. This behavior arises from the ferroelectric polymer’s polarization, which induces the depletion of free electrons in the SnS2 channel via a high-intensity electrostatic field. To evaluate photoresponse efficiency, the ferroelectric gate polarization was deactivated (voltage terminated post-activation), and the device was tested under 520 nm illumination with a 1 V bias (Figure 14b). Photoconductive gain and responsivity scaled with light intensity, reaching maxima of 4.0 × 105 and 2.1 × 105 A W−1 at 0.06 mW cm−2. Suppressed dark current noise enhanced detectivity to 1.3 × 1016 Jones. The time-resolved photoresponse was conducted under 520 nm illumination, shown in Figure 14c. The SnS2 NW device exhibited good switching periodicity and stability, and its Ilight/Idark ratio was as high as 107. The ferroelectric polarization effect further accelerated electron–hole pair dynamics, yielding rapid transient response characteristics with a 56 ms rise time and 91 ms decay time.
The inherent built-in electric fields within heterostructures drive the directional migration of photogenerated electrons and holes, forming depletion layers while leveraging bandgap disparities to broaden spectral absorption ranges, thereby significantly enhancing photoresponse and detection capabilities. Das et al. [173] fabricated Si/SnS2 nanowire heterojunctions via hydrothermal synthesis, integrating Au/Al electrodes through thermal evaporation (Figure 14d). The devices demonstrated ultralow dark currents (2.86 × 10−12 A at 0 V), surpassing their spin-coated SnS2/Si counterparts (2 × 10−10 A) [174]. Under 340 nm illumination (20 nW mm−2), the detectors achieved exceptional responsivity (383 A W−1) and external quantum efficiency (1.3 × 105%), attributed to bias-activated (2 V) defect-state electron release, which induced multiplicative interface carrier generation and expanded photogeneration zones. Temporal response analysis (Figure 14f) revealed rapid 10–90% rise (0.55 s) and 90–10% decay (0.33 s) times, confirming superior response kinetics. These advancements position 1D SnS2 heterostructures as promising candidates for next-generation UV-vis photodetection technologies.
Figure 14. Ferroelectric single-side door control SnS2 NWs-based PD: (a) schematic, (b) photoconductor gain and responsivity at V = 1 V, (c) photocurrent response of the device under optical chopping (520 nm, 11 mW cm−2) at Vds = 1 V, without additional gate voltage. Reprinted with permission from Ref. [172]. Copyright 2021, Walter de Gruyter. Si/SnS2 NWs heterojunction devices: (d) PD schematic, (e) I-V curve under logarithmic conditions, (f) time-resolved optical response, rise (tr) and fall time (tf). Reprinted with permission from Ref. [175]. Copyright 2021, American Institute of Physics.
Figure 14. Ferroelectric single-side door control SnS2 NWs-based PD: (a) schematic, (b) photoconductor gain and responsivity at V = 1 V, (c) photocurrent response of the device under optical chopping (520 nm, 11 mW cm−2) at Vds = 1 V, without additional gate voltage. Reprinted with permission from Ref. [172]. Copyright 2021, Walter de Gruyter. Si/SnS2 NWs heterojunction devices: (d) PD schematic, (e) I-V curve under logarithmic conditions, (f) time-resolved optical response, rise (tr) and fall time (tf). Reprinted with permission from Ref. [175]. Copyright 2021, American Institute of Physics.
Jcs 09 00262 g014

Tungsten Sulfide (WS2)-Based Photodetectors

One-dimensional WS2, with similar hexagonal symmetry to the ZnS structure, is a direct bandgap semiconductor (1.3–1.95 eV) [176] which is widely used in photodiodes [177], photodetectors [178], and solar devices [179], among others. The high electron mobility, optical responsiveness, and luminescence-enhancing properties of one-dimensional WS2 reveal potential promise for nanostructured devices and applications [180,181].
To investigate the light detection performance of photodetectors based on WS2 nanotubes (NTs), Zhang et al. [54] synthesized WS2 nanotubes (NTs) via hydrogen reduction and the sulfidation of WO3 nanoparticles in a scalable fluidized bed reactor. Photodetectors were fabricated through electron-beam lithography and the thin-film deposition of Ti/Au electrodes onto individual WS2 NTs aligned on SiO2/Si substrates. The devices demonstrated wavelength-dependent photoresponses, exhibiting superior photocurrents under 633 nm illumination (532 W cm−2, 0.5 V bias) compared to 785 nm excitation (600 W cm−2) (Figure 15a). Notably, the on/off ratio reached 336 under 633 nm (5.3 kW cm−2) illumination—ten-fold higher than the 32 ratio observed at 785 nm (6 kW cm−2). This disparity originates from 633 nm photons (1.96 eV) inducing direct bandgap transitions in WS2 NTs, while 785 nm photons (1.58 eV) only activate indirect transitions. The devices achieved peak responsivity (2360 A W−1) and external quantum efficiency (4.6 × 106%) at 633 nm (0.5 V bias), with retained functionality in the near-infrared region (785 nm: 2.88 A W−1, 454% EQE). The superior crystallinity of WS2 NTs minimizes surface dangling bonds, effectively suppressing persistent charge trapping. Cyclic illumination testing at a 633 nm wavelength (Figure 15b) revealed ultrafast the response dynamics and cyclic stability of the devices, with precisely measured rise time (tr = 256 μs) and decay time (tf = 286 μs), as quantified in Figure 15c. The exceptional properties of organic materials—flexibility, optical transparency, and cost-effectiveness—have driven significant interest in organic–inorganic hybrid photodetection systems. Liu et al. [182] developed hybrid photodetectors by solution-processing a composite of WS2 nanotubes (NTs) and poly(N-vinylcarbazole) (PVK), spin-coated onto Au-patterned SiO2/Si substrates. Current-voltage (I-V) characterization under varied illumination (Figure 15d) revealed semiconducting behavior in the WS2-PVK hybrid film, with nonlinear curves bypassing the origin, indicative of self-powered operation. Under 552 nm (0.8 mW cm−2) and 661 nm (0.2 mW cm−2) illumination, the device exhibited faster response dynamics (0.92 s rise/0.09 s decay) for green light compared to red light (2 s/3.34 s) (Figure 15e,f), correlating with enhanced shorter-wavelength absorption. At 0.06 mW cm−2 (zero bias), the detector achieved 0.2 mA W−1 responsivity and 2.23 × 104 Jones detectivity. These results position WS2 NT-based hybrids as promising platforms for visible/NIR high-speed photodetection and optical-switching applications.
In other works, Chaudhary and Khanuja [183] synthesized heterostructures based on MoS2-NFs/WS2-NRs using a hydrothermal technique and prepared a photodetector structure using a drop-casting technique with silver as the electrode; the schematic diagram is reported in Figure 16a. The photocurrents under different irradiation wavelengths are plotted in Figure 16b for an applied power supply of 12 V and a power of 50 mW/cm2. As opposed to the photocurrents at 635 nm and 1064 nm (0.29 μA, 0.55 μA), the photocurrent can be as high as 0.75 μA at NIR (785 nm), and this high photocurrent gives the device an excellent photoresponse, with a value of 15 μA/W, demonstrating excellent near-infrared (785 nm) light sensitivity. The device also has the minimum near-infrared Noise Equivalent Power (16.89 × 10−6) under the above external conditions, which enables it to achieve the best detection performance. For example, the external quantum efficiency and specific detection rate have extremely high values compared to visible (635 nm) and far-infrared (1064 nm) light, as seen in Figure 16c,d, of 16.89 × 10−6% and 24 × 106 Jones, respectively. In addition, the composite device has a response speed of 0.82 s/1.59 s. This suggests that the heterostructure can be used as a photodetector for broadband spectroscopy, especially in the near-infrared region.

Other Metal Disulfide (MoS2, ReS2)-Based Photodetectors

Despite advancements in layered MoS2-based photodetectors, their suboptimal light absorption and complex fabrication processes have constrained their commercial viability. To address this, Gao et al. [184] successfully prepared high-purity n-type MoS2 square nanotubes by a simple hydrothermal method and constructed a photodetector by spin-coating a solution containing MoS2 nanotubes on a PET substrate and uniformly coating both ends of the nanotubes as electrodes with silver conductive paste. As illustrated in Figure 17a, when light is irradiated on MoS2 nanotubes, the device exhibits excellent optoelectronic performance in the near-infrared (NIR) light band, which is mainly due to the fact that the surface of the nanotubes consists of nanosheets of S-vacancy-rich nanosheets, which provide high responsivity in the NIR region, and in addition, the unique luminal structure of the nanotubes also reflects and scatters the incident light multiple times and absorbs the incident light more completely. The I-V curves of the device at 915 nm under different optical powers are presented in Figure 17b, and the photocurrent is as high as about 2 μA at the same bias voltage when the light intensity reaches 100 mW/cm2. Under fixed voltage and wavelength with varying optical power levels, the device maintains consistent dynamic photoresponse and initial current through multiple illumination cycles, demonstrating remarkable photoresponse reproducibility and exceptional operational stability (Figure 17c). Performing a high-resolution display of a switching cycle, as described in Figure 17d, the device shows a good optical response speed with a rise/recovery time value of 5.3 s/1.53 s. Its high photogenerated current and response speed enable the device to have excellent detection performance at a 915 nm near-infrared light wavelength (100 mW/cm2 intensity), including an excellent photoresponse, specific detectivity, and external quantum effect, as shown in Figure 17e,f with the bias voltage gradually increasing to 3 V and values as high as 2.33 mA/W, 7.55 × 108 Jones, and 3.33 × 10−1%, and these results provide new possibilities for the development of novel one-dimensional MoS2-based photodetectors.
Li et al. [185] fabricated MoS2 nanotubes (NTs) on ITO substrates via anion-exchange synthesis, subsequently integrating Y-TiOPc nanoparticles (NPs) synthesized by microemulsion phase-transfer and p-TPD (N,N’-diphenyl-N,N’-di-p-toluylbenzidine) layers via spin-coating to form a photoactive heterostructure. Au electrodes were deposited via magnetron sputtering, yielding a Y-TiOPc/MoS2 photodetector with a broadband UV-vis-NIR spectral response. Compared to pristine MoS2 devices, Y-TiOPc-induced interfacial band bending enhances electron trapping and hole injection efficiency, significantly improving carrier generation, separation, and transport dynamics. At 20 V reverse bias under 0.01 mW cm−2 illumination (365–850 nm), the device achieves peak responsivity (20,588 mA W−1), external quantum efficiency (4947.6%), and detectivity (1.94 × 1012 Jones). Notably, it demonstrates an ultrafast temporal response (134 ms rise/143 ms decay), surpassing conventional MoS2 NT-based photodetectors in both speed and stability. In addition, Varshney et al. [186] synthesized β-Ga2O3 by thermally oxidizing a GaN epitaxial layer on a sapphire substrate, followed by sputter-depositing MoS2 nanorods to form MoS2/β-Ga2O3 heterojunctions. High-performance broadband (UV-vis-NIR) self-powered photodetectors were fabricated by patterning Pt electrodes via DC magnetron sputtering at both ends of the heterostructure, as schematically illustrated in Figure 18a. The devices exhibits ultralow dark currents of approximately 79 nA under dark conditions. The bias-dependent I-T characteristics were systematically evaluated under ultraviolet (266 nm) and near-infrared (950 nm) illuminations, with the applied biases ranging from 0.5 to 5 V and a constant optical power of 180 μW (Figure 18b,c). The devices maintained exceptional current stability and measurement reproducibility through multiple switching cycles under various environmental conditions. The maximum photocurrents and response speeds (tr and tf) reached 16.92 μA, 9.42 μA, 0.29/0.3 s, and 0.15/0.19 s, respectively. The disparity arose because 266 nm illumination required a longer duration for photogenerated carriers to excite from the valence band to the conduction band and establish photocurrent. Under a 5 V bias (Figure 18d), the devices exhibited power-dependent responsivity enhancement across a broadband UV-vis-NIR spectral range. At 266 nm UV illumination (30 μW cm−2), peak responsivity reached 42.11 A W−1. Concurrently, the detectors achieved a specific detectivity of 3.2 × 1011 Jones and an external quantum efficiency (EQE) of 1.97 × 104% under identical conditions, maintaining high-performance detection from UV to NIR wavelengths (Figure 18e,f). These results demonstrate a viable pathway for developing broadband 1D MoS2-based photodetectors with tailored spectral operability.
ReS2 has emerged as a highly promising optoelectronic material [187], though current research predominantly focuses on its two-dimensional (2D) configurations, with investigations into one-dimensional (1D) ReS2 architectures remaining scarce. Addressing this gap, An et al. [188] pioneered the synthesis of single-crystalline ReS2 nanowires via chemical vapor deposition (CVD), fabricating photodetectors with Ag nanowire networks as transparent conductive electrodes. Under 3.0 V bias, the devices demonstrated substantial photocurrent switching ratios, with currents rising from 0.42 nA (dark) to 4.95 nA (illuminated). Temporal response analysis revealed rise and decay times of 1.8 s and 3.9 s, respectively, indicating competitive temporal response characteristics for emerging low-dimensional optoelectronics. To comprehensively assess the photodetection performance of the ReS2 nanowire (NW)-based photodetectors (PDs), the devices were evaluated under both ambient and vacuum conditions under 500 nm illumination (0.42 nW cm−2). In air, the PDs achieved exceptional responsivity (5.08 × 105 A W−1), external quantum efficiency (EQE: 1.07 × 106%), and specific detectivity (6.1 × 1015 Jones). Under vacuum, these metrics decreased to 2.03 × 105 A W−1 (Rλ), 4.3 × 105% (EQE), and 6.1 × 1014 Jones (D*), respectively. This performance discrepancy is attributed to the absence of oxygen molecules in the vacuum, which suppresses surface oxygen adsorption and the associated electron scavenging, thereby promoting bulk carrier recombination. The superior optoelectronic metrics position ReS2 NWs as leading candidates for next-generation high-performance nanostructured photodetectors.

Comparative Analysis

This section comprehensively examines photodetectors utilizing one-dimensional disulfide metal compounds, emphasizing structural design and performance optimization through comparative device analysis. Optimized nanostructures with tunable bandgaps and enhanced carrier mobility prove critical for achieving superior photodetection metrics. Notably, 1D disulfide architectures demonstrate unique broadband spectral responsivity, exemplified by chemically stabilized SnS2 nanowire-based devices exhibiting exceptional UV-vis performance in the following metrics: responsivity (2.1 × 105 A W−1), external quantum efficiency (4.0 × 105%), and detectivity (1.3 × 1016 Jones). Hollow-structured WS2 nanotube detectors showcase visible-light specificity with ultrafast response dynamics (256 μs rise/286 μs decay), outperforming conventional nanowire systems by orders of magnitude. Heterostructured designs, such as thermally oxidized β-Ga2O3/MoS2 configurations, achieve noise suppression (dark current: 79 nA) and accelerated response kinetics through interfacial band engineering. Single-crystalline CVD-grown ReS2 nanowire detectors further highlight material potential with record metrics: 5.08 × 105 A W−1 responsivity, 6.1 × 1015 Jones detectivity, and 1.07 × 106% EQE. Table 5 systematically benchmarks these advancements, underscoring 1D disulfide architectures as transformative platforms for high-sensitivity, broadband photodetection technologies.

3.2.3. One-Dimensional Polysulfide Metal Compound-Based Photodetectors

Variations in sulfur stoichiometry induce substantial structural and functional divergences in metal sulfide compounds. In contrast to low-sulfur-content materials like ZnS and SnS2, higher-sulfur-stoichiometry systems such as Sb2S3, Bi2S3, and In2S3 (metal-to-sulfur ratio 2:3) exhibit reduced bandgap energies (<2 eV), rendering them ideal for broadband spectral photoconversion with accelerated carrier dynamics. Beyond bandgap modulation, increased sulfur content enhances crystallinity, which passivates surface defects and optimizes charge transport. These synergistic effects significantly elevate device performance metrics, particularly carrier mobility and temporal photoresponse characteristics.

Indium Sulfide (In2S3)-Based Photodetectors

Butanov’s group [189] synthesized In2S3 nanowires on Au-catalyzed Si/SiO2 substrates via atmospheric-pressure chemical vapor transport (AP-CVT), fabricating single-nanowire photodetectors through photolithographic patterning and thermal evaporation of Cr/Au electrodes. These devices exhibited ultralow dark currents (9.95 nA at 1 V bias), enabling superior photoresponsive behavior. Under 405 nm illumination (2 W cm−2), they achieved a switching ratio of 11.3, with responsivity and external quantum efficiency (EQE) reaching 16.01 A W−1 and 4903%, respectively, demonstrating robust photovoltaic conversion capabilities.
In addition, Xie and Shen [190] synthesized single-crystalline In2S3 nanowires via chemical vapor deposition (CVD), fabricating flexible photodetectors by lithographically patterning Au/Ni source/drain electrodes on individual nanowires transferred to PET substrates (3D schematic: Figure 19a). These devices demonstrated pronounced visible-range photoresponses, exhibiting strong wavelength-dependent photocurrent generation (300–600 nm, Figure 19b). Under 450 nm illumination (176.7 W cm−2), photoinduced carrier multiplication significantly enhanced the photocurrent to 293 nA at 5 V bias, yielding an exceptional switching ratio (~106) relative to the dark current (0.29 pA). Photoresponsivity and external quantum efficiency (EQE) peaked at 7.35 × 104 A W−1 and 2.28 × 107% (400 nm, 5 V bias), respectively, with concurrent detectivity reaching 2.4 × 1014 Jones (Figure 19c), establishing In2S3 nanowires as high-performance flexible photodetection platforms. Moreover, the devices demonstrate highly stable and reproducible characteristics under different bias voltages. The time-dependent photoresponse of the detectors, as shown in Figure 19d, was measured by toggling 450 nm light illumination with a 10 s on/off cycle. Under a constant optical power density of 177 μW/cm2 and different bias voltages (2 V, 4 V, and 8 V), the current values remained nearly identical to their initial magnitudes, demonstrating highly stable and reproducible characteristics across all three operational conditions. After repeating the switching many times, the photocurrent and the dark current were almost at the same level as in the initial state, and the photoresponse time and the attenuation time were only 6.5 ms/9.5 ms, an excellent photoresponse speed. When the devices were bent at bend angles from 0° to 120°, the photocurrents generated were almost the same as that of the unbent state (0°). These results show that In2S3 NW devices have a promising future in the application of high-performance and wearable optoelectronic devices.

Antimony Sulfide (Sb2S3)-Based Photodetectors

Antimony sulfide (Sb2S3), an inorganic n-type semiconductor crystallizing in an orthorhombic system, possesses a direct bandgap ranging between 1.7 and 1.8 eV [191,192]. This unique combination of structural and electronic characteristics enables broad-spectrum photoresponsivity, spanning ultraviolet–visible to near-infrared regions, which has driven its versatile utilization across multiple technological domains. Its current applications encompass photovoltaic devices, photocatalytic hydrogen evolution, Li/Na/K-ion battery systems, photodetection architectures, and related advanced functional materials [193,194,195,196].
One-dimensional Sb2S3 nanostructures have garnered significant research interest, owing to their well-defined crystallinity, exceptional thermal stability, and enhanced optical absorption coefficients relative to other 1D nanomaterials [197]. In this context, Zhong et al. [198] synthesized Sb2S3 nanowires (NWs) on SiO2/Si substrates via sulfur-assisted vapor-phase deposition, subsequently fabricating NW-based photodetectors through a sequence of lithographic patterning, the electron-beam deposition of Au source/drain electrodes, and lift-off processes. Under 638 nm illumination (1 V bias, 0.03 mW/cm2 optical power), the devices demonstrated remarkable optoelectronic performance: an ultrahigh responsivity of 1.15 × 103 A/W, a specific detectivity of 2 × 1013 Jones, and a low dark current of 2.0 × 10−10 A. At an elevated optical power (45.2 mW/cm2), a substantial photocurrent of 4.2 × 10−8 A was achieved, yielding an impressive switching ratio of 210. Temporal response characterization revealed rapid rise/fall times of 37/38 ms, confirming their efficient photo-switching capabilities. Concurrently, Zhao’s team [199] developed optically anisotropic quasi-1D Sb2S3 NWs on SiO2/Si using sulfur-assisted gas-phase transport, integrating polarization-sensitive photodetectors via lithographically defined Au contacts. The optimized devices exhibited a maximum current on/off ratio of 69.6 under 450 nm illumination (2 V bias, 40 μW/cm2), coupled with a responsivity of 343.4 mA/W. At 520 nm excitation, ultrafast photoresponse dynamics were observed, with rise/fall times (tr/tf) of 470/680 μs, highlighting sub-millisecond operational speeds.
To mitigate surface-adsorbed impurities (e.g., metal–organic precursors, surfactants, and reducing agents) that degrade photodetector performance, Ye et al. [200] developed high-crystallinity Sb2S3 nanowires (NWs) on SiO2/Si substrates via chemical vapor deposition (CVD). Individual NW devices were fabricated by thermally evaporating Ni/Au electrodes at both termini. As illustrated in Figure 20a, the devices exhibit broadband spectral responsivity (360–785 nm) under 28 mW/cm2 illumination, with a peak photocurrent of 2.5 nA at 532 nm. This phenomenon arises from the bandgap of Sb2S3 being approximately 1.5–1.6 eV (corresponding to an absorption edge wavelength of ~600–700 nm). When the incident light wavelength is shorter than this critical wavelength (i.e., the photon energy exceeds the bandgap), the material efficiently absorbs photons and generates electron–hole pairs, leading to an enhanced photocurrent as the wavelength decreases (energy increases). When the wavelength exceeds 700 nm, the photon energy becomes insufficient to excite electronic transitions, causing a sharp decline in absorption efficiency and a significant reduction in photocurrent. At ultralow power (0.03 mW/cm2, 5 V bias), the device demonstrates exceptional charge transport efficiency (responsivity: 65 A/W) and superior figures of merit—specific detectivity (2.1 × 1014 Jones), external quantum efficiency (1.5 × 104%), and sensitivity (2.2 × 104 Jones). Cyclic stability testing reveals < 15.6% photocurrent degradation after 104 on/off cycles, coupled with rapid response dynamics (rise/fall times: 76/82 ms). Concurrently, Ma et al. [201] synthesized single-crystalline Sb2S3 NWs via atmospheric-pressure CVD on SiO2/Si, integrating Ti/Au electrodes to construct polarization-sensitive UV-vis photodetectors. As illustrated in Figure 20b, the devices demonstrate excellent operational stability and reproducibility under an irradiation of 318 mW/cm2, with the on/off ratio maintained at approximately 2800. At 635 nm (0.3 μW/cm2), they deliver an ultrahigh responsivity (270 A/W) and record external quantum efficiency (5.3 × 104%) and specific detectivity (4.37 × 1013 Jones). Notably, the device exhibits ultrafast response/recovery kinetics (10/12 ms) and polarization-dependent photosensitivity at 405 nm, achieving a dichroic ratio of 7.2—a benchmark performance among low-dimensional polarized photodetectors.
To achieve precise control over crystallinity, purity, and defect density in Sb2S3 nanomaterials for high-performance photodetectors, Liu’s group [202] engineered single-crystalline Sb2S3 nanorods (NRs) with tunable aspect ratios via hydrothermal synthesis followed by H2/Ar ambient thermal annealing. Photodetectors were fabricated through electron-beam deposition of Au electrodes onto Sb2S3 NRs arrays on SiO2/Si substrates. Under 560 nm illumination (110 μW/cm2), the devices demonstrated a pronounced switching ratio of 109.8. As illustrated in Figure 20c,d, the photoresponse metrics exhibited an inverse correlation with optical power density, attributed primarily to enhanced carrier recombination losses at higher excitation intensities. At 0.038 mW/cm2, the detector achieved peak performance parameters in the following metrics: responsivity (5.10 A/W), external quantum efficiency (1130.68%), specific detectivity (2.16 × 1010 Jones), and ultrafast response kinetics (rise/decay times: tr/tf = 4.03/4.08 ms). These results underscore the direct correlation between crystallographic perfection and enhanced optoelectronic functionality in nanostructured devices.
Tubular architectures demonstrate enhanced photon-harvesting efficiency compared to nanowire and nanorod configurations through their augmented light-trapping capabilities. Zhang et al. [203] prepared single-crystalline Sb2S3 microtubes via chemical vapor deposition (CVD). Despite micrometer-scale tube dimensions, the fabricated devices exhibited exceptional photoresponsivity under 808 nm laser illumination (8 V bias, 300 mW/cm2), achieving a photocurrent of 185 nA, a switching ratio of ~200, and rapid response/recovery times (tr/tf = 22/24 ms) in two-terminal measurements. The detector further demonstrated superior figures of merit—responsivity (8.5 mA/W), sensitivity (667 cm2/W), and specific detectivity (1.33 × 106 Jones)—indicating the efficient charge transport mechanisms of Sb2S3 microtubes and providing foundational insights for tubular photodetector optimization. In a separate study, Chao et al. [204] developed Sb2S3 nanowire arrays (NWAs) on fluorine-doped tin oxide (FTO) substrates through a templateless solvothermal reflow process, constructing photodetectors via FTO overlayer deposition. At 0.5 V bias under 640 mW/cm2 illumination, the devices exhibited 15.3-fold photocurrent enhancement (113.5 nA vs. dark current 7.4 nA). Rapid switching tests revealed sub-second response kinetics (rise/recovery: 0.52/1.1 s), underscoring the devices’ operational agility in dynamic light environments.
To enhance photodetector responsivity, Zhang et al. [205] engineered Au nanoparticle (NP)-functionalized Sb2S3 nanowires (NWs) through chemical vapor deposition (CVD) on silicon substrates. The NWs were transferred via contact printing to SiO2/Si or flexible polyethylene terephthalate (PET) substrates, followed by Cr/Au electrode deposition using lithography and thermal evaporation. Au NPs were subsequently integrated by solution-phase deposition to fabricate rigid/flexible UV-vis detectors. As shown in Figure 21a, under 700 nm illumination (10 V bias, 510 μW cm−2), the Au-NP-modified rigid device exhibited a 286-fold photocurrent enhancement (163 nA vs. 57 pA for pristine NWs). Transient response analysis (Figure 21b,c) revealed retained rise times (0.2 s) but significantly accelerated recovery kinetics (0.3 s vs. 2 s for unmodified devices), demonstrating plasmon-enhanced charge extraction. The flexible variant maintained spectral parity with its rigid counterpart across 350–700 nm (Figure 21d), exhibiting peak white-light responsivity (59.5 A/W) and detectivity (4.29 × 1010 Jones) at 680 μW cm−2. The dynamic testing of the device under a 10 V bias (Figure 21e) revealed a stable cycling performance under 350 nm (154 μW cm−2) and 700 nm (150 μW cm−2) light switching, with the photocurrent maintaining nearly identical magnitudes to its initial values after multiple cycles, particularly at 700 nm. Additionally, wavelength-dependent response/recovery times were observed: 0.29/0.59 s (350 nm) and 0.25/0.13 s (750 nm) (Figure 21f). The detector achieved superior figures of merit at 350 nm (detectivity: 3.24 × 109 Jones; photomultiplication: 2.02) and 750 nm (1.09 × 109 Jones; 16.0), confirming Au-NP-induced broadband performance enhancement.
The abundant p-n heterointerfaces in composite systems facilitate the efficient separation of photogenerated electron–hole pairs and enhanced carrier transport kinetics. Leveraging this principle, Wang et al. [206] developed Sb2S3 nanorod (NR)/CuSCN heterojunctions via sequential hydrothermal synthesis, constructing ITO/composite/ITO sandwich-structured devices for optoelectronic characterization. Under white-light illumination at zero bias, the photovoltaic effect-driven devices exhibited ultrafast response dynamics (rise/fall times: tr/tf = 0.18/0.15 s), confirming the superior self-powered photodetection capabilities of CuSCN/Sb2S3 heterostructures. In a parallel development, Jiang and Meng [207] synthesized Sb2S3 nanowire array (NWA)/Mo2C-C composites hydrothermally, fabricating photodetectors with Au interdigitated electrodes. Compared to pristine Sb2S3 NW devices, the heterojunction system demonstrated accelerated response kinetics (tr/tf = 52.7/79.2 ms) and a 150-fold enhancement in current switching ratio (1.5 × 102), underscoring the critical role of interfacial engineering in optoelectronic device optimization.

Bismuth Sulfide (Bi2S3)-Based Photodetectors

An ideal n-type semiconductor with excellent light absorption and a high absorption coefficient (~105 cm−1) and photon conversion efficiency, Bi2S3 has a bandgap of 1.3–1.7 eV [208,209]. In particular, the low cost, non-toxicity, and good electrical conductivity of the material make Bi2S3 show great potential for applications in UV, visible, and NIR broad-spectrum photodetection [70,210,211].
Bi2S3 emerged as a pivotal photoactive material due to its narrow bandgap, high absorption coefficient, and efficient photon-to-carrier conversion capabilities [212]. Ding et al. [213] synthesized phase-pure Bi2S3 nanorods via a facile hydrothermal route, fabricating photodetectors through the dip-coating of nanorod films onto Al2O3 substrates with Au interdigitated electrodes. The devices exhibited pronounced photovoltaic behavior under illumination (Figure 22a), demonstrating a two-order-of-magnitude enhancement in photocurrent versus dark current across the applied biases. The cyclic current–time curve under a 5 V bias voltage (Figure 22b) demonstrated exceptional operational reproducibility and stability (98% photocurrent retention over 100 cycles) and rapid response dynamics (tr/tf = 371.66/386 ms). Complementing these findings, Yu et al. [214] developed colloidal Bi2S3 nanorods through sulfur-stoichiometry-controlled synthesis, constructing ITO/Bi2S3/ITO photodetectors. Current–voltage characteristics (Figure 22c) under visible illumination (475/550/650 nm, 4.1 mW cm−2) showed their wavelength-dependent photoconductivity, with maximal red-light responsivity (650 nm) indicating spectral selectivity. Dynamic testing at 1 V bias (Figure 22d) confirmed their robust photostability (<5% signal decay) and competitive response kinetics (tr/tf = 0.3/0.6 s), positioning Bi2S3 as a high-performance candidate for fast-switching optoelectronic systems.
Li’s group [215] synthesized single-crystalline Bi2S3 nanowires (NWs) via chemical vapor deposition (CVD), fabricating photodetectors through e-beam lithography and thermal evaporation techniques with Cr/Au electrode deposition at NW termini. Under 700 nm illumination (1 V bias, 1.54 mW cm−2), the devices exhibited rapid photoresponse activation (<0.1 s rise time) during cyclic on/off modulation. The detectors achieved a responsivity of 3.57 A/W and an external quantum efficiency (EQE) of 633%, demonstrating sub 100 ms temporal resolution and repeatable on/off cycling stability. To develop near-infrared (NIR)-polarized photodetectors with superior crystallinity, Wang et al. [216] synthesized high-crystallinity Bi2S3 nanowires (NWs) via chemical vapor deposition (CVD), fabricating devices by sputtering Pt electrodes on SiO2-supported NWs. As illustrated in Figure 23a, the optimized NW photodetector demonstrates strong NIR photoresponsivity at 830 nm, outperforming previously reported nanorod-based analogs. Figure 23b displays the photoresponse of a single-Bi2S3-nanowire photodetector measured under laser illumination at wavelengths of 400 nm, 635 nm, and 830 nm (optical power density ≈ 250 mW/cm2). The device exhibits excellent signal stability and reproducibility across different wavelengths and power conditions, with rise/fall times of merely 12.25 ms. At an equivalent laser intensity (~250 mW/cm2), 830 nm illumination delivers higher photon flux compared to 400 nm and 635 nm light, generating more electron–hole pairs in the single Bi2S3 nanowire, consequently resulting in an enhanced photocurrent being observed under 830 nm irradiation. Power-dependent studies (Figure 23c,d) exhibit a trade-off between photocurrent amplification and responsivity degradation at elevated power densities (>52 mW cm−2), stemming from saturating carrier transport pathways under high photon flux. At 52 mW cm−2 (830 nm, 1 V bias), the device achieves peak performance metrics: responsivity (4.21 A/W), detectivity (1.64 × 1010 Jones), and external quantum efficiency (981.76%). Polarization-resolved measurements demonstrate anisotropic photocurrent modulation, with the maximal response at 0°/180° alignment (parallel to NW axis) and the minimal response at 90°/270° (orthogonally polarized), yielding a dichroic ratio of 1.79. Similarly, Yi’s group [217] synthesized phase-pure Bi2S3 nanowires via atmospheric-pressure chemical vapor deposition (APCVD), employing a damage-free transfer method to assemble devices by positioning individual nanowires on SiO2/Si substrates with lithographically patterned Ti/Au electrodes. The gate-modulated device (Vg = 30 V) demonstrated record optoelectronic metrics under 532 nm illumination (23.8 μW cm−2)—responsivity (2.376 × 104 A/W), external quantum efficiency (5.55 × 106%), and specific detectivity (3.68 × 1013 Jones)—surpassing those of previous architectures by orders of magnitude. Ultrafast response kinetics (tr/tf = 1/4.5 ms) and polarization-selective operation (dichroic ratio: 2.4 at 405 nm) were concurrently achieved, establishing Bi2S3 nanowires as a benchmark material for high-speed polarized photodetection in the visible spectrum.
To optimize photodetector performance through enhanced photogenerated carrier density, Xu et al. [218] fabricated Bi2S3 nanowire arrays (NWAs) on mica substrates via physical vapor deposition (PVD), engineering photodetectors with Au source/drain electrodes. The Bi2S3 NWA-based devices demonstrated exceptional photoresponsivity, achieving 5.23 × 103 A W−1 responsivity and 1.8 × 1012 Jones detectivity under 830 nm illumination at ultralow intensity (64 nW cm−2). Notably, the architecture exhibited a photomultiplication factor of 7.8 × 103, coupled with rapid response dynamics (rise/fall times: tr/tf = 21 μs/7.8 ms). Simultaneously, Maria et al. [219] engineered Bi2S3 homojunctions comprising nanorod arrays and gradient-thickness films (film thickness inversely proportional to deposition time) on carbon substrates via hot-plate-assisted physical vapor deposition (PVD), as schematized in Figure 24a. Electrical characterization was conducted using Ta microprobes to interface the Bi2S3 nanostructures with the carbon substrate. Under white-light illumination (22.7 W m−2, 1 V bias), the homojunction exhibited a 4.5-fold photocurrent enhancement (4.5 μA vs. dark current 0.96 μA, Figure 24b), attributed to the parallel conduction pathways in nanorod arrays that minimized series resistance. The device demonstrates exceptional operational stability under high-power optical-switching conditions (Figure 24c), maintaining consistent dark/photocurrent levels with sub-second response dynamics (tr/tf = 192/270 ms). During prolonged switching cycle tests, the device exhibits minimal performance degradation, highlighting its superior repeatable stability. These results fully demonstrate the device’s capability to maintain a stable performance output under extreme operating conditions. The synergistic combination of high-surface-area nanorods and optically active thin films yielded superior optoelectronic metrics—responsivity (749.3 mA/W) and detectivity (5.61 × 108 Jones)—validating the efficient interfacial charge harvesting in this hierarchical architecture. The integration of semiconductor nanostructures is an effective way to improve the performance of photodetection with the development of the technology. Here, Ghosh et al. [220], used an in situ hydrothermal method to anchor MXene (Ti3C2Tx) nanosheets in situ on one-dimensional Bi2S3 (BS) nanorods and utilized IDE-patterned gold (Au) electrodes to prepare a Bi2S3/Ti3C2Tx (TC) photodetector based on an Al substrate with excellent performance. To investigate the light absorption ability of their samples, the UV-Vis absorption spectra of different samples are shown in Figure 24d. The heterojunction material shows stronger absorption over the whole range (300–1800 nm) with a peak at about 770 nm compared to the pure BS nanorods and TC nanosheets. Comparing the photovoltaic performance of the devices, the PD made with the BS/TC material exhibits a high switching ratio (ILight/IDark) of about 255, which is two orders of magnitude higher compared to the switching ratio of the pristine BS PD (~10), suggesting that the BSTC device has a higher photocurrent. The BSTC nanorod-based photodetector (PD) exhibits ultrafast response dynamics under 808 nm pulsed illumination (6.3 mW cm−2), achieving rise/fall times of tr/tf = 0.3/2.1 ms. Long-term stability tests (Figure 24e) reveal merely 0.6% photocurrent degradation after four months of ambient storage. The device demonstrates exceptional environmental stability and reproducible performance. Power-dependent characterization (Figure 24f) demonstrated its responsivity (R), external quantum efficiency (EQE), and detectivity (D*) degradation at elevated intensities, attributed to trap-state-mediated recombination in the BS nanorods. High photon flux exacerbates carrier trapping, accelerating electron–hole recombination and diminishing photocurrent yields. At ultralow power (0.03 mW cm−2, 3 V bias), the device achieved peak responsivity (25.5 A/W), EQE (3.9 × 103%), and detectivity (3.9 × 1012 Jones), outperforming conventional BS-based architectures by two orders of magnitude.
Distinctly, Singh et al. [221] reported a Bi2S3/PANI hybrid photodetector by functionalizing hydrothermally synthesized Bi2S3 nanorods (NRs) with polyaniline (PANI) and integrating silver paste electrodes. This hybrid architecture enables interfacial charge transfer acceleration at Bi2S3-PANI junctions, yielding an ultralow dark current (0.003 μA) and enhanced photocurrent (0.58 μA) under 365 nm illumination (1 V bias, 50 μW cm−2). The device achieved the following benchmark optoelectronic metrics: responsivity (2.676 × 104 A/W), detectivity (5.24 × 1013 Jones), external quantum efficiency (9073%), and rapid response kinetics (tr/tf = 50/63 ms). The synergistic interplay between Bi2S3’s narrow-bandgap absorption and PANI’s conductive framework suppresses dark carrier injection while facilitating photogenerated charge extraction, establishing a paradigm for organic–inorganic hybrid photodetector design. Remarkably, Devasia et al. [222] synthesized in situ-grown Cs3Bi2I₉:Bi2S3 nanorod (NR) composites on fluorine-doped tin oxide (FTO) substrates via single-step ultrasonic spray deposition (USD), fabricating photodiodes with spray-coated Ag-C electrodes. The Cs3Bi2I₉:Bi2S3 heterostructure exhibits intrinsic photosensitivity and operational photostability, achieving a photocurrent sensitivity of 198% (1.35 nA photocurrent vs. 0.68 nA dark current) at 1 V bias. In self-powered mode under 532 nm illumination (790 μW cm−2), the device demonstrated competitive photoresponsivity (0.59 mA W−1) and specific detectivity (8.18 × 109 Jones). The ultrabroadband spectral response (300–1550 nm) of this Bi2S3-based architecture positions it as a versatile candidate for next-generation broadband optoelectronic systems.
Titanium trisulfide (TiS3), as a new material for a new generation of electronic and optical devices, exhibits an ultrabroadband optical absorption range with its small direct bandgap (1.1 eV). Currently, Randle et al. [223] and Liu’s group [224], based on quasi-one-dimensional titanium trisulfide (TiS3) photovoltaic devices, have demonstrated that one-dimensional TiS3 has unique and excellent photovoltaic properties, which provides a development direction for the fabrication of high-performance TiS3-based photodetectors.

Comparative Analysis

Finally, we evaluate the photodetection capabilities of diverse one-dimensional (1D) metal trisulfide nanostructures. Among these, 1D Sb2S3, In2S3, and their composites have been extensively employed in high-sensitivity UV–visible photodetection. Notably, photodetectors (PDs) utilizing CVD-synthesized Sb2S3 NRs exhibit exceptional detection performance, achieving a specific detectivity of 2.1 × 1014 Jones. Similarly, single-crystalline In2S3 NW-based PDs fabricated via the same methodology demonstrate outstanding photoresponsivity (7.35 × 104 A/W) and external quantum efficiency (EQE: 2.88 × 107%). In contrast, Bi2S3-based photoconverters, while exhibiting weaker photoresponse magnitudes compared to Sb2S3 and In2S3, enable significantly accelerated optoelectronic signal transduction. For instance, Ti3C2Tₓ/Bi2S3 NR-based PDs achieve ultrafast photoresponses, with rise/decay times of 0.3/2.1 ms, respectively. These superior characteristics primarily originate from the narrow bandgaps inherent to metal trisulfides, which impart high carrier mobility and efficient charge separation/transport dynamics. Fundamentally, quantum confinement effects serve as dominant factors in the unique photoresponse behaviors of 1D nanostructures. A comparative summary of the 1D metal trisulfide-based PD performance is provided in Table 6. In collective comparison with standard photodetectors, this review underscores the exceptional responsivity (e.g., ZnS/CdS heterojunctions: >104 A W−1), detectivity (e.g., SnS2 nanowires: ~1016 Jones), and spectral tunability of one-dimensional metal sulfide photodetectors. When benchmarked against conventional Si (R: 0.4–0.6 A W−1; D: 1012–1013 Jones) and InGaAs (R: 0.8–1.2 A W−1; D: 1011–1012 Jones) systems, these 1D metal sulfide architectures demonstrate orders-of-magnitude superiority in responsivity and detectivity metrics, coupled with broad spectral adaptability [24,25]. However, they encounter persistent challenges in their industrial-scale reproducibility and environmental resilience under extreme operational conditions. While metal sulfides excel in mechanical flexibility, visible-to-near-infrared (Vis-NIR) broadband operation, and low-power/self-powered functionality, they exhibit deficiencies in response uniformity and substrate integration maturity relative to their established counterparts. Future endeavors must bridge these gaps through compositional engineering and interface optimization, aiming to achieve commercial viability comparable to Si/InGaAs technologies in high-speed photonic applications such as imaging and telecommunications.

4. Challenges and Future Prospects

Although one-dimensional metal sulfide photodetectors demonstrate remarkable advantages in response speed, sensitivity, and spectral range, their practical applications still face multiple challenges. This section systematically analyzes the current technical bottlenecks and proposes critical research pathways for future investigations. Therefore, this section summarizes recent developments regarding self-powered and wide-spectrum thermally stable photodetectors for the reference of researchers.

4.1. Self-Powered Photodetector

Conventional photodetectors predominantly rely on external bias voltages, which constrain their applications in wearable and IoT scenarios. In contrast, self-powered devices require the simultaneous optimization of light absorption efficiency and built-in electric field strength [225]. This is exemplified by p-n junctions, which generate self-driven photovoltaic effects through interfacial band bending [226].
The research group led by Gao fabricated a Type-II CsPbBr3/ZnS heterostructure photodetector with Al/ITO electrodes using hot-injection and spin-coating techniques (Figure 25a) [227]. Figure 25b presents the current–voltage (I-V) characteristics of the device under dark conditions and 365 nm illumination with varying intensities. The dark current curve exhibits a typical origin-passing behavior, while under illumination, the I-V curves show pronounced offsets at 0 V bias, with photocurrent enhancement proportional to light intensity, confirming the device’s photovoltaic conversion capability without external bias. This self-powered operation originates from the built-in potential-induced photovoltaic effect at the heterojunction interface. Figure 25c illustrates the photoresponse under varying optical intensities at 0 V bias, where rapid switching between illumination on/off states demonstrates efficient detection in self-powered mode. The device exhibits outstanding repeatable stability, enabling reliable operation across diverse power conditions. The measured rise (tr) and fall (tf) times are determined as 150 ms and 30 ms, respectively. Under 365 nm illumination (2 mW/cm2), the device achieves remarkable zero-bias performance: a photosensitivity of 1.38 × 104, responsivity of 37.5 mA/W, and specific detectivity of 1.21 × 1012 Jones.
Li et al. [228] fabricated a self-powered device (Figure 26a) through a combined chemical bath deposition (CBD) and spin-coating approach, where a uniform MAPbI3/CdS thin film was deposited on an ITO substrate followed by Au electrode formation via vacuum thermal evaporation. Figure 26b displays the current–voltage (I-V) characteristics of the MAPbI3/CdS photodetector under dark conditions and white-light illumination with varying power densities. A pronounced photocurrent emerges at zero bias, exhibiting linear enhancement with increasing optical power density, which substantiates the self-biased photovoltaic behavior. At 730 nm illumination (10 mW/cm2), the device achieves optimal performance, with a specific detectivity of 2.3 × 1011 Jones (1 Jones = 1 cm·Hz1/2·W−1) and responsivity of 0.43 A/W. Figure 26c demonstrates the exceptional reproducibility and stability across multiple light on/off cycles in the tested devices. Specifically, the device exhibits highly consistent photoresponses under periodic illumination, confirming its reliability in prolonged operational scenarios. Compared to single-component devices, the heterojunction exhibits accelerated response characteristics, manifested by rise and decay times of 3.2 ms and 9.6 ms, respectively, attributed to efficient electron–hole separation and charge transfer at the heterointerface. Notably, Zhou’s group [229] recently developed a high-responsivity self-powered UV photodetector based on a ZnO nanoarray/CdS/GaN architecture employing ITO and In as top/bottom electrodes, expanding the research paradigm of heterostructure devices. Figure 26d presents the current–voltage (I-V) characteristics of the photodetector (PD) under dark conditions and 300 nm UV illumination (0.61 mW/cm2), with the inset showing Ohmic contact behavior at In/P-GaN interfaces. The dark current curve reveals a rectification ratio of 365 at 3 V bias, while pronounced photocurrent generation at zero bias under illumination confirms its self-powered operation. As shown in Figure 26e, periodic UV light switching tests under zero-bias conditions demonstrate repeatable photoresponse stability with minimal current fluctuation and rapid response characteristics (rise time tr and fall time tf < 0.35 s). Notably, under zero-bias operation (Figure 26f), the device achieves superior spectral responsivity (176 mA W−1) and specific detectivity (1012 Jones) compared to its CdS-free ZnO/GaN counterparts (27 mA W−1), benefiting from an ultralow dark current (2 nA). This performance enhancement originates from the effective suppression of interfacial charge recombination and optimized electron transport pathways within the device architecture.
To develop photodetectors with concurrent high sensitivity, high responsivity, low power consumption, polarization sensitivity, and broadband detection capabilities, Xie et al. [230] fabricated vertically stacked Nb-WS2/Ta2NiSe₅ heterostructures via mechanical exfoliation and dry transfer techniques. Au electrodes were deposited by electron-beam evaporation to construct the low-power device shown in Figure 27a. The wavelength-dependent photocurrent measurements (Figure 27b) demonstrate rapid and stable optical responses across 405–1550 nm, confirming broadband operation. As illustrated in Figure 27c, under 600 nm illumination (0.035–275.67 mW/cm2), the source current at zero bias significantly exceeds the dark leakage current, verifying the device’s self-powered operation through photogenerated carrier separation driven by the built-in electric field. Under 660 nm laser illumination, the device has excellent response and recovery capabilities, with response and recovery times of 118 and 13 μs, respectively. At low light intensity (0.45 mW/cm2), remarkable figures of merit are achieved: an R of 57.64 A W−1, D* of 6 × 1010 Jones, and EQE reaching 10,854%. In addition, Kumar’s group [231] successfully fabricated a self-powered broadband photodetector by spinning WS2 nanoparticles on a MoS2 substrate synthesized by chemical vapor deposition (CVD) and depositing Au/Cr cross-finger electrodes on the surface of thin films by shadow-assisted thermal evaporation technology. Figure 27d–f comparatively illustrate the photoelectronic characteristics of pristine MoS2 and MoS2-WS2 heterojunctions under zero-bias (0 V) operation across visible to near-infrared (VIS-NIR) wavelengths (405–800 nm). Under 580 nm illumination (120 μW cm−2), the heterojunction device achieves a responsivity of 282.77 mA W−1 and an EQE of 60.59%, representing 12-fold and 11-fold enhancements compared to MoS2-based photodetectors (23.55 mA W−1, 5.01%). Notably, the device exhibits improved response dynamics, with a 375 μs rise time and a 6 ms decay time, demonstrating significantly accelerated response speeds compared to pristine MoS2 and WS2 devices.
Its atomic-scale thickness and superior hetero-integration capability endow MoS2 with significant potential for self-powered reconfigurable devices. Zhou et al. [232] demonstrated an Ar plasma-treated metal–semiconductor–metal (Ag-MoS2-Pt) two-terminal structure, where reversible built-in electric field modulation via voltage-controlled sulfur vacancy concentration enabled a bidirectional self-powered photocurrent. Under 532 nm illumination (0.519 mW/mm2), the device achieved record response speeds of 224 μs (rise)/293 μs (decay) and a peak responsivity of 280 mA W−1 at zero bias, with a self-biased detectivity of 0.76 × 1011 Jones and an EQE of 65%. While emerging active materials advance photodetector technology, the trade-off between spectral sensitivity and operational efficiency persists. Vashishtha’s group [158] engineered a MoS2/GaN epitaxial heterostructure-based self-powered detector (Au/Ni electrodes) featuring dual-band UV-Vis detection in a two-terminal configuration (Figure 28a). Its absorption spectra (Figure 28b) reveal enhanced UV absorption and a broadened visible response compared to bare GaN. As depicted in Figure 28c,d, the device maintains consistent switching dynamics under UV–visible illumination, exhibiting minimal current degradation and retaining alignment with initial current values across multiple periodic illumination cycles at varying power levels, demonstrating exceptional repeatable stability in extreme operational environments. With rise/decay times of 1.04/0.8 ms and 0.9/0.8 ms, respectively, the device achieves dual-mode operation featuring responsivities of 631/37 mA W−1, detectivities of 8.5 × 1010/0.5 × 1010 Jones, and EQE values of 214%/8%, providing an innovative solution for cutting-edge applications.
The development of self-powered three-dimensional photodetectors (3DPDs) based on heterostructures faces significant challenges due to technical bottlenecks in the three-dimensional assembly of heterojunctions and the precise control of electrode interface matching. To address this issue, Shen’s research group [233] pioneered a novel approach by employing molecular beam epitaxy (MBE) technology combined with a controlled rolling assembly strategy for tubular MoS2/GaAs/InGaAs heterostructures and the meticulous fabrication of Ni/Au composite electrodes, successfully demonstrating the first polarization-sensitive self-powered 3D photodetector (Figure 29a). To investigate the light intensity dependence of device performance under self-bias conditions, Figure 29b presents the current–voltage characteristics under 650 nm illumination. Experimental results reveal a monotonic increase in photocurrent with light intensity, ranging from dark-state to 970 mW/cm2, while maintaining a dark current (Idark) of 2 × 10−11 A and achieving a maximum on/off ratio of 2.0 × 103, indicating superior photoresponse tolerance. Further analysis through responsivity (R) and external quantum efficiency (EQE) versus light intensity curves (Figure 29c) demonstrates the exponential growth characteristics of both R and EQE under zero bias as light intensity decreases, reaching 86 mA W−1 and 16%, respectively, at the minimum light intensity of 3.48 mW cm−2. Furthermore, the environmentally benign layered semiconductor SnS2 has attracted considerable attention due to its tunable bandgap and potential in high-performance photodetection. Deng’s team [234] successfully fabricated SnS2 nano-wall arrays on FTO substrates via a facile hydrothermal method, implementing them as photoanodes in self-powered photodetectors. Remarkably, the devices exhibit outstanding photocurrent density (39.06 μA cm−2), responsivity (1460 μA/W), and cycling stability under 475 nm illumination (10.8 mW cm−2), providing new insights for expanding the applications of two-dimensional materials in optoelectronic detection.
Antimony trisulfide (Sb2S3) photodetectors (PDs) demonstrate promising application potential. Efficient carrier transport in PDs critically governs their detectivity and response speed. Kang’s group [235] synthesized a fully inorganic self-powered Sb2S3/TiO2 PD with FTO/Au electrodes (Figure 30a) via the hydrothermal growth of vertically aligned TiO2 nanorod arrays followed by the spin-coating of Sb2S3. The one-dimensional TiO2 nanorod architecture facilitated the effective separation and transport of photogenerated carriers, thereby significantly enhancing device performance. Figure 30b displays the current–time (I-t) curve of the device under zero bias at an illumination power density of 400 W cm−2. The PD exhibits stable photoresponses across wavelengths within the visible spectrum (380–760 nm), with a maximum photocurrent of 4.5 μA observed under red light (625 nm). Figure 30c presents the current–voltage (I-V) characteristics of the Sb2S3 PD under dark conditions and varying optical power densities at 625 nm. The photocurrent increases progressively with incident power density, highlighting the device’s superior self-powered functionality. At 625 nm illumination (9 W cm−2), the device achieves a responsivity (R) of 0.29 A W−1 and specific detectivity (D*) of 3.37 × 1012 Jones. The temporal response of the c-TiO2/TiO2-NRs/Sb2S3 PD under 625 nm illumination (Figure 30d) reveals rapid rise and decay times of 9.3 μs and 7.8 μs, respectively, indicating substantial commercial viability. For comparison, Zamani et al. [236] fabricated a p-Si/Bi2S3 heterojunction via chemical vapor deposition (CVD) and explored the role of Au coating in optoelectronic applications. The sample with Au deposited at 420 °C exhibited the highest carrier concentration (7.67 × 1018 cm−3). The p-Si/n-Bi2S3 heterojunction PD demonstrated remarkable self-biased detection performance in the red wavelength regime, achieving a responsivity of 0.94 mA W−1 and detectivity of 6.40 × 108 Jones.
Recent advances in self-powered photodetectors (PDs) highlight the critical role of heterojunction engineering and interface optimization in achieving high responsivity, rapid responses, and broadband detection without external bias, as summarized in Table 7, in recent performance studies. However, challenges remain in these devices’ scalable fabrication, spectral selectivity, and long-term stability.

4.2. Thermal Stability and Broadband Testing

Under high-temperature conditions, sulfide materials are prone to sulfur volatilization (e.g., sulfur loss in Bi2S3 at temperatures > 200 °C) and lattice distortion, leading to device performance degradation. Furthermore, broadband detection (spanning ultraviolet to near-infrared wavelengths) imposes stringent requirements on material properties, typically necessitating gradient bandgap designs or heterojunction architectures to achieve wide-spectrum responsivity. However, interfacial defects can significantly enhance carrier recombination, severely limiting the photoelectric conversion efficiency of devices. Thus, the development of thermally stable broadband detection materials, coupled with interface engineering strategies to address temperature fluctuations and harsh environmental conditions, has become an urgent research priority.
Yadav et al. [237] synthesized hexagonal-phase SnS2 nanosheets via a facile and low-cost solvothermal method. A SnS2 nanosheet-based ultraviolet photodetector (PD) was fabricated using silver (Ag) as contact electrodes, with the device architecture illustrated in Figure 31a. The device exhibited remarkable thermal stability and high-precision UV detection capability: under a 5 V bias, the dark current increased merely 1.2-fold as the temperature rose from 60 °C to 120 °C (Figure 31b). Photoelectronic characterization at a 365 nm wavelength with 5 V bias revealed the device’s superior sensitivity (~400), responsivity (~5.5 A W−1), external quantum efficiency (EQE, ~1868%), and detectivity (~1.72 × 1013 Jones). Temporal response measurements under UV illumination (1.112 mW cm−2) with a 20 s light switching cycle (Figure 31c) showed rise and fall times of 2.2 s and 6.3 s, respectively (inset of Figure 31c). The device’s current stability in harsh environments demonstrates its reproducible resilient operation across varying temperatures, likely originating from the enhanced carrier vibration and migration induced by elevated thermal conditions. The prolonged rise time was attributed to the wide inter-electrode spacing between Ag contacts, which extended the transit time of photogenerated carriers. Similarly, Jin et al. [238] fabricated a conjugated polymer PThTPTI/WS2 heterojunction film via chemical vapor deposition (CVD) and constructed Au-electrode devices using shadow mask thermal evaporation (Figure 31d). The resulting PD demonstrated stable operation in ambient air at temperatures up to 300 °C. For the non-annealed PThTPTI/WS2 PD, both rise and decay times increased significantly with temperature (Figure 31e: from 17 ms to 6.89 s for rise time and 38 ms to 8.43 s for decay time between 25 °C and 300 °C). This behavior originated from enhanced electron accumulation at the PThTPTI/WS2 interface under elevated temperatures, prolonging the carrier redistribution equilibrium time (thus increasing rise time). During light-off phases, the trapped electrons in WS2 nanosheets required extended diffusion periods to return to the interface at higher temperatures, leading to longer decay times. Notably, post-annealing treatment preserved the device performance (Figure 31f), confirming exceptional thermal stability in ambient air at 300 °C.
Furthermore, Vashishtha et al. [239] deposited MoS2 thin films on Si3N₄ substrates via radio-frequency (RF) sputtering and fabricated Pt electrode contacts on the MoS2 surface through DC sputtering, resulting in the optoelectronic device shown in Figure 32a. High-temperature stability evaluation at 100 °C revealed a dark current of 2.5 µA in current–voltage (I-V) measurements (Figure 32b), slightly higher than the room-temperature value (1 µA). This increase in dark current originated from thermionic emission (Edison effect). Since the charge carrier mobility (μ) is positively correlated with temperature (μ α KBT), the current increases with increasing temperature. The transient photoresponse under 950 nm illumination (2 µW) at 100 °C (Figure 32c) demonstrated stable operation over multiple light switching cycles, with a photocurrent (2.3 µA) comparable to the room-temperature value (2.7 µA). The device also exhibited superior optoelectronic parameters at 100 °C, including responsivity (1170 mA W−1), external quantum efficiency (409%), and specific detectivity (1.6 × 1010 Jones). Similarly, the MoS2/GaN detector developed by Vashishtha’s group [158] achieved high-efficiency detection under zero bias and maintained stable operation at 100 °C, highlighting the potential of self-powered thermally robust devices for cutting-edge applications. The inherent flexibility and unique fluidity of organic materials not only suit wearable technologies but also enhance thermal dissipation. Leveraging this, Wahalathantrige’s group [240] synthesized a vinyl-functionalized diketopyrrolopyrrole polymer (PDPPVTT)/MoS2 hybrid via chemical vapor transport (CVT) and solution processing, followed by Au/Cr electrode fabrication. Owing to the polymer’s high absorption coefficient (or molar absorptivity) in the 800 nm region, the device delivered a high photocurrent of 325 nA under 0.9 nW illumination (20 V), with its responsivity (R) reaching 103 A W−1. To further enhance MoS2-based photodetector performance, Li et al. [241] introduced an Al2O3 atomic-layer stress buffer into MoS2/GaN films through CVD and atomic layer deposition (ALD), constructing a device with Au/Cr electrodes. The Al2O3-induced tensile strain softened phonon vibrations and reduced thermal conductivity in MoS2, endowing the device with exceptional thermal stability. Under 365 nm illumination (3.141 μW), it achieved a responsivity of 24.62 A W−1, accompanied by a photoconductive gain of 520 and EQE of 8381%.
Gallium oxide (Ga2O3), characterized by its polymorphic nature and exceptional thermal stability, demonstrates significant potential for high-temperature optoelectronic applications [242]. The Xiong research group [243] epitaxially grew Ga2O3 thin films on aluminum nitride substrates via chemical vapor deposition (CVD), subsequently fabricating Au/Ni interdigital electrodes through photolithography and magnetron sputtering to construct a Ga2O3-based ultraviolet photodetector (Figure 33a). Under 274 µW cm−2 illumination at 250 nm, Figure 33b presents the comparative current–voltage (I-V) characteristics under dark and illuminated conditions. At an operational temperature of 200 °C, the device achieved a photocurrent-to-dark-current ratio exceeding 102. The temporal photoresponse characteristics shown in Figure 33c demonstrate consistent response times across the 50–200 °C temperature range under 250 nm illumination, confirming excellent thermal stability. Compared to the room-temperature (25 °C) responsivity (3.2 × 10–4 A W−1) of IZO/β-Ga2O3/IZO photodetectors, this device exhibited a significantly enhanced responsivity of 0.518 A W−1 at 200 °C. Zhang et al. [244] developed a self-powered deep-ultraviolet photodetector with an all-oxide configuration (Figure 33d) by constructing ε-Ga2O3/ZnO heterojunctions through CVD and depositing Ti/Au electrodes via magnetron sputtering. Figure 33e illustrates the device’s robust rectification characteristics under 254 nm illumination (400 µW cm−2) across various temperatures. At room temperature, photocurrents of 0.3 µA and 5.5 nA were obtained under −5 V and 0 V biases, respectively, indicating efficient photoconversion. Remarkably, the device maintained a high photocurrent-to-dark-current ratio of 102 even at extreme temperatures exceeding 500 K. Figure 33f reveals the temperature dependence of response times, where the rise/decay times showed negligible correlation with temperature despite amplitude variations in current spikes, suggesting temperature-independent carrier generation–recombination dynamics. The room-temperature self-powered operation yielded a responsivity of 2.4 mA W−1, detectivity of 5 Jones, and external quantum efficiency of 1.5%, providing critical insights for developing thermally stable chalcogenide-based photodetectors.
In recent years, photodetectors have made remarkable progress in the fields of thermal stability and wide-spectrum detection, as shown in Table 8. The devices still exhibit good detection performance at high temperatures. However, the volatilization of sulfur at high temperatures and interface defects in sulfides remain bottlenecks to efficiency. It is necessary to further optimize the gradient bandgap design, stress regulation (such as an Al2O3 buffer layer) and the integration process of sulfides/oxides (Ga2O3) to expand the application in extreme environments.

4.3. Active Region

The active region constitutes the critical domain in photodetectors where photogenerated electron–hole pairs undergo generation and recombination processes. Its fundamental operational mechanism involves efficient photon absorption and subsequent carrier generation, followed by directional charge collection to establish a measurable photocurrent. The photoelectric conversion efficiency is critically determined by strategic material selection and the optimized architectural design of this functional zone [245]. Notably, the thickness of the active region significantly impacts light absorption efficiency. Studies show that an active region thickness of 100 nm concentrates the optical field distribution within the absorption layer, thereby enhancing photodetector performance. However, increasing the thickness to 150 nm introduces multiple minima in the optical field distribution, thereby reducing the average optical energy density and degrading photoelectric conversion efficiency [246,247]. Additionally, reducing barrier thickness (e.g., from 10 nm to 5 nm) significantly mitigates gain inhomogeneity in p-side quantum wells, improving device performance [248]. Thus, optimizing the active region thickness is critical for advancing photodetector capabilities. Future research should focus on developing low-cost, scalable fabrication techniques, such as solution-based methods and chemical vapor deposition (CVD), while expanding the active region through structural and material optimization in order to enhance the overall performance of photodetectors.

4.4. Current Limitations and Future Directions

While one-dimensional metal sulfide photodetectors exhibit remarkable advantages in response speed, sensitivity, and spectral range, their practical implementation encounters multiple challenges. This section systematically analyzes existing technical bottlenecks and proposes critical research directions for future investigations. The discussion specifically addresses fundamental challenges—environmental stability, scalability, and the toxicity of cadmium-based materials—all of which are critical for the commercial viability of photodetectors.

4.4.1. Environmental Stability

The environmental stability of metal sulfide nanomaterials constitutes a critical limitation for their practical deployment. Materials such as CdS and SnS2 are prone to surface sulfide hydrolysis or oxidation in high-humidity or oxidative environments (e.g., CdS + H2O → Cd(OH)2 + H2S ↑), resulting in photoresponsive performance degradation [234]. Furthermore, two-dimensional derivatives like MoS2 and WS2 undergo accelerated degradation under prolonged illumination due to photocatalytic effects [249]. Mitigation strategies against this include surface passivation (e.g., Al2O3 encapsulation via atomic layer deposition) and heterojunction engineering (e.g., ZnS/CdS core–shell architectures), which effectively suppress environmental degradation [250]. For instance, Zhang et al. demonstrated that SiO2-coated CdS nanowires extended device operational lifetime from 72 h to 1000 h under 85% relative humidity [251].

4.4.2. Scalability

The current synthesis methods for 1D metal sulfides (e.g., chemical vapor deposition, hydrothermal synthesis) predominantly rely on precision-controlled conditions (e.g., high temperatures, vacuum), posing challenges for cost-effective large-scale production. For example, CdS nanowires exhibit vapor-phase growth rates limited to micrometers per hour, with substrate sizes typically constrained to <10 cm2 [252]. Emerging techniques such as template-assisted electrochemical deposition or inkjet printing enable high-throughput fabrication but compromise crystallographic quality and device performance (e.g., 30–50% mobility reduction) [253]. Future efforts should explore compromise strategies balancing scalability and performance, such as self-assembled monolayer modulation or continuous-flow reactor designs.

4.4.3. Toxicity of Cadmium-Based Materials

Cadmium-based compounds like CdS and CdSe face stringent environmental regulations (e.g., EU RoHS Directive) due to their high toxicity (carcinogenicity, bioaccumulation). Despite their superior photoresponsive performance (e.g., CdS nanowire detectivity up to 1016 Jones), industrial adoption necessitates risk–performance trade-offs. Significant progress has been made in alternative materials: SnS2 (low toxicity, tunable bandgap to 2.1 eV) and Sb2S3 (heavy-metal-free) achieve detectivities approaching 1016 Jones [254], though their response speeds (ms) lag behind CdS (μs) [139,172]. Furthermore, inadequate recycling technologies exacerbate environmental burdens; currently, only 30% of cadmium is recoverable via hydrometallurgical processes [220], urgently demanding closed-loop manufacturing processes.

4.4.4. Future Research Directions

To address future societal development demands, interdisciplinary efforts should prioritize the following aspects: (1) Developing low-temperature synthesis methods (e.g., inkjet printing, roll-to-roll processing) for depositing one-dimensional sulfides on stretchable substrates (e.g., PET, PDMS) to fabricate flexible and wearable integrated products. Recent advances in Ag nanowires/ZnS nanotubes have demonstrated exceptional mechanical durability, maintaining 95% of their initial response after 1000 bending cycles [255]. (2) Employing artificial intelligence to predict optimal material compositions (e.g., doping ratios, heterojunction pairings) and device architectures. Neural networks have successfully reduced experimental optimization cycles by 70% for MoS2-based detectors, while simultaneously maximizing EQE [256]. (3) Integrating one-dimensional sulfides with two-dimensional materials (e.g., graphene, MXenes) or perovskites to achieve synergistic effects. A recently developed Bi2S3/MXene composite has achieved broadband detection (300–1800 nm) with an ultrafast response (0.3 ms) [257], highlighting the potential of this strategy.

5. Conclusions and Outlook

This review comprehensively examines recent advances in the photodetection properties of one-dimensional (1D) metal sulfide nanostructures and their corresponding devices. We first outline the operational mechanisms of photodetectors, their critical performance parameters, and their dominant device architectures. Subsequently, we systematically analyze device structures, synthesis methodologies, and performance metrics through comparative evaluation. Notably, photodetector performance demonstrates heightened sensitivity to carrier mobility, surface-to-volume ratios, and interfacial contact resistance between photosensitive layers and electrodes. Furthermore, morphological modifications, electrode selections, and synthetic approaches introduce substantial performance variations across devices. Nevertheless, 1D metal sulfide-based photodetectors demonstrate superior specific surface areas and abundant surface oxygen desorption sites compared to their thin-film and metal oxide counterparts, thereby enabling enhanced optoelectronic functionalities.
One-dimensional (1D) single-chalcogenide metal compound nanostructures have garnered significant research attention due to their structural tunability and optoelectronic performance. For instance, chemical vapor deposition (CVD)-synthesized CdS nanowire (NW)-based photodetectors (PDs) demonstrate exceptional responsivity (2.6 × 105 A/W) and specific detectivity (2.3 × 1016 Jones) upon ferroelectric polymer polarization, attributed to the drastic suppression of dark current via intrinsic carrier depletion and photon-generated current dominance. Transition metal disulfides (WS2, SnS2, MoS2, ReS2) in 1D configurations exhibit broad spectral absorption, high porosity, and optimal bandgaps, enabling high carrier mobility and rapid response speeds for ultrasensitive photodetection. Notably, emerging ReS2 NW-PDs achieve record responsivity (5.08 × 105 A/W), detectivity (6.1 × 1015 Jones), and EQE (1.07 × 106%), facilitated by ultralow ReS2/Ag interfacial contact resistance that enhances charge extraction efficiency. Enhanced sulfur stoichiometry promotes crystallinity, passivating surface defects and substantially improving device performance. Among 1D trisulfides (Sb2S3, Bi2S3, In2S3), high-crystallinity Sb2S3 NWs exhibit superior detectivity (2.1 × 1014 Jones), EQE (1.5 × 104%), and sensitivity (2.2 × 104), primarily due to their minimized impurity incorporation (e.g., organometallic precursors, surfactants) via crystallinity-driven surface stabilization. These advances underscore the potential of 1D metal sulfides in next-generation photodetectors. By synthesizing the recent progress in self-powered and thermally stable PDs, this review establishes critical benchmarks for future research. Prioritizing material/structural optimization, multifunctional integration, and environmental resilience will drive advancements in low-power, high-stability photodetection technologies.
Despite notable breakthroughs in the synthesis and design of one-dimensional (1D) metal sulfide-based photodetectors, substantial advancements in both synthesis methodologies and performance metrics—relative to 1D metal oxides and 2D metal sulfides—remain essential in order to address emerging demands across diverse applications. To further advance 1D metal sulfide photodetectors, priority should be given to the following considerations:
(1)
When utilizing doped elements or constructing heterostructures to improve the sensitivity and responsivity of photodetectors, there remains a need to develop new, effective methods for achieving efficient detection, for example, via the polarization effect of ferroelectric polymers on the device; however, the specific effect of ferroelectric polarization on the optical response and the underlying principles need to be further studied.
(2)
The exploration of homogeneous hybridized nanostructures as photodetectors and the in-depth investigation of the principles at work are essential, as the complex interfaces and carrier migration pathways in hybridized nanostructures often give rise to unique optical and photoelectronic properties.
(3)
Due to the flexibility, transparency, and low cost of organic materials, the study of nanostructured material photodetectors with different polymer modifications can lead to the development of more stable and flexible devices while reducing costs.
(4)
Devices with high thermal stability and self-powering capabilities not only contribute significantly to energy conservation and emissions reduction but also offer critical utility in harsh environmental exploration. Consequently, the development of high-performance thermally stable materials and the realization of efficient devices under zero-bias conditions remain pressing challenges that demand accelerated exploration by researchers.
(5)
Finally, although numerous methodologies exist for material synthesis and device fabrication, synthesis processes remain challenging to control effectively. Furthermore, device manufacturing typically involves integrated processes combining photolithography, etching, and deposition. Undoubtedly, such inherently complex and time-consuming procedures hinder economic viability. Consequently, advancing fabrication techniques to streamline these processes remains a critical area for exploration.

Author Contributions

Conceptualization, H.L. (Haowei Lin), J.C. and M.L.; writing—original draft preparation, J.C., M.L. and W.C.; writing—review and editing, H.L. (Haowei Lin), J.C., M.L., C.Z., Z.W. and H.L. (Huiying Li); project administration, H.L. (Haowei Lin); funding acquisition, H.L. (Haowei Lin). All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Henan Province Science and Technology Research Project (252102230093), Henan Province College Students’ Innovation and Entrepreneurship Training Program Project (202410463057), Cultivation Programme for Young Backbone Teachers in Henan University of Technology, and the Science and Technology Key Project from Education Department of Henan Province (21A430011).

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to thank the financial support from the Henan Province Science and Technology Research Project (252102230093), Henan Province College Students’ Innovation and Entrepreneurship Training Program Project (202410463057), Cultivation Programme for Young Backbone Teachers in Henan University of Technology, and the Science and Technology Key Project from Education Department of Henan Province (21A430011).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Kim, J.; Lee, J.; Lee, J.M.; Facchetti, A.; Marks, T.J.; Park, S.K. Recent Advances in Low-Dimensional Nanomaterials for Photodetectors. Small Methods 2023, 8, 2300246. [Google Scholar] [CrossRef]
  2. Paras; Yadav, K.; Kumar, P.; Teja, D.R.; Chakraborty, S.; Chakraborty, M.; Mohapatra, S.S.; Sahoo, A.; Chou, M.M.C.; Liang, C.T.; et al. A Review on Low-Dimensional Nanomaterials: Nanofabrication, Characterization and Applications. Nanomaterials 2022, 13, 160. [Google Scholar] [CrossRef] [PubMed]
  3. Zhao, Y.S.; Fu, H.; Peng, A.; Ma, Y.; Xiao, D.; Yao, J. Low-Dimensional Nanomaterials Based on Small Organic Molecules: Preparation and Optoelectronic Properties. Adv. Mater. 2008, 20, 2859–2876. [Google Scholar] [CrossRef]
  4. Chen, G.; Shrestha, L.K.; Ariga, K. Zero-to-Two Nanoarchitectonics: Fabrication of Two-Dimensional Materials from Zero-Dimensional Fullerene. Molecules 2021, 26, 4636. [Google Scholar] [CrossRef]
  5. Chen, Y.; Lai, Z.; Zhang, X.; Fan, Z.; He, Q.; Tan, C.; Zhang, H. Phase engineering of nanomaterials. Nat. Rev. Chem. 2020, 4, 243–256. [Google Scholar] [CrossRef] [PubMed]
  6. Liu, S.; Pan, X.; Liu, H. Two-Dimensional Nanomaterials for Photothermal Therapy. Angew. Chem. Int. Ed. 2020, 59, 5890–5900. [Google Scholar] [CrossRef]
  7. Shulga, A.G.; Derenskyi, V.; Salazar-Rios, J.M.; Dirin, D.N.; Fritsch, M.; Kovalenko, M.V.; Scherf, U.; Loi, M.A. An All-Solution-Based Hybrid CMOS-Like Quantum Dot/Carbon Nanotube Inverter. Adv. Mater. 2017, 29, 1701764. [Google Scholar] [CrossRef]
  8. Wang, H.P.; Li, S.; Liu, X.; Shi, Z.; Fang, X.; He, J.H. Low-Dimensional Metal Halide Perovskite Photodetectors. Adv. Mater. 2020, 33, 2003309. [Google Scholar] [CrossRef]
  9. Zhai, W.; Zhou, K. Nanomaterials in Superlubricity. Adv. Funct. Mater. 2019, 29, 1806395. [Google Scholar] [CrossRef]
  10. Zhang, H.; Fan, T.; Chen, W.; Li, Y.; Wang, B. Recent advances of two-dimensional materials in smart drug delivery nano-systems. Bioact. Mater. 2020, 5, 1071–1086. [Google Scholar] [CrossRef]
  11. Chi, X.; Lan, Z.A.; Chen, Q.; Zhang, X.; Chen, X.; Zhang, G.; Wang, X. Electronic Transmission Channels Promoting Charge Separation of Conjugated Polymers for Photocatalytic CO2 Reduction with Controllable Selectivity. Angew. Chem. Int. Ed. 2023, 62, e202303785. [Google Scholar] [CrossRef] [PubMed]
  12. Das, S.; Senapati, S.; Naik, R. 1D metal telluride heterostructure: A review on its synthesis for multifunctional applications. J. Alloys Compd. 2023, 968, 171923. [Google Scholar] [CrossRef]
  13. Lin, H.; Yang, L.; Zhang, X.; Liu, G.; Zhuo, S.; Chen, J.; Song, J. Emerging Low-Dimensional Nanoagents for Bio-Microimaging. Adv. Funct. Mater. 2020, 30, 2003147. [Google Scholar] [CrossRef]
  14. Sun, H.; Deng, J.; Qiu, L.; Fang, X.; Peng, H. Recent progress in solar cells based on one-dimensional nanomaterials. Energy Environ. Sci. 2015, 8, 1139–1159. [Google Scholar] [CrossRef]
  15. Xu, J.; Li, Q.; Liu, X.; Yang, Q. P-doped nanorod MoO3 and nanoflower NiAl-LDH construct S-type heterojunction for photocatalytic high-efficiency hydrogen evolution. Surf. Interfaces 2023, 43, 103593. [Google Scholar] [CrossRef]
  16. Peng, G.; Wu, J.; Wu, S.; Xu, X.; Ellis, J.E.; Xu, G.; Star, A.; Gao, D. Perovskite solar cells based on bottom-fused TiO2 nanocones. J. Mater. Chem. A 2016, 4, 1520–1530. [Google Scholar] [CrossRef]
  17. Shen, Q.; Xu, T.; Zhuang, G.-l.; Zhuang, Y.; Sun, L.; Han, X.; Wang, X.; Zhan, W. Spatially Separated Photoinduced Charge Carriers for the Enhanced Photocatalysis Over the One-Dimensional Yolk–Shell In2Se3@N-C Nanoreactor. ACS Catal. 2021, 11, 12931–12939. [Google Scholar] [CrossRef]
  18. Wei, D.; Dadey, A.A.; McArthur, J.A.; Bank, S.R.; Campbell, J.C. Enhancing Extended SWIR Al0.3InAsSb PIN Photodetectors with All-Dielectric Amorphous Germanium Photon-Capturing Gratings. ACS Photonics 2024, 11, 484–488. [Google Scholar] [CrossRef]
  19. Chen, K.; Wang, X.; Zou, C.; Liu, Q.; Chen, K.; Shi, Y.; Xu, T.; Zhao, W.; He, L.; Gao, F.; et al. Two-In-One: End-Emitting Blue LED and Self-Powered UV Photodetector based on Single Trapezoidal PIN GaN Microwire for Ambient Light UV Monitoring and Feedback. Small Methods 2023, 7, 2300138. [Google Scholar] [CrossRef]
  20. Chen, Y.C.; Lu, Y.J.; Liu, Q.; Lin, C.N.; Guo, J.; Zang, J.H.; Tian, Y.Z.; Shan, C.X. Ga2O3 photodetector arrays for solar-blind imaging. J. Mater. Chem. C 2019, 7, 2557–2562. [Google Scholar] [CrossRef]
  21. Li, G.; Wang, Y.; Huang, L.; Sun, W. Research Progress of High-Sensitivity Perovskite Photodetectors: A Review of Photodetectors: Noise, Structure, and Materials. ACS Appl. Electron. Mater. 2022, 4, 1485–1505. [Google Scholar] [CrossRef]
  22. Luo, M.C.; Ren, F.F.; Gagrani, N.; Qiu, K.; Wang, Q.; Yu, L.; Ye, J.; Yan, F.; Zhang, R.; Tan, H.H.; et al. Polarization-Independent Indium Phosphide Nanowire Photodetectors. Adv. Opt. Mater. 2020, 8, 2000514. [Google Scholar] [CrossRef]
  23. Yang, D.; Ma, D. Development of Organic Semiconductor Photodetectors: From Mechanism to Applications. Adv. Opt. Mater. 2018, 7, 1800522. [Google Scholar] [CrossRef]
  24. Huang, Y.C.; Parimi, V.; Chang, W.C.; Syu, H.J.; Su, Z.C.; Lin, C.F. Silicon-Based Photodetector for Infrared Telecommunication Applications. IEEE Photonics J. 2021, 13, 1–7. [Google Scholar] [CrossRef]
  25. Zhang, H.; Wang, W.; Yip, S.; Li, D.; Li, F.; Lan, C.; Wang, F.; Liu, C.; Ho, J.C. Enhanced performance of near-infrared photodetectors based on InGaAs nanowires enabled by a two-step growth method. J. Mater. Chem. C 2020, 8, 17025–17033. [Google Scholar] [CrossRef]
  26. Ding, M.; Liang, K.; Yu, S.; Zhao, X.; Ren, H.; Zhu, B.; Hou, X.; Wang, Z.; Tan, P.; Huang, H.; et al. Aqueous-Printed Ga2O3 Films for High-Performance Flexible and Heat-Resistant Deep Ultraviolet Photodetector and Array. Adv. Opt. Mater. 2022, 10, 2200512. [Google Scholar] [CrossRef]
  27. Ghosh, C.; Dey, A.; Biswas, I.; Gupta, R.K.; Yadav, V.S.; Yadav, A.; Yadav, N.; Zheng, H.; Henini, M.; Mondal, A. CuO–TiO2 based self-powered broad band photodetector. Nano Mater. Sci. 2024, 6, 345–354. [Google Scholar] [CrossRef]
  28. Guo, Y.; Zhou, J.; Zhao, F.; Wu, Y.; Tao, J.; Zuo, S.; Jiang, J.; Hu, Z.; Chu, J. Carbon-based 2D-layered Rb0.15Cs2.85Sb2ClxI9−x solar cells with superior open-voltage up to 0.88 V. Nano Energy 2021, 88, 106281. [Google Scholar] [CrossRef]
  29. Jain, R.K.; Khanna, A. CuO-doped WO3 thin film H2S sensors. Sens. Actuators B Chem. 2021, 343, 130153. [Google Scholar] [CrossRef]
  30. Li, M.; Mao, D.; Wan, J.; Wang, F.; Zhai, T.; Wang, D. Hollow multi-shell structured SnO2with enhanced performance for ultraviolet photodetectors. Inorg. Chem. Front. 2019, 6, 1968–1972. [Google Scholar] [CrossRef]
  31. Long, Z.; Xu, X.; Yang, W.; Hu, M.; Shtansky, D.V.; Golberg, D.; Fang, X. Cross-Bar SnO2-NiO Nanofiber-Array-Based Transparent Photodetectors with High Detectivity. Adv. Electron. Mater. 2019, 6, 1901048. [Google Scholar] [CrossRef]
  32. Lu, Y.C.; Zhang, Z.F.; Yang, X.; He, G.H.; Lin, C.N.; Chen, X.X.; Zang, J.H.; Zhao, W.B.; Chen, Y.C.; Zhang, L.L.; et al. High-performance solar-blind photodetector arrays constructed from Sn-doped Ga2O3 microwires via patterned electrodes. Nano Res. 2022, 15, 7631–7638. [Google Scholar] [CrossRef]
  33. Mondal, S.; Basak, D. Very high photoresponse towards low-powered UV light under low-biased condition by nanocrystal assembled TiO2 film. Appl. Surf. Sci. 2018, 427, 814–822. [Google Scholar] [CrossRef]
  34. Ouyang, W.; Chen, J.; Shi, Z.; Fang, X. Self-powered UV photodetectors based on ZnO nanomaterials. Appl. Phys. Rev. 2021, 8, 031315. [Google Scholar] [CrossRef]
  35. Puntsagdorj, S.; Koirala, A.R.; Gombovanjil, J.; Khanh, N.N.; Sung, S.D.; Lee, W.I.; Yoon, K.B. Increase in Photocurrent Density of WO3 Photoanode by Placing a Layer of an Ordered Array of Mesoporous WO3 Micropillars on Top of a WO3 Sheet Layer. ACS Appl. Mater. Interfaces 2022, 14, 31838–31850. [Google Scholar] [CrossRef]
  36. Yin, B.; Zhang, H.; Qiu, Y.; Luo, Y.; Zhao, Y.; Hu, L. Piezo-phototronic effect enhanced self-powered and broadband photodetectors based on Si/ZnO/CdO three-component heterojunctions. Nano Energy 2017, 40, 440–446. [Google Scholar] [CrossRef]
  37. Nam, C.Y.; Stein, A. Extreme Carrier Depletion and Superlinear Photoconductivity in Ultrathin Parallel-Aligned ZnO Nanowire Array Photodetectors Fabricated by Infiltration Synthesis. Adv. Opt. Mater. 2017, 5, 1700807. [Google Scholar] [CrossRef]
  38. Ouyang, W.; Teng, F.; He, J.H.; Fang, X. Enhancing the Photoelectric Performance of Photodetectors Based on Metal Oxide Semiconductors by Charge-Carrier Engineering. Adv. Funct. Mater. 2019, 29, 1807672. [Google Scholar] [CrossRef]
  39. Tian, W.; Lu, H.; Li, L. Nanoscale ultraviolet photodetectors based on onedimensional metal oxide nanostructures. Nano Res. 2015, 8, 382–405. [Google Scholar] [CrossRef]
  40. Wang, Y.; Yang, L.; Huang, M.; Kang, Y.; Wang, L.; Zhang, N. A Novel Self-Powered and High Switching Ratio UV–vis Photodetector Based on Multilayer Bi2O3/In2O3 Heterojunction. Phys. Status Solidi (RRL) Rapid Res. Lett. 2024, 18, 2300456. [Google Scholar] [CrossRef]
  41. Wang, Z.; Wang, P.; Wang, F.; Ye, J.; He, T.; Wu, F.; Peng, M.; Wu, P.; Chen, Y.; Zhong, F.; et al. A Noble Metal Dichalcogenide for High-Performance Field-Effect Transistors and Broadband Photodetectors. Adv. Funct. Mater. 2019, 30, 1907945. [Google Scholar] [CrossRef]
  42. Xu, Y.; Shen, H.; Zhang, J.; Zhao, Q.; Wang, Z.; Xu, B.; Zhang, W. High-performance CuS/n-Si heterojunction photodetectors prepared by e-beam evaporation of Cu films as precursor layers. J. Alloys Compd. 2021, 884, 161121. [Google Scholar] [CrossRef]
  43. Singh, S.V.; Sharma, A.; Biring, S.; Pal, B.N. Solution processed Cu2S/TiO2 heterojunction for visible-near infrared photodetector. Thin Solid Film. 2020, 710, 138275. [Google Scholar] [CrossRef]
  44. Wu, J.J.; Tao, Y.R.; Wu, Y.; Wu, X.C. Ultrathin SnS2 nanosheets of ultrasonic synthesis and their photoresponses from ultraviolet to near-infrared. Sens. Actuators B Chem. 2016, 231, 211–217. [Google Scholar] [CrossRef]
  45. Jiang, J.; Xu, W.; Guo, F.; Yang, S.; Ge, W.; Shen, B.; Tang, N. Polarization-Resolved Near-Infrared PdSe2 p-i-n Homojunction Photodetector. Nano Lett. 2023, 23, 9522–9528. [Google Scholar] [CrossRef] [PubMed]
  46. Li, T.; Miao, J.; Fu, X.; Song, B.; Cai, B.; Ge, X.; Zhou, X.; Zhou, P.; Wang, X.; Jariwala, D.; et al. Reconfigurable, non-volatile neuromorphic photovoltaics. Nat. Nanotechnol. 2023, 18, 1303–1310. [Google Scholar] [CrossRef]
  47. Toe, C.Y.; Zhou, S.; Gunawan, M.; Lu, X.; Ng, Y.H.; Amal, R. Recent advances and the design criteria of metal sulfide photocathodes and photoanodes for photoelectrocatalysis. J. Mater. Chem. A 2021, 9, 20277–20319. [Google Scholar] [CrossRef]
  48. Hong, J.; Kim, B.S.; Yang, S.; Jang, A.R.; Lee, Y.W.; Pak, S.; Lee, S.; Cho, Y.; Kang, D.; Shin, H.S.; et al. Chalcogenide solution-mediated activation protocol for scalable and ultrafast synthesis of single-crystalline 1-D copper sulfide for supercapacitors. J. Mater. Chem. A 2019, 7, 2529–2535. [Google Scholar] [CrossRef]
  49. Sun, L.; Leong, W.S.; Yang, S.; Chisholm, M.F.; Liang, S.J.; Ang, L.K.; Tang, Y.; Mao, Y.; Kong, J.; Yang, H.Y. Concurrent Synthesis of High-Performance Monolayer Transition Metal Disulfides. Adv. Funct. Mater. 2017, 27, 1605896. [Google Scholar] [CrossRef]
  50. Wang, L.; Fan, Y.; Han, G.; Ben, H.; Lin, C.; Ma, C. Engineering fast diffusion channels in metal Sulfide/Selenide heterostructures via confinement effect for high-performance sodium storage. Chem. Eng. J. 2024, 493, 152732. [Google Scholar] [CrossRef]
  51. Wang, P.; Zhang, X.; Zhang, J.; Wan, S.; Guo, S.; Lu, G.; Yao, J.; Huang, X. Precise tuning in platinum-nickel/nickel sulfide interface nanowires for synergistic hydrogen evolution catalysis. Nat. Commun. 2017, 8, 14580. [Google Scholar] [CrossRef] [PubMed]
  52. Li, Y.; Tang, L.; Li, R.; Xiang, J.; Teng, K.S.; Lau, S.P. SnS2 quantum dots: Facile synthesis, properties, and applications in ultraviolet photodetector. Chin. Phys. B 2019, 28, 037801. [Google Scholar] [CrossRef]
  53. Gao, W.; Zhang, S.; Zhang, F.; Wen, P.; Zhang, L.; Sun, Y.; Chen, H.; Zheng, Z.; Yang, M.; Luo, D. 2D WS2 based asymmetric Schottky photodetector with high performance. Adv. Electron. Mater. 2021, 7, 2000964. [Google Scholar] [CrossRef]
  54. Zhang, C.; Wang, S.; Yang, L.; Liu, Y.; Xu, T.; Ning, Z.; Zak, A.; Zhang, Z.; Tenne, R.; Chen, Q. High-performance photodetectors for visible and near-infrared lights based on individual WS2 nanotubes. Appl. Phys. Lett. 2012, 100, 243101. [Google Scholar] [CrossRef]
  55. Freitag, M.; Low, T.; Xia, F.; Avouris, P. Photoconductivity of biased graphene. Nat. Photonics 2012, 7, 53–59. [Google Scholar] [CrossRef]
  56. Han, S. New developments in photoconductive detectors (invited). Rev. Sci. Instrum. 1997, 68, 647–652. [Google Scholar] [CrossRef]
  57. Silva, J.P.B.; Vieira, E.M.F.; Gwozdz, K.; Silva, N.E.; Kaim, A.; Istrate, M.C.; Ghica, C.; Correia, J.H.; Pereira, M.; Marques, L.; et al. High-performance and self-powered visible light photodetector using multiple coupled synergetic effects. Mater. Horiz. 2024, 11, 803–812. [Google Scholar] [CrossRef] [PubMed]
  58. Won, U.Y.; Lee, B.H.; Kim, Y.R.; Kang, W.T.; Lee, I.; Kim, J.E.; Lee, Y.H.; Yu, W.J. Efficient photovoltaic effect in graphene/h-BN/silicon heterostructure self-powered photodetector. Nano Res. 2020, 14, 1967–1972. [Google Scholar] [CrossRef]
  59. Liu, S.; Tian, J.; Wu, S.; Zhang, W.; Luo, M. A bioinspired broadband self-powered photodetector based on photo-pyroelectric-thermoelectric effect able to detect human radiation. Nano Energy 2022, 93, 106812. [Google Scholar] [CrossRef]
  60. Liu, Y.; Hu, Q.; Cao, Y.; Wang, P.; Wei, J.; Wu, W.; Wang, J.; Huang, F.; Sun, J.L. High-Performance Ultrabroadband Photodetector Based on Photothermoelectric Effect. ACS Appl. Mater. Interfaces 2022, 14, 29077–29086. [Google Scholar] [CrossRef]
  61. Ma, N.; Yang, Y. Enhanced self-powered UV photoresponse of ferroelectric BaTiO3 materials by pyroelectric effect. Nano Energy 2017, 40, 352–359. [Google Scholar] [CrossRef]
  62. Wu, M.; Jiang, Z.; Lou, X.; Zhang, F.; Song, D.; Ning, S.; Guo, M.; Pennycook, S.J.; Dai, J.Y.; Wen, Z. Flexoelectric Thin-Film Photodetectors. Nano Lett. 2021, 21, 2946–2952. [Google Scholar] [CrossRef] [PubMed]
  63. Jiang, A.; Xing, S.; Lin, H.; Chen, Q.; Li, M. Role of Pyramidal Low-Dimensional Semiconductors in Advancing the Field of Optoelectronics. Photonics 2024, 11, 370. [Google Scholar] [CrossRef]
  64. Morkoc, H.; Omnès, F.; Litton, C.W.; Monroy, E.; Muñoz, E.; Reverchon, J.L. Wide bandgap UV photodetectors: A short review of devices and applications. Gallium Nitride Mater. Devices II 2007, 6473, 111–125. [Google Scholar]
  65. Yadav, P.V.K.; Ajitha, B.; Kumar Reddy, Y.A.; Sreedhar, A. Recent advances in development of nanostructured photodetectors from ultraviolet to infrared region: A review. Chemosphere 2021, 279, 130473. [Google Scholar] [CrossRef]
  66. Vosk, R.; Huse, D.A.; Altman, E. Theory of the Many-Body Localization Transition in One-Dimensional Systems. Phys. Rev. X 2015, 5, 031032. [Google Scholar] [CrossRef]
  67. Yang, Y.; Lin, J.; Huang, Y.; Zhang, Y.; Wang, M.; Zhu, X.; Guo, J.; Fan, K.; Ao, J.; Li, J.; et al. Vertical van der Waals Integrated p-W0.09Re0.91S2 /GaN Heterojunction for Ultra-High Detectivity UV Image Sensing. Laser Photonics Rev. 2024, 19, 2401142. [Google Scholar] [CrossRef]
  68. Ahmad, M.; Pan, C.; Zhao, J.; Zhu, J. Impact of Pb Doping on the Optical and Electrical Properties of ZnO Nanowires. J. Nanosci. Nanotechnol. 2011, 11, 1950–1957. [Google Scholar] [CrossRef]
  69. Kim, C.; Han, S.; Kim, T.; Lee, S. Implantable pH Sensing System Using Vertically Stacked Silicon Nanowire Arrays and Body Channel Communication for Gastroesophageal Reflux Monitoring. Sensors 2024, 24, 861. [Google Scholar] [CrossRef]
  70. Xu, J.; Li, H.; Fang, S.; Jiang, K.; Yao, H.; Fang, F.; Chen, F.; Wang, Y.; Shi, Y. Synthesis of bismuth sulfide nanobelts for high performance broadband photodetectors. J. Mater. Chem. C 2020, 8, 2102–2108. [Google Scholar] [CrossRef]
  71. Gao, H.; Cao, J.; Li, T.; Luo, W.; Gray, M.; Kumar, N.; Burch, K.S.; Ling, X. Phase-Controllable Synthesis of Ultrathin Molybdenum Nitride Crystals Via Atomic Substitution of MoS2. Chem. Mater. 2021, 34, 351–357. [Google Scholar] [CrossRef]
  72. Lei, H.; Yang, G.; Zheng, X.; Zhang, Z.G.; Chen, C.; Ma, J.; Guo, Y.; Chen, Z.; Qin, P.; Li, Y.; et al. Incorporation of High-Mobility and Room-Temperature-Deposited CuxS as a Hole Transport Layer for Efficient and Stable Organo-Lead Halide Perovskite Solar Cells. Sol. RRL 2017, 1, 1700038. [Google Scholar] [CrossRef]
  73. Gao, J.; Chen, R.; Li, D.H.; Jiang, L.; Ye, J.C.; Ma, X.C.; Chen, X.D.; Xiong, Q.H.; Sun, H.D.; Wu, T. UV light emitting transparent conducting tin-doped indium oxide (ITO) nanowires. Nanotechnology 2011, 22, 195706. [Google Scholar] [CrossRef] [PubMed]
  74. Chen, Z.; Lin, X.; Lin, S.; Ren, J.; Wan, L.; Peng, B. Inverted-Structural Self-Powered Gan/PZT/ITO UV Photodetector Enhanced by Ferroelectric Modulation. Adv. Electron. Mater. 2023, 10, 2300588. [Google Scholar] [CrossRef]
  75. Wu, P.; Song, X.; Si, S.; Ke, Z.; Cheng, L.; Li, W.; Xiao, X.; Jiang, C. Significantly enhanced visible light response in single TiO2 nanowire by nitrogen ion implantation. Nanotechnology 2018, 29, 184005. [Google Scholar] [CrossRef]
  76. Sirkeli, V.P.; Yilmazoglu, O.; Hajo, A.S.; Nedeoglo, N.D.; Nedeoglo, D.D.; Preu, S.; Küppers, F.; Hartnagel, H.L. Enhanced responsivity of ZnSe-based metal–semiconductor–metal near-ultraviolet photodetector via impact ionization. Phys. Status Solidi (RRL)–Rapid Res. Lett. 2018, 12, 1700418. [Google Scholar] [CrossRef]
  77. Singh, A.K.; Chowdhury, N.K.; Hazra, A.; Bhowmik, B. Room-Temperature Au/TiO2 Nanorods/Ti TFT Butanone Sensor: Role of Surface States. J. Electron. Mater. 2023, 52, 3622–3632. [Google Scholar] [CrossRef]
  78. Sirkeli, V.P.; Yilmazoglu, O.; Hajo, A.S.; Nedeoglo, N.D.; Nedeoglo, D.D.; Küppers, F.; Hartnagel, H.L. High Performance ZnSe-Based Metal–Semiconductor–Metal Ultraviolet Photodetectors with Different Schottky Contacts. Phys. Solid State 2024, 66, 257–264. [Google Scholar] [CrossRef]
  79. Dahiya, S.; Hazra, S.; Pandey, U.; Pramanik, S.; Dahiya, P.; Singh, S.V.; Kumari, N.; Pal, B.N. Enhancement in photo-response of CuZnS nanocrystals-based photodetector using asymmetric work function electrodes. Opt. Mater. 2024, 157, 116182. [Google Scholar] [CrossRef]
  80. Yang, Z.; Liu, M.; Li, W.; Xu, J.; Wan, P.; Xu, T.; Shi, D.; Kan, C.; Jiang, M. Boosting photoresponses in a SnO2 microwire heterojunction ultraviolet self-biased photodetector through tailoring heterointerface. Surf. Interfaces 2024, 46, 104001. [Google Scholar] [CrossRef]
  81. Zhang, L.; Geng, X.; Zha, G.; Xu, J.; Wei, S.; Ma, B.; Chen, Z.; Shang, X.; Ni, H.; Niu, Z. Self-catalyzed molecular beam epitaxy growth and their optoelectronic properties of vertical GaAs nanowires on Si(111). Mater. Sci. Semicond. Process. 2016, 52, 68–74. [Google Scholar] [CrossRef]
  82. Hsieh, Y.-Y.; Tuan, H.-Y. Architectural van der Waals Bi2S3/Bi2Se3 topological heterostructure as a superior potassium-ion storage material. Energy Storage Mater. 2022, 51, 789–805. [Google Scholar] [CrossRef]
  83. Li, D.; Liu, Y.; Shi, W.; Shao, C.; Wang, S.; Ding, C.; Liu, T.; Fan, F.; Shi, J.; Li, C. Crystallographic-orientation-dependent charge separation of BiVO4 for solar water oxidation. ACS Energy Lett. 2019, 4, 825–831. [Google Scholar] [CrossRef]
  84. He, B.; Jia, S.; Zhao, M.; Wang, Y.; Chen, T.; Zhao, S.; Li, Z.; Lin, Z.; Zhao, Y.; Liu, X. General and Robust Photothermal-Heating-Enabled High-Efficiency Photoelectrochemical Water Splitting. Adv. Mater. 2021, 33, 2004406. [Google Scholar] [CrossRef] [PubMed]
  85. Hu, Z.; Ying, L.; Huang, F.; Cao, Y. Towards a bright future: Polymer solar cells with power conversion efficiencies over 10%. Sci. China Chem. 2017, 60, 571–582. [Google Scholar] [CrossRef]
  86. Ren, Z.; Wang, Q.; Zhang, G.; Zhang, T.; Liu, J.; Wang, S.; Qiao, S. Self-powered CdS nanorods/planar-Si photodetector and its performance optimization by fully developing pyro-phototronic effect. Surf. Interfaces 2023, 42, 103495. [Google Scholar] [CrossRef]
  87. Li, J.; Li, R.; Wang, W.; Lan, K.; Zhao, D. Ordered Mesoporous Crystalline Frameworks Toward Promising Energy Applications. Adv. Mater. 2024, 36, 2311460. [Google Scholar] [CrossRef]
  88. Jamal, F.; Rafique, A.; Moeen, S.; Haider, J.; Nabgan, W.; Haider, A.; Imran, M.; Nazir, G.; Alhassan, M.; Ikram, M.; et al. Review of Metal Sulfide Nanostructures and their Applications. ACS Appl. Nano Mater. 2023, 6, 7077–7106. [Google Scholar] [CrossRef]
  89. Wang, C.L.; Wu, Q.Q.; Ding, Y.; Zhang, X.M.; Wang, W.H.; Guo, X.T.; Ni, Z.H.; Lin, L.G.; Cai, Z.Y.; Gu, X.F.; et al. High-Responsivity and Broadband MoS2 Photodetector Using Interfacial Engineering. ACS Appl. Mater. Interfaces 2023, 15, 46236–46246. [Google Scholar] [CrossRef]
  90. Zhang, G.; Guan, Z.; Yang, J.; Li, Q.; Zhou, Y.; Zou, Z. Metal Sulfides for Photocatalytic Hydrogen Production: Current Development and Future Challenges. Sol. RRL 2022, 6, 2200587. [Google Scholar] [CrossRef]
  91. Yoneda, K.; Toda, T.; Hishida, Y.; Niina, T. The preparation of conductive ZnS films by using MBE with a single effusion source. J. Cryst. Growth 1984, 67, 125–134. [Google Scholar] [CrossRef]
  92. Arsad, A.Z.; Bahrudin, M.S.; Arzaee, N.A.; Rahman, M.N.A.; Chau, C.F.; Abdullah, S.F.; Zuhdi, A.W.M. Zinc sulfide thin films deposited by chemical bath: Tuning consideration of structural, optical band gap, and electrical properties for CIGS solar cells application. Ceram. Int. 2024, 50, 11776–11786. [Google Scholar] [CrossRef]
  93. Azmand, A.; Kafashan, H. Physical and electrochemical properties of electrodeposited undoped and Se-doped ZnS thin films. Ceram. Int. 2018, 44, 17124–17137. [Google Scholar] [CrossRef]
  94. Zhao, W.; Wei, Z.; Zhang, L.; Wu, X.; Wang, X.; Jiang, J. Optical and magnetic properties of Co and Ni co-doped ZnS nanorods prepared by hydrothermal method. J. Alloys Compd. 2017, 698, 754–760. [Google Scholar] [CrossRef]
  95. Alzaid, M.; Alwshih, M.; Abd-el Salam, M.N.; Hadia, N.M.A. Role of Cu dilute on microstructures, optical, photoluminescence, magnetic and electrical properties of CdS film. Mater. Sci. Semicond. Process. 2021, 127, 105687. [Google Scholar] [CrossRef]
  96. Deng, K.; Li, L. CdS Nanoscale Photodetectors. Adv. Mater. 2014, 26, 2619–2635. [Google Scholar] [CrossRef]
  97. Salas-Villasenor, A.L.; Mejia, I.; Sotelo-Lerma, M.; Guo, Z.B.; Alshareef, H.N.; Quevedo-Lopez, M.A. Improved electrical stability of CdS thin film transistors through hydrogen-based thermal treatments. Semicond. Sci. Technol. 2014, 29, 085001. [Google Scholar] [CrossRef]
  98. Cui, Y.; Zhou, Z.; Wang, X.; Wang, X.; Ren, Z.; Pan, L.; Yang, J. Wavelength-selectivity polarization dependence of optical absorption and photoresponse in SnS nanosheets. Nano Res. 2021, 14, 2224–2230. [Google Scholar] [CrossRef]
  99. Gotoh, T. Control of carrier concentration in SnS films by annealing with S and Sn. Phys. Status Solidi (A) 2016, 213, 1869–1872. [Google Scholar] [CrossRef]
  100. Lim, D.; Suh, H.; Suryawanshi, M.; Song, G.Y.; Cho, J.Y.; Kim, J.H.; Jang, J.H.; Jeon, C.W.; Cho, A.; Ahn, S.; et al. Kinetically Controlled Growth of Phase-Pure SnS Absorbers for Thin Film Solar Cells: Achieving Efficiency Near 3% with Long-Term Stability Using an SnS/CdS Heterojunction. Adv. Energy Mater. 2018, 8, 1702605. [Google Scholar] [CrossRef]
  101. Jeff, J.P.M.S.; Cindy, H.X.L.; Rebecca, A.D.; Li, R.X.; Jin, L.; Ma, T.; Wilhelmus, M.M.K.; Steven, J.; Ageeth, A.B. Correction to Nb Doping and Alloying of 2D WS2 by Atomic Layer Deposition for 2D Transition Metal Dichalcogenide Transistors and HER Electrocatalysts. ACS Appl. Nano Mater 2024, 7, 12205–12206. [Google Scholar]
  102. Mohamed, A.B.; Shrouk, E.Z.; Yasin, R.E. Fast optoelectronic gas sensing with p-type V2O5/WS2/Si heterojunction thin film. Mater. Chem. Phys. 2023, 310, 128491. [Google Scholar]
  103. Aji, A.S.; Solís-Fernández, P.; Ji, H.G.; Fukuda, K.; Ago, H. High Mobility WS2 Transistors Realized by Multilayer Graphene Electrodes and Application to High Responsivity Flexible Photodetectors. Adv. Funct. Mater. 2017, 27, 1703448. [Google Scholar] [CrossRef]
  104. Hieu, N.N.; Ilyasov, V.V.; Vu, T.V.; Poklonski, N.A.; Phuc, H.V.; Phuong, L.T.T.; Hoi, B.D.; Nguyen, C.V. First principles study of optical properties of molybdenum disulfide: From bulk to monolayer. Superlattices Microstruct. 2018, 115, 10–18. [Google Scholar] [CrossRef]
  105. Lin, J.; Monaghan, S.; Sakhuja, N.; Gity, F.; Jha, R.K.; Coleman, E.M.; Connolly, J.; Cullen, C.P.; Walsh, L.A.; Mannarino, T.; et al. Large-area growth of MoS2 at temperatures compatible with integrating back-end-of-line functionality. 2D Mater. 2020, 8, 025008. [Google Scholar] [CrossRef]
  106. Wang, J.W.; Liu, Y.P.; Chen, P.H.; Chuang, M.H.; Pezeshki, A.; Ling, D.C.; Chen, J.C.; Chen, Y.F.; Lee, Y.H. Controlled Low-Frequency Electrical Noise of Monolayer MoS2 with Ohmic Contact and Tunable Carrier Concentration. Adv. Electron. Mater. 2017, 4, 1700340. [Google Scholar] [CrossRef]
  107. Yang, L.; Yuan, X.; Liu, R.; Wu, P.; Zhong, Y.; Zhu, F.; Chang, W. Intrinsic properties of metallic edge states in MoS2 nanobelt. J. Mater. Sci. Mater. Electron. 2022, 33, 23722–23728. [Google Scholar] [CrossRef]
  108. Huang, Y.; Sutter, E.; Sadowski, J.T.; Cotlet, M.; Monti, O.L.A.; Racke, D.A.; Neupane, M.R.; Wickramaratne, D.; Lake, R.K.; Parkinson, B.A. Tin Disulfide-An Emerging Layered Metal Dichalcogenide Semiconductor: Materials Properties and Device Characteristics. ACS Nano 2014, 8, 10743–10755. [Google Scholar] [CrossRef]
  109. Arulanantham, A.M.S.; Raj, P.M.; Gunavathy, K.V.; Begam, M.R.; Ganesh, V.; Algarni, H. Photosensitivity properties of Eu-doped SnS2 thin films deposited by cost-effective nebulizer spray pyrolysis technique. Appl. Phys. A 2022, 128, 364. [Google Scholar] [CrossRef]
  110. Pradhan, N.R.; McCreary, A.; Rhodes, D.; Lu, Z.; Feng, S.; Manousakis, E.; Smirnov, D.; Namburu, R.; Dubey, M.; Walker, A.R. Metal to Insulator Quantum-Phase Transition in Few-Layered ReS2. Nano Lett 2015, 15, 8377–8384. [Google Scholar] [CrossRef]
  111. Leicht, G.; Berger, H.; Levy, F. The growth of n- and p-type ReS2 and ReSe2 single crystals and their electrical properties. Solid State Commun. 1987, 61, 531–534. [Google Scholar] [CrossRef]
  112. Bangolla, H.K.; Yusuf Fakhri, M.; Lin, C.H.; Cheng, C.M.; Lu, Y.H.; Fu, T.Y.; Selvarasu, P.; Ulaganathan, R.K.; Sankar, R.; Chen, R.S. Electrical and optoelectronic anisotropy and surface electron accumulation in ReS2 nanostructures. Nanoscale 2023, 15, 19735–19745. [Google Scholar] [CrossRef] [PubMed]
  113. Liu, M.; Gong, Y.; Li, Z.; Dou, M.; Wang, F. A green and facile hydrothermal approach for the synthesis of high-quality semi-conducting Sb2S3 thin films. Appl. Surf. Sci. 2016, 387, 790–795. [Google Scholar] [CrossRef]
  114. Zhang, H.; Yuan, S.; Deng, H.; Ishaq, M.; Yang, X.; Hou, T.; Shah, U.A.; Song, H.; Tang, J. Controllable orientations for Sb2S3 solar cells by vertical VTD method. Prog. Photovolt. Res. Appl. 2020, 28, 823–832. [Google Scholar] [CrossRef]
  115. Zhong, P.; Xie, J.; Bagheri, R.; Yi, Q.; Chen, Q.; Tan, J.; He, L.; Zhang, F.; Zhang, L.; Zou, G. An aqueous solution method towards Sb2S3 thin films for photoanodes. Chem. Commun. 2019, 55, 14530–14533. [Google Scholar] [CrossRef]
  116. Gao, C.; Shen, H.; Sun, L.; Shen, Z. Chemical bath deposition of Bi2S3 films by a novel deposition system. Appl. Surf. Sci. 2011, 257, 7529–7533. [Google Scholar] [CrossRef]
  117. Ji, W.; Shi, X.L.; Liu, W.D.; Yuan, H.; Zheng, K.; Wan, B.; Shen, W.; Zhang, Z.; Fang, C.; Wang, Q.; et al. Boosting the thermoelectric performance of n-type Bi2S3 by hierarchical structure manipulation and carrier density optimization. Nano Energy 2021, 87, 106171. [Google Scholar] [CrossRef]
  118. Medles, M.; Benramdane, N.; Bouzidi, A.; Nakrela, A.; Tabet-Derraz, H.; Kebbab, Z.; Mathieu, C.; Khelifa, B.; Desfeux, R. Optical and electrical properties of Bi2S3 films deposited by spray pyrolysis. Thin Solid Film. 2006, 497, 58–64. [Google Scholar] [CrossRef]
  119. Liu, M.; Li, Z.; Ji, J.; Dou, M.; Wang, F. Properties of nanostructured pure β-In2S3 thin films prepared by sulfurization-assisted electrodeposition. J. Mater. Sci. Mater. Electron. 2016, 28, 5044–5052. [Google Scholar] [CrossRef]
  120. Soni, V.; Raizada, P.; Kumar, A.; Hasija, V.; Singal, S.; Singh, P.; Hosseini-Bandegharaei, A.; Thakur, V.K.; Nguyen, V.H. Indium sulfide-based photocatalysts for hydrogen production and water cleaning: A review. Environ. Chem. Lett. 2021, 19, 1065–1095. [Google Scholar] [CrossRef]
  121. Ferrer, I.J.; Ares, J.R.; Clamagirand, J.M.; Barawi, M.; Sánchez, C. Optical properties of titanium trisulphide (TiS3) thin films. Thin Solid Film. 2013, 535, 398–401. [Google Scholar] [CrossRef]
  122. Muratov, D.S.; Vanyushin, V.O.; Vorobeva, N.S.; Jukova, P.; Lipatov, A.; Kolesnikov, E.A.; Karpenkov, D.; Kuznetsov, D.V.; Sinitskii, A. Synthesis and exfoliation of quasi-1D (Zr,Ti)S3 solid solutions for device measurements. J. Alloys Compd. 2020, 815, 152316. [Google Scholar] [CrossRef]
  123. Tripathi, N.; Pavelyev, V.; Sharma, P.; Kumar, S.; Rymzhina, A.; Mishra, P. Review of titanium trisulfide (TiS3): A novel material for next generation electronic and optical devices. Mater. Sci. Semicond. Process. 2021, 127, 105699. [Google Scholar] [CrossRef]
  124. Ebrahimi, S.; Yarmand, B.; Naderi, N. ZnS-Based UV Detectors. In Handbook of II-VI Semiconductor-Based Sensors and Radiation Detectors; Springer: Berlin/Heidelberg, Germany, 2023; Volume 2, pp. 333–347. [Google Scholar]
  125. Peng, H.; Liuyang, B.; Lingjie, Y.; Jinlin, L.; Fangli, Y.; Yunfa, C. Shape-Controlled Synthesis of ZnS Nanostructures: A Simple and Rapid Method for One-Dimensional Materials by Plasma. Nanoscale Res. Lett. 2009, 4, 1047–1053. [Google Scholar] [CrossRef]
  126. Fang, X.; Wu, L.; Hu, L. ZnS Nanostructure Arrays: A Developing Material Star. Adv. Mater. 2010, 23, 585–598. [Google Scholar] [CrossRef] [PubMed]
  127. Wei, M.; Yang, L.; Yan, Y.; Ni, L. Preparation of ZnS quantum dot photocatalyst and study on photocatalytic degradation of antibiotics. Mater. Express 2019, 9, 413–418. [Google Scholar] [CrossRef]
  128. Xu, X.; Li, S.; Chen, J.; Cai, S.; Long, Z.; Fang, X. Design Principles and Material Engineering of ZnS for Optoelectronic Devices and Catalysis. Adv. Funct. Mater. 2018, 28, 1802029. [Google Scholar] [CrossRef]
  129. Sue, Y.S.; Pan, K.Y.; Wei, D.H. Optoelectronic and photocatalytic properties of zinc sulfide nanowires synthesized by vapor-liquid-solid process. Appl. Surf. Sci. 2019, 471, 435–444. [Google Scholar] [CrossRef]
  130. Dai, J.; Song, X.; Zheng, H.; Wu, C. Excitonic photoluminescence and photoresponse of ZnS nanowires. Mater. Chem. Phys. 2016, 174, 204–208. [Google Scholar] [CrossRef]
  131. An, Q.; Meng, X.; Xiong, K.; Qiu, Y.; Lin, W. One-step fabrication of single-crystalline ZnS nanotubes with a novel hollow structure and large surface area for photodetector devices. Nanotechnology 2017, 28, 105502. [Google Scholar] [CrossRef]
  132. An, Q.; Meng, X.; Xiong, K.; Qiu, Y. Self-powered ZnS Nanotubes/Ag Nanowires MSM UV Photodetector with High On/Off Ratio and Fast Response Speed. Sci. Rep. 2017, 7, 4885. [Google Scholar] [CrossRef] [PubMed]
  133. Liang, Y.; Liang, H.; Xiao, X.; Hark, S. The epitaxial growth of ZnS nanowire arrays and their applications in UV-light detection. J. Mater. Chem. 2012, 22, 1199–1205. [Google Scholar] [CrossRef]
  134. Saeed, S.; Dai, R.; Janjua, R.A.; Huang, D.; Wang, H.; Wang, Z.; Ding, Z.; Zhang, Z. Fast-Response Metal–Semiconductor–Metal Junction Ultraviolet Photodetector Based on ZnS:Mn Nanorod Networks via a Cost-Effective Method. ACS Omega 2021, 6, 32930–32937. [Google Scholar] [CrossRef] [PubMed]
  135. Jiang, P.; Jie, J.; Yu, Y.; Wang, Z.; Xie, C.; Zhang, X.; Wu, C.; Wang, L.; Zhu, Z.; Luo, L. Aluminium-doped n-type ZnS nanowires as high-performance UV and humidity sensors. J. Mater. Chem. 2012, 22, 6856–6861. [Google Scholar] [CrossRef]
  136. Zhang, K.; Ding, J.; Lou, Z.; Chai, R.; Zhong, M.; Shen, G. Heterostructured ZnS/InP nanowires for rigid/flexible ultraviolet photodetectors with enhanced performance. Nanoscale 2017, 9, 15416–15422. [Google Scholar] [CrossRef]
  137. Lou, Z.; Li, L.; Shen, G. Ultraviolet/visible photodetectors with ultrafast, high photosensitivity based on 1D ZnS/CdS heterostructures. Nanoscale 2016, 8, 5219–5225. [Google Scholar] [CrossRef]
  138. Lee, J.K.; Sun, G.J.; Lee, W.S.; Hyun, S.K.; Kim, K.K.; Choi, S.B.; Lee, C. Ultraintense UV emission from ZnO-sheathed ZnS nanorods. Sci. Rep. 2017, 7, 13034. [Google Scholar] [CrossRef]
  139. Ren, B.; Cao, M.; Zhang, Q.; Huang, J.; Zhao, Z.; Jin, X.; Li, C.; Shen, Y.; Wang, L. Controllable synthesis of CdS nanowire by a facile solvothermal method and its temperature dependent photoluminescent property. J. Alloys Compd. 2016, 659, 74–81. [Google Scholar] [CrossRef]
  140. Kumar, V.; Kumar, K.; Jeon, H.C.; Kang, T.W.; Lee, D.; Kumar, S. Effect of Cu-doping on the photoluminescence and photoconductivity of template synthesized CdS nanowires. J. Phys. Chem. Solids 2019, 124, 1–6. [Google Scholar] [CrossRef]
  141. Zhao, X.; Wang, X.; Guo, S.; Wu, M.; Quan, S.; Ma, Z. Enhanced photoresponse of the CdS microwire photodetectors based on indium ion implantation. J. Phys. D Appl. Phys. 2023, 57, 115102. [Google Scholar] [CrossRef]
  142. Chavan, P.G.; Kashid, R.V.; Badhade, S.S.; Mulla, I.S.; More, M.A.; Joag, D.S. CdS nanowires: Ultra-long growth and enhanced field emission properties. Vacuum 2014, 101, 38–45. [Google Scholar] [CrossRef]
  143. Weng, B.; Liu, S.; Zhang, N.; Tang, Z.R.; Xu, Y.J. A simple yet efficient visible-light-driven CdS nanowires-carbon nanotube 1D–1D nanocomposite photocatalyst. J. Catal. 2014, 309, 146–155. [Google Scholar] [CrossRef]
  144. Li, L.; Wu, P.; Fang, X.; Zhai, T.; Dai, L.; Liao, M.; Koide, Y.; Wang, H.; Bando, Y.; Golberg, D. Single-Crystalline CdS Nanobelts for Excellent Field-Emitters and Ultrahigh Quantum-Efficiency Photodetectors. Adv. Mater. 2010, 22, 3161–3165. [Google Scholar] [CrossRef]
  145. Zhou, W.; Peng, Y.; Yin, Y.; Zhou, Y.; Zhang, Y.; Tang, D. Broad spectral response photodetector based on individual tin-doped CdS nanowire. AIP Adv. 2014, 4, 123005. [Google Scholar] [CrossRef]
  146. Li, G.; Jiang, Y.; Zhang, Y.; Lan, X.; Zhai, T.; Yi, G.C. High-performance photodetectors and enhanced field-emission of CdS nanowire arrays on CdSe single-crystalline sheets. J. Mater. Chem. C 2014, 2, 8252–8258. [Google Scholar] [CrossRef]
  147. Li, L.; Lou, Z.; Shen, G. Hierarchical CdS Nanowires Based Rigid and Flexible Photodetectors with Ultrahigh Sensitivity. ACS Appl. Mater. Interfaces 2015, 7, 23507–23514. [Google Scholar] [CrossRef]
  148. Zheng, D.; Fang, H.; Wang, P.; Luo, W.; Gong, F.; Ho, J.C.; Chen, X.; Lu, W.; Liao, L.; Wang, J.; et al. High-Performance Ferroelectric Polymer Side-Gated CdS Nanowire Ultraviolet Photodetectors. Adv. Funct. Mater. 2016, 26, 7690–7696. [Google Scholar] [CrossRef]
  149. Zhao, W.; Liu, L.; Xu, M.; Wang, X.; Zhang, T.; Wang, Y.; Zhang, Z.; Qin, S.; Liu, Z. Single CdS Nanorod for High Responsivity UV–Visible Photodetector. Adv. Opt. Mater. 2017, 5, 1700159. [Google Scholar] [CrossRef]
  150. Chai, R.; Lou, Z.; Shen, G. Highly flexible self-powered photodetectors based on core–shell Sb/CdS nanowires. J. Mater. Chem. C 2019, 7, 4581–4586. [Google Scholar] [CrossRef]
  151. Chang, Y.; Wang, J.; Wu, F.; Tian, W.; Zhai, W. Structural Design and Pyroelectric Property of SnS/CdS Heterojunctions Contrived for Low-Temperature Visible Photodetectors. Adv. Funct. Mater. 2020, 30, 2001450. [Google Scholar] [CrossRef]
  152. Zhang, M.; Hu, Q.; Ma, K.; Ding, Y.; Li, C. Pyroelectric effect in CdS nanorods decorated with a molecular Co-catalyst for hydrogen evolution. Nano Energy 2020, 73, 104810. [Google Scholar] [CrossRef]
  153. Lin, H.; Liu, H.; Qian, X.; Lai, S.-W.; Li, Y.; Chen, N.; Ouyang, C.; Che, C.-M.; Li, Y. Constructing a Blue Light Photodetector on Inorganic/Organic p–n Heterojunction Nanowire Arrays. Inorg. Chem. 2011, 50, 7749–7753. [Google Scholar] [CrossRef]
  154. Lin, H.; Liu, H.; Qian, X.; Chen, S.; Li, Y.; Li, Y. Synthesizing Axial Inserting p–n Heterojunction Nanowire Arrays for Realizing Synergistic Performance. Inorg. Chem. 2013, 52, 6969–6974. [Google Scholar] [CrossRef]
  155. Lin, H.; Jiang, A.; Xing, S.; Li, L.; Cheng, W.; Li, J.; Miao, W.; Zhou, X.; Tian, L. Advances in Self-Powered Ultraviolet Photodetectors Based on P-N Heterojunction Low-Dimensional Nanostructures. Nanomaterials 2022, 12, 910. [Google Scholar] [CrossRef] [PubMed]
  156. Lin, H.; Xing, S.; Jiang, A.; Li, M.; Chen, Q.; Wang, Z.; Jiang, L.; Li, H.; Wang, J.; Zhou, C. Controlled Synthesis of Large-Area Oriented ZnO Nanoarrays. Nanomaterials 2024, 14, 1028. [Google Scholar] [CrossRef] [PubMed]
  157. Rani, A.; Verma, A.; Singh, A.; Yadav, B.C. Monitoring of UV-A radiation by TiO2/CdS nanohybrid along with the high on-off ratio. Sens. Actuators A Phys. 2024, 367, 115060. [Google Scholar] [CrossRef]
  158. Vashishtha, P.; Abidi, I.H.; Giridhar, S.P.; Verma, A.K.; Prajapat, P.; Bhoriya, A.; Murdoch, B.J.; Tollerud, J.O.; Xu, C.; Davis, J.A.; et al. CVD-Grown Monolayer MoS2 and GaN Thin Film Heterostructure for a Self-Powered and Bidirectional Photodetector with an Extended Active Spectrum. ACS Appl. Mater. Interfaces 2024, 16, 31294–31303. [Google Scholar] [CrossRef] [PubMed]
  159. Krishnamurthi, V.; Khan, H.; Ahmed, T.; Zavabeti, A.; Tawfik, S.A.; Jain, S.K.; Spencer, M.J.S.; Balendhran, S.; Crozier, K.B.; Li, Z.; et al. Liquid-Metal Synthesized Ultrathin SnS Layers for High-Performance Broadband Photodetectors. Adv. Mater. 2020, 32, 2004247. [Google Scholar] [CrossRef]
  160. Tian, H.; Fan, C.; Liu, G.; Yuan, S.; Zhang, Y.; Wang, M.; Li, E. Ultrafast broadband photodetector based on SnS synthesized by hydrothermal method. Appl. Surf. Sci. 2019, 487, 1043–1048. [Google Scholar] [CrossRef]
  161. Patel, M.; Kim, J.; Kim, Y.K. Growth of Large-Area SnS Films with Oriented 2D SnS Layers for Energy-Efficient Broadband Optoelectronics. Adv. Funct. Mater. 2018, 28, 1804737. [Google Scholar] [CrossRef]
  162. Singh, P.; Rhee, D.; Baek, S.; Yoo, H.H.; Niu, J.; Jung, M.; Kang, J.; Lee, S. SnS/MoS2 van der Waals heterojunction for in-plane ferroelectric field-effect transistors with multibit memory and logic characteristics. EcoMat 2023, 5, e12333. [Google Scholar] [CrossRef]
  163. Ouyang, B.; He, W.; Wu, L.; Zhao, L.D.; Yang, Y. Thermo-phototronic effect in p-type Na-doped SnS single crystals for enhanced self-powered photodetectors. Nano Energy 2021, 88, 106268. [Google Scholar] [CrossRef]
  164. Zheng, D.; Fang, H.; Long, M.; Wu, F.; Wang, P.; Gong, F.; Wu, X.; Ho, J.C.; Liao, L.; Hu, W. High-Performance Near-Infrared Photodetectors Based on p-Type SnX (X = S, Se) Nanowires Grown via Chemical Vapor Deposition. ACS Nano 2018, 12, 7239–7245. [Google Scholar] [CrossRef]
  165. Chen, R.; Li, L.; Jiang, L.; Yu, X.; Zhu, D.; Xiong, Y.; Zheng, D.; Yang, W. Small-diameter p-type SnS nanowire photodetectors and phototransistors with low-noise and high-performance. Nanotechnology 2022, 33, 135707. [Google Scholar] [CrossRef] [PubMed]
  166. Fang, F.; Li, H.; Yao, H.; Jiang, K.; Liu, Z.; Lin, C.; Chen, F.; Wang, Y.; Liu, L. Two-Dimensional Hybrid Composites of SnS2 Nanosheets Array Film with Graphene for Enhanced Photoelectric Performance. Nanomaterials 2019, 9, 1122. [Google Scholar] [CrossRef]
  167. Zhang, L.; Wei, Z.; Wang, X.; Zhang, L.; Wang, Y.; Xie, C.; Han, T.; Li, F.; Luo, W.; Zhao, D.; et al. Ultrahigh-Sensitivity and Fast-Speed Solar-Blind Ultraviolet Photodetector Based on a Broken-Gap van der Waals Heterodiode. ACS Appl. Mater. Interfaces 2023, 15, 14513–14522. [Google Scholar] [CrossRef]
  168. Zhao, X.; Liu, S.; Zhang, H.; Chang, S.Y.; Huang, W.; Zhu, B.; Shen, Y.; Shen, C.; Wang, D.; Yang, Y.; et al. 20% Efficient Perovskite Solar Cells with 2D Electron Transporting Layer. Adv. Funct. Mater. 2018, 29, 1805168. [Google Scholar] [CrossRef]
  169. Kovacic, M.; Papac, J.; Kusic, H.; Karamanis, P.; Loncaric Bozic, A. Degradation of polar and non-polar pharmaceutical pollutants in water by solar assisted photocatalysis using hydrothermal TiO2-SnS2. Chem. Eng. J. 2020, 382, 122826. [Google Scholar] [CrossRef]
  170. Yu, J.; Suleiman, A.A.; Zheng, Z.; Zhou, X.; Zhai, T. Giant-Enhanced SnS2 Photodetectors with Broadband Response through Oxygen Plasma Treatment. Adv. Funct. Mater. 2020, 30, 2001650. [Google Scholar] [CrossRef]
  171. Zervos, M.; Mihailescu, C.N.; Giapintzakis, J.; Othonos, A.; Luculescu, C.R. Surface passivation and conversion of SnO2 to SnS2 nanowires. Mater. Sci. Eng. B 2015, 198, 10–13. [Google Scholar] [CrossRef]
  172. Zheng, D.; Wang, H.; Chen, R.; Li, L.; Guo, J.; Gu, Y.; Zubair, M.M.; Yu, X.; Jiang, L.; Zhu, D.; et al. High-detectivity tin disulfide nanowire photodetectors with manipulation of localized ferroelectric polarization field. Nanophotonics 2021, 10, 4637–4644. [Google Scholar] [CrossRef]
  173. Das, S.; Pal, S.; Larsson, K.; Mandal, D.; Giri, S.; Banerji, P.; Chandra, A.; Basori, R. Hydrothermally grown SnS2/Si nanowire core-shell heterostructure photodetector with excellent optoelectronic performances. Appl. Surf. Sci. 2023, 624, 157094. [Google Scholar] [CrossRef]
  174. Das, S.; Pal, S.; Mandal, D.; Banerji, P.; Chandra, A.; Basori, R. Enhanced Optoelectronic Performance of Silicon Nanowire/SnS2 Core-Shell Heterostructure With Defect Passivation in SnS2 by UV Treatment. IEEE Trans. Electron Devices 2023, 70, 4008–4013. [Google Scholar] [CrossRef]
  175. Das, S.; Sarkar, K.J.; Pal, B.; Mondal, H.; Pal, S.; Basori, R.; Banerji, P. SnS2/Si nanowire vertical heterostructure for high performance ultra-low power broadband photodetector with excellent detectivity. J. Appl. Phys. 2021, 129, 144671. [Google Scholar] [CrossRef]
  176. Kang, S.; Eshete, Y.A.; Lee, S.; Won, D.; Im, S.; Lee, S.; Cho, S.; Yang, H. Bandgap modulation in the two-dimensional core-shell-structured monolayers of WS2. iScience 2022, 25, 103563. [Google Scholar] [CrossRef] [PubMed]
  177. Gherabli, R.; Indukuri, S.R.K.C.; Zektzer, R.; Frydendahl, C.; Levy, U. MoSe2/WS2 heterojunction photodiode integrated with a silicon nitride waveguide for near infrared light detection with high responsivity. Light Sci. Appl. 2023, 12, 60. [Google Scholar] [CrossRef] [PubMed]
  178. Tan, H.; Fan, Y.; Zhou, Y.; Chen, Q.; Xu, W.; Warner, J.H. Ultrathin 2D Photodetectors Utilizing Chemical Vapor Deposition Grown WS2 With Graphene Electrodes. ACS Nano 2016, 10, 7866–7873. [Google Scholar] [CrossRef]
  179. Lin, Y.; Adilbekova, B.; Firdaus, Y.; Yengel, E.; Faber, H.; Sajjad, M.; Zheng, X.; Yarali, E.; Seitkhan, A.; Bakr, O.M.; et al. 17% Efficient Organic Solar Cells Based on Liquid Exfoliated WS2 as a Replacement for PEDOT:PSS. Adv. Mater. 2019, 31, 1902965. [Google Scholar] [CrossRef]
  180. Ghosh, S.; Brüser, V.; Kaplan-Ashiri, I.; Popovitz-Biro, R.; Peglow, S.; Martínez, J.I.; Alonso, J.A.; Zak, A. Cathodoluminescence in single and multiwall WS2 nanotubes: Evidence for quantum confinement and strain effect. Appl. Phys. Rev. 2020, 7, 041401. [Google Scholar] [CrossRef]
  181. Kumar, P.; Balakrishnan, V. Growth and microstructural evolution of WS2 nanostructures with tunable field and light modulated electrical transport. Appl. Surf. Sci. 2018, 436, 846–853. [Google Scholar] [CrossRef]
  182. Liu, F.; Zhou, H.; Gu, Y.; Dong, Z.; Yang, Y.; Wang, Z.; Zhang, T.; Wu, W. Solution Processed Photodetectors with PVK-WS2 Nanotube/Nanofullerene Organic–Inorganic Hybrid Films. ACS Appl. Mater. Interfaces 2022, 14, 43612–43620. [Google Scholar] [CrossRef] [PubMed]
  183. Chaudhary, N.; Khanuja, M. Architectural Design of Photodetector Based on 2D (MoS2 Nanosheets)/1D (WS2 Nanorods) Heterostructure Synthesized by Facile Hydrothermal Method. J. Electrochem. Soc. 2019, 166, B1276–B1285. [Google Scholar] [CrossRef]
  184. Gao, Y.; Du, Z.; Jiang, X.; Zheng, J.; Jiang, J.; Ye, S.; Zou, B.; Yang, T.; Zhao, J. Broadband MoS2 Square Nanotube-Based Photodetectors. ACS Appl. Nano Mater. 2023, 6, 7044–7054. [Google Scholar] [CrossRef]
  185. Li, X.; Wan, J.; Tang, Y.; Wang, C.; Zhang, Y.; Lv, D.; Guo, M.; Ma, Y.; Yang, Y. Boosting the UV–vis–NIR Photodetection Performance of MoS2 through the Cavity Enhancement Effect and Bulk Heterojunction Strategy. ACS Appl. Mater. Interfaces 2024, 16, 29003–29015. [Google Scholar] [CrossRef] [PubMed]
  186. Varshney, U.; Sharma, A.; Vashishtha, P.; Singh, P.; Gupta, G. Highly responsive self-driven broadband photodetector based on MoS2 nanorods/β-Ga2O3 heterojunction. Mater. Sci. Semicond. Process. 2023, 164, 107612. [Google Scholar] [CrossRef]
  187. Hafeez, M.; Gan, L.; Li, H.; Ma, Y.; Zhai, T. Large-Area Bilayer ReS2 Film/Multilayer ReS2 Flakes Synthesized by Chemical Vapor Deposition for High Performance Photodetectors. Adv. Funct. Mater. 2016, 26, 4551–4560. [Google Scholar] [CrossRef]
  188. An, Q.; Liu, Y.; Jiang, R.; Meng, X. Chemical vapor deposition growth of ReS2 nanowires for high-performance nanostructured photodetector. Nanoscale 2018, 10, 14976–14983. [Google Scholar] [CrossRef]
  189. Butanovs, E.; Butikova, J.; Zolotarjovs, A.; Polyakov, B. Towards metal chalcogenide nanowire-based colour-sensitive photodetectors. Opt. Mater. 2018, 75, 501–507. [Google Scholar] [CrossRef]
  190. Xie, X.; Shen, G. Single-crystalline In2S3nanowire-based flexible visible-light photodetectors with an ultra-high photoresponse. Nanoscale 2015, 7, 5046–5052. [Google Scholar] [CrossRef]
  191. Wang, T.; Jiang, T.; Meng, X. PVP-assisted synthesis of Sb2S3 nanowire for self-driven photodetector by using MXenes and Ag nanowires as electrodes. Appl. Nanosci. 2020, 10, 1845–1851. [Google Scholar] [CrossRef]
  192. Bera, A.; Das Mahapatra, A.; Mondal, S.; Basak, D. Sb2S3/Spiro-OMeTAD Inorganic–Organic Hybrid p–n Junction Diode for High Performance Self-Powered Photodetector. ACS Appl. Mater. Interfaces 2016, 8, 34506–34512. [Google Scholar] [CrossRef] [PubMed]
  193. Ismail, R.A.; Abeduljabbar, N.F.; Fatehi, M.W. Effect of dipping time on the properties of Sb2S3/Si heterojunction prepared by chemical bath deposition. Mater. Res. Express 2019, 6, 045915. [Google Scholar] [CrossRef]
  194. Wen, S.; Zhao, J.; Zhao, Y.; Xu, T.; Xu, J. Reduced graphene oxide (RGO) decorated Sb2S3 nanorods as anode material for sodium-ion batteries. Chem. Phys. Lett. 2019, 716, 171–176. [Google Scholar] [CrossRef]
  195. Kondrotas, R.; Chen, C.; Tang, J. Sb2S3 Solar Cells. Joule 2018, 2, 857–878. [Google Scholar] [CrossRef]
  196. Yu, Y.; Yan, W.; Wang, X.; Li, P.; Gao, W.; Zou, H.; Wu, S.; Ding, K. Surface Engineering for Extremely Enhanced Charge Separation and Photocatalytic Hydrogen Evolution on g-C3N4. Adv. Mater. 2018, 30, 1705060. [Google Scholar] [CrossRef]
  197. Nishikubo, R.; Li, S.; Saeki, A. Unprecedented Wavelength Dependence of an Antimony Chalcohalide Photovoltaic Device. Adv. Funct. Mater. 2022, 32, 2201577. [Google Scholar] [CrossRef]
  198. Zhong, M.; Wang, X.; Liu, S.; Li, B.; Huang, L.; Cui, Y.; Li, J.; Wei, Z. High-performance photodetectors based on Sb2S3 nanowires: Wavelength dependence and wide temperature range utilization. Nanoscale 2017, 9, 12364–12371. [Google Scholar] [CrossRef]
  199. Zhao, K.; Yang, J.; Zhong, M.; Gao, Q.; Wang, Y.; Wang, X.; Shen, W.; Hu, C.; Wang, K.; Shen, G.; et al. Direct Polarimetric Image Sensor and Wide Spectral Response Based on Quasi-1D Sb2S3 Nanowire. Adv. Funct. Mater. 2020, 31, 2006601. [Google Scholar] [CrossRef]
  200. Ye, K.; Wang, B.; Nie, A.; Zhai, K.; Wen, F.; Mu, C.; Zhao, Z.; Xiang, J.; Tian, Y.; Liu, Z. Broadband photodetector of high quality Sb2S3 nanowire grown by chemical vapor deposition. J. Mater. Sci. Technol. 2021, 75, 14–20. [Google Scholar] [CrossRef]
  201. Ma, Y.; Ma, C.; Yi, H.; Liang, H.; Wang, W.; Zheng, Z.; Zou, Y.; Deng, Z.; Yao, J.; Yang, G. 1D Sb2S3 with Strong Mie Resonance Toward Highly-Sensitive Polarization-Discriminating Photodetection and Its Application in High-Temperature-Proof Imaging and Dual-Channel Communications. Adv. Opt. Mater. 2023, 12, 2302039. [Google Scholar] [CrossRef]
  202. Liu, X.; Dong, S.; Zheng, X.; Zhang, Y.; Yao, Y.; Zhang, W.; Liu, Z.; Zhu, T.; Wang, H.E. Improving the optoelectronic properties of single-crystalline antimony sulfide rods through simultaneous defect suppression and surface cleaning. J. Mater. Chem. A 2023, 11, 8826–8835. [Google Scholar] [CrossRef]
  203. Zhang, J.; Hou, K.; Huang, T.; Yang, J.; Xu, B.; Li, S.; Yang, B.; Ma, W. Single-Crystal Sb2S3 Tube Prepared by Chemical Vapor Deposition for a 1 cm Photodetector Application. ACS Appl. Electron. Mater. 2022, 4, 2455–2462. [Google Scholar] [CrossRef]
  204. Chao, J.; Xing, S.; Zhang, J.; Qin, C.; Duan, D.; Wu, X.; Shen, Q. Synthesis of Sb2S3 nanowall arrays for high performance visible light photodetectors. Mater. Res. Bull. 2014, 57, 300–305. [Google Scholar] [CrossRef]
  205. Zhang, K.; Luo, T.; Chen, H.; Lou, Z.; Shen, G. Au-nanoparticles-decorated Sb2S3 nanowire-based flexible ultraviolet/visible photodetectors. J. Mater. Chem. C 2017, 5, 3330–3335. [Google Scholar] [CrossRef]
  206. Wang, J.; Qiao, Y.; Wang, T.; Yu, H.; Feng, Y.; Zhang, J. Isovalent bismuth ion-induced growth of highly-disperse Sb2S3 nanorods and their composite with p-CuSCN for self-powered photodetectors. CrystEngComm 2019, 21, 554–562. [Google Scholar] [CrossRef]
  207. Jiang, T.; Meng, X. Surface and Interface Engineering Enhanced Photodetector Based on Mo2C-C/Sb2S3 Composites. Nano 2021, 16, 2150004. [Google Scholar] [CrossRef]
  208. Shkir, M.; Alshahrani, T. Impact of Nd doping in Bi2S3 thin films coated by nebulizer spray pyrolysis technique for photodetector applications. Opt. Mater. 2023, 140, 113837. [Google Scholar] [CrossRef]
  209. Panwar, V.; Dey, M.; Sharma, P.; Sundar, K.; Nandi, S.; Tripathi, R.; Mondal, A.; Makineni, S.K.; Shukla, A.; Singh, A.; et al. Ultrahigh Photo-Responsivity and Detectivity in 2D Bismuth Sulfide Photodetector for Vis–NIR Radiation. Small 2024, 20, 2309428. [Google Scholar] [CrossRef]
  210. Rong, P.; Gao, S.; Ren, S.; Lu, H.; Yan, J.; Li, L.; Zhang, M.; Han, Y.; Jiao, S.; Wang, J. Large-Area Freestanding Bi2S3 Nanofibrous Membranes for Fast Photoresponse Flexible IR Imaging Photodetector. Adv. Funct. Mater. 2023, 33, 2300159. [Google Scholar] [CrossRef]
  211. Chitara, B.; Kolli, B.S.C.; Yan, F. Near-Infrared photodetectors based on 2D Bi2S3. Chem. Phys. Lett. 2022, 804, 139876. [Google Scholar] [CrossRef]
  212. Liu, Y.; Chen, P.; Dai, G.; Su, W.; Sun, Y.; Hou, J.; Zhang, N.; Zhao, G.; Fang, Y.; Dai, N. Single Bi2S3/Bi2S3-xOx nanowire photodetector with broadband response from ultraviolet to near-infrared range. Phys. E Low-Dimens. Syst. Nanostruct. 2020, 120, 114041. [Google Scholar] [CrossRef]
  213. Ding, T.; Tian, Y.; Dai, J.; Chen, C. Building one-dimensional Bi2S3 nanorods as enhanced photoresponding materials for photodetectors. Front. Optoelectron. 2015, 8, 282–288. [Google Scholar] [CrossRef]
  214. Yu, H.; Wang, J.; Wang, T.; Yu, H.; Yang, J.; Liu, G.; Qiao, G.; Yang, Q.; Cheng, X. Scalable colloidal synthesis of uniform Bi2S3nanorods as sensitive materials for visible-light photodetectors. CrystEngComm 2017, 19, 727–733. [Google Scholar] [CrossRef]
  215. Li, X.; Wu, Y.; Ying, H.; Xu, M.; Jin, C.; He, Z.; Zhang, Q.; Su, W.; Zhao, S. In situ physical examination of Bi2S3 nanowires with a microscope. J. Alloys Compd. 2019, 798, 628–634. [Google Scholar] [CrossRef]
  216. Wang, H.; Liu, R.; Zhang, S.; Wang, Y.; Luo, H.; Sun, X.; Ren, Y.; Lei, W. Controlled growth of high-quality Bi2S3 nanowires and their application in near-infrared photodetection. Opt. Mater. 2022, 134, 113174. [Google Scholar] [CrossRef]
  217. Yi, H.; Ma, C.; Wang, W.; Liang, H.; Cui, R.; Cao, W.; Yang, H.; Ma, Y.; Huang, W.; Zheng, Z.; et al. Quantum tailoring for polarization-discriminating Bi2S3 nanowire photodetectors and their multiplexing optical communication and imaging applications. Mater. Horiz. 2023, 10, 3369–3381. [Google Scholar] [CrossRef]
  218. Xu, X.; Fan, C.; Wang, Y.; Qi, Z.; Dai, B.; Jiang, H.; Wang, S.; Zhang, Q. In-plane Epitaxy of Bi2S3 Nanowire Arrays for Ultrasensitive NIR Photodetectors. Phys. Status Solidi (RRL) Rapid Res. Lett. 2020, 14, 2000384. [Google Scholar] [CrossRef]
  219. Maria, C.C.S.; Patil, R.A.; Hasibuan, D.P.; Saragih, C.S.; Lai, C.C.; Liou, Y.; Ma, Y.R. White-light photodetection enhancement and thin film impediment in Bi2S3 nanorods/thin-films homojunction photodetectors. Appl. Surf. Sci. 2022, 584, 152608. [Google Scholar] [CrossRef]
  220. Ghosh, K.; Giri, P.K. Experimental and theoretical study on the role of 2D Ti3C2Tx MXenes on superior charge transport and ultra-broadband photodetection in MXene/Bi2S3 nanorod composite through local Schottky junctions. Carbon 2024, 216, 118515. [Google Scholar] [CrossRef]
  221. Singh, A.; Chauhan, P.; Verma, A.; Yadav, B.C. Interfacial engineering enables polyaniline-decorated bismuth sulfide nanorods towards ultrafast metal–semiconductor-metal UV-Vis broad spectra photodetector. Adv. Compos. Hybrid Mater. 2024, 7, 88. [Google Scholar] [CrossRef]
  222. Devasia, S.; Shaji, S.; Avellaneda, D.A.; Aguilar Martinez, J.A.; Krishnan, B. In situ grown Bi2S3 nanorods in Cs3Bi2I9 thin films as broadband self-driven photodetector with improved photostability. Opt. Mater. 2024, 152, 115532. [Google Scholar] [CrossRef]
  223. Randle, M.; Lipatov, A.; Kumar, A.; Kwan, C.P.; Nathawat, J.; Barut, B.; Yin, S.; He, K.; Arabchigavkani, N.; Dixit, R.; et al. Gate-Controlled Metal–Insulator Transition in TiS3 Nanowire Field-Effect Transistors. ACS Nano 2018, 13, 803–811. [Google Scholar] [CrossRef]
  224. Liu, S.; Xiao, W.; Zhong, M.; Pan, L.; Wang, X.; Deng, H.X.; Liu, J.; Li, J.; Wei, Z. Highly polarization sensitive photodetectors based on quasi-1D titanium trisulfide (TiS3). Nanotechnology 2018, 29, 184002. [Google Scholar] [CrossRef] [PubMed]
  225. Ding, Y.; Xu, X.; Zhuang, Z.; Sang, Y.; Cui, M.; Li, W.; Yan, Y.; Tao, T.; Xu, W.; Ren, F.; et al. Self-powered MXene/GaN van der Waals Schottky ultraviolet photodetectors with exceptional responsivity and stability. Appl. Phys. Rev. 2024, 14, 041413. [Google Scholar] [CrossRef]
  226. Zhang, Y.; Chen, J.; Zhu, L.; Wang, Z.L. Self-Powered High-Responsivity Photodetectors Enhanced by the Pyro-Phototronic Effect Based on a BaTiO3/GaN Heterojunction. Nano Lett. 2021, 21, 8808–8816. [Google Scholar] [CrossRef]
  227. Gao, Z.; Chen, L.; Li, S.; Pan, X.; Liu, Y.; Gao, G.; Zou, B. Type-II CsPbBr3/ZnS Heterostructured Nanocrystals for High-Performance Self-Powered Photodetector. ACS Appl. Nano Mater. 2025, 8, 5623–5631. [Google Scholar] [CrossRef]
  228. Li, Z.; Li, H.; Jiang, K.; Ding, D.; Li, J.; Ma, C.; Jiang, S.; Wang, Y.; Anthopoulos, T.D.; Shi, Y. Self-Powered Perovskite/CdS Heterostructure Photodetectors. ACS Appl. Mater. Interfaces 2019, 11, 40204–40213. [Google Scholar] [CrossRef]
  229. Zhou, H.; Gui, P.; Yang, L.; Ye, C.; Xue, M.; Mei, J.; Song, Z.; Wang, H. High performance, self-powered ultraviolet photodetector based on a ZnO nanoarrays/GaN structure with a CdS insert layer. New J. Chem. 2017, 41, 4901–4907. [Google Scholar] [CrossRef]
  230. Xie, J.; Han, J.; Jiang, J.; Liu, L.; Fu, Z.; Wang, J. Self-Powered, Broadband, and Polarization-Sensitive Photodetector with Nb-WS2/Ta2NiSe5 Van der Waals Heterostructure. Adv. Opt. Mater. 2025, 13, 2403463. [Google Scholar] [CrossRef]
  231. Kumar, A.; Chakkar, A.G.; Das, C.; Kumar, P.; Sahu, S.; Saliba, M.; Kumar, M. Self-Powered Broadband Photodetectors Based on WS2-Anchored MoS2 with Enhanced Responsivity and Detectivity. Small 2025, 21, e2502900. [Google Scholar] [CrossRef]
  232. Zhou, C.; Liu, Q.; Fu, X.; Niu, W.; Wang, J.; Liu, X.; Zou, X.; Miao, J.; Shan, F.; Yang, Z. Self-Powered and Reconfigurable Double-Terminal MoS2 Photodetector for Image Recognition. Nano Lett. 2025, 25, 3515–3523. [Google Scholar] [CrossRef]
  233. Shen, M.; Wang, Q.; Liu, H.; Ye, H.; Yuan, X.; Zhang, Y.; Wei, B.; He, X.; Liu, K.; Cai, S.; et al. Self-Powered Polarization-Sensitive 3D Photodetector with Fast Response Enabled by Mechanically-Guided Assembly of Tubular MoS2/GaAs/InGaAs Heterostructure. Adv. Opt. Mater. 2025, 13, 2402848. [Google Scholar] [CrossRef]
  234. Deng, S.; Chen, Y.; Li, Q.; Sun, J.; Lei, Z.; Hu, P.; Liu, Z.-H.; He, X.; Ma, R. Direct growth of SnS2 nanowall photoanode for high responsivity self-powered photodetectors. Nanoscale 2022, 14, 14097–14105. [Google Scholar] [CrossRef]
  235. Kang, Y.; Deng, H.; Wu, K.; Wang, G.; Hong, J.; Wang, W.; Wu, J.; Zhou, H.; Xia, Y.; Cheng, S. Self-powered Sb2S3 photodetectors with efficient carrier transport enabled by compact vertical TiO2 nanorods. Opt. Lett. 2024, 49, 6533–6536. [Google Scholar] [CrossRef]
  236. Zamani, M.; Jamali-Sheini, F.; Cheraghizade, M. Visible-range and self-powered bilayer p-Si/n-Bi2S3 heterojunction photodetector: The effect of Au buffer layer on the optoelectronics performance. J. Alloys Compd. 2022, 905, 164119. [Google Scholar] [CrossRef]
  237. Yadav, S.M.; Pandey, A. 2D-SnS2 Nanoflakes Based Efficient Ultraviolet Photodetector. IEEE Trans. Nanotechnol. 2020, 19, 301–307. [Google Scholar] [CrossRef]
  238. Jin, Z.; He, D.; Zhou, Q.; Mao, P.; Ding, L.; Wang, J. Bilayer Heterostructured PThTPTI/WS2 Photodetectors with High Thermal Stability in Ambient Environment. ACS Appl. Mater. Interfaces 2016, 8, 33043–33050. [Google Scholar] [CrossRef]
  239. Vashishtha, P.; Prajapat, P.; Sharma, A.; Goswami, P.; Walia, S.; Gupta, G. Self-driven and thermally resilient highly responsive nano-fenced MoS2 based photodetector for near-infrared optical signal. Mater. Res. Bull. 2023, 164, 112260. [Google Scholar] [CrossRef]
  240. Wahalathantrige Don, R.; Das, P.; Ma, Z.; Kuruppu, U.M.; Feng, D.; Shook, B.; Gangishetty, M.K.; Stefan, M.C.; Pradhan, N.R.; Scott, C.N. Vinyl-Flanked Diketopyrrolopyrrole Polymer/MoS2 Hybrid for Donor–Acceptor Semiconductor Photodetection. Chem. Mater. 2023, 35, 4691–4704. [Google Scholar] [CrossRef]
  241. Li, Z.; Luo, J.; Hu, S.; Liu, Q.; Yu, W.; Lu, Y.; Liu, X. Strain enhancement for a MoS2-on-GaN photodetector with an Al2O3 stress liner grown by atomic layer deposition. Photonics Res. 2020, 8, 799–805. [Google Scholar] [CrossRef]
  242. Kaur, D.; Dahiya, R.; Sheokand, V.; Bassi, G.; Kumar, M. Ultrafast, self-powered and highly-stable PtS-Ga2O3 heterojunction photodetector for broad-spectrum sensing. Surf. Interfaces 2025, 61, 106125. [Google Scholar] [CrossRef]
  243. Xiong, L.; Zhang, L.; Lv, Q.; Li, T.; Song, W.; Si, J.; Zhu, W.; Wang, L. Amorphous gallium oxide (a-Ga2O3)-based high-temperature bendable solar-blind ultraviolet photodetector. Semicond. Sci. Technol. 2021, 36, 045010. [Google Scholar] [CrossRef]
  244. Zhang, M.; Liu, Z.; Yang, L.; Yao, J.; Chen, J.; Zhang, J.; Wei, W.; Guo, Y.; Tang, W. High-temperature reliability of all-oxide self-powered deep UV photodetector based on ϵ-Ga2O3/ZnO heterojunction. J. Phys. D Appl. Phys. 2022, 55, 375106. [Google Scholar] [CrossRef]
  245. Polyakov, A.Y.; Yun, J.H.; Usikov, A.S.; Yakimov, E.B.; Smirnov, N.B.; Shcherbachev, K.D.; Helava, H.; Makarov, Y.N.; Kurin, S.Y.; Shmidt, N.M.; et al. Structural, electrical and luminescent characteristics of ultraviolet light emitting structures grown by hydride vapor phase epitaxy. Mod. Electron. Mater. 2017, 3, 32–39. [Google Scholar] [CrossRef]
  246. Loser, S.; Valle, B.; Luck, K.A.; Song, C.K.; Ogien, G.; Hersam, M.C.; Singer, K.D.; Marks, T.J. High-Efficiency Inverted Polymer Photovoltaics via Spectrally Tuned Absorption Enhancement. Adv. Energy Mater. 2014, 4, 1301938. [Google Scholar] [CrossRef]
  247. Wang, X.F.; Zhao, H.M.; Shen, S.H.; Pang, Y.; Shao, P.-Z.; Li, Y.-T.; Deng, N.-Q.; Li, Y.-X.; Yang, Y.; Ren, T.-L. High performance photodetector based on Pd-single layer MoS2 Schottky junction. Appl. Phys. Lett. 2016, 109, 201904. [Google Scholar] [CrossRef]
  248. Yan, F.; Yang, X.; Liu, J.; Tang, H.; Tan, Y.A.; Li, Y. Optimizing the restoration performance of deduplication systems through an energy-saving data layout. Ann. Telecommun. 2019, 74, 461–471. [Google Scholar] [CrossRef]
  249. Ullah, S.; Bouich, A.; Ullah, H.; Mari, B.; Mollar, M. Comparative study of binary cadmium sulfide (CdS) and tin disulfide (SnS2) thin buffer layers. Sol. Energy 2020, 208, 637–642. [Google Scholar] [CrossRef]
  250. Wang, H.; Liu, Y.; Liu, H.; Chen, Z.; Xiong, P.; Xu, X.; Chen, F.; Li, K.; Duan, Y. Effect of various oxidants on reaction mechanisms, self-limiting natures and structural characteristics of Al2O3 films grown by atomic layer deposition. Adv. Mater. Interfaces 2018, 5, 1701248. [Google Scholar] [CrossRef]
  251. Zhang, Q.; Liu, H.; Guo, P.; Li, D.; Fan, P.; Zheng, W.; Zhu, X.; Jiang, Y.; Zhou, H.; Hu, W.; et al. Vapor growth and interfacial carrier dynamics of high-quality CdS-CdSSe-CdS axial nanowire heterostructures. Nano Energy 2017, 32, 28–35. [Google Scholar] [CrossRef]
  252. Kim, M.J.; Alvarez, S.; Chen, Z.; Fichthorn, K.A.; Wiley, B.J. Single-crystal electrochemistry reveals why metal nanowires grow. J. Am. Chem. Soc. 2018, 140, 14740–14746. [Google Scholar] [CrossRef] [PubMed]
  253. Zhang, T.; Liu, X.; Qu, G.; Lu, P.; Wang, J.; Wu, F.; Ren, Y. Secondary resource utilization of metallurgical solid waste: Current status and future prospects of wet extraction of valuable metals. Sep. Purif. Technol. 2024, 361, 131278. [Google Scholar] [CrossRef]
  254. Aung, T.; Giridhar, S.P.; Abidi, I.H.; Ahmed, T.; AI-Hourani, A.; Walia, S. Photoactive Monolayer MoS2 for Spiking Neural Networks Enabled Machine Vision Applications. Adv. Mater. Technol. 2025, 2401677. [Google Scholar] [CrossRef]
  255. Gogoi, K.; Chattopadhyay, A. Surface engineering of quantum dots for self-powered ultraviolet photodetection and information encryption. Langmuir 2022, 38, 2668–2676. [Google Scholar] [CrossRef] [PubMed]
  256. Li, Y.; Li, L.; Li, S.; Sun, J.; Fang, Y.; Deng, T. Highly sensitive photodetectors based on monolayer MoS2 field-effect transistors. ACS Omega 2022, 7, 13615–13621. [Google Scholar] [CrossRef] [PubMed]
  257. Li, J.; Li, Z.; Liu, X.; Li, C.; Zheng, Y.; Yeung, K.W.K.; Cui, Z.; Liang, Y.; Zhu, S.; Hu, W.; et al. Interfacial engineering of Bi2S3/Ti3C2Tx MXene based on work function for rapid photo-excited bacteria-killing. Nat. Commun. 2021, 12, 1224. [Google Scholar] [CrossRef]
Figure 1. (a) The I-V characteristics of a Bi2S3-based photodetector with Au contacts. Reprinted with permission from Ref. [70]. Copyright 2020, The Royal Society of Chemistry. (b) The I-V characteristics of a MoNx-based photodetector with Cr contacts. Reprinted with permission from Ref. [71]. Copyright 2021, American Chemical Society.
Figure 1. (a) The I-V characteristics of a Bi2S3-based photodetector with Au contacts. Reprinted with permission from Ref. [70]. Copyright 2020, The Royal Society of Chemistry. (b) The I-V characteristics of a MoNx-based photodetector with Cr contacts. Reprinted with permission from Ref. [71]. Copyright 2021, American Chemical Society.
Jcs 09 00262 g001
Figure 2. Schematic structures of ZnSe-based MSM UV photodetectors with conventional (a), interdigitated (b), and hybrid (c) contacts. Reprinted with permission from Ref. [78]. Copyright 2024, Pleiades Publishing.
Figure 2. Schematic structures of ZnSe-based MSM UV photodetectors with conventional (a), interdigitated (b), and hybrid (c) contacts. Reprinted with permission from Ref. [78]. Copyright 2024, Pleiades Publishing.
Jcs 09 00262 g002
Figure 3. A diagram of the formation of the built-in electric field at the p-n heterojunction. Reprinted with permission from Ref. [82]. Copyright 2022, Elsevier.
Figure 3. A diagram of the formation of the built-in electric field at the p-n heterojunction. Reprinted with permission from Ref. [82]. Copyright 2022, Elsevier.
Jcs 09 00262 g003
Figure 4. (a) Mechanism of charge separation under illumination, (b) energy band diagram of n-CdS nanorods/p-Si heterojunction. Reprinted with permission from Ref. [86]. Copyright 2023, Elsevier.
Figure 4. (a) Mechanism of charge separation under illumination, (b) energy band diagram of n-CdS nanorods/p-Si heterojunction. Reprinted with permission from Ref. [86]. Copyright 2023, Elsevier.
Jcs 09 00262 g004
Figure 5. A ZnS NWs-based photodetector: (a) The I-V curves in the dark and under UV irradiation; the inset is an SEM image of the device. (b) The multiple switching of light and dark current. Reprinted with permission from Ref. [130]. Copyright 2016, Elsevier. ZnS NTs-based devices: (c) I-V curves; (d) the natural logarithmic plot of the time response spectrum. Reprinted with permission from Ref. [131]. Copyright 2017, IOP Publishing. (e) The I-T curve for on/off switching; (f) a single on/off cycle. Reprinted with permission from Ref. [132]. Copyright 2016, Springer Nature.
Figure 5. A ZnS NWs-based photodetector: (a) The I-V curves in the dark and under UV irradiation; the inset is an SEM image of the device. (b) The multiple switching of light and dark current. Reprinted with permission from Ref. [130]. Copyright 2016, Elsevier. ZnS NTs-based devices: (c) I-V curves; (d) the natural logarithmic plot of the time response spectrum. Reprinted with permission from Ref. [131]. Copyright 2017, IOP Publishing. (e) The I-T curve for on/off switching; (f) a single on/off cycle. Reprinted with permission from Ref. [132]. Copyright 2016, Springer Nature.
Jcs 09 00262 g005
Figure 6. Device schematic: (a) ZnS NWs-based; (b) ZnS NWA-based; (c) I-V curves of (a) in dark and under 325 nm and 442 nm illumination; (d) I-V curves of (b); (e) time response at 325 nm for device (a) with 60 s switching period 60 s and 100 μW cm−2 power; (f) time response at 325 nm for device (b) with 3.9 mW cm−2 power density and 10 Hz pulse. Reprinted with permission from Ref. [133]. Copyright 2012, The Royal Society of Chemistry.
Figure 6. Device schematic: (a) ZnS NWs-based; (b) ZnS NWA-based; (c) I-V curves of (a) in dark and under 325 nm and 442 nm illumination; (d) I-V curves of (b); (e) time response at 325 nm for device (a) with 60 s switching period 60 s and 100 μW cm−2 power; (f) time response at 325 nm for device (b) with 3.9 mW cm−2 power density and 10 Hz pulse. Reprinted with permission from Ref. [133]. Copyright 2012, The Royal Society of Chemistry.
Jcs 09 00262 g006
Figure 7. ZnS:Mn NRs-based PD: (a) I-V characteristic curve, (b) light response under fast-switching 310 nm illumination. Reprinted with permission from Ref. [134]. Copyright 2021, American Chemical Society. ZnS:Al NWs-based PD: (c) I-V curve at 100 mW/cm2 power, (d) I-V curves at 332 nm. Reprinted with permission from Ref. [135]. Copyright 2012, The Royal Society of Chemistry. ZnS/InP NWs-based PD: (e) Spectral photoresponse at 5 V bias; inset: spectral photoresponse of bare ZnS NWs-based PDs under same conditions. (f) I-T curves at different bias pressures. Reprinted with permission from Ref. [136]. Copyright 2017, The Royal Society of Chemistry.
Figure 7. ZnS:Mn NRs-based PD: (a) I-V characteristic curve, (b) light response under fast-switching 310 nm illumination. Reprinted with permission from Ref. [134]. Copyright 2021, American Chemical Society. ZnS:Al NWs-based PD: (c) I-V curve at 100 mW/cm2 power, (d) I-V curves at 332 nm. Reprinted with permission from Ref. [135]. Copyright 2012, The Royal Society of Chemistry. ZnS/InP NWs-based PD: (e) Spectral photoresponse at 5 V bias; inset: spectral photoresponse of bare ZnS NWs-based PDs under same conditions. (f) I-T curves at different bias pressures. Reprinted with permission from Ref. [136]. Copyright 2017, The Royal Society of Chemistry.
Jcs 09 00262 g007
Figure 8. CdS NWs-based PD: (a) Schematic, (b) I-V test results at different power intensities under 470 nm light irradiation and dark conditions, (c) light response characteristics at bias voltage of 2 V and intensity of 0.934 mW/cm2, (d) enlarged view of one cycle in (c). Reprinted with permission from Ref. [147]. Copyright 2015, American Chemical Society.
Figure 8. CdS NWs-based PD: (a) Schematic, (b) I-V test results at different power intensities under 470 nm light irradiation and dark conditions, (c) light response characteristics at bias voltage of 2 V and intensity of 0.934 mW/cm2, (d) enlarged view of one cycle in (c). Reprinted with permission from Ref. [147]. Copyright 2015, American Chemical Society.
Jcs 09 00262 g008
Figure 9. Ferroelectric side-gating single CdS NW-based PD: (a) Schematic, (b) I-V characteristics before and after depletion by ferroelectric polymers in dark and under UV irradiation (375 nm, 18 mW cm−2), (c) time-resolved optical response of device at 1 V. Reprinted with permission from Ref. [148]. Copyright 2016, Wiley-VCH Verlag. (d) Responsiveness of CdS NRs-based PDs in unconditioned conditions. Reprinted with permission from Ref. [149]. Copyright 2017, John Wiley and Sons.
Figure 9. Ferroelectric side-gating single CdS NW-based PD: (a) Schematic, (b) I-V characteristics before and after depletion by ferroelectric polymers in dark and under UV irradiation (375 nm, 18 mW cm−2), (c) time-resolved optical response of device at 1 V. Reprinted with permission from Ref. [148]. Copyright 2016, Wiley-VCH Verlag. (d) Responsiveness of CdS NRs-based PDs in unconditioned conditions. Reprinted with permission from Ref. [149]. Copyright 2017, John Wiley and Sons.
Jcs 09 00262 g009
Figure 10. Sb/CdS NWs-based PD: (a) Structure schematic, (b) I-V characteristics under different intensities of light at 470 nm, (c) optical response characteristics at 470 nm bias 0 V, (d) single optical on/off cycle transient response. Reprinted with permission from Ref. [150]. Copyright 2019, The Royal Society of Chemistry.
Figure 10. Sb/CdS NWs-based PD: (a) Structure schematic, (b) I-V characteristics under different intensities of light at 470 nm, (c) optical response characteristics at 470 nm bias 0 V, (d) single optical on/off cycle transient response. Reprinted with permission from Ref. [150]. Copyright 2019, The Royal Society of Chemistry.
Jcs 09 00262 g010
Figure 11. SnS/CdS heterojunction PD: (a) Structural schematic, (b) three-dimensional mapping images of temperature, optical power intensity, and pyroelectric current, (c) Ipyro/Iphoto at different optical power densities and temperatures. Reprinted with permission from Ref. [151]. Copyright 2020, Wiley Online Library.
Figure 11. SnS/CdS heterojunction PD: (a) Structural schematic, (b) three-dimensional mapping images of temperature, optical power intensity, and pyroelectric current, (c) Ipyro/Iphoto at different optical power densities and temperatures. Reprinted with permission from Ref. [151]. Copyright 2020, Wiley Online Library.
Jcs 09 00262 g011
Figure 12. PANI/CdS NWA-based PD: (a) Nanowire array device, (b) typical I-V curves in dark and under illumination of different wavelength light, (c) on/off switching with 420 nm wavelength light and power density of 5.21 mW/cm2. Reprinted with permission from Ref. [153]. Copyright 2011, American Chemical Society. CdS/PPV NWA-based PD: (d) Typical I-V characteristics in dark and under 545 nm illumination; inset is schematic of device. (e) On/off toggles with 545 nm light pulse at power density of 4.2 mW/cm2. (f) Typical I-V curves for different light intensities at 545 nm illumination. Reprinted with permission from Ref. [155]. Copyright 2022, MDPI AG.
Figure 12. PANI/CdS NWA-based PD: (a) Nanowire array device, (b) typical I-V curves in dark and under illumination of different wavelength light, (c) on/off switching with 420 nm wavelength light and power density of 5.21 mW/cm2. Reprinted with permission from Ref. [153]. Copyright 2011, American Chemical Society. CdS/PPV NWA-based PD: (d) Typical I-V characteristics in dark and under 545 nm illumination; inset is schematic of device. (e) On/off toggles with 545 nm light pulse at power density of 4.2 mW/cm2. (f) Typical I-V curves for different light intensities at 545 nm illumination. Reprinted with permission from Ref. [155]. Copyright 2022, MDPI AG.
Jcs 09 00262 g012
Figure 13. TiO2/CdS NRs-based PD: (a) Photocurrent and dark current I-V characteristics, (b) sample responsiveness and EQE, (c) detection rate and NEP, (d) rise and recovery time. Reprinted with permission from Ref. [157]. Copyright 2024, Elsevier.
Figure 13. TiO2/CdS NRs-based PD: (a) Photocurrent and dark current I-V characteristics, (b) sample responsiveness and EQE, (c) detection rate and NEP, (d) rise and recovery time. Reprinted with permission from Ref. [157]. Copyright 2024, Elsevier.
Jcs 09 00262 g013
Figure 15. WS2 NTs-based PD: (a) I-V curves measured in darkness, red light (633 nm, light intensity of 532 W/cm2) and blue light (blue light 785 nm, light intensity of 600 W/cm2); (b) magnified view of current–time (I-T) curve; (c) single-cycle I-T curve in (b). Reprinted with permission from Ref. [54]. Copyright 2012, American Institute of Physics. (d) I-V curves of photodetector illuminated by light with different wavelengths and in dark for WS2 (FLs)-PVK; (e,f) photoresponse of WS2 (NTs)-PVK films to light pulses (λ = 661,552 nm) at 0.2 and 0.8 mW power. Reprinted with permission from Ref. [182]. Copyright 2022, American Chemical Society.
Figure 15. WS2 NTs-based PD: (a) I-V curves measured in darkness, red light (633 nm, light intensity of 532 W/cm2) and blue light (blue light 785 nm, light intensity of 600 W/cm2); (b) magnified view of current–time (I-T) curve; (c) single-cycle I-T curve in (b). Reprinted with permission from Ref. [54]. Copyright 2012, American Institute of Physics. (d) I-V curves of photodetector illuminated by light with different wavelengths and in dark for WS2 (FLs)-PVK; (e,f) photoresponse of WS2 (NTs)-PVK films to light pulses (λ = 661,552 nm) at 0.2 and 0.8 mW power. Reprinted with permission from Ref. [182]. Copyright 2022, American Chemical Society.
Jcs 09 00262 g015
Figure 16. MoS2/WS2-NRs: (a) light-sensing detector; (b) photocurrent at 635, 785, and 1064 nm laser sources; (c) EQE; (d) D*, measured at wavelengths (λex) = 635, 785, and 1064 nm. Reprinted with permission from Ref. [183]. Copyright 2019, IOPscience.
Figure 16. MoS2/WS2-NRs: (a) light-sensing detector; (b) photocurrent at 635, 785, and 1064 nm laser sources; (c) EQE; (d) D*, measured at wavelengths (λex) = 635, 785, and 1064 nm. Reprinted with permission from Ref. [183]. Copyright 2019, IOPscience.
Jcs 09 00262 g016
Figure 17. (a) Schematic diagram of light shining onto nanotube; (b) typical I-V curves at different wavelengths in darkness and 915 nm illumination; (c) characterization of photoelectric switches with different light intensities (biased at 3 V) in darkness and 915 nm light; (d) high resolution time response; (e) R and D* function plots; (f) plot of NEP and EQE functions. Reprinted with permission from Ref. [184]. Copyright 2023, American Chemical Society.
Figure 17. (a) Schematic diagram of light shining onto nanotube; (b) typical I-V curves at different wavelengths in darkness and 915 nm illumination; (c) characterization of photoelectric switches with different light intensities (biased at 3 V) in darkness and 915 nm light; (d) high resolution time response; (e) R and D* function plots; (f) plot of NEP and EQE functions. Reprinted with permission from Ref. [184]. Copyright 2023, American Chemical Society.
Jcs 09 00262 g017
Figure 18. NRs-MoS2/Ga2O3-based PD: (a) Schematic under UVC-Vis-NIR illumination, (b) bias-correlated plot of developed device under 266 nm and (c) 950 nm light illumination, (d) effect of optical power on device under applied 5 V bias R, (e) D*, (f) EQE. Reprinted with permission from Ref. [186]. Copyright 2023, Elsevier.
Figure 18. NRs-MoS2/Ga2O3-based PD: (a) Schematic under UVC-Vis-NIR illumination, (b) bias-correlated plot of developed device under 266 nm and (c) 950 nm light illumination, (d) effect of optical power on device under applied 5 V bias R, (e) D*, (f) EQE. Reprinted with permission from Ref. [186]. Copyright 2023, Elsevier.
Jcs 09 00262 g018
Figure 19. In2S3 nanowires: (a) the 3D simulation of a flexible PD on a PET substrate, (b) the determined I-V curves in the dark and in the light, (c) the R and EQE with incident wavelength, (d) a photocurrent versus time plot of the device under 450 nm, 177 μ W/cm2 light irradiation. Reprinted with permission from Ref. [190]. Copyright 2025, Royal Society of Chemistry.
Figure 19. In2S3 nanowires: (a) the 3D simulation of a flexible PD on a PET substrate, (b) the determined I-V curves in the dark and in the light, (c) the R and EQE with incident wavelength, (d) a photocurrent versus time plot of the device under 450 nm, 177 μ W/cm2 light irradiation. Reprinted with permission from Ref. [190]. Copyright 2025, Royal Society of Chemistry.
Jcs 09 00262 g019
Figure 20. Sb2S3 NWs-based PD: (a) I-V output curve. Reprinted with permission from Ref. [200]. Copyright 2021, Chinese Society of Metals. (b) I-t on/off curves under periodic illumination [201]. Sb2S3 NRs-based PD: (c) R (blue) and EQE (red), (d) detectability. Reprinted with permission from Ref. [202]. Copyright 2023, Royal Society of Chemistry.
Figure 20. Sb2S3 NWs-based PD: (a) I-V output curve. Reprinted with permission from Ref. [200]. Copyright 2021, Chinese Society of Metals. (b) I-t on/off curves under periodic illumination [201]. Sb2S3 NRs-based PD: (c) R (blue) and EQE (red), (d) detectability. Reprinted with permission from Ref. [202]. Copyright 2023, Royal Society of Chemistry.
Jcs 09 00262 g020
Figure 21. Pristine and Au-modified Sb2S3 NWs: (a) I-V curve; (b) response and recovery time of Au-decorated and (c) pristine Sb2S3 NWs under 700 nm light with power intensity of 150 μw cm−2 at bias of 10 V; (d) variation in I with V at different wavelengths; (e) transient photoresponse of flexible device under on/off cycling for 350 nm and 700 nm light, respectively; (f) response and recovery times of flexible device for 350 nm and 700 nm light, respectively. Reprinted with permission from Ref. [205]. Copyright 2017, Royal Society of Chemistry.
Figure 21. Pristine and Au-modified Sb2S3 NWs: (a) I-V curve; (b) response and recovery time of Au-decorated and (c) pristine Sb2S3 NWs under 700 nm light with power intensity of 150 μw cm−2 at bias of 10 V; (d) variation in I with V at different wavelengths; (e) transient photoresponse of flexible device under on/off cycling for 350 nm and 700 nm light, respectively; (f) response and recovery times of flexible device for 350 nm and 700 nm light, respectively. Reprinted with permission from Ref. [205]. Copyright 2017, Royal Society of Chemistry.
Jcs 09 00262 g021
Figure 22. Bi2S3 NRs: (a) Logarithmic plot of I-V curve of device under dark and illumination conditions; (b) current–time dependence at 5 V bias. Reprinted with permission from Ref. [213]. Copyright 2015, Springer Nature. (c) I-V curves of photodetectors under blue light (475 nm) and red light (650 nm) (inset: ITO/Bi2S3/ITO device model); (d) I-t curve. Reprinted with permission from Ref. [214]. Copyright 2017, Royal Society of Chemistry.
Figure 22. Bi2S3 NRs: (a) Logarithmic plot of I-V curve of device under dark and illumination conditions; (b) current–time dependence at 5 V bias. Reprinted with permission from Ref. [213]. Copyright 2015, Springer Nature. (c) I-V curves of photodetectors under blue light (475 nm) and red light (650 nm) (inset: ITO/Bi2S3/ITO device model); (d) I-t curve. Reprinted with permission from Ref. [214]. Copyright 2017, Royal Society of Chemistry.
Jcs 09 00262 g022
Figure 23. Bi2S3 NWs-base PD: (a) Photoresponse behavior under laser irradiation at 400 nm, 635 nm, and 830 nm wavelengths (power ~250 mW/cm2); (b) time-dependent photo-switching behavior under 830 nm laser illumination at four different power intensities; (c) D* and EQE; (d) photocurrent and R. Reprinted with permission from Ref. [216]. Copyright 2022, Elsevier.
Figure 23. Bi2S3 NWs-base PD: (a) Photoresponse behavior under laser irradiation at 400 nm, 635 nm, and 830 nm wavelengths (power ~250 mW/cm2); (b) time-dependent photo-switching behavior under 830 nm laser illumination at four different power intensities; (c) D* and EQE; (d) photocurrent and R. Reprinted with permission from Ref. [216]. Copyright 2022, Elsevier.
Jcs 09 00262 g023
Figure 24. Bi2S3 nanorod/thin film homojunction PD: (a) schematic structure, (b) current–voltage characteristics of optical OFF and ON states at irradiation power of 22.74 W/m2 and potential bias of ±1 V; (c) light detection under dark/light irradiation (22.74 W/m2) cycle with exposure times of 60 s each time. Reprinted with permission from Ref. [219]. Copyright 2022, Elsevier. (d) Comparison of UV-Vis absorption spectra of BSTC (blue), BS (red), and TC (black); (e) storage stability of BSTC4 photodetector after four months without any encapsulation (under 808 nm excitation and power density of 6.3 mW/cm2); (f) R, EQE, and D* versus power density curves of BSTC PD under 808 nm laser excitation. Reprinted with permission from Ref. [220]. Copyright 2024, Elsevier.
Figure 24. Bi2S3 nanorod/thin film homojunction PD: (a) schematic structure, (b) current–voltage characteristics of optical OFF and ON states at irradiation power of 22.74 W/m2 and potential bias of ±1 V; (c) light detection under dark/light irradiation (22.74 W/m2) cycle with exposure times of 60 s each time. Reprinted with permission from Ref. [219]. Copyright 2022, Elsevier. (d) Comparison of UV-Vis absorption spectra of BSTC (blue), BS (red), and TC (black); (e) storage stability of BSTC4 photodetector after four months without any encapsulation (under 808 nm excitation and power density of 6.3 mW/cm2); (f) R, EQE, and D* versus power density curves of BSTC PD under 808 nm laser excitation. Reprinted with permission from Ref. [220]. Copyright 2024, Elsevier.
Jcs 09 00262 g024
Figure 25. (a) Schematic diagram of self-powered photodetector based on CsPbBr3/ZnS heterostructure nanocrystals, (b) current–voltage (I-V) characteristics of device under varying light intensities, (c) time-dependent photoresponse curve at 0 V bias under 2 mW/cm2 laser illumination. Reprinted with permission from Ref. [227]. Copyright 2025, ACS Publications.
Figure 25. (a) Schematic diagram of self-powered photodetector based on CsPbBr3/ZnS heterostructure nanocrystals, (b) current–voltage (I-V) characteristics of device under varying light intensities, (c) time-dependent photoresponse curve at 0 V bias under 2 mW/cm2 laser illumination. Reprinted with permission from Ref. [227]. Copyright 2025, ACS Publications.
Jcs 09 00262 g025
Figure 26. MAPbI3/CdS heterojunction PD: (a) Schematic diagram, (b) I-V curves under dark and different light intensities, (c) photoresponse characteristics of three photodetectors under chopped light illumination. Reprinted with permission from Ref. [228]. Copyright 2019, ACS Publications. ZnO/CdS/GaN-based PD: (d) I-V curves under dark and UV illumination, inset shows I-V Ohmic contact characteristics of In and P-GaN contacts; (e) dependence of photocurrent on operation time at zero bias under UV illumination; (f) spectral response curve obtained at zero bias. Reprinted with permission from Ref. [229]. Copyright 2017, RSC Publishing.
Figure 26. MAPbI3/CdS heterojunction PD: (a) Schematic diagram, (b) I-V curves under dark and different light intensities, (c) photoresponse characteristics of three photodetectors under chopped light illumination. Reprinted with permission from Ref. [228]. Copyright 2019, ACS Publications. ZnO/CdS/GaN-based PD: (d) I-V curves under dark and UV illumination, inset shows I-V Ohmic contact characteristics of In and P-GaN contacts; (e) dependence of photocurrent on operation time at zero bias under UV illumination; (f) spectral response curve obtained at zero bias. Reprinted with permission from Ref. [229]. Copyright 2017, RSC Publishing.
Jcs 09 00262 g026
Figure 27. Nb-WS2/Ta2NiSe5-based PD: (a) schematic diagram of device, (b) photocurrent response of device across wavelengths from 405 to 1550 nm, (c) I-V curves under different light intensities at 660 nm. Reprinted with permission from Ref. [230]. Copyright 2025, ACS Publications. For MoS2 and MoS2-WS2 PDs under exposure to 580 nm light: (d) responsivity, (e) detectivity, (f) external quantum efficiency (EQE), measured at an input power density of 120 μW cm−2. Reprinted with permission from Ref. [231]. Copyright 2025, Wiley Online Library.
Figure 27. Nb-WS2/Ta2NiSe5-based PD: (a) schematic diagram of device, (b) photocurrent response of device across wavelengths from 405 to 1550 nm, (c) I-V curves under different light intensities at 660 nm. Reprinted with permission from Ref. [230]. Copyright 2025, ACS Publications. For MoS2 and MoS2-WS2 PDs under exposure to 580 nm light: (d) responsivity, (e) detectivity, (f) external quantum efficiency (EQE), measured at an input power density of 120 μW cm−2. Reprinted with permission from Ref. [231]. Copyright 2025, Wiley Online Library.
Jcs 09 00262 g027
Figure 28. MoS2/GaN-based PD: (a) microscopic image (top left), equivalent circuit for charge transport (top right), and schematic diagram of device; (b) absorption spectra at zero bias; (c) power-dependent I-T curves under ultraviolet (UV) light; (d) power-dependent I-T curves under visible light. Reprinted with permission from Ref. [158]. Copyright 2024, ACS Publications.
Figure 28. MoS2/GaN-based PD: (a) microscopic image (top left), equivalent circuit for charge transport (top right), and schematic diagram of device; (b) absorption spectra at zero bias; (c) power-dependent I-T curves under ultraviolet (UV) light; (d) power-dependent I-T curves under visible light. Reprinted with permission from Ref. [158]. Copyright 2024, ACS Publications.
Jcs 09 00262 g028
Figure 29. MoS2/GaAs/InGaAs-based PD: (a) schematic diagram, (b) I-V curves of device under 650 nm illumination at different light intensities, (c) relationship between device’s responsivity, external quantum efficiency (EQE), and light intensity at zero bias. Reprinted with permission from Ref. [233]. Copyright 2025, John Wiley and Sons.
Figure 29. MoS2/GaAs/InGaAs-based PD: (a) schematic diagram, (b) I-V curves of device under 650 nm illumination at different light intensities, (c) relationship between device’s responsivity, external quantum efficiency (EQE), and light intensity at zero bias. Reprinted with permission from Ref. [233]. Copyright 2025, John Wiley and Sons.
Jcs 09 00262 g029
Figure 30. TiO2/Sb2S3-based PD: (a) Structural diagram, (b) I-t curves under illumination at different wavelengths, (c) I-V curves under dark conditions and 625 nm illumination, (d) response time. Reprinted with permission from Ref. [235]. Copyright 2024, Optical Society of America.
Figure 30. TiO2/Sb2S3-based PD: (a) Structural diagram, (b) I-t curves under illumination at different wavelengths, (c) I-V curves under dark conditions and 625 nm illumination, (d) response time. Reprinted with permission from Ref. [235]. Copyright 2024, Optical Society of America.
Jcs 09 00262 g030
Figure 31. SnS2-based PD: (a) Device schematic diagram, (b) ln(I) −V characteristics of device at different temperatures, (c) time-dependent response of device at wavelength of 365 nm with optical power density of 1.112 mW/cm2. Reprinted with permission from Ref. [237]. Copyright 2020, IEE Xplore. WS2-based PD: (d) Device schematic diagram, (e) time-resolved photocurrent of PThTPTI/WS2 PD, (f) time-resolved photocurrent of PThTPTI/WS2 PD; properties of PDs measured after being annealed at different temperatures. Reprinted with permission from Ref. [238]. Copyright 2016, ACS Publications.
Figure 31. SnS2-based PD: (a) Device schematic diagram, (b) ln(I) −V characteristics of device at different temperatures, (c) time-dependent response of device at wavelength of 365 nm with optical power density of 1.112 mW/cm2. Reprinted with permission from Ref. [237]. Copyright 2020, IEE Xplore. WS2-based PD: (d) Device schematic diagram, (e) time-resolved photocurrent of PThTPTI/WS2 PD, (f) time-resolved photocurrent of PThTPTI/WS2 PD; properties of PDs measured after being annealed at different temperatures. Reprinted with permission from Ref. [238]. Copyright 2016, ACS Publications.
Jcs 09 00262 g031
Figure 32. Nf-MoS2-based PD: (a) Schematic diagram of device, (b) I-V characteristics of device under self-driven conditions (applied 0 V bias) at 100 °C and 2 W optical power, (c) I-T curve. Reprinted with permission from Ref. [239]. Copyright 2023, ScienceDirect.
Figure 32. Nf-MoS2-based PD: (a) Schematic diagram of device, (b) I-V characteristics of device under self-driven conditions (applied 0 V bias) at 100 °C and 2 W optical power, (c) I-T curve. Reprinted with permission from Ref. [239]. Copyright 2023, ScienceDirect.
Jcs 09 00262 g032
Figure 33. A-Ga2O3-based PD: (a) Schematic diagram, (b) dark current and photocurrent at different measurement temperatures, (c) time-dependent photoresponse under 250 nm illumination at various measurement temperatures. Reprinted with permission from Ref. [243]. Copyright 2021, IOP Publishing. Ε-Ga2O3/ZnO-based PD: (d) Device diagram, (e) I-V curves of device at different temperatures, (f) rise time and decay time of device at different temperatures. Reprinted with permission from Ref. [244]. Copyright 2022, IOPscience.
Figure 33. A-Ga2O3-based PD: (a) Schematic diagram, (b) dark current and photocurrent at different measurement temperatures, (c) time-dependent photoresponse under 250 nm illumination at various measurement temperatures. Reprinted with permission from Ref. [243]. Copyright 2021, IOP Publishing. Ε-Ga2O3/ZnO-based PD: (d) Device diagram, (e) I-V curves of device at different temperatures, (f) rise time and decay time of device at different temperatures. Reprinted with permission from Ref. [244]. Copyright 2022, IOPscience.
Jcs 09 00262 g033
Table 1. Work functions of different electrode materials.
Table 1. Work functions of different electrode materials.
Electrode MaterialWork Functions (eV)Contact TypeI-V Curve FeaturesRef.
Au5.1Schottky contactNonlinear[74]
Cr4.5Ohmic contactLinear[75]
ITO4.4–4.5Quasi-Ohmic contactClose to linear[75]
Table 2. Performance characteristics of ZnSe-based MSM UV photodetectors with different Schottky contacts. Reprinted with permission from Ref. [78]. Copyright 2024, Pleiades Publishing.
Table 2. Performance characteristics of ZnSe-based MSM UV photodetectors with different Schottky contacts. Reprinted with permission from Ref. [78]. Copyright 2024, Pleiades Publishing.
Type of ContactsDark Current, nA (15 V)Rejection Rate (15 V)Responsivity,
A W−1 (15 V)
Photo-Conductivity Gain (15 V)Detectivity, cm W−1 HZ1/2 (15 V)
MSM, conventional Cr/Au0.715.4 × 1031.25~148.3 × 1010
MSM, interdigitated Cr/Au1.644.2 × 1032.23~269.9 × 1010
MSM, conventional Ni/Au0.594.1 × 1030.79~95.8 × 1010
MSM, interdigitated Ni/Au0.822.0 × 1045.40~623.4 × 1011
MSM, hybrid Ni/Au and Ag-NW0.365.0 × 1030.58~75.5 × 1010
Table 3. Crystal structure, semiconductor type, resistivity, carrier mobility, and carrier concentration of various metal sulfide nanostructures.
Table 3. Crystal structure, semiconductor type, resistivity, carrier mobility, and carrier concentration of various metal sulfide nanostructures.
MaterialCrystal StructureBandgap
(eV)
Direct- Indirect-
Type of ConductivityElectrical Resistivity
(Ω·cm)
Carrier Concentration
(cm−3)
Carrier Mobility
(cm2/v·s)
Ref.
ZnSCubic3.7~n-type1.79 × 1036.04 × 101990[91,92,93,94]
CdSHexagonal2.4~n-type8.11 × 10−22.96 × 102014.5[95,96,97]
SnSOrthorhombic1.321.09p-type5.4 × 10310 1737.75[98,99,100]
WS2Hexagonal2.11.35n-type5 × 1041018~101950[101,102,103]
MoS2Hexagonal1.91.3n-type8.59 × 1041014~10150.5–17[104,105,106,107]
SnS2Trigonal2.92.1n-type1.2 × 1025.43 × 1017230[108,109]
ReS2Triclinic1.5~n-type2~5101916–30[110,111,112]
Sb2S3Orthorhombic1.78~n-type1.3 × 1047.3 × 10136.4[113,114,115]
Bi2S3Orthorhombic1.3~n-type7 × 1031017~1020~[116,117,118]
In2S3Hexagonal2.0~n-type38.83.8 × 101542.3[119,120]
TiS3Monoclinic1.1~n-type2~6~80[121,122,123]
Table 4. Synthesis methods of 1D monosulfide metal compounds and their photodetector properties.
Table 4. Synthesis methods of 1D monosulfide metal compounds and their photodetector properties.
MaterialContact ElectrodesMethodologyWavelength
Power
Responsivity
(A/W)
Detectivity (Jones)EQE
(%)
Rise Time
Fall Time
Photo-I
Dark-I
Ref.
ZnS NWsPtTCVD365 nm
38 mW/cm2
2.761~0.5723.2 s
3.6 s
~[129]
ZnS NWsAuCVT325 nm
1.25 mW/cm2
~~~<0.1 s~[130]
ZnS NTsAg Thermal Evaporation~
40 mW/cm2
16.51.41 × 1098.920.12 s
0.4 s
8.44 pA
1.32 μA
[131]
ZnS NTsAgThermal Evaporation297 nm
40 mW/cm2
2.561.67 × 101013.60.09 s
0.07 s
14.9 pA
0.29 μA
[132]
ZnS NWACr/AuMOCVD325 nm
16 mW/cm2
>1.87~>7105 ms
40 ms
~
508 pA
[133]
ZnS:Mn NRsAgHydrothermal310 nm
0.5 W/cm2
1.8~71916 ms
1.1 ms
~[134]
ZnS:Al NWsITOThermal Co-Evaporation254 nm
300 mW/cm2
3.1 × 106~1.5 × 107153 s
445 s
~[135]
ZnS/InP NWsCr/AuCVD323 nm
1.87 μW/cm2
2951.65 × 10131.10 × 1030.75 s
0.5 s
~
10.9 pA
[136]
ZnS/CdS NWsCr/AuThermal Evaporation450 nm
213 μW/cm2
~2.23 × 1014~5 ms
7 ms
10 fA
1200 pA
[137]
CdS NWsCr/AuVapor Transfer470 nm
0.93 mW/cm2
~4.27 × 1012~0.3 s
0.4 s
4.7 fA
92.2 pA
[147]
CdS NWsCr/AuCVD375 nm
0.01 mW/cm2
2.6 × 1052.3 × 10168.6 × 10512.6 ms
180 ms
10−12 A
1.13 μA
[148]
CdS NRsTi/AuCVD450 nm
0.5 mW/cm2
1.2 × 1042.8 × 10113.5 × 1060.82 s
0.84 s
~[149]
Sb/CdS NWsCr/AuTwo-Step CVD470 nm
19.1 μW/cm2
93.622.33 × 10142.47 × 1020.384 s
0.312 s
12.85 fA
45.5 pA
[150]
SnS/CdS NWAAuHydrothermal-Thermal Evaporation650 nm
0.08 mW/cm2
10.4 m3.56 × 1011~<30 ms
<30 ms
14.9 pA
0.29 μA
[151]
Si/CdS NRsAgHydrothermal~
2.55 mW/cm2
64.8 m1.31 × 1010~190.8 μs
298.4 μs
~[139]
PPV/CdS NWAAuElectrochemical Co-Deposition545 nm
4.2 mW/cm2
~~~~0.027 μA
1.457 μA
[155]
TiO2/CdS NRsAgHydrothermal365 nm
40 μW/cm2
2.8659.9 × 1012971.360.99 s
0.49 s
~[157]
SnS NWsCr/AuCVD838 nm
0.05 mW/cm2
1.6 × 1042.4 × 1012~1.2 ms
15.1 ms
~[164]
SnS NWsCr/AuCVD830 nm
0.12 mW/cm2
2.6 × 1021.8 × 1013~9.6 ms
14 ms
~[165]
Table 5. Synthesis of one-dimensional metal disulfide compounds and their photodetector properties.
Table 5. Synthesis of one-dimensional metal disulfide compounds and their photodetector properties.
MaterialContact ElectrodesMethodologyWavelength
Power
Responsivity
(A/W)
Detectivity (Jones)EQE
(%)
Rise Time
Fall Time
Photo-I
/Dark-I
Ref.
SnS2 NWsCr/AuCVD520 nm
0.06 mW/cm2
2.1 × 1051.3 × 10164.0 × 10556 ms
91 ms
3 × 10−13 A
2.74 μA
[172]
Si/SnS2 NWsAu/AlHydrothermal 340 nm
20 nW/cm2
383~1.3 × 1050.55 s
0.33 s
2.9 × 10−12 A
~
[173]
WS2 NTsTi/AuHigh temperature 633 nm
532 W/cm2
2360~4.6 × 106256 μs
286 μs
336 times lower[54]
WS2 NTs-AuSolution synthesis 552 nm
0.06 mW/cm2
0.2 m2.23 × 104~0.92 s
0.09 s
~[182]
MoS2/WS2 NRsAgHydrothermal785 nm
50 mW/cm2
15 μ24 × 10616.9 × 10−60.82 s
1.59 s
~[183]
MoS2 NTsAgHydrothermal915 nm
100 mW/cm2
2.33 m7.55 × 1083.33 × 10−15.3 s
1.53 s
~
2 μA
[184]
Y-TiOPc/MoS2 NTsAuAnion exchange365–850 nm
0.01 mW/cm2
20,588 m1.94 × 1012494,736134 ms
143 ms
~[185]
β-Ga2O3/MoS2 NRsPtMagnetron sputtering266 nm
30 μW/cm2
42.113.2 × 10111.97 × 1040.29 s
0.3 s
79 nA
16.92 μA
[186]
ReS2 NWsAgCVD500 nm
0.42 nW/cm2
5.08 × 1056.1 × 10151.07 × 1061.8 s
3.9 s
0.42 pA
4.95 nA
[188]
Table 6. Properties of one-dimensional polysulfide metal compound-based photodetectors and methods of material synthesis.
Table 6. Properties of one-dimensional polysulfide metal compound-based photodetectors and methods of material synthesis.
MaterialContact ElectrodesMethodologyWavelength PowerResponsivity
(A/W)
Detectivity (Jones)EQE
(%)
Rise Time
Fall Time
Photo-I
Dark-I
Ref.
In2S3 NWsCr/AuNPT-CVT405 nm
2 W/cm2
16.01~4903~9.95 nA
112.4 μA
[189]
In2S3 NWsAu/NiCVD450 nm
176.7 W/cm2
7.35 × 1042.4 × 10142.88 × 1076.5 ms
9.5 ms
0.29 pA
293 nA
[190]
Sb2S3 NWsAuSteam Transport638 nm
0.03 mW/cm2
11522 × 1013~37 ms
38 ms
2 × 10−10 A
4.2 × 10−8 A
[198]
Sb2S3 NWsAuSteam Transport450 nm
40 μW/cm2
343.4 m~~470 μs
680 μs
69.6 times lower[199]
Sb2S3 NWsNi/AuCVD532 nm
0.03 mW/cm2
652.1 × 10141.5 × 10476 ms
83 ms
~
2.5 nA
[200]
Sb2S3 NWsTi/AuAPCVD635 nm
318 μW/cm2
2704.37 × 10135.3 × 10410 ms
12 ms
2800 times lower[201]
Sb2S3 NRsAuHydrothermal560 nm
0.38 mW/cm2
5.12.16 × 10101130.684.03 ms
4.08 ms
109.8 times lower[202]
Sb2S3 NTs CVD808 nm
300 mW/cm2
8.5 m1.33 × 106~22 ms
24 ms
0.93 nA
185 nA
[203]
Sb2S3 NWAFTOPolyol Reflux808 nm
640 mW/cm2
~~~0.52 s
1.1 s
7.4 nA
113.5 nA
[204]
Au:Sb2S3 NWsCr/AuCVD600 nm
680 μW/cm2
59.54.29 × 1010~0.2 s
0.3 s
57 pA
163 nA
[205]
CuSCN/Sb2S3 NRsITOTwo-Step Hydrothermal600 nm
680 μW/cm2
~~~0.18 s
0.15 s
102 times lower[206]
Mo2C-C/Sb2S3 NRsAuHydrothermal400 nm
320 W/cm2
~~~52.7 ms
79.2 ms
150 times lower[207]
Bi2S3 NRsAuOne-Pot Hydrothermal~~~~371.6 ms
386 ms
102 times lower[213]
Bi2S3 NRsITOColloidal Chemical650 nm
4.1 mW/cm2
~~~0.3 s
0.6 s
10 times lower[214]
Bi2S3 NWsCr/AuCVD700 nm
1.54 mW/cm2
3.57~6330.1
0.1
~[215]
Bi2S3 NWsPtCVD830 nm
52 mW/cm2
4.211.64 × 1010981.7612.25 ms
12.25 ms
~[216]
Bi2S3 NWsTi/AuAPCVD532 nm
23.8 μW/cm2
23,7603.68 × 10135.5 × 1061 ms
4.5 ms
~[217]
Bi2S3 NRAAuPVD830 nm
64 nW/cm2
52331.8 × 10127.8 × 10321 μs
7.8 ms
~[218]
Bi2S3 NRATaHot Plate
PVD
~
22.7 W/cm2
749.65.61 × 108~192 ms
270 ms
0.96 μA
4.5 μA
[219]
Ti3C2Tx/Bi2S3 NRsAuIn Situ Hydrothermal808 nm
0.03 mW/cm2
2.55 × 10−23.9 × 10123.9 × 1030.3 ms
2.1 ms
255 times lower[220]
Bi2S3/PANI NRsAgHydrothermal365 nm
50 μW/cm2
26,7605.24 × 1013907350 ms
60 ms
0.003 μA
0.58 μA
[221]
Cs3Bi2I9:Bi2S3 NRsAg/CUSD532 nm
790 μW/cm2
0.59 × 10−38.18 × 109198~0.68 nA
1.35 nA
[222]
Table 7. Performance comparison of self-powered photodetectors.
Table 7. Performance comparison of self-powered photodetectors.
MaterialContact ElectrodesMethodologyWavelength
Power
Responsivity
(A/W)
Detectivity (Jones)EQE
(%)
Rise Time
Fall Time
Ref.
CsPbBr3/ZnSAl/ITOHot injection and spin coating365 nm
2 mW/cm2
37.5 m1.21 × 1012~150 ms
30 ms
[227]
MAPbI3/CdSAu/ITOCBD325 nm
10 mW/cm2
0.43 2.3 × 1011 ~3.2 ms
9.6 ms
[228]
ZnO/CdS/GaNIn/ITO~300 nm
0.61 mW/cm2
176 m1012 719<0.35 s
<0.35 s
[229]
Nb-WS2/Ta2NiSe5AuMechanical exfoliation and dry transfer techniques660 nm
0.45 mW/cm2
57.64 6 × 1010 10,854118 μs
13 μs
[230]
MoS2/WS2Au/CrSpin coating and CVD580 nm
120 μW/cm2
282 m~60.59375 μs
6 ms
[231]
Ag-MoS2PtAr plasma532 nm
0.519 W/cm2
280 m0.76 × 101165224 μs
293 μs
[232]
MoS2/GaNAu/NiVapor transferUV
0.93 mW/cm2
631 m8.5 × 10102141.04 ms
0.8 ms
[158]
MoS2/GaAs/InGaAsAu/NiMBE650 nm
3.48 mW/cm2
86 m2.3 × 10161612.6 ms
18 ms
[233]
SnS2FTOHydrothermal475 nm
10 mW/cm2
1460 μ~~~[234]
TiO2/Sb2S3FTO/AuHydrothermal and spin coating625 nm
9 W/cm2
0.29 3.37 × 1012·~9.3µs
7.8µs
[235]
p-Si/Bi2S3AuCVDRed
~
0.94 m8.92 × 10829,685~[236]
Table 8. Performance comparison of high-temperature photodetector devices.
Table 8. Performance comparison of high-temperature photodetector devices.
MaterialContact ElectrodesMethodologyOperating Temperature (℃)Responsivity
(A/W)
Detectivity (Jones)EQE
(%)
Rise Time
Fall Time
Ref.
SnS2AgSolvothermal method1205.51.72 × 101318682.2 s
6.3 s
[237]
PThTPTI/WS2AuCVD300~~~6.89 ms
8.43 ms
[238]
MoS2PtRF sputtering technology1001170 m1.6 × 1010 409~[239]
MoS2/GaNAuVapor transfer100631 m8.5 × 10102141.04 ms
0.8 ms
[158]
PDPPVTT/MoS2Au/CrCVT~~~~~[240]
MoS2/GaN/Al2O3Au/CrCVD and ALD~24.62~8381~[241]
Ga2O3Au/NiCVD2000.518~~~[243]
ε-Ga2 O3/ZnOTi/AuCVD>500 K2.4 m51.5~[244]
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

Chen, J.; Li, M.; Lin, H.; Zhou, C.; Chen, W.; Wang, Z.; Li, H. Advances in One-Dimensional Metal Sulfide Nanostructure-Based Photodetectors with Different Compositions. J. Compos. Sci. 2025, 9, 262. https://doi.org/10.3390/jcs9060262

AMA Style

Chen J, Li M, Lin H, Zhou C, Chen W, Wang Z, Li H. Advances in One-Dimensional Metal Sulfide Nanostructure-Based Photodetectors with Different Compositions. Journal of Composites Science. 2025; 9(6):262. https://doi.org/10.3390/jcs9060262

Chicago/Turabian Style

Chen, Jing, Mingxuan Li, Haowei Lin, Chenchen Zhou, Wenbo Chen, Zhenling Wang, and Huiying Li. 2025. "Advances in One-Dimensional Metal Sulfide Nanostructure-Based Photodetectors with Different Compositions" Journal of Composites Science 9, no. 6: 262. https://doi.org/10.3390/jcs9060262

APA Style

Chen, J., Li, M., Lin, H., Zhou, C., Chen, W., Wang, Z., & Li, H. (2025). Advances in One-Dimensional Metal Sulfide Nanostructure-Based Photodetectors with Different Compositions. Journal of Composites Science, 9(6), 262. https://doi.org/10.3390/jcs9060262

Article Metrics

Back to TopTop