Abstract
Due to their exceptional performance across a wide range of wavelength regions, from UV to broadband detection, 2D materials (2DMs)-based photodetectors have expanded quickly in recent years. 2DMs gained a reputation for playing important roles in photodetection because of their unique physical characteristics, with layer number engineering, low cost and simple preparation, wet chemical processing, deterministic all-dry transfer process, surface passivation for dangle bond-free surface, band gap engineering, valleytronics, and so forth. Given these benefits, the preparation of 2DMs, specifically molybdenum disulfide (MoS2), paves several advantages, and the use of near-infrared detection was the main topic of discussion in this chapter. MoS2 is a well-known transition metal dichalcogenide (TMD) with a tunable bandgap (from ∼1.8 eV in monolayer to ∼1.2 eV in bulk). Heterostructures of MoS2 enable band alignment optimization for extended NIR absorption. Further, the van der Waals heterostructures facilitate efficient charge transfer and reduce recombination losses. In addition, Type-II band alignment in MoS2-based heterostructures promotes carrier separation, enhancing photocurrent generation. Due to these exceptional characters, the proposed chapter covers the fundamentals, operation principles, and fabrication of MoS2 heterostructure-based photodetectors (heterojunction photodetectors), overcoming the generation of dark current, low quantum efficiency, and enhancing detectivity. This chapter offers a thorough guide to achieving high-performance near-infrared photodetectors based on MoS2.
Keywords
- MoS2
- NIR detection
- van der Waals heterostructures
- photogenerated carriers
- carrier lifetime
1. Introduction
In the modern era of the electronics industry, photodetectors are essential parts. It is essentially an optoelectronic device that generates electricity from light. They are also called photosensors, which work primarily as the optical receivers to convert information from the light into electrical signals. They are crucial for capture, identification, and visualization of optical information [1]. The first photodetector emerged during World War II (1933) by Edgar Kutzscher, who found that lead sulfide (PbS) can be used as a photoconducting detector to detect hot emissions (infrared radiation) from bombers and ships. Our exploration of photo detectors began relatively recently, in the early 1990s [2]. In stark contrast, humanity’s fascination with the nature of light stretches back to as early as 300 BC. This timeline shows our long-standing effort to understand light, laying the groundwork for the development of advanced photodetector technologies. The research on light started at 300 BC with Aristotle, followed by Huygens, who found the wave nature of light in the seventeenth century. Later in the eighteenth century, Maxwell predicted that those waves were electromagnetic waves with both electric and magnetic components. After many remarkable findings and theories in the nineteenth century, Einstein proposed that “light is composed of small packets of electromagnetic energy called photons.” He even proved it with the help of the photoelectric effect, which is the basic principle behind the working of photodetectors [3]. Photodetectors work based on the principle of creating electron-hole pairs when exposed to light. When photons with energies equal to or greater than the material’s bandgap hit the material, they excite electrons from the valence band to the conduction band, leaving behind positively charged holes. These conduction band electrons and valence band holes can move due to concentration gradients or electric fields, whether intrinsic or externally applied. These photogenerated electron-hole pairs may recombine and release light in the absence of an electric field. They can, however, be separated as free charge carriers by an electric field, and some of these free charges will gather at the electrode of the photodetector to form a photocurrent. This phenomenon of photocurrent generation is depicted in Figure 1. The strength of the incoming light has a direct correlation with the photocurrent’s magnitude at a particular wavelength [4]. In 1800, Sir Frederick William Herschel paved the way to let us know about the wide spectrum of light. He was the first man to find that there is light beyond the red portion of the spectrum, and he named it “Calorific rays” which is now called “Infrared radiation.” In 1801, followed by his finding, Johann Wilhelm Ritter found that there is also light beyond the violet portion, and he named it “Chemical rays” which is now called “Ultraviolet.” Ultraviolet to infrared light is all the electromagnetic radiation with different wavelengths that constitutes the electromagnetic spectrum. The ability to detect light in different sub-bands has many uses in industries, the military, agriculture, and the biological sciences. Displays, video imaging, electronic eyes, optoelectronic storage, industrial safety, and other applications use visible (Vis) light detection, while ultraviolet (UV) light detection is used in military surveillance, space exploration, border security, environmental monitoring, and sterilization procedures, among other applications. Fire alarms, precision strikes, satellite remote sensing, optical communication, night vision technologies, remote control systems, analytical science, and navigational aids are just a few of the applications for infrared (IR) light detection [5]. Figure 2 illustrates the broad electromagnetic spectrum, highlighting the wavelength ranges and the materials typically employed for detecting radiation in each spectral region. An ideal photodetector should efficiently convert light into electrical signals, detect weak optical signals, respond quickly, and utilize photons for charge carrier generation. These qualities are evaluated through figures of merit like responsivity, response time, noise equivalent power, detectivity, quantum efficiency, and gain, which are influenced by material properties such as bandgap, absorption coefficient, and charge carrier dynamics. Understanding these metrics helps assess material performance and optimization in photodetector applications. The formulas for each figure of merit and how they relate to material behavior and performance have been discussed in the following sections.

Figure 1.
The schematics of the operation principle of the photodetection (left), photon absorption and excitation (right), and electron-hole pair generation by the photoexcitation.

Figure 2.
Illustration of the range of energy of the electromagnetic spectrum with respective material classification.
1.1 Responsivity
The ratio of the output photocurrent of the photodetector for a given incident optical power at a certain wavelength on the active region of the device.
1.2 Response time
It is typically defined in terms of rise time (τr) and fall or decay time (τf). The rise time corresponds to the time required for the photocurrent to increase from 10 to 90% of its peak value, while the fall time represents the duration needed for the photocurrent to decrease from 90–10% of its peak value [7]. In the frequency domain, the response time is associated with the bandwidth (
1.3 Noise equivalent power (NEP)
It is described as the lowest optical power needed to obtain a signal-to-noise ratio (SNR) of 1 across a bandwidth of 1 Hz.
Where
A highly sensitive photodetector requires a high value of SNR [6]. In some cases, the NEP can be written as,
where
Where
Shot noise is derived from random thermal agitation of charge carriers.
where
Where
Where
1.4 Detectivity
It represents the sensitivity of the photodetector by characterizing its ability to distinguish and detect weak optical signals from noise.
Where
Where
Where
1.5 Quantum efficiency
Quantum efficiency measures the effectiveness of a photodetector in converting incident photons into charge carriers. The two classifications are external quantum efficiency (EQE) and internal quantum efficiency (IQE). EQE (
Where
The number of absorbed photons
In the absence of gain, EQE and IQE can be related as
Here,
1.6 Gain
It evaluates the capability of the photodetector to generate multiple charge carriers from a single incident photon. It quantifies the extent of photo-exciton recycling through a multiplication mechanism and denotes the number of charge carriers detected per incident photon.
Where c is the velocity of light,
The gain can also be expressed as
The EQE for a photodiode is typically less than 100% when no carrier multiplication is occurring. In the case of an ideal photodiode with an EQE of 100%, the responsivity is expressed as,
1.6.1 Gain in photoconductors
In photoconductors with a perfect semiconductor, the gain arises due to the difference in mobility between electrons and holes. Electrons, having higher mobility, move faster under an external field. When an electron-hole pair is generated in the device’s center, the photogenerated electron is collected at the drain electrode before the hole reaches the source electrode. To maintain charge neutrality, an electron is injected from the source electrode. This cycle continues until the hole reaches the source electrode or recombines with an electron, allowing multiple electrons to be collected for each photon, resulting in gain. In imperfect semiconductors, gain can also result from charge traps. The photogenerated holes get trapped by these defect traps. In order to maintain the charge neutrality, electrons have to be injected from the source electrode, which will be collected at the drain. When a photon causes more than one electron to be collected, then gain will occur.
1.6.2 Gain in photodiodes
Photodiodes operate by separating and collecting photogenerated carriers using the built-in electric field. The injection of electrons into the p-type semiconductor is not possible because of its minority nature despite being more mobile than holes. For holes, the p-type semiconductor and electrode make an Ohmic contact; for electrons, it creates a blocking contact, which will also lead to the restriction of electrons. For these reasons, even in the reverse bias condition, no gain is obtained.
Without permitting carrier multiplication, the reverse bias voltage merely increases the depletion region and the electric field inside it, increasing light absorption and carrier drift velocity. Gain cannot be induced by charge traps in the depletion zone because carrier injection is prevented by the blocking contact. Instead, these traps may cause recombination of photogenerated carriers, reducing the photo response. To achieve a gain in photodiodes, alternative carrier multiplication mechanisms are needed. One such mechanism is avalanche breakdown, which involves impact ionization. This process leads to carrier multiplication and gain, which is widely employed in avalanche photodiodes. However, avalanche breakdown requires a high reverse bias. Additionally, photogenerated hot carriers with sufficient energy can also cause carrier multiplication, resulting in external quantum efficiencies (EQE) greater than 100%. The impact ionization can be obtained only when the energy of the photon surpasses the bandgap energy [9]. Silicon (CMOS-based) photodetectors dominate the photodetector market due to their outstanding photoelectric characteristics, affordability, and seamless integration with electronic systems. However, Si cannot absorb the majority of the infrared spectrum due to its bandgap of about 1.1 eV [12]. In addition, due to its indirect bandgap nature, Si has a relatively weak absorption. Such inherent properties of Si make it challenging to develop active photodetectors. Recent years have seen a bottleneck in the fabrication of silicon-based devices, which would result in the end of Moore’s Law [13]. To address these limitations, researchers have focused on integrating silicon with other semiconductor materials such as germanium; Group III-V semiconductors: gallium arsenide (GaAs), indium phosphide (InP), gallium nitride (GaN) etc.; Group II-VI semiconductors: cadmium telluride (CdTe), cadmium sulfide (CdS), etc.; and two-dimensional materials. Over the past decade, there has been considerable effort to design and fabricate photodetectors that combine silicon and various semiconductor materials. This strategy has earned significant attention due to its potential to improve the performance of silicon-based photodetectors. However, these materials had drawbacks such as limited flexibility, irreversible dangling bonds, a short bandgap, and low optical transparency. In their quest for superior materials, scientists have discovered the promising potential of two-dimensional (2D) materials in the realm of photodetectors. These materials, with their unique properties, have opened new chapters for innovation, offering enhanced performance and capabilities in light detection technologies. Since the isolation of graphene (Gr), by Novoselov and Geim in 2004, 2D materials (2DMs) have evolved rapidly during the last 15 years. With hundreds of members, 2DMs have so far expanded into a large family encompassing graphene, group VA semiconductors, group IVB transition metal dichalcogenides (TMDCs), group VIB TMDCs, group IVA metal chalcogenides (MCs), black phosphorus (b-P), hexagonal boron nitride (h-BN), etc. Two-dimensional (2D) materials are inherently layered, making them outstanding candidates for photodetectors. Their high charge carrier mobility, combined with the extraordinary tunability of the band gap through variations in thickness, external strain, and electric fields, sets them apart [6]. Additionally, these materials can be easily stacked to form van der Waals heterostructures, effectively reducing lattice mismatch. These unique characteristics position 2D materials as superior choices for cutting-edge photodetector technology. Graphene, the first two-dimensional (2D) material to gain attention for photodetection applications, owes its prominence to its exceptional electrical properties. Notably, its remarkable planar mobility, which reaches up to 200,000 cm2/(V·s), facilitates the creation of photodetectors with bandwidths as high as 40 GHz. However, one major limitation of graphene as an active layer in photodetectors is its lack of an energy band gap, which results in significant noise contributions from dark currents [14]. Consequently, the scientific community has increasingly focused on 2D materials with finite band gaps. Among the array of two-dimensional materials, transition metal dichalcogenides (TMDCs) have gained significant attention. Their unique properties, such as the transition from an indirect to a direct band gap when their dimensions are reduced from the bulk material to the monolayer limit, and this dimensional reduction enhances their light-matter interaction and tunable electronic characteristics, make them stand out in the field of photodetectors. These materials promise enhanced performance, with intriguing electro-optical properties despite their relatively modest mobility of 200 cm2/(V·s). TMDCs exhibit absorbance values that are an order of magnitude higher than those of silicon (Si) and gallium arsenide (GaAs), enabling efficient light absorption with a very thin layer of the photoactive material. These properties position TMDCs as promising candidates for advanced photodetector applications, offering significant advantages over traditional materials. Among the TMDCs, a group VI transition metal dichalcogenide, molybdenum disulfide (MoS2) attracted much interest in the last decade for photodetection because of its efficient absorption and emission of light. MoS2 structure consists of 3 atomic layers, where a layer of transition metal atoms (Mo, W) is positioned between two layers of chalcogen atoms (S, Se, Te) with strong in-plane covalent and weak out-of-plane van der Waals interactions, resulting in easy exfoliation into monolayers with atomically thin thickness [15]. The molybdenite is relatively abundant in nature; this widespread availability of molybdenite makes MoS2 an economically viable option for large-scale production, ensuring a steady supply of raw material. MoS2 has a tunable bandgap that changes from 1.3 eV for multilayer (ML) MoS2 (indirect bandgap) to 1.9 eV for single layer (SL). This change in bandgap value is arising from d-orbital-related interactions [16, 17]. MoS2 exhibits a high absorption coefficient of up to 2.8 × 106 cm−1, which is at least one order of magnitude higher than that of conventional semiconductors like Si and GaAs. This is a reflection of the strong light-matter interaction that distinguishes MoS2 (19). Two separate low-energy peaks in the MoS2 absorption spectra at 1.88 eV (A excitons) and 1.06 eV (B excitons) were obtained because of the direct and indirect transitions at the K point. The PL peaks appear approximately at 620 and 688 nm, corresponding to the indirect-gap and direct-gap luminescence. MoS2 monolayers exhibit exceptional photoluminescence quantum yield (PLQY), reaching over 95%. Additionally, they have a remarkable minority carrier lifetime, with the longest-observed duration being 10.8 ± 0.6 nanoseconds by means of surface passivation [18, 19]. Van Hove singularities in the electronic density of states of TMDC guarantee enhanced light-matter interactions, leading to enhanced photon absorption and electron-hole creation (which are collected in transparent graphene electrodes). This allows development of extremely efficient flexible photovoltaic devices with photoresponsivity above 0.1 ampere per watt (corresponding to an external quantum efficiency of above 30%) [20]. In ultrathin 2D MoS2, reduced electronic Coulomb screening due to charge confinement significantly increases exciton binding energy. This enhancement alters the wavelengths of light absorbed or emitted, resulting in an exciton binding energy increase of hundreds of meV. These characteristics position MoS2 as a leading candidate for photodetection applications [21].
2. Fabrication techniques of MoS2 heterostructure photodetectors
The concept of assembling and fabricating 2D materials, including MoS2, into heterostructures was first proposed by Wang et al. in their work on high-quality graphene electronics [22], leading to the discovery of numerous novel properties in these engineered systems. These heterostructures can be broadly classified into two categories: vertically stacked heterostructures, where layers are assembled via van der Waals interactions, and epitaxially grown planar heterostructures, where materials are seamlessly integrated through direct growth, ensuring strong in-plane bonding. Among the various methods of fabrication of these heterostructures, the Chemical Vapor Deposition (CVD) and deterministic transfer methods are the powerful, efficient way to attain the better devices with remarkable properties. These methods were briefly discussed here. A high-precision method for creating MoS2 heterostructures, the deterministic transfer method guarantees regulated monolayer stacking free from contamination or misalignment. The schematic experimental setup and the deterministic transfer process have been illustrated in the Figure 3(a) and (b) respectively. The development of high-performance van der Waals (vdW) heterostructures, where exact interlayer coupling directly affects electronic and optical properties, requires this technique. First, bulk crystals of monolayer or few-layer MoS2 are usually exfoliated by applying liquid exfoliation, Nitto tape, or scotch tape on SiO2/Si, sapphire, or other substrates. A viscoelastic polymer, like polydimethylsiloxane (PDMS) or poly(methyl methacrylate) (PMMA), is applied to the monolayer as a support layer during the transfer process. With the aid of micromanipulators and an optical microscope, the MoS2 flake is precisely positioned over the target substrate or another two-dimensional material (e.g., WS2, graphene). A temperature-controlled stamp-assisted technique (≈ 40–100°C) can be used to achieve the clean adhesion while reducing interfacial contamination. Ultimately, the pristine interface is restored by annealing (approximately 300–400°C in Ar/H2 atmosphere) or acetone removal of the polymer layer [23].

Figure 3.
Deterministic setup and procedure for transfers. (a) schematic representation of the experimental configuration used for the all-dry transfer procedure. The viscoelastic stamp preparation and the deterministic transfer of an atomically thin flake onto a user-defined location (such as another atomically thin flake) are depicted in the diagram (b) reproduced with permission from [23].
Chemical Vapor Deposition (CVD) is a widely used bottom-up approach for the fabrication of vertical 2D heterostructures, enabling the growth of planar multi-junction heterostructures. Several strategies have been explored to achieve scalable and controllable fabrication of both vertical and planar heterostructures with atomically sharp heterojunctions and clean interfaces. The two-step CVD method, commonly used for this purpose, involves growing layered crystals as a substrate for the second layer. The rate of gas flow and synthesis time play crucial roles in the formation of vertically stacked and lateral heterostructures. For instance, the growth of a 2D GaSe/MoSe2 heterostructure was achieved using this method, though large lattice misfits between the layers can result in incommensurate superstructures [24]. Additionally, a stacked TMD/hBN heterostructure was synthesized using a Ni-Ga alloy and Mo foil as substrates, where the alloy promoted the formation of the hBN honeycomb lattice while Mo served as a source of Mo [25]. The multi-step CVD method offers greater flexibility and control by switching the direction and modulating the components of the gas flow. This approach allows for sharper heterojunction boundaries and enables sequential growth of multi-junction heterostructures. For example, a WS2 − WS2(1 − x) Se2x monolayer planar heterostructure with tunable band alignment was successfully grown by modulating the gas flow direction and temperature. Similarly, a one-pot synthesis strategy was developed by placing two precursor powders (MoX2 and WX2) in the same boat, with different gas flows promoting the selective growth of MoX2 and WX2, respectively [26]. This multi-step approach, compared to one-step or two-step CVD methods, offers more precise control, enabling the creation of spatially distinct optoelectronic devices by separating electrons and holes into different materials. Dong Hee Shin et al. fabricated MoS2/SQD:SiO2 photodiodes by direct chemical vapor deposition (CVD) of MoS2 onto SiO2-embedded silicon quantum dot (SQD) multilayers. The SQD:SiO2 multilayers were first prepared by reactive ion beam sputtering, followed by high-temperature annealing at 1100°C in a nitrogen atmosphere to form well-embedded quantum dots. High-resolution transmission electron microscopy (HRTEM) and photoluminescence (PL) confirmed the presence of SQDs. For MoS2 deposition, SQD: SiO2 substrates were placed downstream of MoO3 powder, and hydrogen sulfide (H2S) was introduced to synthesize single- and multilayer MoS2 films at 600°C. The MoO3 density was controlled to regulate the film thickness, which was analyzed using Raman spectroscopy. Atomic bonding states of MoS2 were characterized using X-ray photoelectron spectroscopy (XPS), while its optical absorption was evaluated using a UV-Vis-NIR spectrophotometer. To enhance photodetector (PD) performance, a bathocuproine (BCP) layer was spin-coated onto the Si surface, serving as an electron-selective contact to minimize leakage current. Finally, Au and InGa electrodes were deposited using photolithography and a lift-off process. This fabrication method ensures high-quality MoS2 growth with minimized defects at the MoS2/SQD:SiO2 interface, leading to enhanced carrier transport. The resulting photodetector achieves exceptional performance, including a high external quantum efficiency (EQE) of 92%, a detectivity (D*) of 6.1 × 1013 Jones, and an ultrafast response time (60 ns rise/756 ns fall), making it one of the fastest MoS2-based PDs reported. The integration of SQDs further boosts light absorption and carrier extraction, making this fabrication approach highly effective for high-speed and broadband photodetection Figure 4 [27].

Figure 4.
(a). A schematic of a typical MoS2/SQDs:SiO2 MLs/n-Si PD with Au and InGa as top and bottom electrodes, respectively. (b)Dark/photo J-V curves at λ = 500 nm in the range of bias from 5 to 5 V. (1) (c) Schematic of the device indicating the direction of light illumination. (d) Dark and illuminated I-V characteristics of the MoS2/GaN/Si-based device (inset shows the I-V characteristics near zero bias) reproduced with permission from [27].
The MoS2/GaN/Si heterostructure photodetector (PD) was fabricated by Deependra Kumar Singh et al. to achieve self-powered, broadband, and spectrally distinctive photodetection. A 250 nm GaN thin film was first grown on Si(111) substrates using plasma-assisted molecular beam epitaxy (PAMBE). The GaN layer plays a crucial role in forming a heterojunction with MoS2, enabling efficient carrier transport and built-in potential formation. Following GaN deposition, MoS2 thin films were grown using pulsed laser deposition (PLD). An MoS2 pellet (2 cm diameter) served as the target, positioned 4.5 cm from the substrate. The deposition was performed at a base pressure of 5.0 × 10−6 mbar, using a pulsed laser source with a repetition rate of 3 Hz, 1000 laser pulses, and a laser energy of 200 mJ, resulting in a laser fluence of 1 J/cm2. During deposition, the chamber pressure was maintained at 1.5 × 10−5 mbar. The sample underwent in situ annealing at 700°C for 20 minutes, followed by slow cooling at a ramp rate of 5°C/min to enhance film crystallinity. This fabrication approach enables the realization of a high-performance photodetector with a broadband response (300–1100 nm) and an ultrahigh responsivity of 23.81 A/W at 995 nm under low-intensity illumination (0.075 mW/cm2) in a self-powered mode. The device exhibits a unique feature of photocurrent polarity inversion in the NIR region due to a competitive mechanism between the photothermal electric (PTE) effect in MoS2 and the built-in potential at the MoS2/GaN/Si heterojunction. The demonstrated wavelength-dependent polarity switching opens new possibilities for low-power optoelectronic applications, making this PD one of the most advanced MoS2-based devices for future filter-less and energy-efficient photodetection technologies [28].
Minkyun Son et al. fabricated MoS2/HfS2 heterojunction photodetector. Firstly, MoS2 was synthesized via a two-step CVD process in a 2-inch tubular furnace. A SiO2/p-Si substrate was pre-deposited with MoO3 and NaCl using thermal evaporation, then placed in a crucible at the center of the furnace, while 4 g of sulfur was positioned 20 cm upstream. The system was heated at 30°C/min under 200 sccm Ar flow at 760 Torr, held at 150°C for 15 min, and then heated at 25°C/min to 500°C. At this point, sulfur vaporization began, and when the chamber temperature reached 750°C, the sulfur heating zone reached 200°C, ensuring uniform MoS2 growth. The temperature was maintained for 20 min before natural cooling. The grown MoS2 exhibited large flake sizes (∼200 μm) with 86% coverage, providing a high-quality platform for heterostructure fabrication. The HfS2/MoS2 vertical heterojunction was synthesized in a 1-inch tubular furnace with MoS2/SiO2 placed at the center and a mixture of HfCl4 (0.02 g) and NaCl (0.01 g) positioned further away in a crucible. Upstream, 4 g of sulfur was kept 15 cm from MoS2, with H2 (15 sccm) and Ar (20 sccm) as carrier gases. The system was heated at 30°C/min to 150°C (held for 15 min), followed by a faster ramp at 40°C/min to 500°C, after which the sulfur heating zone was activated. When the main heater reached 950°C, sulfur reached 150°C, maintaining these conditions for 30 min before natural cooling. This controlled synthesis enabled the selective growth of HfS2 on MoS2 due to strong interfacial interactions, forming a vertical heterojunction with a large interface area, enhancing interlayer exciton generation. The conduction band edge of HfS2 (5.2 eV) resulted in a narrow interlayer bandgap, facilitating IR detection. The built-in potential at the interface suppressed dark current and created gradient band bending, ensuring efficient carrier drift. The resulting HfS2/MoS2 photodetector demonstrated exceptional performance, achieving a photoresponsivity of ≈600 A/W and detectivity (D*) of ≈7 × 1013 Jones at 1550 nm and ≈1500 A/W with D* ≈ 2 × 1014 Jones at 980 nm. The device exhibited rapid response times (rise/decay: 60/71 μs), surpassing conventional MoS2-based detectors. This interlayer exciton-driven IR photodetector, fabricated via a scalable two-step CVD process, marks a significant advancement in high-performance 2D optoelectronic devices Figure 5 [30].

Figure 5.
(a). Schematic of overall two-step HfS2/MoS2 synthesis procedures. The inset image is of synthesized MoS2 and the scale bar denotes 200 μm. (b) Schematic of HfS2/MoS2 photodetector. (c) IDS versus VDS of the HfS2/MoS2 photodetector in the dark and at a different P (λ = 532 nm). (3) (d) I-V characteristics of fabricated heterojunction, using 2 nm MoS2 QDs on Si, under dark and illumination conditions. (e) J-V characteristics of fabricated p-n heterojunctions under dark conditions, at room temperature. The inset shows the logarithmic J-V curve for the heterojunction fitted with a power law J .Vm corresponding to the trap-charge-limited current transport. Reproduced with permission from Ref. [29].
Subhrajit Mukherjee et al. synthesized MoS2 quantum dots (QDs) via solvent-assisted sonochemical exfoliation, modified to ensure uniform film deposition on Si for vertical Si/MoS2 p-n heterojunction fabrication. Since high-boiling DMF caused QD aggregation, prolonged sonication was followed by 2 days of vigorous stirring for size uniformity. Gradual centrifugation yielded QD precipitates of different sizes, which were dried under Ar at ∼60°C and redispersed in ethanol for spin-coating onto H-passivated p-type Si (0.77 Ω-cm). Device fabrication included Au (∼70 nm) as the top electrode and Al (∼80 nm) on the Si backside via thermal evaporation (base pressure ∼ 1 × 10−6 Torr). For LED fabrication, a ∼ 20 nm Al-doped ZnO (AZO) film was deposited via pulsed laser deposition (PLD) at ∼300°C (base pressure ∼ 5 × 10−6 Torr, energy density ∼ 2 J/cm2, repetition rate ∼ 5 Hz), exhibiting a resistivity of ∼10−3 Ω·cm and ∼ 92% transmittance. MoS2 QDs exhibited size-dependent emission (2–26 nm) with nearly stoichiometric 2H-phase and n-type doping due to Cl, confirmed via XPS. Time-resolved PL showed superior carrier lifetimes (1.5–2.5 ns) compared to MoS2 flakes (∼10 ps). The 0D/3D heterostructures demonstrated rectifying characteristics (ideality factor ∼ 9.0) and bias-dependent electroluminescence (450–800 nm), indicating potential for LED applications. Spectral responsivity increased with decreasing QD size, reaching 0.85 A/W and a peak detectivity of ∼8 × 1011 Jones at −2 V for ∼2 nm QDs, outperforming commercial Si homojunctions and graphene/Si devices. The study highlights the potential of colloidal n-MoS2 QDs for Si-compatible, large-area multifunctional optoelectronic devices using 0D/3D heterojunctions [29].
3. MoS2 heterostructure for near IR detection
Our recent work has employed metal passivation by Zn and Fe on MoS2 to achieve high-performance in the NIR regime and low dark current. As seen in Figure 6d, the synthesized pure MoS2 (MoS), Zn-MoS, and Fe-MoS are tested in both light and dark conditions. The thermally persuaded charge carriers, also known as leakage current, are typically responsible for the dark current (Idark), which is measured at 1.5x10−6 A, 3.98x10−6 A, and 2.81x10−6 A for MoS, Zn-MoS, and Fe-MoS devices, respectively. Ag (metal electrode)-MoS2 (semiconductor) photodetectors with leakage current (Ileak) exhibit Schottky behavior in dark conditions, as demonstrated by the obtained I-V curve. The carrier injection, which regulates the flow of carriers into semiconductors, is directly impacted by the metal-semiconductor interface. Surface passivation of the (Zn and Fe)-MoS2 surface, which reduces the surface states, could counteract the fermi level pinning (FLP) effect that would deteriorate the carrier injection through the metal-semiconductor (M-S) interface. Band bending (uplift/downshift) is facilitated by the gap states formed by the metal contact between the semiconductor energy bands or by crystal defects or distortion that induce the surface states. Additionally, in order to prevent leakage, the FLP effect is suppressed by altering the surface of MoS2 with Fe and Zn, which passivates the dangling bonds on the surface and allows for the establishment of a downward shift in fermi energy. Therefore, as illustrated in Figure 6d, the enhanced photoresponse of the devices is indicated by the increased current (Iillum) of 1.89x10−6 A, 9.77x10−6 A, and 6.25x10−6 A for MoS, Zn-MoS, and Fe-MoS under the light illumination. Iph = Idark - Iillum is used to calculate the photocurrent (Iph) at a bias voltage of 3 V. For MoS, Zn-MoS, and Fe-MoS, the enhanced Iph values are 0.3x10−6 A, 5.79x10−6 A, and 3.44x10−6 A, respectively. The trapping mechanism can be clearly understood from the temporal photoresponse measurement. Figure 6a shows non-exponential behavior in the light ON–OFF condition in the Illum -t curve, indicating. The trapping of carriers induced photocurrent. For 10 seconds, the MoS first cycle maintains a dark steady-state current of 1.59x10−6 A. Light is illuminated in such a situation, and an increase in current (1.89x10−6 A) is observed with a similar steady state for 10 s. The behavior returns with a similar trend as seen in the previous cycle when the dark and light conditions are switched. Comparably, the dark current for (Zn and Fe)-MoS is 3.98x10−6 and 2.81x10−6 A, respectively, and it abruptly rises to 9.77x10−6 and 6.25x10−6 A when illuminated. Figure 6 illustrates the response’s improved stability across various cycles (a, b, and c). For MoS, Zn-MoS, and Fe-MoS, the computed rise time (τ

Figure 6.
Temporal photocurrent measurement for pure MoS2 (a) and metal-passivated MoS2 photodetectors with (b) Fe, (c) Zn, and (d) photocurrent of pure, Zn and Fe passivated MoS2 photodetectors [31].

Figure 7.
(a) Band alignment of graphene/MoS2 heterostructure panel; (b) responsivity as a function of the wavelength of the incident light at a bias of -2 V for IR photodiode; (c) responsivity and the specific detectivity variation for different wavelengths of the incident illumination; and (d) transient frame of source-drain current under illumination with various wavelengths [32].
Long et al. introduced an advanced shortwave infrared (SWIR) photodiode based on a MoS2/graphene/WSe2 van der Waals (vdW) heterojunction, in which a monolayer of graphene was sandwiched between a bottom p-type MoS2 layer and a top n-type WSe2 layer, all deposited onto a SiO2/Si substrate. This unique layered architecture enabled the photodiode to leverage the distinct optical and electronic properties of each material, leading to enhanced performance in terms of broadband absorption and carrier transport. One of the primary advantages of this device configuration was the inclusion of a graphene interlayer, which contributed significantly to broadband light absorption [33]. Unlike conventional semiconductors with a fixed bandgap, graphene, being a zero-bandgap material, possesses a continuous density of electronic states that allows it to absorb a wide spectrum of incident light, from visible to infrared wavelengths. This characteristic was instrumental in extending the photodetection range of the device far beyond that of MoS2 or WSe2 alone. The incorporation of MoS2 (p-type) and WSe2 (n-type) formed a well-defined p–n junction, which played a crucial role in the efficient separation and transport of photogenerated charge carriers. Upon illumination, electron-hole pairs were generated at different regions of the heterostructure. The built-in electric field at the MoS2/WSe2 junction facilitated the rapid separation of these charge carriers, reducing recombination losses and improving the overall quantum efficiency of the photodetector.
The combination of broadband absorption in graphene and the efficient charge separation mechanism in the p–n heterojunction allowed the device to achieve a remarkably wide spectral response. The photodetector exhibited sensitivity across a broad wavelength range, spanning from 0.4 μm (visible) to 2.4 μm (SWIR), as illustrated in Figure 7c. Furthermore, the synergistic effects of the vdW heterojunction led to an improvement in specific detectivity, enabling the device to operate with high sensitivity even in low-light conditions. The MoS2/graphene/WSe2 vdW heterojunction photodiode demonstrated a significant advancement in optoelectronic device engineering by integrating the unique properties of 2D materials to achieve high-performance broadband photodetection. Its ability to detect wavelengths from the visible to the SWIR region, coupled with high carrier mobility and efficient charge separation, makes it highly suitable for applications in optical sensing, imaging, and telecommunications. More notably, the MoS2/graphene/WSe2 vdW heterojunction photodiode exhibited an impressive responsivity of approximately 2.64 × 103 A/W at a forward bias of 1 V when illuminated with 1.064 μm wavelength light. This high responsivity was accompanied by a remarkable specific detectivity of 1.1 × 1013 Jones and an external quantum efficiency (EQE) reaching nearly 104. An EQE exceeding unity indicates the presence of a photoconductive gain in the device, which can be attributed to several key mechanisms.
First, the short transit distance in the Vertical Junction: The vertical charge transport between MoS2, graphene, and WSe2 allowed photogenerated carriers to traverse the junction efficiently, reducing recombination losses. Then, the interlayer inelastic tunneling effect is the presence of graphene as an intermediate layer that facilitated charge transfer across the heterojunction through inelastic tunneling, further enhancing carrier collection efficiency, and finally, the rapid carrier transport, which is the high carrier mobility within the vdW heterostructure, enabled efficient charge separation and extraction, contributing to the significant photoconductive gain. Additionally, the photodetector demonstrated a fast response time across various wavelengths. For instance, at 1.064 μm, the device exhibited a rise time of 24.1 ms, as shown in Figure 7d. This rapid response, combined with its high responsivity and detectivity, underscores the potential of the MoS2/graphene/WSe2 heterojunction photodiode for high-performance optoelectronic applications, including infrared imaging and optical communication [36]. Long et al. demonstrated the fabrication of a mid-wave infrared (MWIR) photodiode by utilizing the narrow optical bandgap and the ease of heterojunction formation in two-dimensional black arsenic-phosphorus (b-AsP), specifically in the form of AsxP1−x. The resulting b-AsP/MoS2 stacked van der Waals (vdW) heterojunction photodiode exhibited outstanding MWIR photodetection capabilities. Under broadband illumination in the 3–5 μm range, the device achieved an exceptional specific detectivity exceeding 4.9 × 109 Jones at zero bias and room temperature. This performance significantly surpassed that of certain commercially available infrared photodetectors, including PbSe-based and thermistor-based IR photodetectors, as illustrated in Figure 8a. In conclusion, the b-AsP/MoS2 vdW heterojunction photodiode demonstrated superior MWIR photodetection performance, achieving a high specific detectivity of 4.9 × 109 Jones at zero bias and room temperature. Its broadband response in the 3–5 μm range outperformed conventional IR photodetectors, highlighting the potential of 2D materials for next-generation infrared sensing applications. The exceptional specific detectivity of the b-AsP/MoS2 vdW heterojunction photodiode was primarily attributed to its suppressed dark current, resulting from the potential barrier at the p-type b-AsxP1−x and n-type MoS2 heterojunction, as illustrated in Figure 8(b) and (c). This barrier effectively reduced noise, leading to a two-order magnitude improvement in detectivity compared to a bare b-AsxP1−x photoconductor. In addition to its high detectivity, the photodiode exhibited an ultrafast response time of approximately 0.5 ms at 4 μm, as shown in Figure 8b [37]. Several black phosphorus (bP)-based vdW heterojunction photodiodes, such as the MoS2/bP heterojunction for 1.55 μm detection, have also shown excellent performance in the shortwave infrared (SWIR) region in addition to MWIR detection [26], the WSe2/bP/MoS2 sandwich-like heterojunction for 1.55 μm [38], and the PdSe2/bP heterojunction for 1.31 μm detection [35]. These results highlight the versatility of bP-based vdW heterojunctions for broadband infrared photodetection. 2D layered materials-based heterojunctions and CNT-based heterojunctions were fabricated, and both of their properties were compared, from which it is proven that the problems of organic NIR detectors and CNT-based NIR detectors, such as environmental instability and low sensitivity, are said to be overcome with the help of layered materials-based heterojunctions. The graphene/MoS2 heterojunction is fabricated using a two-step process in which a single layer of graphene is grown on the Cu foil by the chemical vapor deposition method, and then it is transferred to the SiO2/Si substrate by patterning it as an 800 nm nanoribbon using electron beam lithography, and then MoS2 flakes are exfoliated and transferred onto the graphene channel by means of PMMA-mediated transfer technique. Finally, source and drain electrodes were also constructed using electron beam lithography. Further, the WeSe2/MoS2 heterojunction is fabricated by means of mechanical exfoliation and micromanipulation methods. WSe2 was transferred to the SiO2/Si substrate after getting exfoliated from the crystal. MoS2 flake was placed on WSe2 by means of a micromanipulation method with the aid of polydimethylsiloxane film and a 3-axis manipulator. Finally, Au/Cr and Pd electrodes were constructed on MoS2 and WSe2 flakes, respectively, using electron beam lithography and lift-off techniques (Figure 9a and b). Both the heterojunctions show the diode rectifying behavior (forward bias) in the dark, and the photocurrent was increased when the heterojunctions were illuminated with NIR light (785 cw laser), which shows its response in the NIR region. (Figure 9c and d) The inset figure represents the on/off current ratio at V = ±1 V, which was about 10 and 200, respectively. In particular, WSe2/MoS2 heterojunction exhibited photovoltaic effect. The photocurrent density is measured at V = - 0.5 V for both the heterojunctions by using the laser diode of wavelength 808 nm which is turned on and off repeatedly (Figure 9e and f). The photocurrent rise and decay times were shorter than 1 s, which makes it better than CNT/ZnO heterojunction with the rise and fall times of 90 and 70 s, respectively [35].

Figure 8.
(a) Specific detectivity as a function of the wavelength of incident light for b-AsP/MoS2 photodiode; (b) time transient response behavior of 4 μm illumination; (c) current noise and frequency for the IR photodetector based on the heterojunction [34]; and (d) band diagram of the MoTe2/MoS2 heterostructure photodiode [34].

Figure 9.
Field emission scanning electron microscope images of the graphene/MoS2 (a), WSe2/MoS2 (b), graphene/MoS2 (c), and WSe2/MoS2 (d) I-V curves measured across the junction under NIR light and in the dark are displayed in the insets on semi-logarithmic scales. When a laser diode (λ = 808 nm) with a power of 5.3 mW mm−2 was turned on and off repeatedly for the graphene/MoS2 (e), WSe2/MoS2, the temporal photoresponse was measured at V = - 0.5 V (f) [35].
4. Summary and outlook
2D material-based infrared (IR) photodetectors have emerged as promising candidates for next-generation optoelectronic devices due to their exceptional properties, including tunable bandgaps, high carrier mobility, and strong light-matter interactions. These photodetectors can be classified into three main categories based on device configurations: photoconductors, photodiodes, and nanophotonic-integrated photodetectors. Photoconductors, particularly graphene-based ones, offer high-speed performance, while photodiodes exhibit improved photoresponsivity and noise reduction due to p–n junctions. Additionally, integrating optical nanostructures enhances light absorption, leading to higher responsivity and detectivity. Despite significant advancements, several challenges include limited spectral coverage: most 2D material-based IR photodetectors operate in the SWIR and MWIR ranges, with only a few materials extending to LWIR/FIR, crucial for thermal imaging and biosensing. The Incomplete performance characterization in various photodetector evaluations focus on single-wavelength illumination, lacking full spectral response data needed for optimization. In addition to this, the real-world applicability concerns rely on laser illumination, making practical deployment uncertain. Using blackbody illumination would provide a more realistic assessment. Finally, the scalability issues in current fabrication methods, primarily mechanical exfoliation, hinder large-area and uniform production, limiting commercial viability. To advance 2D material-based IR photodetectors, the area that requires further exploration is bandgap engineering through 2D material alloys—adjusting material composition (e.g., InxGa1 − xP alloys) could enable LWIR to FIR detection. Further, developing CVD growth methods for high-quality, large-area 2D materials tailored for IR photodetectors and comprehensive performance evaluation, i.e., systematic characterization across multiple wavelengths to optimize spectral response. Remarkably, integration with optical nanostructures unambiguously enhances light absorption via waveguides, cavities, plasmonic structures, and large-scale detector arrays, which address integration challenges for high-resolution imaging applications requiring CMOS-compatible processing. By overcoming these challenges, 2D material-based IR photodetectors could revolutionize imaging, sensing, and optoelectronic technologies, paving the way for high-performance, scalable, and broadband photodetection systems.
References
- 1.
Chee KWA. Introductory chapter: Photodetectors. In: Advances in Photodetectors. London, UK: IntechOpen; 2019 - 2.
Rogalski A. History of infrared detectors. Opto-Electronics Review. 2012; 20 :279-308 - 3.
Yajnik UA. The conception of photons - Part I. Resonance. 2015; 20 :1085-1110 - 4.
Donati S. Photodetectors: Devices, Circuits and Applications. 2nd ed. IEEE Press Wiley; 2021 - 5.
Romoli M et al. The ultraviolet and visible-light coronagraph of the HERSCHEL experiment. AIP Conf. Proc. 2003; 679 :846-849 - 6.
Wang F et al. 2D library beyond graphene and transition metal dichalcogenides: A focus on photodetection. Chemical Society Reviews. 2018; 47 :6296-6341 - 7.
Long M et al. Progress, challenges, and opportunities for 2D material-based photodetectors. Advanced Functional Materials. 2018; 28 :1803807 - 8.
Xie C et al. Photodetectors based on two-dimensional layered materials beyond graphene. Advanced Functional Materials. 2016; 27 :1603886 - 9.
Wei Y et al. Recent advances in photodetectors based on two-dimensional material/Si heterojunctions. Applied Sciences. 2023; 13 (19):11037 - 10.
Yao J et al. 2D material broadband photodetectors. Nanoscale. 2020; 12 :454-476 - 11.
Wang J et al. Design strategies for two-dimensional material photodetectors to enhance device performance. InfoMat. 2019; 1 :251-290 - 12.
Wadhwa R et al. A strategic review of recent progress, prospects and challenges of MoS2-based photodetectors. Journal of Physics D: Applied Physics. 2022; 55 :063002 - 13.
Cheng Y et al. Research progress on improving the performance of MoS2 photodetector. Journal of Optics. 2022; 24 :105001 - 14.
Bernardi M et al. Optical and electronic properties of two-dimensional layered materials. Nano. 2017; 6 (2):479-493 - 15.
Li H et al. From bulk to monolayer MoS2: Evolution of Raman scattering. Advanced Functional Materials. 2012; 22 :1385-1390 - 16.
Tian H et al. Optoelectronic devices based on two-dimensional transition metal dichalcogenides. Nano Research. 2016; 9 :1543-1560 - 17.
Ahmad S et al. A comparative study of electronic properties of bulk MoS2 and its monolayer using DFT technique: Application of mechanical strain on MoS2 monolayer. Graphene. 2014; 3 (4):73-81 - 18.
Kwak JY et al. Absorption coefficient estimation of thin MoS2 film using attenuation of silicon substrate Raman signal. Results in Physics. 2019; 13 :102202 - 19.
Amani M et al. Near-unity photoluminescence quantum yield in MoS2. Science. 2015; 350 (6264):1065-1068 - 20.
Britnell L et al. Strong light-matter interactions in heterostructures of atomically thin films. Science. 2013; 340 (6138):1311-1314 - 21.
Chaves A et al. Bandgap engineering of two-dimensional semiconductor materials. npj 2D Materials and Applications. 2020; 4 :29 - 22.
Dean CR, Young AF, Meric I, Lee C, Wang L, et al. Boron nitride substrates for high-quality graphene electronics. Nature Nanotechnology. 2010; 5 (10):722-726 - 23.
Castellanos-Gomez A, Buscema M, Molenaar R, Singh V, Janssen L, van der Zant HSJ, et al. Deterministic transfer of two-dimensional materials by all-dry viscoelastic stamping. 2D Materials. 2014; 1 :011002 - 24.
Gong YJ, Lin JH, Wang XL, Shi G, Lei SD, et al. Vertical and in-plane heterostructures from WS2/MoS2 monolayers. Nature Materials. 2014; 13 (12):1135-1142 - 25.
Fu L, Sun YY, Wu N, Mendes RG, Chen LF, et al. Direct growth of MoS2/h-BN heterostructures via a sulfide-resistant alloy. ACS Nano. 2016; 10 (2):2063-2070 - 26.
Zheng BY, Ma C, Li D, Lan JY, Zhang Z, et al. Band alignment engineering in two-dimensional lateral heterostructures. Journal of the American Chemical Society. 2018; 140 (36):11193-11197 - 27.
Shin DH et al. High-speed heterojunction photodiodes made of single- or multiple-layer MoS2 directly-grown on Si quantum dots. Journal of Alloys and Compounds. 2020; 820 :153074 - 28.
KumarSingh D et al. Differentiation of ultraviolet/visible photons from near infrared photons by MoS2/GaN/Si-based photodetector. Applied Physics Letters. 2021; 119 :121102 - 29.
Mukherjee S et al. Novel colloidal MoS2 quantum dot heterojunctions on silicon platforms for multifunctional optoelectronic devices. Scientific Reports. 2016; 6 :29016 - 30.
Son M et al. High-performance infrared photodetectors driven by interlayer exciton in a Van Der Waals epitaxy grown HfS2/MoS2 vertical heterojunction. Advanced Optical Materials. 2023; 34 :2308906 - 31.
Abinaya R et al. Modulating Fermi energy in few-layer MoS2 via metal passivation with enhanced detectivity for near IR photodetector. Journal of Materials Chemistry C. 2024; 12 :5247-5254 - 32.
Vabbina P et al. Highly sensitive wide bandwidth photodetector based on internal photoemission in CVD grown p-type MoS2/graphene Schottky junction. ACS Applied Materials & Interfaces. 2015; 7 :15206-15212 - 33.
Long M et al. Broadband photovoltaic detectors based on an atomically thin heterostructure. Nano Letters. 2016; 16 (4):2254-2259 - 34.
Long M et al. Room temperature high-detectivity mid-infrared photodetectors based on black arsenic phosphorus. Science Advances. 2017; 3 (6):e1700589 - 35.
Park MJ et al. Near-infrared photodetectors utilizing MoS2-based heterojunctions. Journal of Applied Physics. 2015; 118 :044504-044510 - 36.
Ye L et al. Near-infrared photodetector based on MoS2/black phosphorus heterojunction. ACS Photonics. 2016; 3 (4):692-699 - 37.
Li H et al. High-performance broadband floating-base bipolar phototransistor based on WSe2/BP/MoS2 heterostructure. ACS Photonics. 2017; 4 (4):823-829 - 38.
Afzal AM et al. High-performance p-BP/n-PdSe2 near-infrared photodiodes with a fast and gate-tunable photoresponse. ACS Applied Materials & Interfaces. 2020; 12 (17):19625-19634