Open Access
4 January 2023 Graphene/molybdenum disulfide nanocomposites: characterization and optoelectronic application
Nagham M. Obaid, Amer Al-Nafiey, Ghaleb Al-Dahash
Author Affiliations +
Abstract

The ability to modify and shape the surface of polymer and composite materials is crucial for a number of biological and electronics applications. Molybdenum disulfide (MoS2) and graphene are two-dimensional materials that have distinctive electrical and optical properties that are useful in many optoelectronic applications. The latest applications of graphene / MoS2, as well as heterostructure manufacturing, properties, and applications, are discussed. Heterostructured materials, as opposed to single-component materials, are designed to provide additional functionality or flexibility. Our study focuses on their unique traits and capabilities, as well as applications, notably in the field of photodetector technology.

1.

Introduction

The chemical type and structure of a material’s components are recognized to have a significant impact on its properties.1 Bulk materials that contain a large number of atoms, in particular, can be identified by the energy bands that are responsible for most of the physical and chemical properties of solid materials due to the overlapping of atomic and molecular orbitals. As for nanomaterials with dimensions ranging from 10 to 100 nm, the number of atoms is so small that the electronic energy bands change dramatically, affecting all physical properties of materials.24

Silicon and germanium structures are nanostructured semiconductors that emit visible light luminescence despite having modest and indirect band gaps in their bulk state.4 The PL signals appear to be dependent on the size of the nanostructures, implying that the quantum confinement process is involved.4,5

On the other hand, metals with nanostructures are known to have a variety of unique characteristics. Metallic nanoparticles (NPs) show marked differences in color when their size and shape differ, as well as changes in transmitted and reflected spectra.6 These color effects are linked to individual metal NPs’ efficient resonant absorption of light and the stimulation of localized surface plasmons, which are associated with metal electron collective oscillations.6

This review will look at the many ways of producing nanomaterials based on MOS2 and graphene for device manufacturing and optoelectronics applications, such as mechanical exfoliation, liquid exfoliation, and chemical vapor deposition (CVD).

2.

Nanocomposite

Metallic NPs implanted in a dielectric multitude are significantly influenced by their physical and chemical characteristics, such as their distribution within the matrix and attributes such size, shape, or size dispersion. As a result, controlling these characteristics during manufacturing has long been a difficulty, and one of the biggest obstacles to the development of useful applications based on these materials has been the lack of production techniques with the required level of control.7 This review will focus on the following nanomaterials.

2.1.

Molybdenum Disulfide

Among 2D materials, the TMD family has received the most attention.8 TMDs are hexagonal in nature and absorb visible- to infrared-ranged light.9 They are promising semiconducting materials with extremely thin bodies that enhance carrier confinement and electrostatic gate control as a result of programmable bandgaps and the absence of surface dangling bonds.8 This study focuses on molybdenum disulfide (MoS2), a well-known and thoroughly investigated TMD material with Mo atoms sandwiched between two layers of S atoms that has environmental stability and natural availability (Fig. 1).8,10

Fig. 1

MOS2 structure.8

JNP_17_1_010901_f001.png

Moreover, it appears in nature as an n-type material,11 with visible to near-infrared light absorption,9 a large work function of 5.1 eV,12 and mechanical stability and transparency for flexible devices.8

The most attractive feature of MOS2 is the variable energy gap from 1.2 to 1.8 eV as the material moves from the bulk indirect bandgap to the direct bandgap monolayer due to the quantum confinement effect.

Furthermore, MoS2 has optical properties that are thickness dependent.1315 The number of layers in MoS2 influences its electron affinity. The monolayer MoS2 had an electron affinity of 4.0 eV,16 three layers of MoS2 had an affinity of 4.0 eV,17 and multilayers (28 nm) had an affinity of 4.1 eV.18 Thus the MoS2, with a monolayer thickness of 0.65 nm, has more than 80% of its transparency, making it ideal for wearable applications.19 The MoS2’s physical properties are given in Table 1.

Table 1

Physical properties of the MoS2.

PropertyValue
Charge carrier mobilityUp to 200  cm2/V·s20
Molar mass160.07 g/mol21
Band gap1.8 eV (direct)22
Specific density5.06  g/cm321
Melting point2648 K23

As mentioned, the contact Schottky barrier height, contact resistivity, and carrier mobility are all impacted by thickness.8 Depending on the application, the MoS2 layer’s thickness needs to be carefully adjusted for maximum performance. MoS2 nanofiber bundles were made by the hydrothermal technique in 2007 by Nagaraju et al.,24 using ammonium molybdate in an acidic medium and maintained at 180°C for several hours. The sample’s PXRD pattern may be easily identified as hexagonal 2H-MoS2. According to the MoS2 FTIR spectra, the band at 480  cm1 is called (Mo-S). MoS2 nanofibers are found in bundles that are 20 to 25 m long and 120 to 300 nm in diameter.

According to Li et al.,25 advanced materials for photoelectrochemical water splitting are essential for the field of renewable energy. This study demonstrated selective solvothermal synthesis of MoS2 NPs on sheets of reduced graphene oxide (rGO) floating in solution. The resulting MoS2/rGO hybrid material, in contrast to the enormous aggregated MoS2 particles grown freely in solution without GO, had nanoscopic few-layer MoS2 structures with a lot of exposed edges piled on graphene. The MoS2/rGO hybrid demonstrated greater electrocatalytic activity in the hydrogen evolution process when compared with conventional MoS2 catalysts [hydrogen evolution reaction (HER)].

The study by Helmly et al.26 investigated the production and characterization of molybdenum (IV) chalcogenide NPs. The process involves the interaction of molybdenum hexacarbonyl with sulfur or selenium. The NPs absorb light in the UV area of the spectrum, whereas bulk molybdenum (IV) chalcogenides absorb light in the near-IR. When stimulated in the UV range, the particles also emit a blue fluorescence.

Also Li et al.27 used temporally structured femtosecond laser ablation of bulk MoS2 targets in water to create monolayer MoS2 QDs. The findings of the investigation reveal that very pure, homogeneous, and monolayer MoS2 QDs may be produced effectively.

A combination of multilayer photoexfoliation of monolayer MoS2 and water photoionization enhanced light absorption is the underlying process. Because of the many active edge sites, high specific surface area, and superior electrical conductivity, the as-prepared MoS2QDs have remarkable electrocatalytic activity for hydrogen evolution processes. The second important material for postsilicon electronics that are used in optoelectronics applications is MoS2.27

MoS2 NPs were generated using the laser ablation technique (Nd:YAG=1064  nm) in the polymer ethylene glycol solution as a function of laser fluence, according to Moniri and Hantehzadeh.28 The resulting NPs are hexagonal crystals with diameters ranging from 4 to 160 nm, according to the findings. Furthermore, according to the laser fluence, the majority of them had a maximum wavelength that ranged from 212.5 to 216.5 nm. As the laser flux increases, the size of these particles will increase from 10 to 18 nm.

2.2.

Graphene

Graphene, as shown in Fig. 2, is the basic structure of all forms of graphite that is now used in most modern fields in the last decade. As a result of its unusual physical and chemical features, it has piqued the interest of researchers all over the world, proving its theoretical significance.30 The physical characteristics of graphene are provided in Table 2.

Fig. 2

Structure of graphene.29

JNP_17_1_010901_f002.png

Table 2

Physical properties of graphene.31

PropertyValue
Charge carrier mobility200,000  cm2/V·s
Thermal conductivity(3to5)  103  W/m·K
Transparency97.4%
Specific surface area2630  m2/g
Young’s modulus1  TPa
Tensile strength1100  GPa
Band gap0
Specific density2.26  g/cm3
Resistivity2.26  μΩ·cm
Thermal stability450°C to 650°C
Melting point3800 K
Current density>108  A/cm2

3.

Synthesis Methods for 2D Materials

To obtain high performance, two-dimensional materials must be synthesized with large dimensions, adjustable thicknesses, and few crystal defects.32 Mechanical exfoliation, liquid exfoliation, CVD, laser ablation, and pulsed-laser deposition (PLD) have all been used to make 2D materials. A quick description of how these methods work is given below.

3.1.

Mechanical Exfoliation Technique

The mechanical exfoliation technique is a method that involves extending layered materials by outside forces (such as chemical or physical forces) to lessen interlayer noncovalent interactions connections, allowing them to gradually separate into a single structure of very few layers.33 The first synthesis method for making graphene was the method of mechanical exfoliation, known as the adhesive tape method; it has since been successfully used for multilayer 2D materials. The few-layer MoS2 nanosheets were generated by Li et al.34 utilizing an exfoliation technique, as shown in Fig. 3(a). A second sticky side of the tape was applied to the scotch tape with crystal after the first one was pushed onto the crystal surface for about 5 s. The scotch tape is then carefully removed at the tiniest possible angles. The two tapes are then torn, and crystals of the same size are glued to each. When this procedure is repeated, the bulk MoS2 flakes become thinner and thinner until they are reduced to one or a few layers. The MoS2 layers on tape were next successfully adhered to a spotless Al2O3 ceramic substrate. Before removing the scotch tape, the adhesion state was maintained for 6 h. After that, acetone was poured over the substrate to eliminate the scotch tape’s adhesive residue. It is a simple, inexpensive, and straightforward technology for creating high-quality 2D materials, but it has limitations due to the size of the created material being <100  m and the inability to change the layer thickness. As a result, it is unsuitable for materials research and device demonstration at the industrial level in pioneering work.

Fig. 3

(a) Diagram of the mechanically exfoliated MoS2 nanosheets,30 (b) a schematic of the liquid exfoliation process’s mechanism, and (c) diagram of the CVD method used to create MoSe2 crystals on molten glass.35

JNP_17_1_010901_f003.png

3.2.

Liquid Exfoliation Technique

The liquid exfoliation method is an inexpensive method. Hernandez36 used sonication to exfoliate graphite in an organic solution containing N-methyl pyrrolidone. The graphene dispersion was obtained at a concentration of 0.01  mg/ml, and a monolayer graphene yield was 1 wt.%. Various liquid exfoliation procedures, such as high shear mixing,35 microoxidization,37 and jet cavitation,38 have been documented. As illustrated in Fig. 3(b), sonication-assisted liquid exfoliation is widely used in synthesis. Nanosheets are created simply by dissolving the beginning bulk material in a solvent to disperse the two-dimensional components in the solution, and ultrasound or centrifugation is used.

3.3.

Chemical Vapor Depositions Technique

Using active chemical reactions of precursors in a designed environment, CVD is used to create 2D materials. Numerous operational factors, such as substrates, the position of substrates, carrier gas, pressure, temperature, and flow rate, require in-depth knowledge and engineering to produce crystals of high quality.32,39 Chen et al.40 used CVD to rapidly create high-quality millimeter-sized monolayer MoSe2 crystals on molten glass. As shown in Fig. 3(c), glass substrates were placed inside a quartz tube that was installed inside an oven. To contain the molten glass, a piece of SiO2/Si was used with Mo sheets, which are placed the above the MoO3 source.

3.4.

Pulse Laser Ablation Technique

The laser ablation technique is one of the most effective physical methods for producing NPs. The principle of this technique is summarized by absorbing the laser beam falling on the surface of a solid target, which leads to the expulsion of its components and the formation of nanostructures. When the solid target is removed in a vacuum or gas, nanoclusters are deposited on a substrate placed at a distance from the solid target (usually a few centimeters). When the ablation is carried out in a liquid environment (e.g., water, ethanol, and acetone), the NPs are released into the liquid medium and form a colloidal nanosolution. The properties of the resulting nanostructures can be controlled by different laser parameters (wavelength, pulse duration, laser energy, and number of pulses) as well as by the properties of the medium liquid.41

3.5.

Pulsed-Laser Deposition Technique

When material is ablated from a solid target in a gaseous environment, it can be recovered as a powder or as a deposit on a substrate. Although some uses, such as fuel cell applications of carbon nanotubes and nanohorns,42 necessitate the manufacture of nanostructured powders, the majority of applications necessitate the deposition of thin nanostructured films.43 The traditional PLD technology was found to be well suited for these jobs as it does not need any synthetic cluster size selection. The ablation of material is commonly conducted in PLD nanofabrication investigations using a nanosecond excimer laser (KrF and ArF) operating at UV wavelengths (248 and 193 nm, respectively) or the second harmonic of a Nd:YAG laser, as illustrated in Fig. 3 (532 nm). Because these wavelengths are lightly absorbed by plasma, they have little impact on the ablation process. At an incidence angle of roughly 45 deg, the radiation is focused on the surface of a revolving target (e.g., Si). The radiation intensity is typically between 108 and 109 W/cm. The plume of the laser-induced plasma extends perpendicular to the target surface. The substrates are placed at a distance (usually a few centimeters) from the target on a rotatable substrate holder. To increase the adherence of the deposited substance, the substrate is either held at room temperature or heated. The deposition is done in the presence of residual inert (He and Ar) or reactive (O and N) gases, at a low pressure (about 0.01 to 20 Torr), prior to filling with gases (Fig. 4).

Fig. 4

Schematic of the PLD experiment for Si-based nanostructured film deposition.

JNP_17_1_010901_f004.png

4.

Optoelectronics

A branch of photonics known as optoelectronics (or optronics) studies uses electronic devices and systems that source, detect, and control light. Light is often used to refer to both visible and invisible radiations in this context and includes gamma rays, x-rays, ultraviolet (UV), and infrared rays. Optoelectronic devices are electrical-to-optical or optical-to-electrical transducers or instruments that use such devices.44 The quantum mechanical effects of light on electronic materials, particularly semiconductors, in the presence of electric fields are the basis of optoelectronics.45

5.

Photodetector Parameters

In general, the performance of a photodetector is determined by some of the key parameters listed below.4649

5.1.

Sensitivity (R)

The ratio of photocurrent (Iph) via the detector to input light power Poptical, where Iph=IilluminatedIdark is

IilluminatedIdarkPoptical  (A/W).

5.2.

External Quantum Effectiveness

The proportion of charge carriers captured by the photodetector to incident light is the external quantum effectiveness (EQE), which is influenced by the material coefficient and the thickness of the absorbing material:

hceλR.

5.3.

On/Off Ratio

The proportion of photocurrent to dark current is the on/off ratio. The device should have a high on/off ratio because this indicates an improved photocurrent and reduced dark current, both of which increase device noise. The on/off ratio is calculated as

IphIdark.

5.4.

Detection Specificity (D*)

The photodetector’s responsivity and dark current determine how feeble a light the gadget can detect. Dark current should be reduced to the greatest extent possible because it makes it difficult to distinguish very weak optical signals and adds to the device’s noise; it is expressed as follows, where Δf is the electrical bandwidth, n is the noise current, and A is the effective area of the detector:

(AΔf)1/2.Rin  (Jones).

5.5.

Response Time

The rise and fall times of the optical signal response range from 10% to 90% and 90% to 10% of the maximum photocurrent, respectively.

5.6.

Power Equivalent to Noise

The NEP is the incident power required to achieve a signal-to-noise ratio of 1 at 1 Hz bandwidth, as determined by the noise spectral density.

6.

Photodetectors Based on Two Dimensional Materials

6.1.

Graphene-Based Photodetectors

The photoconductor and photovoltaic effects lie at the heart of graphene PDs. The external bias voltage is not required for graphene PDs because the high conductivity of graphene may create a huge dark current. A metal–graphene Schottky junction can be used to add built-in electric fields. Xia et al.50 was the first to announce graphene PD with a bandwidth of 40 GHz and a photoresponsivity of 0.5  mA/W. Most PDs have a low response51 or high dark current due to the graphene channel bias52 or fabrication method.53

In addition, Huang et al.54 discovered that the graphene/HfO/Si photodetector has a significantly greater interfacial gating effect than the graphene/SiO/Si photodetector. The graphene/HfO/Si photodetector has a photo responsivity of 45.8 A/W.

By modifying the rGO film with hydrophilic polymers, Shih et al.55 demonstrated a feasible method for designing surface homogeneity and improving photodetector performance. The on/off ratio of the PVA/rGO photodetector is 3.5 times higher than that of the rGO photodetector, and the detective improves by 53% even when the photodetector is operated at a low bias of 0.3 V.

6.2.

MoS2-Based Photodetectors

Graphene PDs when compared with 2D MoS2-based PDs rely primarily on photovoltaic effects, resulting in higher responsivity and smaller dark currents. MoS2 is one of the most thoroughly investigated TMDs because of its high carrier mobility of up to 200  cm2/Vs, high absorption efficiency, and electrical operation (on/off) ratio of more than 108. As a result of quantum confinement, the bulk MoS2 band gap can be converted from indirect to direct 1.2 to 1.8 eV by reducing it to a single layer.

Yin et al.56 demonstrated a single-layer phototransistor prepared by the mechanical exfoliation method with a responsivity to light of 7.5  mA/W at a wavelength of 750 nm and a response time of about 50 ms. Following that, researchers observed that charge trapping in MoS2, surface modification, preparation techniques, layer number, and the electrode contact all play a key role in MoS2 PD performance.5760 Meanwhile, Selamneni et al.61 established flexible broadband for all MoS2 PD based on MoS2 QDs and MoS2 heterojunctions recently, with a responsivity (R) value of 12.8  mA/W and specific detectivity (D*) value of 7.2×109  Jones. Furthermore, Han et al.62 created a MoS2 PD with a responsivity (R) of 1.4×105  A/W and specific detectivity (D*) of 9×1015  Jones, both of which were many magnitudes greater than Si and Ge PDs.

Also Kumar et al.63 constructed a highly effective UV photodetector based on MoS2 layers using the PLD technology, with a D* of 1.8×1014  Jones.

Recently, the photoelectric performance of the MoS2/PSS photodetector was increased under three layers of 365, 460, and 660 nm.64 The photocurrent rose thrice under the 365-nm laser, the noise equivalent power decreased to 1.77×1014  W/Hz1/2, and the specific detectivity (D*) increased by 1.2×1010  Jones.

Also Li et al.65 described a waveguide-integrated MoS2 photodetector functioning for hot electron-assisted photodetection to enable the telecom band. The low bias voltage of 0.3 V resulted in a photo responsiveness of 15.7  mAW1 at a wavelength of 1550 nm, which was also fairly uniform throughout the entire telecom band. Table 3 displays a comparison of MoS2 production methods.66

Table 3

Comparison of MoS2 synthesis techniques.

MethodsStructures and layer numbersAdvantagesDisadvantages
Mechanical sloughingA single layer of nanosheetsAppropriate for making individual devicesLow output
Questionable layer numbers
Exfoliation using a laserNanosheetsLocation and energy that can be controllableUnknown scale
Sonic exfoliationQuantum dotsHigh quality flexible layerUnknown scale
Large scaleLarge size distribution
Shear exfoliationNanosheets (40 to 220 nm)Easy processA wide range of sizes and thicknesses
Length (2 to 12) layersFast procedure
Liquid intercalationNanosheets (1 to 8) layersHigher monolayer energyRequired: flammable chemical (Li compound)
HydrothermalNanoflowersLow temperatureUncertain structure
High yield
CVDSingle or monolayer nanosheets lateral size up to several micrometersControllable layer numbers and sizeHigh temperature
Large scale

6.3.

Photodetectors Based on Polymer–Graphene/MoS2

MoS2 is considered to be a possible 2D material for optoelectronic applications due to its exceptional electrical and optical properties. The response range of MoS2-based photodetectors can be increased using a semimetal material with a zero bandgap, such as graphene. Table 4 illustrates the various strategies for building a polymer/MoS2/graphene-based photodetector.

Table 4

Different methods for fabricating polymer/MoS2/graphene-based photodetector.

DesignYearDevice fabrication approachApplication typeResponsivity (W/A)Specific detectivity (Jones)Reference
MOS2/graphene heterostructure2015Facile hydrothermalPhotodetector67
MOS2/graphene2015G:MEPhotodetector2.06 × 1031.5 × 101517
MOS2:ME
Graphene/MOS2/PE2016CVDPhotodetector45.519
MOS2/graphene2018CVDPhotodetector33.368
Poly (3-hexylthiophene)/graphene2019Spin-coatingPhotodetector0.251.8 × 10869
MOS2-PANI2019Internal polymerization of MOS2Photodetector2570
Graphene/MOS2/graphene2019G:CVDPhotodetector2101351
MOS2:CVD
Graphene/MOS2/graphene2020G:CVDPhotodetector2.2×1053.5×101371
Dry transfer
Polymer–graphene2020Spin-coatingGas detector72
PVA/G/Ag2021PLALPhotodetector9.3654.0473
PMMA/G/Ag2021PLALPhotodetector15.690.0673

6.4.

Solar Cell Based on MoS2/Graphene

Many photovoltaic applications use transition metal dichalcogenides (TMDCs), particularly MoS2 with a 2H crystal structure.74 The bandgap of TMDCs ranges from 1.0 to 2.0 eV, which corresponds to the Sun’s spectrum and with solar cell based on semiconductors, such as Si (1.1 eV), GaAs (1.4 eV), and CdTe (1.5 eV).75 It has been reported that 2D TMDCs with a thickness of 1 nm absorb 5% to 10% of incoming light per unit volume and have the absorption efficiency of the Sun that is greater than commonly used light absorbers such as Si and GaAs. The quantity of light absorbed by the TMDC single layer is equivalent to the amount absorbed by the 50-nm-thick Si layer, resulting in greater photocurrent values.76

In dye-sensitive solar cells, the antpolar plays an important role in the reduction process that occurs by transferring electrons through an external circuit. The electrocatalytic action of the counter electrode substantially influences the photo electrochemical process in dye-sensitized solar cells (DSSCs).77 In DSSCs, MoS2 is employed as a counter electrode instead of the conventionally used high-priced Pt electrodes.7880

Lei et al.81 employed MoS2 as a counter electrode in three different structural: MoS2 NPs, multilayered MoS2, and few-layered MoS2. The photon conversion efficiency of various MoS2 morphologies is (5.41%) for MoS2 NPs, (2.92%) for ML-MoS2, and (1.74%) for FL-MoS2.

Also Lin et al.82 obtained a power conversion efficiency (PCE) of 6.07% as a counter electrode, and MoS2 was implanted with graphene flakes. MoS2 allows for a quick reduction of triiodide ions, and graphene flakes, which are extremely conductive, allow for a quick charge transfer. In 2012, the same team employed a MoS2/rGO nanocomposite as the opposite electrode and achieved a PCE of 6.04%.83

Furthermore, MoS2 was identified as an HTL rather than PEDOT:PSS (poly(4,8-bis[(2-ethylhexyl)oxy)[1,2-b:4,5-b]dithiophene-2,6-diyl-3-fluoro-2-[(2-ethoxyhexyl)-carbonyl][3,4-b]dithiophene-2,6-diyl-3-fluoro-2-[(2-ethoxyhexyl)-carbonyl]thiophenediyl) PTB7:PC71BM solar cells and (3-alkylthiophenes).84 Organic solar cells based on P3HT:PC71BM ([6,6]-phenyl C71 butyric acid methyl ester) demonstrated a PCE of 8.11% and 4.02%, respectively. The PCE varies from 6.97% to 8.11% in PTB7:PC71BM-based solar cells when the layers of MoS2 are changed, demonstrating the reliance of the PCE on the number of layers of MoS2. The solar cell based on monolayer MoS2 achieved the VOC = 0.69 V, the fill factor FF = 64%, the JSC = 15.78  mA/cm2, and the PCE = 6.97%.

Li et al.85 demonstrated that MoS2 may be employed as an anode buffer layer in organic solar cells to avoid current leakage between ITO and PC61BM; the result was a PCE = 3.9%. Solar cells made of ITO without a buffer layer have a PCE = 2.14%. A nanocomposite of PEDOT:PSS/MoS2 may be employed as a hole extraction layer in solar cells based on P3HT:PCBM ([6,6]-phenyl-C61-butyric acid methyl ester) with a PCE = 3.74%.86

Yang et al.87 employed MoS2 as a hole layer in an organic solar cell. The organic solar cell PTB7:PC71BM was employed as an active layer; PFN poly[(9,9-bis(3′-(N-N-dimethylamino)-propyl)-2,7-fluorene)-alt-2,7-(9,9dioctylfluorene)-alt-2,7-(9,9dioctylfluorene)-alt-2,7-9, in comparison with PEDOT:PSS, MoS2 nanosheets and oxygen-containing MoS2 nanosheets demonstrated superior optical transparency in range the wavelength 400 to 900 nm, with a PCE of 4.99%. This is due to the possibility of using graphene as layers for extracting electrons and holes in organic solar cells.

MoS2 was used as both an electron extraction layer and a hole withdrawal layer.88 P-type doped MoS2 was used for the hole withdrawal layer, and n-type doped MoS2 was used for the electron extraction layer. The photon conversion efficiency increased from 2.8% to 3.4% in organic solar cells using p-type doped MoS2 as a hole extraction layer.

MoS2 was used as an interfacial layer, and silver nanowires were used in conjunction with MoS2 to create a transparent electrode.89 At 550 nm, Ag NWs-MoS2 exhibited 93.1% transparency as a transparent electrode and had a reduced sheet resistance 9.8/sq. When compared with the Ag NW electrode alone, MoS2 in the Ag NWs-MoS2 composite electrode enhanced environmental dependability and offered the best oxidation and moisture resistance. A PBDTTT-C-T:A PC70BM-based organic solar cell with a cathode made of Ag NWs-MoS2 nanosheets/n-MoS2 nanosheets and an anode made of p-MoS2 nanosheets/Ag for hole extraction produced a PCE of 8.72%. As seen in Fig. 5(a), the general cell structure is glass/Ag-MoS2/n-MoS2/PBDTTT-C-T:PC70BM/p-MoS2/Ag.

Fig. 5

Tapping-mode AFM images of electrodes. (a) Ag NWs bottom electrode film height image. (b) Ag NWs-MoS2 bottom electrode film height image.

JNP_17_1_010901_f005.png

Qin et al.90 and Yang et al.91 employed MoS2 in conjunction with the MoO3 layer and MoS2 layer with Au NPs as hole extraction layers in solar cells. Also Chuang, et al.92 used MoS2 nanosheets with Au NP as an electron extraction layer, yielding a PCE = 4.91%. Jian et al.93 investigated the role of the MoS2 layer in graphene/silicon organic solar cells. MoS2 was used as the interlayer in this case and provided a 4.4% PCE.

The same group93 investigated the influence of annealing temperature on the PCE of the MoS2 prototype. Tsuboi et al.94 employed MoS2 as an electron blocking and passivation layer between graphene and silicon of heterojunction solar cells based on graphene/MoS2/n-Si in the same year. When comparing the solar cell that contains the MoS2 layer, the PCE of 1.35% is higher than the PCE of the solar cells without the MoS2 layer, with a PCE of 0.07%. The use of three-layer graphene increased the efficiency of PCE to 8.0%.

Rehman et al.95 used the Al2O3 layer on the MoS2 surface to improve the performance of the solar cells based on MoS2/silicon. The PCE = 5.6% was obtained; this is higher than the conversion efficiency of the solar cell based on MoS2/silicon without adding the Al2O3 layer, which was PCE = 2.21%. Also Saraswat et al.96 made a solar cell based on TMDCs/silicon and gave highly efficient results and excellent photovoltaic performance. Meanwhile, Krishnamoorthy and Prakasam97 reported on the simple hydrothermal synthesis of diverse compositions of MoS2/graphene nanocomposites and their subsequent application in DSSCs. The PCE=8.92% was for MoS2/graphene and was higher than the conversion efficiency of MoS2 without graphene (PCE=3.36%) because graphene increased the charge collection faster and sensitized the dye better.

Maa et al.98 described a unique approach to synthesizing layers of MoS2 on diverse substrates such as n-Si that act as photon absorbing layers. This significantly improved the photovoltaic conversion efficiency of graphene-based Schottky junction solar cells. In heterojunction solar cells based on 2D materials, a comparatively high conversion efficiency of 12% was successfully obtained.

Flexible solar cells were manufactured from MoS2, which works as a hole transfer layer with a transparent conductive electrode.99 The PCE of these solar cells depends on the number of MoS2 layers as well as on the concentration of doping, with the value of the PCE being the highest when the number of layers is (Ln=2) and the concentration of doping is (nD=20  mM). When the active layer with quantum dots was added to the solar cell installation, the efficiency of the conversion power was increased to 4.23%, and this efficiency was maintained after conducting 1000 cycles of tests, which indicates the mechanical stability of this solar cell.

7.

Conclusion

We addressed current advances in the basic studies as well as multiapplications of MOS2 with a focus on one MOS2 or multilayer MOS2 as well as on heterostructures. The nanocomposites showed distinctive and unique properties that cannot be obtained in one material. Recently, there have been a large number of studies and applications of technological uses of thin layered MOS2. The MOS2 thin layer is important in various low-cost optoelectronics applications, such as solar cells, light-emitting diodes, photodetectors, and sensors, due to its excellent optical and electronic properties as well as having an adjustable bandgap. In this review, many studies using MOS2, graphene, and polymers that were used in many heterogeneous applications using a single layer or more were discussed. The results showed a significant improvement in many applications as photodetectors and solar cells and showed new features, which led to an increase in the efficiency of these electronic devices. Several critical challenges must yet be solved if development is to be maintained at this rate. First, new techniques for doping MOS2 and similar materials to achieve bipolar and p-type behavior must be devised. Second, knowledge and understanding of how building materials interact optimally with device applications must be investigated. The final challenge is a better understanding of heterojunction structures, especially at the atomic level. To address these challenges, advanced scientific research must be conducted in various disciplines. It is expected that the research and development of MOS2, graphene, and polymers will be effective and have distinctive and promising results that contribute to the development of many electronic applications in the near future.

References

1. 

D. Chrisey and G. Hubler, Pulsed Laser Deposition of Thin Films, John Wiley & Sons, New York (1994). Google Scholar

2. 

Laser Processing and Chemistry, (2000). Google Scholar

3. 

E. Fogarassy and S. Lazare, Laser Ablation of Electronic Materials, North Holland Publishing Company( (1992). Google Scholar

4. 

L. Gallais and J.-Y. Natoli, “Optimized metrology for laser-damage measurement: application to multiparameter study,” Appl. Opt., 42 (6), 960 –971 https://doi.org/10.1364/AO.42.000960 (2003). Google Scholar

5. 

K. Mittal, Particles on Surfaces, Plenum Press, New York (1988). Google Scholar

6. 

X. Wu et al., 70 167 –178 (1999). Google Scholar

7. 

J. Perriere, E. Millon and E. Fogarassy, Recent Advances in Laser Processing of Materials, Elsevier Science( (2006). Google Scholar

8. 

A. Rai et al., “Progress in contact, doping and mobility engineering of MoS2: an atomically thin 2D semiconductor,” Crystals, 8 (8), 316 https://doi.org/10.3390/cryst8080316 CRYSBC 2073-4352 (2018). Google Scholar

9. 

Q. Cui et al., “Material and device architecture engineering toward high performance two-dimensional (2D) photodetectors,” Crystals, 7 (5), 149 https://doi.org/10.3390/cryst7050149 CRYSBC 2073-4352 (2017). Google Scholar

10. 

K. M. Freedy and S. J. McDonnell, “Contacts for molybdenum disulfide: interface chemistry and thermal stability,” Materials (Basel), 13 (3), 693 https://doi.org/10.3390/ma13030693 MATEG9 1996-1944 (2020). Google Scholar

11. 

K. Roy et al., “Graphene–MoS2 hybrid structures for multifunctional photoresponsive memory devices,” Nat. Nanotechnol., 8 (11), 826 –830 https://doi.org/10.1038/nnano.2013.206 NNAABX 1748-3387 (2013). Google Scholar

12. 

M. Rahman et al., “Design and numerical analysis of highly sensitive Au–MoS2–graphene-based hybrid surface plasmon resonance biosensor,” Opt. Commun., 396 36 –43 https://doi.org/10.1016/j.optcom.2017.03.035 OPCOB8 0030-4018 (2017). Google Scholar

13. 

H. R. Nalwa, “A review of molybdenum disulfide (MoS2) based photodetectors: from ultra-broadband, self-powered to flexible devices,” RSC Adv., 10 (51), 30529 –30602 https://doi.org/10.1039/D0RA03183F (2020). Google Scholar

14. 

T. Ahmed et al., “Interplay of charge transfer and disorder in optoelectronic response in graphene/hBN/MoS2 van der Waals heterostructures,” 2D Mater., 7 (2), 025043 https://doi.org/10.1088/2053-1583/ab771f (2020). Google Scholar

15. 

H. Rashid et al., “Prospects of molybdenum disulfide (MoS2) as an alternative absorber layer material in thin film solar cells from numerical modeling,” Chalcogenide Lett., 11 (8), 397 –403 (2014). Google Scholar

16. 

Q. A. Vu et al., “Tuning carrier tunneling in van der Waals heterostructures for ultrahigh detectivity,” Nano Lett., 17 (1), 453 –459 https://doi.org/10.1021/acs.nanolett.6b04449 NALEFD 1530-6984 (2017). Google Scholar

17. 

S. Rathi et al., “Tunable electrical and optical characteristics in monolayer graphene and few-layer MoS2 heterostructure devices,” Nano Lett., 15 (8), 5017 –5024 https://doi.org/10.1021/acs.nanolett.5b01030 NALEFD 1530-6984 (2015). Google Scholar

18. 

H. Li et al., “Restoring the photovoltaic effect in graphene-based van der Waals heterojunctions towards self-powered high-detectivity photodetectors,” Nano Energy, 57 214 –221 https://doi.org/10.1016/j.nanoen.2018.12.004 (2019). Google Scholar

19. 

D. De Fazio et al., “High responsivity, large-area graphene/MoS2 flexible photodetectors,” ACS Nano, 10 (9), 8252 –8262 https://doi.org/10.1021/acsnano.6b05109 ANCAC3 1936-0851 (2016). Google Scholar

20. 

B. Radisavljevic et al., “Single-layer MoS2 transistors,” Nat. Nanotechnol., 6 (3), 147 –150 https://doi.org/10.1038/nnano.2010.279 NNAABX 1748-3387 (2011). Google Scholar

21. 

W. M. Haynes, CRC Handbook of Chemistry and Physics, (Internet Version 2011), Taylor Francis Group, Boca Raton, Florida (2011). Google Scholar

22. 

W. S. Yun et al., “Thickness and strain effects on electronic structures of transition metal dichalcogenides: 2H-MX2 semiconductors (M = Mo, W; X = S, Se, Te),” Phys. Rev. B, 85 (3), 033305 https://doi.org/10.1103/PhysRevB.85.033305 PRBMDO 1098-0121 (2012). Google Scholar

23. 

A. T. Dideikin and A. Y. Vul, “Graphene oxide and derivatives: the place in graphene family,” Front. Phys., 6 149 https://doi.org/10.3389/fphy.2018.00149 FRPHAY 0429-7725 (2019). Google Scholar

24. 

G. Nagaraju et al., “Hydrothermal synthesis of amorphous MoS2 nanofiber bundles via acidification of ammonium heptamolybdate tetrahydrate,” Nanosc. Res. Lett., 2 (9), 461 –468 https://doi.org/10.1007/s11671-007-9087-z NRLAAD 1556-276X (2007). Google Scholar

25. 

Y. Li et al., “MoS2 nanoparticles grown on graphene: an advanced catalyst for the hydrogen evolution reaction,” J. Am. Chem. Soc., 133 (19), 7296 –7299 https://doi.org/10.1021/ja201269b JACSAT 0002-7863 (2011). Google Scholar

26. 

B. C. Helmly et al., “Preparation and spectroscopic characterization of MoS2 and MoSe2 nanoparticles,” Spectrosc. Lett., 40 (3), 483 –492 https://doi.org/10.1080/00387010701295984 SPLEBX 0038-7010 (2007). Google Scholar

27. 

B. Li et al., “Preparation of monolayer MoS2 quantum dots using temporally shaped femtosecond laser ablation of bulk MoS2 targets in water,” Sci. Rep., 7 (1), 1 –12 https://doi.org/10.1038/s41598-017-10632-3 (2017). Google Scholar

28. 

S. Moniri and M. R. Hantehzadeh, “Colloidal synthesis of MoS2 NPs by nanosecond laser ablation of a bulk MoS2 target in ethylene glycol solution,” Opt. Quantum Electron., 53 (5), 1 –17 https://doi.org/10.1007/s11082-021-02872-5 OQELDI 0306-8919 (2021). Google Scholar

29. 

A. K. Geim and K. S. Novoselov, “The rise of graphene,” Nat. Mater., 6 183 –191 https://doi.org/10.1038/nmat1849 NMAACR 1476-1122 (2007). Google Scholar

30. 

Biomedical Applications of Graphene and 2D Nanomaterials, Elsevier( (2019). Google Scholar

31. 

M. K. Mohammed, A. Al-Nafiey and G. Al-Dahash, “Manufacturing graphene and graphene-based nanocomposite for piezoelectric pressure sensor application: a review,” Nano Biomed. Eng., 13 (1), 27 –35 https://doi.org/10.5101/nbe.v13i1.p27-35 (2021). Google Scholar

32. 

A. Zavabeti et al., “Two-dimensional materials in large-areas: synthesis, properties and applications,” Nano-Micro Lett., 12 (1), 1 –34 https://doi.org/10.1007/s40820-020-0402-x (2020). Google Scholar

33. 

H. Tao et al., “Scalable exfoliation and dispersion of two-dimensional materials—an update,” Phys. Chem. Chem. Phys., 19 (2), 921 –960 https://doi.org/10.1039/C6CP06813H PPCPFQ 1463-9076 (2017). Google Scholar

34. 

W. Li et al., “Gas sensors based on mechanically exfoliated MoS2 nanosheets for room-temperature NO2 detection,” Sensors, 19 (9), 2123 https://doi.org/10.3390/s19092123 SNSRES 0746-9462 (2019). Google Scholar

35. 

K. R. Paton et al., “Scalable production of large quantities of defect-free few-layer graphene by shear exfoliation in liquids,” Nat. Mater., 13 (6), 624 –630 https://doi.org/10.1038/nmat3944 NMAACR 1476-1122 (2014). Google Scholar

36. 

Y. Hernandez et al., “High-yield production of graphene by liquid-phase exfoliation of graphite,” Nat. Nanotechnol., 3 (9), 563 –568 https://doi.org/10.1038/nnano.2008.215 NNAABX 1748-3387 (2008). Google Scholar

37. 

P. G. Karagiannidis, “Microfluidization of graphite and formulation of graphene-based conductive inks,” ACS Nano, 11 2742 –2755 https://doi.org/10.1021/acsnano.6b07735 ANCAC3 1936-0851 (2017). Google Scholar

38. 

Z. Shen et al., “Preparation of graphene by jet cavitation,” Nanotechnology, 22 365306 https://doi.org/10.1088/0957-4484/22/36/365306 NNOTER 0957-4484 (2011). Google Scholar

39. 

J. Yu et al., “Synthesis of two-dimensional materials via chemical vapor deposition,” Chem. Sci., 6 6705 –6716 https://doi.org/10.1039/C5SC01941A 1478-6524 (2015). Google Scholar

40. 

J. Chen et al., “Chemical vapor deposition of large-size monolayer MoSe2 crystals on molten glass,” J. Am. Chem. Soc., 139 1073 –1076 https://doi.org/10.1021/jacs.6b12156 (2017). Google Scholar

41. 

Z. Yan and B. C. Douglas, “Pulsed laser ablation in liquid for micro-/nanostructure generation,” J. Photochem. Photobiol., C, 13 (3), 204 –223 https://doi.org/10.1016/j.jphotochemrev.2012.04.004 (2012). Google Scholar

42. 

S. Iijima et al., “Nano-aggregates of single-walled graphitic carbon nano-horns,” Chem. Phys. Lett., 309 165 https://doi.org/10.1016/S0009-2614(99)00642-9 CHPLBC 0009-2614 (1999). Google Scholar

43. 

E. Werwa et al., “Synthesis and processing of silicon nanocrystallites using a pulsed laser ablation supersonic expansion method,” Appl. Phys. Lett., 64 (14), 1821 –1823 https://doi.org/10.1063/1.111766 (1994). Google Scholar

44. 

P. V. Kazakevich et al., “Laser induced synthesis of nanoparticles in liquids,” Appl. Surf. Sci., 252 (13), 4373 –4380 https://doi.org/10.1016/j.apsusc.2005.06.059 ASUSEE 0169-4332 (2006). Google Scholar

45. 

V. S. Naik, V. S. P. Bhanudas and N. N. Ghosh, “A simple aqueous solution based chemical methodology for synthesis of Ag nanoparticles dispersed on mesoporous silicate matrix,” Powder Technol., 199 (2), 197 –201 https://doi.org/10.1016/j.powtec.2010.01.005 POTEBX 0032-5910 (2010). Google Scholar

46. 

Y. Qin et al., “Review of deep ultraviolet photodetector based on gallium oxide,” Chin. Phys. B, 28 018501 https://doi.org/10.1088/1674-1056/28/1/018501 1674-1056 (2019). Google Scholar

47. 

W. Tian et al., “Self-powered nanoscale photodetectors,” Small, 13 1701848 https://doi.org/10.1002/smll.201701848 SMALBC 1613-6810 (2017). Google Scholar

48. 

H. Qiao et al., “Self-powered photodetectors based on 2D materials,” Adv. Opt. Mater., 8 1900765 https://doi.org/10.1002/adom.201900765 2195-1071 (2019). Google Scholar

49. 

A. De Sanctis et al., “Graphene-based light sensing: fabrication, characterisation, physical properties and performance,” Materials (Basel), 11 1762 https://doi.org/10.3390/ma11091762 MATEG9 1996-1944 (2018). Google Scholar

50. 

B. Sun et al., “Progress on crystal growth of two-dimensional semiconductors for optoelectronic applications,” Crystals, 8 252 https://doi.org/10.3390/cryst8060252 CRYSBC 2073-4352 (2018). Google Scholar

51. 

B. Liu et al., “High performance photodetector based on graphene/MoS2/graphene lateral heterostrurcture with Schottky junctions,” J. Alloys Compd., 779 140 –146 https://doi.org/10.1016/j.jallcom.2018.11.165 JALCEU 0925-8388 (2019). Google Scholar

52. 

L. Xu et al., “Two-dimensional MoS2–graphene-based multilayer van Der Waals heterostructures: enhanced charge transfer and optical absorption, and electric-field tunable Dirac point and band gap,” Chem. Mater., 29 5504 –5512 https://doi.org/10.1021/acs.chemmater.7b00281 CMATEX 0897-4756 (2017). Google Scholar

53. 

G. Konstantatos, “Current status and technological prospect of photodetectors based on two-dimensional materials,” Nat. Commun., 9 9 –11 https://doi.org/10.1038/s41467-018-07643-7 NCAOBW 2041-1723 (2018). Google Scholar

54. 

Z. Huang et al., “Interfacial gated graphene photodetector with broadband response,” ACS Appl. Mater. Interfaces, 13 22796 –22805 https://doi.org/10.1021/acsami.1c02738 AAMICK 1944-8244 (2021). Google Scholar

55. 

Y. S. Shih et al., “Low power photodetectors based on PVA modified reduced graphene oxide hybrid solutions,” Macromol. Rapid Commun., 43 2100854 https://doi.org/10.1002/marc.202100854 MRCOE3 1521-3927 (2022). Google Scholar

56. 

Z. Yin et al., “Single-layer MoS2 phototransistors,” ACS Nano, 6 74 –80 https://doi.org/10.1021/nn2024557 ANCAC3 1936-0851 (2012). Google Scholar

57. 

O. Lopez-Sanchez et al., “Ultrasensitive photodetectors based on monolayer MoS2,” Nat. Nanotechnol., 8 497 –501 https://doi.org/10.1038/nnano.2013.100 NNAABX 1748-3387 (2013). Google Scholar

58. 

M. F. Khan et al., “Photocurrent response of MoS2 field-effect transistor by deep ultraviolet light in atmospheric and N2 gas environments,” ACS Appl. Mater. Interfaces, 6 21645 –21651 https://doi.org/10.1021/am506716a AAMICK 1944-8244 (2014). Google Scholar

59. 

D. S. Tsai et al., “Few-layer MoS2 with high broadband photogain and fast optical switching for use in harsh environments,” ACS Nano, 7 3905 –3911 https://doi.org/10.1021/nn305301b ANCAC3 1936-0851 (2013). Google Scholar

60. 

J. Lu et al., “Improved photoelectrical properties of MoS2 films after laser micromachining,” ACS Nano, 8 6334 –6343 https://doi.org/10.1021/nn501821z ANCAC3 1936-0851 (2014). Google Scholar

61. 

V. Selamneni, S. K. Ganeshan and P. Sahatiya, “All MoS2 based 2D/0D localized unipolar heterojunctions as flexible broadband (UV-vis-NIR) photodetectors,” J. Mater. Chem. C, 8 11593 –11602 (2020). Google Scholar

62. 

P. Han et al., “Highly sensitive MoS2 photodetectors with graphene contacts,” Nanotechnology, 29 20LT01 https://doi.org/10.1088/1361-6528/aab4bb NNOTER 0957-4484 (2018). Google Scholar

63. 

S. Kumar et al., “High performance UV photodetector based on MoS2 layers grown by pulsed laser deposition technique,” J. Alloys Compd., 835 155222 https://doi.org/10.1016/j.jallcom.2020.155222 JALCEU 0925-8388 (2020). Google Scholar

64. 

X. Liu et al., “Suspended MoS2 photodetector using patterned sapphire substrate,” Nano-Microsmall, 17 2100246 https://doi.org/10.1002/smll.202100246 (2021). Google Scholar

65. 

Z. Li et al., “Telecom-band waveguide-integrated MoS2 photodetector assisted by hot electrons,” ACS Photonics, 9 (1), 282 –289 https://doi.org/10.1021/acsphotonics.1c01622 (2022). Google Scholar

66. 

W. Zhang et al., “Synthesis and sensor applications of MoS2-based nanocomposites,” Nanoscale, 7 (44), 18364 –18378 https://doi.org/10.1039/C5NR06121K NANOHL 2040-3364 (2015). Google Scholar

67. 

Z. Huang et al., “Photoelectrochemical-type sunlight photodetector based on MoS2/graphene heterostructure,” 2D Mater., 2 (3), 035011 https://doi.org/10.1088/2053-1583/2/3/035011 (2015). Google Scholar

68. 

B. Sun et al., “Large-area flexible photodetector based on atomically thin MoS2/graphene film,” Mater. Des., 154 1 –7 https://doi.org/10.1016/j.matdes.2018.05.017 MADSD2 0264-1275 (2018). Google Scholar

69. 

A. Yadav et al., “Poly-(3-hexylthiophene)/graphene composite based organic photodetectors: the influence of graphene insertion,” Thin Solid Films, 675 128 –135 https://doi.org/10.1016/j.tsf.2019.02.013 THSFAP 0040-6090 (2019). Google Scholar

70. 

N. Chaudhary, M. Khanuja and S. S. Islam, “Broadband photodetector based on 3D architect of MoS2–PANI hybrid structure for high photoresponsive properties,” Polymer, 165 168 –173 https://doi.org/10.1016/j.polymer.2019.01.028 POLMAG 0032-3861 (2019). Google Scholar

71. 

I. Lee et al., “Photoinduced tuning of Schottky barrier height in graphene/MoS2 heterojunction for ultrahigh performance short channel phototransistor,” ACS Nano, 14 7574 –7580 https://doi.org/10.1021/acsnano.0c03425 ANCAC3 1936-0851 (2020). Google Scholar

72. 

E. Soleimani et al., “Tuning the polymer–graphene interfaces by picric acid molecules to improve the sensitivity of a prepared conductive polymer composite gas detector,” Iran. Polym. J., 29 (4), 341 –350 https://doi.org/10.1007/s13726-020-00800-0 (2020). Google Scholar

73. 

M. K. Mohammed et al., “Employment of polymer/graphene-Ag nanocomposite films prepared by PLAL for optoelectric applications,” (2021). Google Scholar

74. 

E. Singh et al., “Atomically thin-layered molybdenum disulfide (MoS2) for bulk-heterojunction solar cells,” ACS Appl. Mater. Interfaces, 9 (4), 3223 –3245 https://doi.org/10.1021/acsami.6b13582 AAMICK 1944-8244 (2017). Google Scholar

75. 

W. Tang, S. S. Rassay and N. M. Ravindra, “Electronic and optical properties of transition-metal dichalcogenides,” Madr. J. Nanotechnol. Nanosci., 2 58 –64 https://doi.org/10.18689/mjnn-1000111 (2017). Google Scholar

76. 

M. Bernardi, M. Palummo and J. C. Grossman, “Extraordinary sunlight absorption and one nanometer thick photovoltaics using two-dimensional monolayer materials,” Nano Lett., 13 3664 –3670 https://doi.org/10.1021/nl401544y NALEFD 1530-6984 (2013). Google Scholar

77. 

Y. Y. Dou et al., “Nickel phosphide-embedded graphene as counter electrode for dye-sensitized solar cells,” Phys. Chem. Chem. Phys., 14 1339 –1342 https://doi.org/10.1039/C2CP23775J PPCPFQ 1463-9076 (2012). Google Scholar

78. 

H. Jeong et al., “Rapid sintering of MoS2 counter electrode using near-infrared pulsed laser for use in highly efficient dye-sensitized solar cells,” J. Power Sour., 330 104 –110 https://doi.org/10.1016/j.jpowsour.2016.09.002 JPSODZ 0378-7753 (2016). Google Scholar

79. 

A. Antonelou et al., “Facile, Substrate-scale growth of mono- and few-layer homogeneous MoS2 films on Mo foils with enhanced catalytic activity as counter electrodes in DSSCs,” Nanotechnology, 27 045404 https://doi.org/10.1088/0957-4484/27/4/045404 NNOTER 0957-4484 (2015). Google Scholar

80. 

C. K. Cheng and C. K. Hsieh, “Electrochemical deposition of molybdenum sulfide thin films on conductive plastic substrates as platinum-free flexible counter electrodes for dye-sensitized solar cells,” Thin Solid Films, 584 52 –60 https://doi.org/10.1016/j.tsf.2015.02.025 THSFAP 0040-6090 (2015). Google Scholar

81. 

B. Lei, G. R. Li and X. P. Gao, “Morphology dependence of molybdenum disulfide transparent counter electrode in dye-sensitized solar cells,” J. Mater. Chem. A, 2 3919 –3925 https://doi.org/10.1039/c3ta14313a (2014). Google Scholar

82. 

J.-Y. Lin et al., “Hydrothermal synthesis of graphene flakeembedded nanosheet-like molybdenum sulfide hybrids as counter electrode catalysts for dye-sensitized solar cells,” Mater. Chem. Phys., 143 53 –59 https://doi.org/10.1016/j.matchemphys.2013.08.010 MCHPDR 0254-0584 (2013). Google Scholar

83. 

C.-J. Liu et al., “Facile synthesis of MoS2/graphene nanocomposite with high catalytic activity toward triiodide reduction in dye-sensitized solar cells,” J. Mater. Chem., 22 21057 –21064 https://doi.org/10.1039/c2jm33679k JMACEP 0959-9428 (2012). Google Scholar

84. 

X. Gu et al., “A solution-processed hole extraction layer made from ultrathin MoS2 nanosheets for efficient organic solar cells,” Adv. Energy Mater., 3 1262 –1268 https://doi.org/10.1002/aenm201300549 ADEMBC 1614-6840 (2013). Google Scholar

85. 

X. Li et al., “Solution-processed MoSx as an efficient anode buffer layer in organic solar cells,” ACS Appl. Mater. Interfaces, 5 8823 –8827 https://doi.org/10.1021/am402105d AAMICK 1944-8244 (2013). Google Scholar

86. 

M. S. Ramasamy et al., “Solution-processed PEDOT: PSS/MoS2 nanocomposites as efficient hole-transporting layers for organic solar cells,” Nanomaterials (Basel), 9 1328 https://doi.org/10.3390/nano9091328 (2019). Google Scholar

87. 

X. Yang et al., “Engineering crystalline structures of two-dimensional MoS2 sheets for high-performance organic solar cells,” J. Mater. Chem. A, 2 7727 –7733 https://doi.org/10.1039/C4TA01336K (2014). Google Scholar

88. 

J.-M. Yun et al., “Efficient work-function engineering of solution-processed MoS2 thin-films for novel hole and electron transport layers leading to high-performance polymer solar cells,” J. Mater. Chem. C, 1 3777 –3783 https://doi.org/10.1039/c3tc30504j (2013). Google Scholar

89. 

X. Hu et al., “Versatile MoS2 nanosheets in ITO-free and semi-transparent polymer power-generating glass,” Sci. Rep., 5 1 –13 https://doi.org/10.1038/srep12161 SRCEC3 2045-2322 (2015). Google Scholar

90. 

P. Qin et al., “In situ growth of double-layer MoO3/MoS2 film from MoS2 for hole-transport layers in organic solar cell,” J. Mater. Chem. A, 2 2742 –2756 https://doi.org/10.1039/c3ta13579a (2014). Google Scholar

91. 

X. Yang et al., “Au nanoparticles on ultrathin MoS2 sheets for plasmonic organic solar cells,” J. Mater. Chem. A, 2 14798 –14806 https://doi.org/10.1039/C4TA03178D (2014). Google Scholar

92. 

M. K. Chuang, S. S. Yang and F. C. Chen, “Metal nanoparticle-decorated two-dimensional molybdenum sulfide for plasmonic-enhanced polymer photovoltaic devices,” Materials (Basel), 8 5414 –5425 https://doi.org/10.3390/ma8085252 (2015). Google Scholar

93. 

K. Jiao et al., “The role of MoS2 as an interfacial layer in graphene/silicon solar cells,” Phys. Chem. Chem. Phys., 17 8182 –8186 https://doi.org/10.1039/C5CP00321K PPCPFQ 1463-9076 (2015). Google Scholar

94. 

Y. Tsuboi et al., “Enhanced photovoltaic performances of graphene/Si solar cells by insertion of a MoS2 thin film,” Nanoscale, 7 14476 –14482 https://doi.org/10.1039/C5NR03046C NANOHL 2040-3364 (2015). Google Scholar

95. 

A. U. Rehman et al., “N-MoS2/p-Si solar cells with Al2O3 passivation for enhanced photogeneration,” ACS Appl. Mater. Interfaces, 8 29383 –29390 https://doi.org/10.1021/acsami.6b07064 AAMICK 1944-8244 (2016). Google Scholar

96. 

K. C. Saraswat et al., Low Cost Silicon (Si)/Transition Metal Dichalcogenides (TMDs) Tandem Solar Cells with Goal of >30% Efficiency, 1 –2 Stanford University, Stanford, California (2019). Google Scholar

97. 

D. Krishnamoorthy and A. Prakasam, “Preparation of MoS2/graphene nanocomposite-based photoanode for dye-sensitized solar cells (DSSCs),” Inorg. Chem. Commun., 118 108016 https://doi.org/10.1016/j.inoche.2020.108016 1387-7003 (2020). Google Scholar

98. 

J. Ma, Y. Yuan and P. Sun, “Performances enhancement of graphene/n-Si Schottky junction solar cells with dual-functional MoS2 interfacial layers,” J. Alloys Compd., 883 160898 https://doi.org/10.1016/j.jallcom.2021.160898 JALCEU 0925-8388 (2021). Google Scholar

99. 

D. H. Shin et al., “Enhancement of efficiency and stability in organic solar cells by employing MoS2 transport layer, graphene electrode, and graphene quantum dots-added active layer,” Appl. Surf. Sci., 538 148155 https://doi.org/10.1016/j.apsusc.2020.148155 ASUSEE 0169-4332 (2021). Google Scholar

Biography

Nagham M. Obaid received her BSc degree in physics from the College of Education at the University of Babylon in 2006 and her MSc degree in laser physics from the College of Science for Women at the University of Babylon in 2016. Now she is a PhD student in the research stage at the College of Science for Women of the University of Babylon. Her scientific interests are in the fields of lasers and their wide applications and she has conducted several research projects in this field.

Amer Al-Nafiey received his BSc degree in physics from the University of Babylon in 1998, his MSc degree in physics from the same university in 2006, and his PhD in nanotechnology from Lille1 University of Science and Technology in 2016. Now, he is an academic researcher and an assistant professor at the College of Science for Women of the University of Babylon. His scientific interests are in the fields focused on nanomaterials, nanomaterials synthesis, material characterization, thin film synthesis, nanostructured materials, nanocomposites, carbon nanomaterials, catalysis, and bio-nanoapplications. He has more than 25 research papers published in international peer-reviewed journals.

Ghalib Al-Dahash received his BSc degree in physics from Al-Mustansiriya University in 1986, his MSc degree in nuclear material physics from the same university in 1992, and his PhD in solid-state physics with a focus on superconductivity from the University of Baghdad in 1998. Now, he is a professor of physics at the College of Science for Women of the University of Babylon. His scientific interests are in the fields of laser material processing and superconductors. He has more than 86 research papers published in international peer-reviewed journals.

© 2023 Society of Photo-Optical Instrumentation Engineers (SPIE)
Nagham M. Obaid, Amer Al-Nafiey, and Ghaleb Al-Dahash "Graphene/molybdenum disulfide nanocomposites: characterization and optoelectronic application," Journal of Nanophotonics 17(1), 010901 (4 January 2023). https://doi.org/10.1117/1.JNP.17.010901
Received: 30 August 2022; Accepted: 8 December 2022; Published: 4 January 2023
Lens.org Logo
CITATIONS
Cited by 1 scholarly publication.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Molybdenum

Graphene

Photodetectors

Solar cells

Nanocomposites

2D materials

Electrodes

Back to Top