Home A comprehensive study of laser irradiated hydrothermally synthesized 2D layered heterostructure V2O5(1−x)MoS2(x) (X = 1–5%) nanocomposites for photocatalytic application
Article Open Access

A comprehensive study of laser irradiated hydrothermally synthesized 2D layered heterostructure V2O5(1−x)MoS2(x) (X = 1–5%) nanocomposites for photocatalytic application

  • Muhammad Hasnain Jameel EMAIL logo , Aqeela Yasin , Samia , Mohd Zul Hilmi Bin Mayzan EMAIL logo , Muhammad Sufi bin Roslan , Fahmiruddin Bin Esa , Mohd Arif Bin Agam , Mohd Hafiz Mohd Zaid , Khaled Althubeiti and Mohammed Aljohani
Published/Copyright: August 7, 2024
Become an author with De Gruyter Brill

Abstract

It has been studied that both two-dimensional (2D) MoS2 and V2O5, which are classified as transition metal dichalcogenides and transition metal oxides, are good photocatalyst materials. For this purpose, the hydrothermal method was practiced to synthesize V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites with different 1–5% w/w weight percent of MoS2 as a prominent photocatalyst under laser irradiation for 2, 4, 6, 8, and 10 min to tune photocatalytic degradation of industrial wastage water. The surface of the 2D molybdenum nanolayered matrix was efficaciously decorated with V2O5 nanoparticles. The crystal phase and layered structures of the V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites samples were verified by X-ray diffraction and scanning electron microscopy, atomic force microscopy, X-ray photoelectron spectroscopy respectively. In the range of the UV visible spectrum, the increment in light absorption from 3.6 to 14.5 Ω−1 cm−1 with an increase of active surface from 108 to 169 μ m 2 with increased MoS2 doping percentage. Furthermore, dielectric findings like the complex dielectric function, tangent loss, electrical conductivity, quality factors, and impedance of V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites are studied. According to photoluminescence studies, the intensity of peaks decreases when laser irradiation time and doping percentages of MoS2 are increased. As a result, a small peak indicates a decrement rate of electron–hole pair recombination, which increases the capacity for separation. Thermo-gravimetric analysis and differential thermal analysis results revealed that weight loss decreased from 0.69 to 0.35 mg and thermal stability increased with increased doping concentrations. Methylene blue was degraded in 150 min, proving that the prepared MoS2-doped V2O5 material was a stable and economically low-cost nanocomposite for photocatalytic activity.

1 Introduction

The world population is increasing, so clean, purified drinkable water demand is also increasing day by day because clean water is essential for the survival of human life. The establishment of industries is also increasing to fulfill the requirements of the global population necessary daily life things but their effluent contains different organic and inorganic pollutants that affect the quality of drinking water and eventually disturb the environment and ecology system [1,2]. Among the different pollutants of industrial discharge, non-biodegradable organic and inorganic dyes are one of the effluents that cause water pollution, which can have serious repercussions. To address this serious problem, different physical, chemical, and biological techniques were used to treat the water; however, these efforts failed due to the hazardous contaminants could not be fully mineralized or expensive setup [3]. The failure of all approaches has caused researchers to focus on the creation of simple, affordable, and innovative techniques that have the potential to completely remediate wastewater [4,5]. Consequently, photocatalysis based on the advanced oxidation process has emerged as an efficient and promising approach for the degradation of both organic and inorganic pigments [6]. For photocatalytic applications, researchers have focused on heterostructure two-dimensional (2D) transition metal dichalcogenides (TMDs) and transition metal oxides (TMOs) such as phosphates, sulfides, carbides, and nitrides [7,8,9].

The diverse range of applications of TMOs in supercapacitors, sensors, transistors, and photocatalysis has garnered significant interest. Vanadium pentoxide (V2O5) is among the most extensive and potential materials considered as a TMOs. V2O5 is a common TMO material that has good ion or molecular interaction, is inexpensive, naturally abundant, and has a high guest cation density [10,11]. However, low electrical conductivity, sluggish electrochemical kinetics, poor electrochemical stability, and substantial volume expansion during cycling are problems with bulk V2O5 with dense shapes [12].

Among the various 2D photocatalyst materials, layered structured molybdenum disulfide (MoS2) has attained the most attention due to earth-abundant composition, increased stability, and proficient activity to the improvement of various MoS2-based photocatalysts [13]. Nevertheless, the transition metal sulfides’ reduced stability and active surface sites severely limited their efficiency to function as effective photocatalysts [14,15]. Therefore, MoS2, in combination with other materials, like V2O5, forms hetero-interfaces that benefit from the abundance of V═O and can enhance active intermediates (OOH*) by controlling the valence electron structure of V element and stabilizing the oxygen sites within the atomic network [16]. This is made possible by the subsistence of multivalent states of the V element. TMD materials such as MoS2 with the general formula AX2 where A is a transition metal such as Ni, Cr, Co, Ru, and Mo, and X indicates Te, Se, or S. Many studies have been found to reduce MoS2 limitations by creating a composite or heterogeneous structure to reduce recombination rates, increasing conductivity by inclusion with anion, and charge carrier transformation [17,18]. The chemical formulation of MoS2 is about 59.94% molybdenum (Mo) and 40.05% sulfur (S) with weight percentages, respectively. MoS2 is 5.069 g cm−3 denser and possesses a molar mass of about 160.07 g mol−1. MoS2 can produce dry lubricating coating and generally possesses outstanding chemical and thermal stability [19,20]. The catalytic activity of MoS2 nanoparticles (NPs) is highly conductive, and their physical characteristics are outstanding [21]. MoS2 is more reactive, has a greater adsorption capacity, and has a larger active surface area than bulk materials. Although MoS2 material is stable in aqueous conditions, its applicability in the solar spectrum is limited. An excellent photocatalyst must have the following essential characteristics: maximal absorption with extensive surface active site area in the visible light spectrum, ideal band edges for initiating reactions, friendliness toward the environment, cheap cost, good stability, and sustainability [22,23,24,25].

When V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites are bombed with a radiant light source (450 nm continuous diode laser), the nanocomposites resistance becomes less intense with small resistance to the nanocomposite pellets. Excessive exposure to laser beams leads to tuning the distinctive properties [26]. The higher binding rate is due to the higher energy given through laser exposure to the nanocomposites, causing its chemical and structural changes. The restructuring process may be enhanced by incorporating MoS2 NPs into the V2O5 matrix for light absorption due to tuning of energy bandgap.

Here, in the present research, the hydrothermal technique was used to synthesize 2D-layered heterostructure V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites with high uniformity, homogeneity, and controllable size. The novelty of this study is that a continuous diode laser has been used to irradiate MoS2-doped V2O5 nanocomposites for 2, 4, 6, 8, and 10 min, respectively, to tune various properties by reconstructing material layers for enhancement of photocatalytic applications. According to photocatalytic studies, the large active sites for light absorption, decrease in the electron–hole pair recombination rate, and increase in charge transportation, respectively, enable the as-prepared V2O5(1−x)MoS2(x) (X = 1–5% w/w) composite catalysts to exhibit effective catalytic performance for water degradation. When compared to its separate materials, the 2D heterostructured layered V2O5(1−x)MoS2(x) (X = 1–5% w/w) composite exhibits low over-potential and high stability. Therefore, due to its excellent conductivity, highly exposed catalytically active sites, and the cooperatively formed reconstructed 2D-layered heterostructures MoS2-doped V2O5, it exhibits superior activity that is easily reproducible for a variety of applications such as carbon dioxide (CO2) reduction, water splitting, hydrogen (H2) production, nitrogen (N2) fixation, etc. Thus, synthesized layered heterostructures V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites demonstrated exceptional increased electrochemical characteristics and great photocatalytic applicability.

2 Experimental method and laser setup

2.1 Preparation of sample

Hydrothermal process was used to synthesize a V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposite. For the preparation of the MoS2-doped V2O5 nanocomposite, MoS2 and V2O5 NPs was prepared separately. First, dissolve 2.285 g ammonium heptamolybdate tetrahydrate and 1.1 g thiourea in 70 mL of de-ionized (DI) water, which was then agitated for 30 min. With the use of a 500 mL Teflon line autoclave, the resultant mixture was placed in the furnace at 180–200°C for 48 h to eliminate residual solution, as shown in Figure 1. After being allowed to cool to ambient temperature, the obtained samples were cleaned with ethanol and DI water and stored in centrifuge tubes. After being baked for 12 h at 60°C, the final MoS2 precipitate samples were dried. For the preparation of V2O5 NPs, 1 g of salt ammonium metavanadate NH4VO3 was dissolved in 20 mL of hydrogen peroxide. Furthermore, this mixture was kept in a magnetic stirrer for stirring for 60 min. After stirring, a homogeneous mixture was obtained. The homogeneous mixture was autoclaved for 48 h at 200˚C, as shown in Figure 1. HNO3 was used to maintain the pH of the solution. The obtained sample was placed in open air to cool. In the end, the obtained precipitate was dried in the oven for about 6 h at 80°C. The final product V2O5 NPs was dried into nanopowder form. For the synthesis of V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposite, V2O5 and MoS2 were first synthesized in a nanomaterial form. These nanomaterials were dissolved in 20 mL ethanol solution. Furthermore, the obtained homogeneous mixture was stirred using a magnetic stirrer for 12 h at room temperature. After stirring, the obtained product was placed in a sonicater for 2 h. The obtained sample was dried in an oven at 60°C for 12 h. In the end, the nanocomposite of MoS2-doped V2O5 precipitate was calcinated for 3 h at 250°C.

Figure 1 
                  Experimental diagram of hydrothermally synthesized V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites.
Figure 1

Experimental diagram of hydrothermally synthesized V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites.

2.2 Laser exposure setup

The continuous diode laser had a maximum mean laser power of 20 W. The continuous pulse produced visible light with a wavelength of 450 nm, as shown in Figure 2. The laser and MoS2-doped-V2O5 nanocomposites pallets were separated by 6 cm, allowing the continuous laser light to cover almost 1 × 1 cm2 of the pallet surface. The continuous laser lights were irradiated at the MoS2-doped V2O5 nanocomposite pallets. The duration of laser exposure was 2, 4, 6, 8, and 10 min. Laser irradiation can tune the energy band gap and light absorption for photocatalytic applications.

Figure 2 
                  The experimental setup of laser irradiation into V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposite for tuning of different properties.
Figure 2

The experimental setup of laser irradiation into V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposite for tuning of different properties.

3 Results and discussion

3.1 XRD phase analysis

A Philips X-ray diffraction was used to capture X-ray diffraction (XRD) patterns using Ni-filtered Cu-Kα radiation. The average crystallite size and phase formation of laser-irradiated pure V2O5 and V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites were observed using a X-ray diffraction, as shown in Figure 3. The crystallite sizes of pure V2O5 and MoS2-doped-V2O5 nanocomposites decreased from 42.53 to 22.89 nm with increasing MoS2 percentage as well as laser irradiation exposure time, respectively, as shown in Table 1. The observed peaks at 2θ values 10.12°, 16.45°, 20.15°, 24.45°, 29.06°, 33.32°, and 44.35° are well coordinated to the (002), (200), (001), (101), (110), (400), (011), and (210) hkl plane of MOS2-doped-V2O5, which are in accordance with JCPDS No. 37-1492 and No. 41-1426, respectively, as shown in Figure 3. There were no impurity peaks, indicating the purity of the V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites and their excellent crystalline nature. Weight percentage increases caused the distinctive peaks to shift toward a lower angle. The Debye–Scherrer equation was utilized to determine the NPs’ crystallite size [27].

(1) D = K λ β cos θ ,

where β is the full width at half maximum (FWHM) of the XRD peak that appears at the diffraction angle θ, D is the crystallite size, θ is the Bragg diffraction angle, and λ is the X-ray wavelength (Cu-Kα radiation = 1.54 Å) [28]

(2) d = λ 2 sin θ ,

where ϴ is the Bragg diffraction angle, λ is the X-ray wavelength (Cu-Kα radiation = 1.54 Å), and d is the d-spacing. Lattice constants values increased slightly due to the increased weight percentage of MoS2 in the V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites. Thus, an increase in the lattice constant is associated with a lattice increment in cell volume (V cell) as the weight percentage of MoS2 in the crystal structure increases.

Figure 3 
                  The phase analysis pattern of laser irradiated V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites.
Figure 3

The phase analysis pattern of laser irradiated V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites.

Table 1

Structural parameters of laser irradiated V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposite

Nanocomposites Irradiation time (min) 2θ (°) FWHM Miller indices (hkl) d-spacing (Å) Average crystallite size D (nm)
Pure-V2O5 0 13 0.1322 002 2.6421 42.53
V2O5(1−x)MoS2(x) (X = 1% w/w) 4 15 0.3685 200 2.5341 36.66
V2O5(1−x)MoS2(x) (X = 2% w/w) 8 20 0.4557 001 2.34567 32.33
V2O5(1−x)MoS2(x) (X = 3% w/w) 12 26 0.7548 101 2.01489 28.44
V2O5(1−x)MoS2(x) (X = 4% w/w) 16 31 0.8811 400 1.96289 25.24
V2O5(1−x)MoS2(x) (X = 5% w/w) 20 32 0.8956 011 1.95161 22.89

Also, as observed by XRD peaks, the average particle size of laser-irradiated pure V2O5 and V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites tends to decrease, increasing surface area and decreasing the rate at which electron–hole pairs recombine. This increases photon light absorption and accelerates the degradation of organic dyes. As a result, photocatalysis would be more effective at gathering light and be better able to break down industrial colors like methylene blue (MB) in contaminated water.

3.2 Elemental compositional analysis (EDX)

The chemical compositions of laser irradiated pure V2O5 and V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites are analyzed hydrothermally using EDX spectroscopy, as illustrated in Figure 4(a–f). The presence of vanadium (V), oxygen (O), and gold (Au) with respective atomic weights of 70.84, 28.36, and 0.80% is shown by the EDX peaks of pure V2O5 in Figure 4(a). The X-ray diffraction peaks in the V2O5 lattice system demonstrate that MoS2 replacements were successful, and this was validated by the acquired EDX data. Vanadium (V), oxygen (O), aluminum (Al), molybdenum (Mo), sulfur (S), and gold (Au) with atomic weight percentages of 80.89, 16.25, 1.87, 2.09, and 0.90% are displayed by the EDX peaks in Figure 4(b). The vanadium (V), oxygen (O), aluminum (Al), molybdenum (Mo), sulfur (S), and gold (Au) EDX peaks in Figure 4(c–f) demonstrate that these elements have varying atomic weight percentages as shown in Table 2.

Figure 4 
                  (a–f) The EDX of laser irradiated V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposite.
Figure 4

(a–f) The EDX of laser irradiated V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposite.

Table 2

The elemental mapping of laser irradiated V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposite

Materials name (V) (O) (Mo) (S) (Au)
Laser-irradiated-pure-V2O5 70.84 28.36 0.00 0.00 0.80
Laser-irradiated V2O5(1−x)MoS2(x) (X = 1% w/w) 78.89 16.25 1.87 2.09 0.90
Laser-irradiated V2O5(1−x)MoS2(x) (X = 2% w/w) 75.89 19.25 2.13 2.09 0.64
Laser-irradiated V2O5(1−x)MoS2(x) (X = 3% w/w) 55.91 35.07 3.90 3.22 0.90
Laser-irradiated V2O5(1−x)MoS2(x) (X = 4% w/w) 46.42 36.09 7.73 9.68 0.08
Laser-irradiated V2O5(1−x)MoS2(x) (X = 5% w/w) 42.54 35.17 10.93 9.80 1.67

3.3 X-ray photoelectron spectroscopy (XPS)

The surface texture of a MoS2-doped V2O5 nanocomposite was examined using XPS in relation to the presence of various components, chemical structures, chemical oxidation states, and electron movement. As seen in Figure 5(a), MoS2 has two peaks at 228.9 and 232.1 eV, which correspond to Mo4+ 3d5/2 and Mo4+ 3d3/2, respectively. These peaks demonstrate the presence of Mo4+, which is an indicator of MoS2. The shift in the binding energy of Mo 3d in MoS2-doped V2O5 composite further indicates a strong synergistic effect between Mo and V metals in MoS2-doped V2O5 composite photocatalyst and suggests a notable electronic structure difference in Mo between MoS2 and MoS2-doped V2O5 composite photocatalyst. The presence of S2− in MoS2 is demonstrated by the peaks for S 2p3/2 at approximately 161.7 eV and S 2p1/2 at 162.9 eV in Figure 5(c). The four peaks are associated with the V 2p spectrum, which is shown in Figure 5(b). Two of the peaks correspond to the V5+ values V 2p1/2 at 524.5 eV and V 2p3/2 at 516.9 eV. The remaining two significant peaks, which are located at 523.1 eV and 515.3 eV, respectively, are consistent with 2 p1/2 and 2 p3/2 of V4+ in V2O5. It is possible to assume that V5+ predominates in the synthesized nanocomposites since V5+ peak intensity and area are much bigger than V4+ peak. Additionally, as demonstrated in Figure 5(d), the notable peaks of O 1 s at 529.7 and 531.2 eV are indicative of the presence of the V−O functional group of V2O5.

Figure 5 
                  (a–d) XPS spectrum of V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposite of (a) Mo 3d (b) V 2p (c) S 2p and (d) O 1s.
Figure 5

(a–d) XPS spectrum of V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposite of (a) Mo 3d (b) V 2p (c) S 2p and (d) O 1s.

3.4 Morphological analysis

Using a JEOL IT800 model of field-emission scanning electron microscopy (FESEM) was used to analyze the surface morphology and shape of hydrothermally prepared laser-irradiated pure V2O5 and V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites, as shown in Figure 6(a–f). Before being examined with a FESEM, the samples were polished to boost the emissivity of the samples and then gold coated. The samples of V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites were shown to have layered structures in the scanning electron microscopy images. The surface morphology of pure V2O5 and V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites shows that part of the V2O5 nanospheres pierced into the MoS2 layered matrix, while the rest of the nanospheres are arranged in bunches on MoS2 layers. The average grain size of the laser-irradiated doped composites V2O5(1−x)MoS2(x) (X = 1–5% w/w) decreased as the doping fraction of MoS2 increased, but the nanocomposites’ shape did not change. Furthermore, small grains emerge as a result of increased nucleation as the V2O5 is integrated into the MoS2 layered structure. The non-homogeneous distribution of V2O5 NPs on the MoS2 layered structure’s surface is seen in Figure 6(e–f), and this phenomenon might be related to electrostatic attraction. Additionally, as seen in Figure 6(e–f), the smaller grain size increases surface area, which in turn causes a decrease in the rate at which photogenerated charge carriers recombine and an increase in photon light absorption, which accelerates the degradation of organic dyes.

Figure 6 
                  (a–f) Surface morphology analysis of laser irradiated V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposite.
Figure 6

(a–f) Surface morphology analysis of laser irradiated V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposite.

3.5 Photocatalytic performance under UV–Visible light

The photocatalytic performances of laser-irradiated pure V2O5 and V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites for MB is shown in Figure 7(a–h), measured using a U-3900H spectrometer. The absorption peaks of MB for 0, 30, 60, 90, 120, and 150 min are described in the UV–Visible investigation. Approximately 95% of the MB was degraded after 150 min, and the sharpness of absorption peaks steadily decreased with increasing degradation time of the organic dyes. The maximum MB absorption wavelength shifted from 0 to 150 min as the degradation time increased, as seen by the UV–Visible absorption peaks. This indicates that the structure of MB within 150 min was degraded and the photocatalytic degradation of MB is a mineralization process. To establish an adsorption–desorption equilibrium before photocatalytic degradation, the catalyst adsorbed MB for 30 min while being protected from light. The “dark” tests’ findings indicated that the small amount of MB degradation that takes place in the absence of UV–Visible light is caused by photocatalytic degradation. From Figure 7(a–h), it can be observed that in the presence of a catalyst of pure V2O5 and V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites, the degradation percentages of MB were 40.11, 51.40, 70.30, 80.75, and 95.40%, respectively. The degradation percentages of MB are increased with an increase of the doping percentage of MoS2 from 1 to 5% w/w as well as laser irradiation. According to this, the amount of MB that is degraded by catalyst V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites is more than it is when using pure V2O5. After 120 min of remediation, the greatest deteriorated percentage (95.40%) was tested using catalyst V2O5(1−x)MoS2(x) (X = 1–5% w/w) composites. This nanocomposite demonstrated improved photocatalytic activity. This is consistent in line with the analysis’s conclusions, which are shown in Figure 7(a–h). Based on the simulated curve, there is a solid linear connection between the degradation time and ln(C t/C 0). A pseudo-first-order kinetic model with a degree of fit of 0.9501 (R 2) illustrates the photocatalytic degradation of MB in the following formulas [29]:

(3) ln C t C o = kt ,

where k is the rate constant, C t is the MB concentration at time t, and C 0 is the MB starting concentration. The k value of MB was calculated to be 0.0021 min−1 for complete V2O5(1−x)MoS2(x) (X = 1–5% w/w) samples.

Figure 7 
                  (a–h) Photocatalytic activity of laser irradiated V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposite.
Figure 7 
                  (a–h) Photocatalytic activity of laser irradiated V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposite.
Figure 7

(a–h) Photocatalytic activity of laser irradiated V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposite.

3.6 Photoluminescence (PL)

PL characteristic was performed, as shown in Figure 8, to investigate the ability of electrons ( e ) and hole ( h + ) pairs separation. According to PL’s findings, a lower photocatalyst intensity (PL) peak is indicative of a slower rate of electron–hole recombination, which increases separation capacity. This result indicates that the dye-degrading capacity of laser irradiated V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites is growing with increasing substitution percentage and shows well-organized charge transfer on the photocatalyst surface. The laser irradiated V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites’ PL peaks clearly show a decrease in intensity as the MoS2 doping percentage increases. This indicates that the laser-irradiated MoS2-doped-V2O5 nanocomposites have higher photogenerated electrons ( e ) and hole ( h + ) pairs, which enhances photocatalytic application.

Figure 8 
                  PL spectra of laser irradiated V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites.
Figure 8

PL spectra of laser irradiated V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites.

3.7 Fourier transformation infrared spectroscopy (FTIR)

A Perkin Elmer spectrum 100 FTIR spectrometer is used to determine potential functional groups present in laser-irradiated V2O5(1−x)MoS2(x) (X = 1–5% w/w) and pure V2O5 nanocomposites. The crystal vibration modes and ion locations of laser-irradiated MoS2-doped-V2O5 nanocomposites may be determined using the FTIR approach. The FTIR method may be used to determine the ion locations and crystal vibration modes of both laser-irradiated pure V2O5 and V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites. Figure 9 shows the FTIR spectra of laser irradiated pure V2O5 and V2O5(1−x)MoS2(x) (X = 1–5% w/w) that were obtained between 250 and 3,200 cm−1. Vanadium–oxygen and molybdenum–sulfur bonds bending and stretching correspond to the unique characteristics of the 250–3,200 cm–1. The large band at 843–859 cm−1 is created by the O–H stretching from water, whereas the absorption bands at 941 cm–1 are ascribed to the OH bending. The absorption bands at 1,014–1,025 cm−1 are responsible for the C–H bending. The small bands at 2,850–2,872 and 3,015–3,028 cm−1 are created by Mo–S and Mo═O by the MoS2 are described by bending bands. The vanadium-oxygen bonds were shifted by the MoS2 doping and laser irradiation. As shown in Table 3, the absorption peaks are shifted toward higher wavelengths as a result of an increase in MoS2 doping concentration and laser irradiation time. While doped MoS2-doped-V2O5 is assigned to the bands at 537 to 617 cm−1 range, associated with symmetric stretching vibration of V–O–V is located at band 635–662 cm−1 for pure-V2O5. The optimal integration of MoS2 into the V2O5 host material results in an increase in transmittance intensities. The increase in intensities leads to the stability of laser-irradiated V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites.

Figure 9 
                  Functional group analysis of laser irradiated V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites.
Figure 9

Functional group analysis of laser irradiated V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites.

Table 3

Different functional groups of laser irradiated V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites

Functional groups Laser-irradiated pure-V2O5 (cm−1) Laser-irradiated MoS2-doped-V2O5 (1%) (cm−1) Laser-irradiated MoS2-doped-V2O5 (2%) (cm−1) Laser-irradiated MoS2-doped-V2O5 (3%) (cm−1) Laser-irradiated MoS2-doped-V2O5 (4%) (cm−1) Laser-irradiated MoS2-doped-V2O5 (5%) (cm−1)
O–H 843 846 849 853 856 859
V–O–V 633 636 640 646 654 658
O═V 456 459 461 465 467 470
Mo–S None 2,850 2,856 2,860 2,868 2,872
Mo═O None 3,015 3,017 3,019 3,025 3,028

3.8 Atomic force microscopy (AFM)

The AFM results of laser irradiated pure V2O5 and V2O5(1–x)MoS2(x) (X = 1–5% w/w) nanocomposites are seen in three dimensions by using a BRUKER model dimension edge with scan system as shown in Figure 10(a–e). Table 4 presents the topographical characteristics, including image surface area, image projected surface area, average roughness (R a), and root mean square roughness (R q). As shown in Table 4, increasing the doping concentrations of V2O5(1−x)MoS2(x) (X = 1–5% w/w) results in a decrease in roughness of 58.70, 41.36, 32.11, 29.23, 26.21, and 21.22 nm. When MoS2 doping weight percent of MoS2 is increased, the surface area of layered structured nanocomposites increases from 108, 125, 133, 141, 151, and 169 μ m 2 . By increasing the weight percent of MoS2 in V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites, the number of layers is likewise growing. According to AFM results, increment in active site surface area, which in turn causes a decrease in the rate at which photogenerated charge carriers recombine and an increase in photon light absorption, which accelerates the degradation of organic dyes.

Figure 10 
                  (a–f) Surface analysis through AFM of laser irradiated V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites.
Figure 10 
                  (a–f) Surface analysis through AFM of laser irradiated V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites.
Figure 10

(a–f) Surface analysis through AFM of laser irradiated V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites.

Table 4

Different surface parameters of laser irradiated V2O5(1−x)MoS2(x) (X = 1–5% w/w)

Samples Image surface area ( μ m ) 2 Image projected surface area ( μ m ) 2 Root mean square roughness R q (nm) Average roughness R a (nm)
Laser-irradiated pure-V2O5 108 100 58.70 71.33
Laser-irradiated V2O5(1−x)MoS2(x) (X = 1% w/w) 125 100 41.36 35.41
Laser-irradiated V2O5(1−x)MoS2(x) (X = 2% w/w) 133 100 32.11 29.41
Laser-irradiated V2O5(1−x)MoS2(x) (X = 3% w/w) 141 100 29.23 27.44
Laser-irradiated V2O5(1−x)MoS2(x) (X = 4% w/w) 151 100 26.21 25.33
Laser-irradiated V2O5(1−x)MoS2(x) (X = 5% w/w) 169 100 21.22 18.14

3.9 Dielectric characteristics

An impedance analyzer called the keysight E49918 was used to evaluate the sample’s electrical conductivity for MoS2-doped-V2O5 (about 76 mm thick and 13 mm diameter) at a frequency range of 1 MHz to 3 GHz. The pressure was adjusted by using Keysight 16453A with 1 MHz–1 GHz dielectric material test fixture and ±42 peak output. The Keysight E4991B test head was used as an RF out port. The conductivity measurements were completed at room temperature by using σ = d/AR. The values of ionic conductivity were calculated using the formula where d is the sample thickness, A is the electrode area, and R is the sample resistance.

3.9.1 Dielectric constants

Generally, dielectric relaxation originates from the mobility of the electric dipole that the applied electric field produces. The Debye–Scherer relaxation model has been used to describe the effects of applied electric field on dielectric materials. The following formula is used to calculate the complex dielectric constant [30]:

(4) e * = e + j ϵ .

Dielectric materials can be used to improve the capacity of charge storage. As a result, the capacitance of the material and its dielectric constant are proportionate [31].

(5) ϵ = Cd o A ,

where A stands for area, C for capacitance, d for thickness, and o for the permittivity of free space. The thickness for V2O5(1−x)MoS2(x) (X = 1–5% w/w) composite pellets with doping different percentages 1–5% w/w were 1.58, 1.41, 1.68, 1.74, and 1.82 nm, respectively, measured by digital Vernier calipers. The variation of real dielectric function (RDF) (ε′) and imaginary dielectric function (IDF) (ε″) with the frequency of laser irradiated of pure V2O5 and MoS2-doped-V2O5 nanocomposites as shown in Figure 11. The dielectric constant (RDF) ε′ is quite high at low frequencies and decreases sharply with increasing frequency and MoS2 doping percentages. Dipolar polarization, interfacial polarization (IP), atomic polarization, and electronic polarization are the four polarization types that affect a material’s dielectric behavior. Dipolar and IPs are highly dependent on frequency and temperature, whereas electronic and ionic polarizations occur at very high frequencies. Charge carrier is affected by applied frequency variations because these polarization processes have a distinct relaxation time. All polarization processes are involved at low applied frequencies, but as the frequency increases, polarization decreases. As the frequency increased, the different dipoles inside the V2O5 and V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites were unable to follow the electric field; thus, the polarization reduced and therefore ε′ decreased. Laser-irradiated pure V2O5 and V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites’ distinct dipoles were unable to follow the electric field as the frequency increased, which resulted in a decrement in polarization and dielectric constants (RDF) ε′ decreased with increased of doping percentages. The dielectric constant (RDF) ε′ values were found to be dependent on the strength of the electrostatic interaction force between the V2O5 and V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites (synthesized at varying doping percentages) and the functional group in the blend chain of V2O5 and V2O5(1−x)MoS2(x) (X = 1–5% w/w), which facilitates/restricts molecular movement. As a result, in comparison to pure V2O5, the doped nanocomposites’ effective dielectric polarization at 5% was either enhanced/decreased.

Figure 11 
                     (a, b) Dielectric constants of laser irradiated of V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites.
Figure 11

(a, b) Dielectric constants of laser irradiated of V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites.

Furthermore, the IP peak, also known as the Maxwell–Wagner–Sillars effect, was visible in the lower frequency range of the IDF (ε″) of laser-irradiated V2O5 and V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites. The values of (IDF) ε″ were increased as laser-irradiated V2O5 while they reduced as V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites at 5% of MoS2 with irradiation of 5 min. The decline in crystalline sizes created the interface area of laser-irradiated V2O5 and V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites which caused the IP reduced and hence the ε″ improved. An external power supply source connected to a parallel plate capacitor results in a phase angle of 90° between the current and voltage, which leads to current leakage and power dissipation. Thus, the following formula can be used to calculate the tangent loss [32]:

(6) tan δ = 1 2 π f C p R p ,

where tan δ represents the loss tangent or dielectric loss, C p is the parallel capacitance, R p is the parallel resistance, and 2πf is the angular frequency. Zero loss angle and zero power consumption characterize the perfect capacitor. Power dissipations, often called dielectric loss in commercial capacitors, would be assessed. The variation in dielectric loss (tan δ) as a function of frequency (f) for laser-irradiated pellets that produced pure V2O5 and MoS2-doped-V2O5 nanocomposites with varying MoS2 doping is shown in Figure 12. According to Koop’s Phenomenological Theory, these patterns are consistent with the Maxwell–Wagner interface polarization model. The smooth grain is more active at high frequencies, whereas the grain boundary with insufficient conductivity is more efficient in the low-frequency range due to internal morphological defects. Figure 3 illustrates the relaxation peaks for various components that appear at various frequencies. Each polarization mechanism has a unique relaxation frequency, and polarization resonance states that resonance occurs when the relaxation frequency and applied frequency are matched. Therefore, the relaxation phenomena of studied samples are responsible for the existence of different component peaks. The relaxation peak changes toward low frequency when the content of MoS2 increases, suggesting that the relaxation duration may increase. These findings of RDF (ε′) and IDF (ε″) and tangent loss of laser-irradiated pure V2O5 and MoS2-doped-V2O5 nanocomposites are appropriate for photocatalytic activity.

Figure 12 
                     Tangent loss of laser irradiated of V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites.
Figure 12

Tangent loss of laser irradiated of V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites.

3.9.2 Conductivity analysis

The movement of a charge carrier in response to an applied field is known as conductivity. At lower 1 MHz frequency, conductivity is at high 1.25 × 10−11 S m−1 for pure V2O5 and 3.6 × 10−11 S m−1 for V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposite for 5% doping percentage. However, conductivity gradually decreases at higher frequencies as shown in Figure 13. A small conductivity is found at higher frequencies due to the complex resistive nature of grain boundaries. Jonsher’s Power Law can be used to determine the net conductivity of ceramic materials [33,34].

(7) σ total = σ dc A ω s ,

where s is the exponent, A is the pre-exponential factor, and σ dc is the DC conductivity. The entire design of A ω s is included in the term “ac conductivity.” The ac conductivity may be computed using the following formula [35]:

(8) σ ac = ε ε 0 ω tan δ .

Figure 13 
                     Electric conductivity of laser irradiated of V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites.
Figure 13

Electric conductivity of laser irradiated of V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites.

Here, angular frequency is represented by (ω = 2πf). Figure 4 displays a variation of conductivity (σ ac) versus frequency (f) for laser-irradiated pure V2O5 and V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites with different doping concentration of MoS2. The change in ac conductivity is large at first, but as the frequency increases, the conductivity decreases gradually with an increment of doping concentration of MoS2.

3.9.3 Quality factors

The Q factor fluctuation with frequency (f) for laser-irradiated pure V2O5 and V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites with varying MoS2 doping concentrations is displayed in Figure 14. The dielectric properties of samples of nanocomposites were studied at a frequency of 1 GHz. The Q factor of laser-irradiated pure V2O5 and MoS2-doped V2O5 nanocomposites at higher frequency increases with increasing MoS2 doping weight percentage.

Figure 14 
                     Quality factor of laser irradiated of V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites.
Figure 14

Quality factor of laser irradiated of V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites.

3.9.4 Complex impedance analysis

An effective method for investigation of the function of grains and grain boundaries as well as the polarization process is impedance spectroscopy. A material’s dielectric response is mostly dependent on the resistance and capacitance values of its microstructures, which affect certain solid characteristics. This method enables determining the relaxation time and frequency as well as the resistance and capacitance provided by the bulk and grain boundaries. In complex form, the frequency dependency of impedance may be expressed as [36,37]

(9) Z * = Z + JZ ,

(10) tan δ = ε ε = Z Z ,

(11) Z * = ε C o ω ( ε 2 + ε 2 ) + j ε C o ω ( ε 2 + ε 2 ) .

Figure 15 demonstrates the frequency dependence of real impedance, which shows a diminishing trend as frequency increases. Real impedance’s frequency dependency is seen in Figure 15, where an increasing frequency is accompanied by a decreasing trend.

Figure 15 
                     (a, b) Impedance analysis of laser irradiation into V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposite.
Figure 15

(a, b) Impedance analysis of laser irradiation into V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposite.

As the frequency increases, the impact of conductive grains causes it to decrease. According to the investigation, as frequency and MoS2 doping percentage increase, Z magnitude decrement, suggesting that the sample’s conductivity ( σ dc ) increases. Real impedance Z gradually drops at low frequencies as MoS2 substitution increases. Furthermore, it is shown that the MoS2 concentration increases at low frequencies at which the real impedance Z peaks meet. Conversely, Figure 15 shows that imaginary impedance Zʺ at a higher value with the increment of the frequency with MoS2 doping concentration.

3.10 Thermogravimetric analysis (TGA) and differential thermal analysis (DTA)

TGA and DTA are efficient studies to understand materials’ thermal stability and relative weight loss, respectively. TGA is a quantitative and qualitative investigation of mass to temperature or time as a function. Figure 16(a–f) shows TGA thermograms of pure V2O5 and 2D heterostructured layered V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites. Thermal investigation revealed a three-phase loss in weight. The first loss in weight below 685°C can be attributed to the removal of physisorbed water molecules and other ions found in pure V2O5 and V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites. The second thermal degradation occurs in between 687 and 710°C temperature range as a result of the elimination of other ions and the dehydroxylation of metal hydroxide. The oxidation and loss cause the third thermal deterioration, which occurs between 712 and 1,000°C. The weight loss was decreased from 0.69 to 0.35 mg, and thermal stability increased with the increased doping percentage in V2O5(1−x)MoS2(x) (X = 1–5%w/w) nanocomposites, as shown in Table 5. According to these findings, the pure V2O5 and V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites were created stable at high temperatures.

Figure 16 
                  (a–f) Thermal gravimetric  and differential thermal analysis of laser irradiated of V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites.
Figure 16

(a–f) Thermal gravimetric and differential thermal analysis of laser irradiated of V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites.

Table 5

Thermal stability and weight loss parameters of laser irradiated of V2O5(1−x)MoS2(x) (X = 1–5% w/w) nanocomposites

Material First exothermic peak 1st total weight loss (mg) Second exothermic peak 2nd total weight loss (mg) Total weight loss (mg) Temperature (°C)
Laser-irradiated pure-V2O5 0.13 0.56 0.69 683
Laser-irradiated V2O5(1−x)MoS2(x) (X = 1% w/w) 0.39 0.19 0.58 677 713
Laser-irradiated V2O5(1−x)MoS2(x) (X = 2% w/w) 0.18 0.35 0.53 678 712
Laser-irradiated V2O5(1−x)MoS2(x) (X = 3% w/w) 0.1 0.33 0.43 673 713
Laser-irradiated V2O5(1−x)MoS2(x) (X = 4% w/w) 0.1 0.27 0.37 679 714
Laser-irradiated V2O5(1−x)MoS2(x) (X = 5% w/w) 0.10 0.25 0.35 677 713

Figure 16(a–f) depicts the results of the DTA of pure V2O5 and V2O5(1−x)MoS2(x) (X = 1–5% w/w), which demonstrate the existence of one minor exothermic peak at 710°C and two endothermic peaks at 677 and 712°C. The following equation are related oxidation and decomposition [38]:

(12) [ V 4 + ] Oxidation 400 700 ° C [ V 5 + ] ,

(13) [ MoS 2 ] Oxidation 200 500 ° C [ MoO 2 ] decomposition 500 1 , 000 ° C [ Mo 2 O 2 ] .

3.11 Mechanism of the photocatalytic reaction

The generation of electrons ( e ) and hole ( h + ) pairs on the photocatalyst’s surface and their transformation to the active sites are produced by the fundamental mechanism of the photocatalytic reaction. The hole ( h + ) in the valance band (VB) and electron ( e ) in the conduction band (CB) when exposed to photons of UV light equal to or greater than E g. This can only be accomplished by having electrons from the VB transfer into the CB. The electron ( e ) in the CB convert oxygen into superoxide ions ( O 2 ) , whereas water ( H 2 O ) in the organic dyes MB oxidizes to hydroxyl free radicals (OH) due to the positively charged hole in the valance band. The oxidation of MB and the reduction of oxygen do not occur at the same time throughout the remediation process, as presented in Figure 17. Recombination of electrons ( e ) and positive hole ( h + ) is caused by the concentration of ( e ) in the CB The electrons are necessary for the photocatalytic redox reactions to produce superoxide ions ( ˙ O 2 ) , and hydroxyl free radicals (−OH). As during this degradation process organic dye molecules split into CO2 and H2O, and produced more −OH and ˙ O 2 due to newly generated active sites as powerful oxidizers with increase of MoS2 dopants in nanocomposites. Consequently, MB remediation increases at a faster rate. The photocatalytic mechanism’s schematic depiction may be explained by applying formulae to determine the energies of MoS2 and V2O5 VB and CB [39]

(14) E VB = χ E e + ( 0.5 ) ,

(15) E CB = E VB E g .

Figure 17 
                  The photocatalytic mechanism of laser irradiated V2O5/MoS2 nanocomposites.
Figure 17

The photocatalytic mechanism of laser irradiated V2O5/MoS2 nanocomposites.

Free electron energy on the hydrogen scale, band gap, CB, and VB edges are represented, respectively, by variables E e , E g , E CB , and E VB . Additionally, χ represents the electro-negativity of the material.

The possible mechanism of photocatalysts (MoS2-doped-V2O5) with p–n heterojunction subjected to sunlight. Photogenerated carriers are kept apart in the p–n heterojunction by the n-type MoS2 and the p-type V2O5. A depletion zone with positively and negatively charged regions in the V2O5 and MoS2 sides, correspondingly, is created at the heterojunction when n-type and p-type nanocomposites are linked because of their Fermi levels aligning. When sunshine photons are exposed to a semiconductor, photogenerated ( e ) are moved from the V2O5 VB to the MoS2-doped-V2O5 CB. The superoxide radical ( ˙ O 2 ) is created when these electrons join with the oxygen molecules in the dissolved solution. This radical then changes into the highly reactive hydroxide radicals (−OH). Nevertheless, hole transfer from the valance band of V2O5 to MoS2-doped-V2O5 was impeded by the high potential barrier. The H2O molecules oxidize into hydroxyl radicals (−OH) due to the holes in the nanocomposites. Strong oxidizing radicals like these readily oxidize the chemical molecules in MB. Compared to metal oxide equivalents, MoS2-doped-V2O5 nanocomposites show superior photocatalytic activity because of their enhanced efficiency of photogenerated charge carrier separation. Based on the above experimental results, a possible process for the photocatalytic degradation of MB via MoS2-doped-V2O5 is proposed in equations (16)–(20) [40].

(16) Mo S 2 doped V 2 O 5 + h v ( visible light ) Mo S 2 ( h + , e ) + V 2 O 5 ( h + , e ) ,

(17) Mo S 2 ( h + , e ) + V 2 O 5 ( h + , e ) V 2 O 5 ( h + ) + Mo S 2 ( e ) ,

(18) e + O 2 ˙ O 2 ,

(19) h + + H 2 O OH ,

(20) OH + ˙ O 2 + Methylene Blue intermediates CO 2 + H 2 O .

In the MoS2-doped-V2O5 nanocomposites, the photogenerated charge carriers increase because of the low recombination rate as a consequence of injecting charge carriers from materials with high band gaps toward the small band gaps. These results demonstrate that the photocatalytic activity is boosted due to MoS2-doped-V2O5 nanocomposites on MB degradation compared to other separated single materials such as MoS2 and V2O5.

4 Conclusion

It has been discovered that both 2D MoS2 and V2O5, which are classified as TMDs and TMOs, are good photocatalyst materials. The surface of the 2D molybdenum nanolayered matrix was efficaciously decorated with V2O5 NPs. In the range of the UV visible spectrum, the increment in optical conductivity from 3.6 to 14.5 Ω−1 cm−1 with an increase of the active surface from 108 to 169 μm2. The synthesized nanocomposites show an increase in absorbance from 3 to 8.3 a.u. at wavelength 310 nm. According to PL studies, the intensity of peaks decreases when laser irradiation time and doping percentages are increased. As a result, a small peak indicates a decrement rate of electron-hole pair recombination, which increases the capacity for separation.TGA and DTA results revealed that weight loss decreased from 0.69 to 0.35 mg and thermal stability increased with increased doping concentrations. MB was degraded in 150 min, proving that the prepared MoS2-doped-V2O5 material was a stable and economically low-cost nanocomposite for photocatalytic activity.

Acknowledgments

The authors extend their appreciation to Taif University, Saudi Arabia, for supporting this work through project number (TU-DSPP-2024-59).

  1. Funding information: This research was funded by Taif University, Saudi Arabia, Project No. (TU-DSPP-2024-59).

  2. Authors contributions: Muhammad Hasnain Jameel: writing – original draft, data curation, experimental work, visualization, writing – review and editing, investigation, formal analysis, methodology, validation; Aqeela Yasin: review and editing, Samia: review and editing; Mohd Zul Hilmi Bin Mayzan: supervision, review and editing; Muhammad Sufi bin Roslan: review and editing, Fahmiruddin Bin Esa: review and editing, Mohd Arif Bin Agam: Supervision, Review & editing. Khaled Althubeiti: review and editing, Mohammed Aljohani: review and editing. All authors have accepted responsibility for the entire content of this manuscript and approved its submission.

  3. Conflict of interest: The authors state no conflict of interest.

  4. Data availability statement: The datasets generated and/or analysed during the current study are available from the corresponding author on reasonable request.

References

[1] Navyashree GR, Hareesh K, Sunitha DV, Nagabhushana H, Nagaraju G. Photocatalytic degradation performance of Nd3 + doped V2O5 nanostructures. Mater Res Express. 2018;5(9):095007. 10.1088/2053-1591/aad373.Search in Google Scholar

[2] Tan D, Long H, Zhou H, Deng Y, Liu E, Wang S, et al. Effective photocatalytic degradation of methyl orange using V2O5@ZnO nanocomposite under UV and visible irradiations. Int J Electrochem Sci. 2020;15:12232–43. 10.20964/2020.12.75.Search in Google Scholar

[3] Jayaraman V, Sarkar D, Rajendran R, Palanivel B, Ayappan C, Chellamuthu M, et al. Synergistic effect of band edge potentials on BiFeO3/V2O5 composite: Enhanced photo catalytic activity. J Environ Manage. 2019;247:104–14. 10.1016/j.jenvman.2019.06.041.Search in Google Scholar PubMed

[4] Rafique M, Hamza M, Tahir MB, Muhammad S, Al-Sehemi AG. Facile hydrothermal synthesis of highly efficient and visible light-driven Ni-doped V2O5 photocatalyst for degradation of Rhodamine B dye. J Mater Sci Mater Electron. 2020;31(15):12913–25. 10.1007/s10854-020-03844-3.Search in Google Scholar

[5] Jayaraj SK, Sadishkumar V, Arun T, Thangadurai P. Enhanced photocatalytic activity of V2O5 nanorods for the photodegradation of organic dyes: A detailed understanding of the mechanism and their antibacterial activity. Mater Sci Semicond Process. 2018;85:122–33. 10.1016/j.mssp.2018.06.006.Search in Google Scholar

[6] Zeleke MA, Kuo DH. Synthesis and application of V2O5-CeO2 nanocomposite catalyst for enhanced degradation of methylene blue under visible light illumination. Chemosphere. 2019;235:935–44. 10.1016/j.chemosphere.2019.06.230.Search in Google Scholar PubMed

[7] Zhang S, Chen S, Luo Y, Yan B, Gu Y, Yang F, et al. Large-scale preparation of solution-processable one-dimensional V2O5 nanobelts with ultrahigh aspect ratio for bifunctional multicolor electrochromic and supercapacitor applications. J Alloys Compd. 2020;842:155882. 10.1016/j.jallcom.2020.155882.Search in Google Scholar

[8] Menezes WG, Reis DM, Benedetti TM, Oliveira MM, Soares JF, Torresi RM, et al. V2O5 nanoparticles obtained from a synthetic bariandite-like vanadium oxide: Synthesis, characterization and electrochemical behavior in an ionic liquid. J Colloid Interface Sci. 2009;337(2):586–93. 10.1016/j.jcis.2009.05.050.Search in Google Scholar PubMed

[9] Zhou F, Yan C, Sun Q, Komarneni S. TiO2/Sepiolite nanocomposites doped with rare earth ions: Preparation, characterization and visible light photocatalytic activity. Microporous Mesoporous Mater. 2019;274:25–32. 10.1016/j.micromeso.2018.07.031.Search in Google Scholar

[10] Zia J, Kashyap J, Riaz U. Facile synthesis of polypyrrole encapsulated V2O5 nanohybrids for visible light driven green sonophotocatalytic degradation of antibiotics. J Mol Liq. 2018;272:834–50. 10.1016/j.molliq.2018.10.091.Search in Google Scholar

[11] Jayaraj SK, Thangadurai P. Surface decorated V2O5 nanorods with Pt nanoparticles for enriched visible light photocatalytic performance for the photodegradation of Rh-6G. J Mol Liq. 2020;319:114368. 10.1016/j.molliq.2020.114368.Search in Google Scholar

[12] Pooseekheaw P, Thongpan W, Panthawan A, Kantarak E, Sroila W, Singjai P. Porous V2O5/TiO2nanoheterostructure films with enhanced visible-light photocatalytic performance prepared by the sparking method. Molecules. 2020;25(15):1–11. 10.3390/molecules25153327.Search in Google Scholar PubMed PubMed Central

[13] Zhang X, Teng SY, Loy ACM, How BS, Leong WD, Tao X. Transition metal dichalcogenides for the application of pollution reduction: A review. Nanomaterials. 2020;10(6):1012. 10.3390/nano10061012.Search in Google Scholar PubMed PubMed Central

[14] Gao XT, Liu YT, Zhu XD, Yan DJ, Wang C, Feng YJ, et al. V2O5 nanoparticles confined in Three−Dimensionally organized, porous Nitrogen−Doped graphene frameworks: Flexible and Free−Standing cathodes for high performance lithium storage. Carbon N Y. 2018;140:218–26. 10.1016/j.carbon.2018.08.060.Search in Google Scholar

[15] Mishra A, Panigrahi A, Mal P, Penta S, Padmaja G, Bera G, et al. Rapid photodegradation of methylene blue dye by rGO- V2O5 nano composite. J Alloys Compd. 2020;842:155746. 10.1016/j.jallcom.2020.155746.Search in Google Scholar

[16] Yuan Y, Guo Rt, Hong Lf, Ji Xy, Li Zs, Lin Zd, et al. Recent advances and perspectives of MoS2-based materials for photocatalytic dyes degradation: A review. Colloids Surf A Physicochem Eng Asp. 2021;611:125836. 10.1016/j.colsurfa.2020.125836.Search in Google Scholar

[17] Huang S, Chen C, Tsai H, Shaya J, Lu C. Photocatalytic degradation of thiobencarb by a visible light-driven MoS2 photocatalyst. Sep Purif Technol. 2018;197(2017):147–55. 10.1016/j.seppur.2018.01.009.Search in Google Scholar

[18] Wu MH, Li L, Liu N, Wang DJ, Xue YC, Tang L. Molybdenum disulfide (MoS2) as a co-catalyst for photocatalytic degradation of organic contaminants: A review. Process Saf Environ Prot. 2018;118:40–58. 10.1016/j.psep.2018.06.025.Search in Google Scholar

[19] Hunge YM, Yadav AA, Kang SW, Jun Lim S, Kim H. Visible light activated MoS2/ZnO composites for photocatalytic degradation of ciprofloxacin antibiotic and hydrogen production. J Photochem Photobiol A Chem. 2023;434(2022):114250. 10.1016/j.jphotochem.2022.114250.Search in Google Scholar

[20] Khaing KK, Yin D, Ouyang Y, Xiao S, Liu B, Deng L, et al. Fabrication of 2D-2D heterojunction catalyst with covalent organic framework (COF) and MoS2 for highly efficient photocatalytic degradation of organic pollutants. Inorg Chem. 2020;59(10):6942–52. 10.1021/acs.inorgchem.0c00422.Search in Google Scholar PubMed

[21] Ahamad T, Naushad M, Al-Saeedi SI, Almotairi S, Alshehri SM. Fabrication of MoS2/ZnS embedded in N/S doped carbon for the photocatalytic degradation of pesticide. Mater Lett. 2020;263:127271. 10.1016/j.matlet.2019.127271.Search in Google Scholar

[22] Cao X, Gan X, Lang H, Peng Y. Impact of the surface and microstructure on the lubricative properties of MoS2 aging under different environments. Langmuir. 2021;37(9):2928–41. 10.1021/acs.langmuir.0c03512.Search in Google Scholar PubMed

[23] Lu D, Yang M, Wang H, Kiran Kumar K, Wu P, Neena D. In situ hydrothermal synthesis of Y-TiO2/graphene heterojunctions with improved visible-light-driven photocatalytic properties. Ceram Int. 2017;43(18):16753–62. 10.1016/j.ceramint.2017.09.070.Search in Google Scholar

[24] Asaithambi S, Sakthivel P, Karuppaiah M, Balamurugan K, Yuvakkumar R, Thambidurai M, et al. Synthesis and characterization of various transition metals doped SnO2@MoS2 composites for supercapacitor and photocatalytic applications. J Alloys Compd. 2021;853:157060. 10.1016/j.jallcom.2020.157060.Search in Google Scholar

[25] Krishnan U, Kaur M, Kaur G, Singh K, Dogra AR, Kumar M, et al. MoS2/ZnO nanocomposites for efficient photocatalytic degradation of industrial pollutants. Mater Res Bull. 2019;111:212–21. 10.1016/j.materresbull.2018.11.029.Search in Google Scholar

[26] Li L, Yin X, Sun Y. Facile synthesized low-cost MoS2/CdS nanodots-on-nanorods heterostructures for highly efficient pollution degradation under visible-light irradiation. Sep Purif Technol. 2019;212:135–41. 10.1016/j.seppur.2018.11.032.Search in Google Scholar

[27] Ramsha K, Adeel R, Sofia J, Rahim J, Muhammad AA, Mohammad M. Synthesis and characterization of MoS2/TiO2 nanocomposites for enhanced photocatalytic degradation of methylene blue under sunlight irradiation. Key Eng Mater. 2018;778:137–43. 10.4028/www.scientific.net/KEM.778.137.Search in Google Scholar

[28] Wu D, Han L. Solvothermal synthesis and characterization of visible-light-active MoO3/MoS2 heterostructure. J Sol-Gel Sci Technol. 2019;91(3):441–5. 10.1007/s10971-019-05056-6.Search in Google Scholar

[29] Xavier JR. High protection performance of vanadium pentoxide-embedded polyfuran/epoxy coatings on mild steel. Polym Bull. 2021;78(10):5713–39. 10.1007/s00289-020-03400-3.Search in Google Scholar

[30] Wang Y, Li J, Wei Z. Transition-metal-oxide-based catalysts for the oxygen reduction reaction. J Mater Chem A. 2018;6(18):8194–209. 10.1039/c8ta01321g.Search in Google Scholar

[31] Gao MR, Liang JX, Zheng YR, Xu YF, Jiang J, Gao Q, et al. An efficient molybdenum disulfide/cobalt diselenide hybrid catalyst for electrochemical hydrogen generation. Nat Commun. 2015;6:1–7. 10.1038/ncomms6982.Search in Google Scholar PubMed PubMed Central

[32] Hasnain M, Muhammad J, Mohd R, Bin A. Effect of external static pressure on structural, electronic, and optical properties of 2 ‑ D hetero ‑ junction ­ MoS 2 for a photocatalytic applications: A DFT study. Opt Quantum Electron. 2023;2:1–14. 10.1007/s11082-023-04853-2.Search in Google Scholar

[33] Shafeeq KM, Athira VP, Kishor CHR, Aneesh PM. Structural and optical properties of V2O5 nanostructures grown by thermal decomposition technique. Appl Phys A Mater Sci Process. 2020;126(8):586. 10.1007/s00339-020-03770-5.Search in Google Scholar

[34] Jameel MH, Roslan M, Mayzan M, Shaaban IA, Rizvi S, Agam M, et al. A comparative DFT study of bandgap engineering and tuning of structural, electronic, and optical properties of 2D WS2, PtS2, and MoS2 between WSe2, PtSe2, and MoSe2 materials for photocatalytic and solar cell applications. J Inorg Organomet Polym Mater. 2023;34:322–35. 10.1007/s10904-023-02828-0.Search in Google Scholar

[35] Shaaban ER, Hassaan MY, Mostafa AG, Abdel-Ghany AM. Crystallization kinetics of new compound of V2O5-PbO-Li2O-Fe2O3 glass using differential thermal analysis. J Alloys Compd. 2009;482(1–2):440–6. 10.1016/j.jallcom.2009.04.062.Search in Google Scholar

[36] Jameel MH, Bin Agam MA, bin Roslan MS, Jabbar AH, Malik RQ, Islam MU, et al. A comparative DFT study of electronic and optical properties of Pb/Cd-doped LaVO4 and Pb/Cd-LuVO4 for electronic device applications. Comput Condens Matter. 2023;34(2022):e00773. 10.1016/j.cocom.2022.e00773.Search in Google Scholar

[37] Grandgirard J, Poinsot D, Krespi L, Nénon JP, Cortesero AM. Costs of secondary parasitism in the facultative hyperparasitoid Pachycrepoideus dubius: Does host size matter? Entomol Exp Appl. 2002;103(3):239–48. 10.1023/A:1021193329749.Search in Google Scholar

[38] Jameel MH, Ahmed S, Jiang ZY, Tahir MB, Akhtar MH, Saleem S, et al. First principal calculations to investigate structural, electronic, optical, and magnetic properties of Fe3O4and Cd-doped Fe2O4. Comput Condens Matter. 2022;30(2021):e00629. 10.1016/j.cocom.2021.e00629.Search in Google Scholar

[39] Joiner RL, Vinson SB, Benskin JB. Teratocytes as a source of juvenile hormone activity in a parasitoid-host relationship. Nat New Biol. 1973;246(152):120–1. 10.1038/newbio246120a0.Search in Google Scholar PubMed

[40] Jameel MH, Rehman A, bin Roslan MS, Bin Agam MA. To investigate the structural, electronic, optical and magnetic properties of Sr-doped KNbO3 for perovskite solar cell applications: A DFT study. Phys Scr. 2023;98(5):055802. 10.1088/1402-4896/acc6fb.Search in Google Scholar

Received: 2024-01-31
Revised: 2024-06-30
Accepted: 2024-07-16
Published Online: 2024-08-07

© 2024 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Research Articles
  2. Tension buckling and postbuckling of nanocomposite laminated plates with in-plane negative Poisson’s ratio
  3. Polyvinylpyrrolidone-stabilised gold nanoparticle coatings inhibit blood protein adsorption
  4. Energy and mass transmission through hybrid nanofluid flow passing over a spinning sphere with magnetic effect and heat source/sink
  5. Surface treatment with nano-silica and magnesium potassium phosphate cement co-action for enhancing recycled aggregate concrete
  6. Numerical investigation of thermal radiation with entropy generation effects in hybrid nanofluid flow over a shrinking/stretching sheet
  7. Enhancing the performance of thermal energy storage by adding nano-particles with paraffin phase change materials
  8. Using nano-CaCO3 and ceramic tile waste to design low-carbon ultra high performance concrete
  9. Numerical analysis of thermophoretic particle deposition in a magneto-Marangoni convective dusty tangent hyperbolic nanofluid flow – Thermal and magnetic features
  10. Dual numerical solutions of Casson SA–hybrid nanofluid toward a stagnation point flow over stretching/shrinking cylinder
  11. Single flake homo p–n diode of MoTe2 enabled by oxygen plasma doping
  12. Electrostatic self-assembly effect of Fe3O4 nanoparticles on performance of carbon nanotubes in cement-based materials
  13. Multi-scale alignment to buried atom-scale devices using Kelvin probe force microscopy
  14. Antibacterial, mechanical, and dielectric properties of hydroxyapatite cordierite/zirconia porous nanocomposites for use in bone tissue engineering applications
  15. Time-dependent Darcy–Forchheimer flow of Casson hybrid nanofluid comprising the CNTs through a Riga plate with nonlinear thermal radiation and viscous dissipation
  16. Durability prediction of geopolymer mortar reinforced with nanoparticles and PVA fiber using particle swarm optimized BP neural network
  17. Utilization of zein nano-based system for promoting antibiofilm and anti-virulence activities of curcumin against Pseudomonas aeruginosa
  18. Antibacterial effect of novel dental resin composites containing rod-like zinc oxide
  19. An extended model to assess Jeffery–Hamel blood flow through arteries with iron-oxide (Fe2O3) nanoparticles and melting effects: Entropy optimization analysis
  20. Comparative study of copper nanoparticles over radially stretching sheet with water and silicone oil
  21. Cementitious composites modified by nanocarbon fillers with cooperation effect possessing excellent self-sensing properties
  22. Confinement size effect on dielectric properties, antimicrobial activity, and recycling of TiO2 quantum dots via photodegradation processes of Congo red dye and real industrial textile wastewater
  23. Biogenic silver nanoparticles of Moringa oleifera leaf extract: Characterization and photocatalytic application
  24. Novel integrated structure and function of Mg–Gd neutron shielding materials
  25. Impact of multiple slips on thermally radiative peristaltic transport of Sisko nanofluid with double diffusion convection, viscous dissipation, and induced magnetic field
  26. Magnetized water-based hybrid nanofluid flow over an exponentially stretching sheet with thermal convective and mass flux conditions: HAM solution
  27. A numerical investigation of the two-dimensional magnetohydrodynamic water-based hybrid nanofluid flow composed of Fe3O4 and Au nanoparticles over a heated surface
  28. Development and modeling of an ultra-robust TPU-MWCNT foam with high flexibility and compressibility
  29. Effects of nanofillers on the physical, mechanical, and tribological behavior of carbon/kenaf fiber–reinforced phenolic composites
  30. Polymer nanocomposite for protecting photovoltaic cells from solar ultraviolet in space
  31. Study on the mechanical properties and microstructure of recycled concrete reinforced with basalt fibers and nano-silica in early low-temperature environments
  32. Synergistic effect of carbon nanotubes and polyvinyl alcohol on the mechanical performance and microstructure of cement mortar
  33. CFD analysis of paraffin-based hybrid (Co–Au) and trihybrid (Co–Au–ZrO2) nanofluid flow through a porous medium
  34. Forced convective tangent hyperbolic nanofluid flow subject to heat source/sink and Lorentz force over a permeable wedge: Numerical exploration
  35. Physiochemical and electrical activities of nano copper oxides synthesised via hydrothermal method utilising natural reduction agents for solar cell application
  36. A homotopic analysis of the blood-based bioconvection Carreau–Yasuda hybrid nanofluid flow over a stretching sheet with convective conditions
  37. In situ synthesis of reduced graphene oxide/SnIn4S8 nanocomposites with enhanced photocatalytic performance for pollutant degradation
  38. A coarse-grained Poisson–Nernst–Planck model for polyelectrolyte-modified nanofluidic diodes
  39. A numerical investigation of the magnetized water-based hybrid nanofluid flow over an extending sheet with a convective condition: Active and passive controls of nanoparticles
  40. The LyP-1 cyclic peptide modified mesoporous polydopamine nanospheres for targeted delivery of triptolide regulate the macrophage repolarization in atherosclerosis
  41. Synergistic effect of hydroxyapatite-magnetite nanocomposites in magnetic hyperthermia for bone cancer treatment
  42. The significance of quadratic thermal radiative scrutinization of a nanofluid flow across a microchannel with thermophoretic particle deposition effects
  43. Ferromagnetic effect on Casson nanofluid flow and transport phenomena across a bi-directional Riga sensor device: Darcy–Forchheimer model
  44. Performance of carbon nanomaterials incorporated with concrete exposed to high temperature
  45. Multicriteria-based optimization of roller compacted concrete pavement containing crumb rubber and nano-silica
  46. Revisiting hydrotalcite synthesis: Efficient combined mechanochemical/coprecipitation synthesis to design advanced tunable basic catalysts
  47. Exploration of irreversibility process and thermal energy of a tetra hybrid radiative binary nanofluid focusing on solar implementations
  48. Effect of graphene oxide on the properties of ternary limestone clay cement paste
  49. Improved mechanical properties of graphene-modified basalt fibre–epoxy composites
  50. Sodium titanate nanostructured modified by green synthesis of iron oxide for highly efficient photodegradation of dye contaminants
  51. Green synthesis of Vitis vinifera extract-appended magnesium oxide NPs for biomedical applications
  52. Differential study on the thermal–physical properties of metal and its oxide nanoparticle-formed nanofluids: Molecular dynamics simulation investigation of argon-based nanofluids
  53. Heat convection and irreversibility of magneto-micropolar hybrid nanofluids within a porous hexagonal-shaped enclosure having heated obstacle
  54. Numerical simulation and optimization of biological nanocomposite system for enhanced oil recovery
  55. Laser ablation and chemical vapor deposition to prepare a nanostructured PPy layer on the Ti surface
  56. Cilostazol niosomes-loaded transdermal gels: An in vitro and in vivo anti-aggregant and skin permeation activity investigations towards preparing an efficient nanoscale formulation
  57. Linear and nonlinear optical studies on successfully mixed vanadium oxide and zinc oxide nanoparticles synthesized by sol–gel technique
  58. Analytical investigation of convective phenomena with nonlinearity characteristics in nanostratified liquid film above an inclined extended sheet
  59. Optimization method for low-velocity impact identification in nanocomposite using genetic algorithm
  60. Analyzing the 3D-MHD flow of a sodium alginate-based nanofluid flow containing alumina nanoparticles over a bi-directional extending sheet using variable porous medium and slip conditions
  61. A comprehensive study of laser irradiated hydrothermally synthesized 2D layered heterostructure V2O5(1−x)MoS2(x) (X = 1–5%) nanocomposites for photocatalytic application
  62. Computational analysis of water-based silver, copper, and alumina hybrid nanoparticles over a stretchable sheet embedded in a porous medium with thermophoretic particle deposition effects
  63. A deep dive into AI integration and advanced nanobiosensor technologies for enhanced bacterial infection monitoring
  64. Effects of normal strain on pyramidal I and II 〈c + a〉 screw dislocation mobility and structure in single-crystal magnesium
  65. Computational study of cross-flow in entropy-optimized nanofluids
  66. Significance of nanoparticle aggregation for thermal transport over magnetized sensor surface
  67. A green and facile synthesis route of nanosize cupric oxide at room temperature
  68. Effect of annealing time on bending performance and microstructure of C19400 alloy strip
  69. Chitosan-based Mupirocin and Alkanna tinctoria extract nanoparticles for the management of burn wound: In vitro and in vivo characterization
  70. Electrospinning of MNZ/PLGA/SF nanofibers for periodontitis
  71. Photocatalytic degradation of methylene blue by Nd-doped titanium dioxide thin films
  72. Shell-core-structured electrospinning film with sequential anti-inflammatory and pro-neurogenic effects for peripheral nerve repairment
  73. Flow and heat transfer insights into a chemically reactive micropolar Williamson ternary hybrid nanofluid with cross-diffusion theory
  74. One-pot fabrication of open-spherical shapes based on the decoration of copper sulfide/poly-O-amino benzenethiol on copper oxide as a promising photocathode for hydrogen generation from the natural source of Red Sea water
  75. A penta-hybrid approach for modeling the nanofluid flow in a spatially dependent magnetic field
  76. Advancing sustainable agriculture: Metal-doped urea–hydroxyapatite hybrid nanofertilizer for agro-industry
  77. Utilizing Ziziphus spina-christi for eco-friendly synthesis of silver nanoparticles: Antimicrobial activity and promising application in wound healing
  78. Plant-mediated synthesis, characterization, and evaluation of a copper oxide/silicon dioxide nanocomposite by an antimicrobial study
  79. Effects of PVA fibers and nano-SiO2 on rheological properties of geopolymer mortar
  80. Investigating silver and alumina nanoparticles’ impact on fluid behavior over porous stretching surface
  81. Potential pharmaceutical applications and molecular docking study for green fabricated ZnO nanoparticles mediated Raphanus sativus: In vitro and in vivo study
  82. Effect of temperature and nanoparticle size on the interfacial layer thickness of TiO2–water nanofluids using molecular dynamics
  83. Characteristics of induced magnetic field on the time-dependent MHD nanofluid flow through parallel plates
  84. Flexural and vibration behaviours of novel covered CFRP composite joints with an MWCNT-modified adhesive
  85. Experimental research on mechanically and thermally activation of nano-kaolin to improve the properties of ultra-high-performance fiber-reinforced concrete
  86. Analysis of variable fluid properties for three-dimensional flow of ternary hybrid nanofluid on a stretching sheet with MHD effects
  87. Biodegradability of corn starch films containing nanocellulose fiber and thymol
  88. Toxicity assessment of copper oxide nanoparticles: In vivo study
  89. Some measures to enhance the energy output performances of triboelectric nanogenerators
  90. Reinforcement of graphene nanoplatelets on water uptake and thermomechanical behaviour of epoxy adhesive subjected to water ageing conditions
  91. Optimization of preparation parameters and testing verification of carbon nanotube suspensions used in concrete
  92. Max-phase Ti3SiC2 and diverse nanoparticle reinforcements for enhancement of the mechanical, dynamic, and microstructural properties of AA5083 aluminum alloy via FSP
  93. Advancing drug delivery: Neural network perspectives on nanoparticle-mediated treatments for cancerous tissues
  94. PEG-PLGA core–shell nanoparticles for the controlled delivery of picoplatin–hydroxypropyl β-cyclodextrin inclusion complex in triple-negative breast cancer: In vitro and in vivo study
  95. Conduction transportation from graphene to an insulative polymer medium: A novel approach for the conductivity of nanocomposites
  96. Review Articles
  97. Developments of terahertz metasurface biosensors: A literature review
  98. Overview of amorphous carbon memristor device, modeling, and applications for neuromorphic computing
  99. Advances in the synthesis of gold nanoclusters (AuNCs) of proteins extracted from nature
  100. A review of ternary polymer nanocomposites containing clay and calcium carbonate and their biomedical applications
  101. Recent advancements in polyoxometalate-functionalized fiber materials: A review
  102. Special contribution of atomic force microscopy in cell death research
  103. A comprehensive review of oral chitosan drug delivery systems: Applications for oral insulin delivery
  104. Cellular senescence and nanoparticle-based therapies: Current developments and perspectives
  105. Cyclodextrins-block copolymer drug delivery systems: From design and development to preclinical studies
  106. Micelle-based nanoparticles with stimuli-responsive properties for drug delivery
  107. Critical assessment of the thermal stability and degradation of chemically functionalized nanocellulose-based polymer nanocomposites
  108. Research progress in preparation technology of micro and nano titanium alloy powder
  109. Nanoformulations for lysozyme-based additives in animal feed: An alternative to fight antibiotic resistance spread
  110. Incorporation of organic photochromic molecules in mesoporous silica materials: Synthesis and applications
  111. A review on modeling of graphene and associated nanostructures reinforced concrete
  112. A review on strengthening mechanisms of carbon quantum dots-reinforced Cu-matrix nanocomposites
  113. Review on nanocellulose composites and CNFs assembled microfiber toward automotive applications
  114. Nanomaterial coating for layered lithium rich transition metal oxide cathode for lithium-ion battery
  115. Application of AgNPs in biomedicine: An overview and current trends
  116. Nanobiotechnology and microbial influence on cold adaptation in plants
  117. Hepatotoxicity of nanomaterials: From mechanism to therapeutic strategy
  118. Applications of micro-nanobubble and its influence on concrete properties: An in-depth review
  119. A comprehensive systematic literature review of ML in nanotechnology for sustainable development
  120. Exploiting the nanotechnological approaches for traditional Chinese medicine in childhood rhinitis: A review of future perspectives
  121. Twisto-photonics in two-dimensional materials: A comprehensive review
  122. Current advances of anticancer drugs based on solubilization technology
  123. Recent process of using nanoparticles in the T cell-based immunometabolic therapy
  124. Future prospects of gold nanoclusters in hydrogen storage systems and sustainable environmental treatment applications
  125. Preparation, types, and applications of one- and two-dimensional nanochannels and their transport properties for water and ions
  126. Microstructural, mechanical, and corrosion characteristics of Mg–Gd–x systems: A review of recent advancements
  127. Functionalized nanostructures and targeted delivery systems with a focus on plant-derived natural agents for COVID-19 therapy: A review and outlook
  128. Mapping evolution and trends of cell membrane-coated nanoparticles: A bibliometric analysis and scoping review
  129. Nanoparticles and their application in the diagnosis of hepatocellular carcinoma
  130. In situ growth of carbon nanotubes on fly ash substrates
  131. Structural performance of boards through nanoparticle reinforcement: An advance review
  132. Reinforcing mechanisms review of the graphene oxide on cement composites
  133. Seed regeneration aided by nanomaterials in a climate change scenario: A comprehensive review
  134. Surface-engineered quantum dot nanocomposites for neurodegenerative disorder remediation and avenue for neuroimaging
  135. Graphitic carbon nitride hybrid thin films for energy conversion: A mini-review on defect activation with different materials
  136. Nanoparticles and the treatment of hepatocellular carcinoma
  137. Special Issue on Advanced Nanomaterials and Composites for Energy Conversion and Storage - Part II
  138. Highly safe lithium vanadium oxide anode for fast-charging dendrite-free lithium-ion batteries
  139. Recent progress in nanomaterials of battery energy storage: A patent landscape analysis, technology updates, and future prospects
  140. Special Issue on Advanced Nanomaterials for Carbon Capture, Environment and Utilization for Energy Sustainability - Part II
  141. Calcium-, magnesium-, and yttrium-doped lithium nickel phosphate nanomaterials as high-performance catalysts for electrochemical water oxidation reaction
  142. Low alkaline vegetation concrete with silica fume and nano-fly ash composites to improve the planting properties and soil ecology
  143. Mesoporous silica-grafted deep eutectic solvent-based mixed matrix membranes for wastewater treatment: Synthesis and emerging pollutant removal performance
  144. Electrochemically prepared ultrathin two-dimensional graphitic nanosheets as cathodes for advanced Zn-based energy storage devices
  145. Enhanced catalytic degradation of amoxicillin by phyto-mediated synthesised ZnO NPs and ZnO-rGO hybrid nanocomposite: Assessment of antioxidant activity, adsorption, and thermodynamic analysis
  146. Incorporating GO in PI matrix to advance nanocomposite coating: An enhancing strategy to prevent corrosion
  147. Synthesis, characterization, thermal stability, and application of microporous hyper cross-linked polyphosphazenes with naphthylamine group for CO2 uptake
  148. Engineering in ceramic albite morphology by the addition of additives: Carbon nanotubes and graphene oxide for energy applications
  149. Nanoscale synergy: Optimizing energy storage with SnO2 quantum dots on ZnO hexagonal prisms for advanced supercapacitors
  150. Aging assessment of silicone rubber materials under corona discharge accompanied by humidity and UV radiation
  151. Tuning structural and electrical properties of Co-precipitated and Cu-incorporated nickel ferrite for energy applications
  152. Sodium alginate-supported AgSr nanoparticles for catalytic degradation of malachite green and methyl orange in aqueous medium
  153. An environmentally greener and reusability approach for bioenergy production using Mallotus philippensis (Kamala) seed oil feedstock via phytonanotechnology
  154. Micro-/nano-alumina trihydrate and -magnesium hydroxide fillers in RTV-SR composites under electrical and environmental stresses
  155. Mechanism exploration of ion-implanted epoxy on surface trap distribution: An approach to augment the vacuum flashover voltages
  156. Nanoscale engineering of semiconductor photocatalysts boosting charge separation for solar-driven H2 production: Recent advances and future perspective
  157. Excellent catalytic performance over reduced graphene-boosted novel nanoparticles for oxidative desulfurization of fuel oil
  158. Special Issue on Advances in Nanotechnology for Agriculture
  159. Deciphering the synergistic potential of mycogenic zinc oxide nanoparticles and bio-slurry formulation on phenology and physiology of Vigna radiata
  160. Nanomaterials: Cross-disciplinary applications in ornamental plants
  161. Special Issue on Catechol Based Nano and Microstructures
  162. Polydopamine films: Versatile but interface-dependent coatings
  163. In vitro anticancer activity of melanin-like nanoparticles for multimodal therapy of glioblastoma
  164. Poly-3,4-dihydroxybenzylidenhydrazine, a different analogue of polydopamine
  165. Chirality and self-assembly of structures derived from optically active 1,2-diaminocyclohexane and catecholamines
  166. Advancing resource sustainability with green photothermal materials: Insights from organic waste-derived and bioderived sources
  167. Bioinspired neuromelanin-like Pt(iv) polymeric nanoparticles for cancer treatment
  168. Special Issue on Implementing Nanotechnology for Smart Healthcare System
  169. Intelligent explainable optical sensing on Internet of nanorobots for disease detection
  170. Special Issue on Green Mono, Bi and Tri Metallic Nanoparticles for Biological and Environmental Applications
  171. Tracking success of interaction of green-synthesized Carbopol nanoemulgel (neomycin-decorated Ag/ZnO nanocomposite) with wound-based MDR bacteria
  172. Green synthesis of copper oxide nanoparticles using genus Inula and evaluation of biological therapeutics and environmental applications
  173. Biogenic fabrication and multifunctional therapeutic applications of silver nanoparticles synthesized from rose petal extract
  174. Metal oxides on the frontlines: Antimicrobial activity in plant-derived biometallic nanoparticles
  175. Controlling pore size during the synthesis of hydroxyapatite nanoparticles using CTAB by the sol–gel hydrothermal method and their biological activities
  176. Special Issue on State-of-Art Advanced Nanotechnology for Healthcare
  177. Applications of nanomedicine-integrated phototherapeutic agents in cancer theranostics: A comprehensive review of the current state of research
  178. Smart bionanomaterials for treatment and diagnosis of inflammatory bowel disease
  179. Beyond conventional therapy: Synthesis of multifunctional nanoparticles for rheumatoid arthritis therapy
Downloaded on 10.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ntrev-2024-0078/html
Scroll to top button