Next Article in Journal
Antimicrobial Natural Products from Plant Pathogenic Fungi
Next Article in Special Issue
Determining the Photoelectrical Behavior and Photocatalytic Activity of an h-YMnO3 New Type of Obelisk-like Perovskite in the Degradation of Malachite Green Dye
Previous Article in Journal
In Vitro Anticancer Activity of Novel Ciprofloxacin Mannich Base in Lung Adenocarcinoma and High-Grade Serous Ovarian Cancer Cell Lines via Attenuating MAPK Signaling Pathway
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

In Situ Hydrothermal Synthesis of Ni1−xMnxWO4 Nanoheterostructure for Enhanced Photodegradation of Methyl Orange

by
Imran Hasan
*,
Mohammed Abdullah Albaeejan
,
Alanoud Abdullah Alshayiqi
,
Wedyan Saud Al-Nafaei
and
Fahad A. Alharthi
*
Department of Chemistry, College of Science, King Saud University, Riyadh 11451, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(3), 1140; https://doi.org/10.3390/molecules28031140
Submission received: 25 December 2022 / Revised: 19 January 2023 / Accepted: 19 January 2023 / Published: 23 January 2023

Abstract

:
The monoclinic nanocrystalline Ni1−xMnxWO4 heterostructure has been successfully synthesized by the hydrothermal technique for achieving better sensitive and photocatalytic performances. Different characterization techniques such as X-ray diffraction (XRD), Fourier transform infrared spectroscopy (FTIR), ultraviolet-visible (UV–Vis), and photoluminescence (PL) spectroscopy have been employed to investigate their structural, microstructural, and optical properties. Mn-ion incorporation in the NiWO4 lattice reduces the particle size of the sample compared with the pure undoped NiWO4 sample, which has been confirmed from the transmission electron microscope image. The Tauc plot of the Ni1−xMnxWO4 sample exhibits a significant decrease in bandgap energy compared with the pure undoped NiWO4 sample due to the quantum confinement effect. Finally, the material was explored as a photocatalyst for the degradation of methyl orange (MO) dye from wastewater under visible light irradiation. Various reaction parameters such as pH, catalyst dose, reaction time, and kinetics of the photodegradation were studied using the batch method. The results showed that the Ni1−xMnxWO4 is highly efficient (94.51%) compared with undoped NiWO4 (65.45%). The rate of photodegradation by Ni1–xMnxWO4 (0.067) was found to be 1.06 times higher than the undoped NiWO4 (0.062).

1. Introduction

The advancement in industrial productions, such as paper, dyeing, and textiles, to boost up the economy has resulted in extreme use of heavy metals, azo dyes, and other organic compounds which are discarded directly into surface and ground water [1,2]. The toxic organic and inorganic compounds that have a non-biodegradable and persistent nature pose a threat to human health and other creatures [3]. These azo dyes are very harmful in their own context, even at ng/L concentrations, and may turn into more harmful products by reacting with other chemical species in the water streams or interacting with UV light [4]. Among various azo dyes, methyl orange (MO) has been categorized as an omnipresent water pollutant which can cause diarrhea, vomiting, and skin allergies; moreover, exposure to higher concentrations can cause death [5,6]. MO and its transformation products in water can reflect certain wavelengths of sunlight and thus hinder the photosynthesis reactions of aquatic flora [7]. This process leads to a decrease in the amount of dissolved oxygen which is mandatory for the sustainability of aquatic life [8]. Thus, it is necessary for researchers to develop efficient methods to remove MO from wastewater completely without producing any secondary pollutants. Various methods currently exist, such as chlorination, ozonation, adsorption, sedimentation, and coagulation [9,10,11]. However, these methods transfer MO from one medium to another medium and produce secondary pollution. So, a green procedure, i.e., photocatalytic degradation, was taken into consideration, which constitutes a more effective and energy-efficient method as compared with its counterparts in the scope of advanced oxidation processes (AOPs). The process forms highly reactive radical species generated by the irradiation of light on a catalyst surface [12]. When interacting with pollutant molecules, these highly active radicals decompose them into small non-toxic molecules or even mineralize them into CO2 and H2O [13]. However, the conversion efficiency of this method depends upon the chemical and optical behavior of the developed catalyst material.
In recent years, vast research has been carried out to explore the novel properties of the metal oxide semiconductor nanomaterials, including TiO2 [13], Cu2O [14], ZnO [15], Fe2O3 [16], etc., and mixed-metal oxide nanomaterials, such as BiFeO3 [17], BiVO4 [18], Bi2MoO6 [19], CuWO4 [20], etc., as a photocatalyst for the degradation of organic pollutants. Among these, nanocrystalline metal tungstates of the general formula MWO4 (M = Zn, Co, Ni, Mn, etc.) have attracted the attention of scientists as an efficient semiconductor photocatalyst because of their excellent optical properties and energy bandgap [21,22]. Among various metal tungstates, NiWO4 has been recognized as one of the important semiconductor photocatalysts, with an energy bandgap ≂3.5 eV [23]. NiWO4 can be easily synthesized by the hydrothermal method, decomposition approach, and electrochemical method, but the hydrothermal route has attracted more interest to synthesize a vast number of nanomaterials due to the formation of a large crystal phase [24,25]. NiWO4 exhibits a monoclinic wolframite-type structure with oxygen deposited around the tungsten-associating [WO6] type octahedron [26]. Photocatalytic degradation involves the formation of electron-hole pairs in the catalyst under radiation, which react with surrounding water and surface-adsorbed oxygen to generate OH and O2 radicals to degrade MO [1,5]. However, the fast recombination of these electron-hole pairs hinders the photocatalytic degradation of pollutants and poses a drawback on the efficiency of the material [27]. To overcome this drawback, various steps have been applied previously in the literature, such as composite forming [28], morphological improvement [29], doping with metal elements [30], and heterojunctions [31]. Therefore, one of the effective strategies, namely heterostructure formation, was applied in this study, which creates a methodical partition of photogenerated electron-hole pairs through band alignment to improve the photocatalytic efficiency [32]. In the present study, manganese (Mn) was doped in the crystal lattice of NiWO4 to improve its photocatalytic activity through changes in morphology, crystal lattice, and optical properties. One of the most important aspects of doping is to substitute the metal ion of the host material through another metal ion of smaller radius, thus creating lattice defects and oxygen vacancy, which hinders the rate of electron-hole pair recombination, reduces the energy bandgap, and thus improves the photocatalytic activity of the material towards the pollutant [26,33,34]. From the literature, the ionic radii of Ni2+ and Mn2+ are 0.069 nm and 0.065 nm; this suggests that Mn2+ can easily substitute some of the Ni2+ in the NiWO4 lattice and thus form a Ni1−xMnxWO4 nanocomposite-type heterostructure, where x is the moles of Mn added [35]. The synthesized material was explored as a photocatalyst for the degradation of MO under a visible light source. The efficiency of the synthesized material, Ni1−xMnxWO4 NC, was optimized by varying different reaction parameters, such as irradiation time, pH of the reaction medium, and catalyst dose. Finally, the kinetics of degradation was outlined depending on the photocatalytic results.

2. Results and Discussion

2.1. Material Characterization

2.1.1. Crystal Structure Studies

The XRD patterns of synthesized NiWO4 and Ni1−xMnxWO4 are given in Figure 1. The pristine NiWO4 exhibited the characteristic peaks at Miller indices (100), (011), (110), (111), (002), (200), (102), (112), (211), (022), (130), (202), (113), (132), and (041), which are indexed for the monoclinic wolframite structure of NiWO4 with standard JCPDS no. 072–0480. The XRD spectra of synthesized Ni1−xMnxWO4 also exhibited maximum peaks from NiWO4 except for the appearance of some new Miller indices values, namely (121), (030), (220), and (023), which were found to have good agreement with the lattice structure of MnWO4 with standard JCPDS no. 72-0478. The XRD results suggested that Mn2+ has been successfully incorporated in the NiWO4 lattice; moreover, there is an increase in the intensity of peaks after Mn incorporation, indicating an increase in the crystallinity of material. The crystallite size was determined using the Debye–Scherrer equation given as [36]:
D c = K λ β × Cos θ
where λ is the wavelength of X-ray source, K is a shape factor and is usually ~0.9, θ is the corresponding angle, and β is the breadth of the observed diffraction line at its half intensity maximum. The results obtained using Equation (1) disclosed that the particle size of NiWO4 was 32 ± 1.05 nm, while for Ni1−xMnxWO4, it was 26 ± 0.65 nm. The outcomes indicated a contraction in the crystallite size in NiWO4 upon Mn2+ incorporation, suggesting an improvement in the optical activity of the material as well.

2.1.2. Functional Group Studies

The functional group study of the materials was done using FTIR spectroscopy and the results are given in Figure 2. From the FTIR spectra, NiWO4 shows characteristic bands at 3414, 1630 cm−1 belonging to stretching and bending vibration bands of –OH groups (surface adsorbed water), 879 and 830 cm−1 antisymmetric stretching vibration bands of W–O bonds, 706, 622 cm−1 stretching and bending vibrations of W–O bonds in [WO6] octahedron, and 528, 434 cm−1 stretching vibrations of Ni–O bonds from the NiO6 octahedron [23,27]. The FTIR spectra of the as-synthesized Ni1−xMnxWO4 NC exhibited the same bands with shifted values, which is due to the orbital hybridization and change of chemical environment of NiWO4 by Mn. The band at 706 cm–1 vanished completely, suggesting the acquirement of the portion associated with the [WO6] octahedron by Mn2+ [37].

2.1.3. Optical Studies

The optical properties of the synthesized material were investigated by observing the UV-Vis spectra given in Figure 3. The UV-Vis spectra of NiWO4 exhibited one peak of very low intensity at 278 nm, suggesting that the material is only UV-light active and the corresponding transition is due to a charge transfer process between Ni2+ (d–d) and octahedron [WO6] clusters [38]. The UV spectra of Ni1−xMnxWO4 NC exhibited three peaks at 286 nm, 355 nm, and 685 nm, suggesting the UV and visible light activity of the synthesized material. So, the incorporation of Mn in the NiWO4 lattice enlarged its light absorbing capacity; moreover, the peaks that appeared are due to Ni2+ (d–d) to octahedron [WO6] clusters CT and secondly Ni2+ (d–d) to octahedron [WO6] clusters to Mn2+ CT, which increased its absorption intensity to visible light from the UV region [39]. The energy bandgap (Eg) value of the synthesized material can be calculated using Tauc’s equation given as [40]:
α h υ = B ( h υ E g ) n
where B is a constant (dependent on the nature of the material), α is the absorption coefficient of the material (calculated from the absorption spectra), ν is the frequency of the radiation, h is the Plank’s constant, and Eg is the energy of the band gap in eV. The value of n (coefficient of transition) decides the type of energy bandgap, i.e., n = 2 value corresponds to allowed indirect energy bandgap, while n = 1/2 for allowed direct energy bandgap. Using Equation (2), the value of energy band gap (Eg) was calculated as 3.35 eV for NiWO4 and 2.13 eV for Ni1−xMnxWO4 NC. The contraction in Eg value upon incorporation of Mn in the NiWO4 lattice resulted in improved optical and photocatalytic activities.

2.1.4. SEM-EDX-Mapping Analysis

The surface morphology, elemental composition, and their distribution in the crystal lattice were assessed by SEM-EDX analysis. Figure 4a represents the SEM image of pristine NiWO4, which shows an aggregation of tiny interconnected particles, whereas the SEM image in Figure 4b represents a porous morphology with interconnected tiny particles. The EDX analysis given in Figure 4c,d confirms the composition of the synthesized material by weight % as O K (20.51), Ni K (19.74), and W M (59.75) for NiWO4, and O K (22.97), Ni K (9.36), Mn K (6.85), and W M (60.82) for Ni1−xMnxWO4 NC. The EDX results suggested a ratio of 1.4:1 of Ni:Mn in the synthesized material. The elemental mapping given in Figure 4e taken from the selected marked area of Ni1–xMnxWO4 NC shows a uniform distribution of O, Ni, Mn, and W in the material.
For further insight into the morphology, including crystallite shape and size, TEM-SAED analysis was used. Figure 5a,b represents the TEM image of Ni1−xMnxWO4 NC, which represents the monodispersed hexagonal crystals with mitigated morphology. The Gaussian particle size distribution given in Figure 5c revealed an average crystallite size of 26.84 nm, which is found to be in good agreement with the crystallite size calculated using the Debye–Scherer equation (26 ± 0.65 nm). Figure 5d consists of the selected area diffraction pattern (SAED) of Ni1−xMnxWO4 NC, with marked yellow rings belonging to the Miller indices (110), (112), and (104) indexed in the XRD pattern of Ni1−xMnxWO4 NC.

2.2. Photocatalytic Applications

2.2.1. Effect of pH

The pH of the reaction medium plays a vital role in controlling the rate of photodegradation of the organic pollutant by varying the charge density on the surface of the catalyst. Photocatalytic experiments were conducted by taking 20 mL of 50 ppm MO dye with 10 mg of the catalyst with varying pH values from 1–7 in cylindrical vessels inside the photocatalytic chamber associated with a tungsten lamp (150 mWcm−2) as the visible light source. The results obtained after completion of irradiation time are given in Figure 6a–c in which it was observed that with an increase in pH value from 1–5, the rate of degradation increases, achieving an efficiency of 97.16% for NiWO4 and 98.32% for Ni1−xMnxWO4 NC; however, further increases in the pH value results in a decrease in the rate of degradation. This trend can be explained based on point of zero charge (pHpzc) value, which is found to be 5.8 for NiWO4 and 6.2 for Ni1−xMnxWO4 NC (Figure S2). So, at pH < pHpzc, the surface of the catalyst is positive and is suitable to form adsorption–desorption equilibrium with a maximum number of anionic MO molecules and thus mineralize it in the presence of light utilizing photogenerated OH or O2 radicals. However, at pH > pHpzc, the surface of the catalyst is negative, which will reflect the anionic MO molecules and thus a lesser number of dye molecules will be available for degradation; this is why a decrease in photocatalytic efficiency appeared at higher pH values [41]. The outcomes suggested that the Ni1−xMnxWO4 NC possessed better photocatalytic efficiency as compared with pristine NiWO4. Thus, Mn decoration in the NiWO4 lattice, leading to the formation of the heterostructure, resulted in an enhancement of photocatalytic efficiency towards MO degradation.

2.2.2. Effect of Catalyst Dose

Experiments were conducted by taking 20 mL of 50 ppm MO with variable catalyst dose (5, 10, 15, 20, and 25 mg) for 70 min of irradiation and results obtained are given in Figure 7a–d. From Figure 7a,b, it was observed that from 5–10 mg of catalyst dose, the absorbance value decreases, indicating the increase in rate of degradation of MO; beyond 10 mg, the rate of degradation increases as supported by Figure 7c. From Figure 7d, the maximum degradation efficiency for NiWO4 and Ni1−xMnxWO4 NC was achieved as 93.58% and 97.46% at 10 mg of catalyst dose, respectively. Initially, with low concentration of catalyst, a greater number of surface-active sites are available to accommodate the MO molecules and thus there is a high rate of photodegradation. The decrease in the rate of photodegradation beyond a certain value of catalyst dose is due to an increase in the degree of aggregation of nanoparticles, which produces turbidity in the solution. The turbidity increases the opacity of the solution and impedes the light penetration intensity towards the catalyst surface. This phenomena decrease the rate of generation of ROS (OH or O2 radicals) and thus cause a decrease in photocatalytic efficiency [42].

2.3. Kinetics of Photodegradation and Effect of Irradiation Time

Figure 8a,b represents the degradation of MO in the presence of pristine NiWO4 and the as-synthesized Ni1−xMnxWO4 NC with respect to variation in irradiation time. It can be seen from the results that with an increase in irradiation time from 5 min to 70 min, the intensity of the absorption maxima peak decreases continuously with any shift in wavelength. Thus, an increase in the rate of photodegradation of MO is observed with an increase in irradiation time until 70 min and the maximum photocatalytic efficiency was calculated as 98.79% for NiWO4 and 99.06% for Ni1−xMnxWO4 NC, respectively. The obtained photocatalytic data was adjusted to the Langmuir–Hinshelwood (L-H) pseudo first-order kinetic model to observe the rate of degradation quantitatively. The model is mathematically given as [12,43]:
ln ( C t C 0 ) = k app × t
where C0 and Ct are the concentration of MO at the initial state (t = 0) and specific time intervals (t) and kapp is the apparent rate constant which can be obtained from the slope of linear graph of -ln (Ct/C0) vs. irradiation time (t). The obtained experimental data was applied to equation (3) and the results obtained are given in Table 1 and Figure 8c. It was found that the rate of degradation of MO was higher for Ni1−xMnxWO4 (0.067 min−1) as compared with pristine NiWO4 (0.062 min−1), which again quantitatively supports the idea of the enhancement of photocatalytic efficiency of material towards MO through Mn incorporation. The value of the correlation coefficient R2 = 0.99 reveals that the simulation of the L-H model to the photocatalytic data has minimum error. The half lifetime of the reaction was calculated using t1/2 = ln 2/kapp, and the values obtained are 11 min for NiWO4 and 10.34 min for Ni1−xMnxWO4 NC.

2.4. Detection of ROS

The scavenger’s test was performed to find out the primary ROS involved in the photodegradation of MO by Ni1−xMnxWO4 NC under visible light source. In the present study, some of the scavengers used include EDTA for h+, benzoic acid (BA for OH), acrylamide (AA for e), and benzoquinone (BQ for O2) and the results are given in Figure 9 [44]. The appearance of high absorption intensity for MO in the presence of BA suggests that the photocatalytic rate of MO is maximally inhibited by benzoic acid, which is responsible for the trapping of OH radicals. So, OH radicals are the primary ROS to degrade MO using Ni1−xMnxWO4 NC.
Based on the trapping experiments, the plausible general mechanism of MO degradation is given by Equations (4–10) [1,4,8]:
Ni 1 x Mn x WO 4 + h ν     Ni 1 x Mn x WO 4 ( h VB + + e CB )
Ni 1 x Mn x WO 4 ( h VB + ) + OH ad     O H
Ni 1 x Mn x WO 4 ( e CB ) + O 2     O 2
O 2 + H +     H O 2
2 H O 2     O 2 + H 2 O 2
H 2 O 2 + e CB     O H   +   OH
O H   +   MO   ( dye )     CO 2 + H 2 O
Figure 10 represents the mechanism of photodegradation of MO by Ni1−xMnxWO4 NC under visible light source. The irradiation of the photocatalyst initiates the promotion of electrons (e) from the valence band (VB) to conduction band (CB) through the photoabsorption process, leaving lots of vacancies in VB, referred to as holes (h+), and photogenerated electrons (e) in CB Equation (4). These photogenerated electrons (e) in CB interact with surface-adsorbed O2 molecules and oxidize to superoxide radicals (O2) Equation (6). The holes interact with hydroxyl anions (OH) to produce hydroxy radicals (OH) Equation (5). The O2 molecule reacts with H+ ions and produces H2O2, which further reacts with e (CB) generating super reactive OH radical reaction (7–10); these are responsible for the degradation of MO under visible light source.

2.5. Reusability and TOC Test

To check the stability of the synthesized material towards treatment of MO in wastewater, a photocatalytic test was performed in repeated mode. In cycle 1, 10 mg of the Ni1−xMnxWO4 NC photocatalyst was dispersed in 20 mL of 50 ppm MO solution under optimized reaction conditions. Then, after the completion of reaction, the photocatalyst was separated out by centrifugation and the supernatant was checked using a UV-Vis spectrophotometer. The photocatalyst was washed and dried and then again dispersed in the MO solution for cycle 2. In this way, the same material underwent 6 cycles of the photocatalytic test to check its stability and reusable efficiency under the given optimized conditions towards MO. The results obtained after the photocatalytic experiment was given in Figure 11a. The outcomes suggested that the synthesized Ni1−xMnxWO4 NC material is highly stable to treat MO dye as only a very small change in photocatalytic efficiency from cycle 1 (98.79%) to cycle 6 (95.23%) was observed.
Similarly, to observe the mineralization process of MO by the photocatalyst, the total organic carbon (TOC) test was taken into consideration and the results are given in Figure 11b. It was observed from the graph that as the irradiation time increases, TOC value decreases continuously, and after 70 min of irradiation, the synthesized material Ni1−xMnxWO4 NC exhibits 67.45% of TOC removal, which suggests that the synthesized material is a very effective catalyst to treat MO-contaminated water under visible light source.

2.6. Photocurrent Measurement

Furthermore, to investigate the photophysical behaviors of NiWO4 and Ni1−xMnxWO4 NC under visible light irradiation, photocurrent measurements were taken into consideration. The obtained results are given in Figure 12. It can be seen that the photocurrent response of NiWO4 is found to be lower than Ni1−xMnxWO4 NC under visible light irradiation, which suggests that the separation efficiency of charger carriers in Ni1−xMnxWO4 NC is higher than NiWO4. The photocatalytic experiments and optical studies are also found to be in close agreement with the photocurrent data which concludes that Ni1−xMnxWO4 NC has higher photocatalytic efficiency than the NiWO4. The photocurrent for NiWO4 and Ni1−xMnxWO4 NC was found to be 0.032 μA and 0.035 μA, respectively. The doping of Mn in the NiWO4 solid matrix leads to the creation of mobile oxygen vacancies, which prompted the separation efficiency of charge carriers; consequently, higher photocurrent was observed for Ni1−xMnxWO4 NC compared with pristine NiWO4 [45,46].

2.7. Comparison with Literature

The present study was compared with the reported studies in the literature and the results are given in Table 2. It was found that the synthesized material and performed studies are purely novel as it is not reported in the literature. The material was found to be highly effective and energy efficient towards the degradation of methyl orange.

3. Materials and Methods

3.1. Chemical and Reagents

All precursor chemicals used for the synthesis, such as sodium tungstate [Na2WO4·2H2O] (Sigma Aldrich, St. Louis, MO, USA), nickel nitrate [Ni (NO3)2.6H2O] (Loba chemicals, Mumbai, India), manganese chloride [MnCl2.4H2O], and urea [CH4N2O, 99.5 %] (Merck chemicals, Rahway, NJ, USA) are of analytical grade (purity > 99%).

3.2. Synthesis of NiWO4 and Ni1−xMnxWO4 Nanocomposite

The nanomaterial was synthesized by the hydrothermal process by taking the precursor mixture in a Teflon-lined autoclave at 150 °C for 24 h [51]. In a beaker, 2 moles of Ni (NO3)2.6H2O and 1 mole of MnCl2.4H2O were dissolved in 40 mL of deionized water and placed on a magnetic stirrer for 30 min to achieve homogeneity. After 30 min, 2 moles of Na2WO4·2H2O in 30 mL deionized water was added dropwise followed by the addition of 1.5 g of urea. The mixture was left on a magnetic stirrer for 3 h at room temperature (22 °C) on continuous stirring. After 3 h, the mixture was transferred to a 100 mL Teflon-lined autoclave and placed under an oven at 120 °C for 24 h to achieve complete nucleation of Mn2+ in the NiWO4 lattice. After the reaction, the mixture was taken out of the oven, cooled down and then the precipitate was collected via centrifugation (9000 rpm). The obtained material was washed with deionized water several times and ethanol (2 times) to remove unreacted entities. Finally, the material was dried in the oven at 100 °C for 4 h and then calcined at 600 °C for 3 h. Similarly, in the same way, pristine NiWO4 was prepared without adding MnCl2.4H2O.

3.3. Characterization Techniques

The synthesized NiWO4 and Ni1−xMnxWO4 nanocomposites were characterized by various instruments to verify the synthesis of the material. The crystal structure of the material and lattice deformation upon Mn2+ doping was observed using powder X-ray diffraction (XRD, Bruker D8 Advance with Cu-Kα radiation, λ = 0.15418 nm, Billerica, MA, USA). The surface composition was assessed by scanning electron spectroscopy (SEM; Hitachi S-4800 Field Emission Scanning Electron Microscope, Ibaraki, Japan) and morphological information, shape and size, and their distribution in the lattice were investigated by transmission electron microscopy (TEM; JEM-2100F, Tokyo, Japan). Chemical structural information of the nanomaterial was obtained through Fourier transform infrared (FTIR; Perkin Elmer spectrum 2 ATR, Waltham, MA, USA). The optical properties of the nanoparticles were measured using ultraviolet visible spectroscopy (UV–1900 Shimadzu, Kyoto, Japan). To investigate the difference between the organic carbon (OC) and inorganic carbon (IC) during the photocatalytic reaction, a TOC analyzer (Shimadzu-00077) was used for the analysis of total organic carbon (TOC).

3.4. Photocatalysis Process

The photocatalytic efficiency of the as-synthesized NiWO4 and Ni1−xMnxWO4 nanocomposites was monitored by the degradation of methyl orange (MO) under visible light irradiation using a tungsten lamp (150 mW/cm−2) in a photoreactor chamber. We dispersed 10 mg each of NiWO4 and Ni1−xMnxWO4 in 10 mL of 50 ppm of MO dye solution; this was then mixed under magnetic stirring for 30 min in dark to establish the adsorption–desorption equilibrium. Finally, the mixture was taken in 20 mL cylindrical vessels (test tube, 20 mL) with a magnetic bar and irradiated in a photoreactor chamber; at a time interval, the vessels were taken out of the chamber, centrifuged to separate the catalyst from the MO solution, and then tested using a UV-Vis spectrophotometer at λmax = 492 nm to evaluate the photocatalytic efficiency as given by Equation (11):
Degradation   Efficieny   ( % ) = ( 1 C t C 0 ) × 100
where C0 is the initial concentration of the MO solution and Ct is the concentration of the MO solution at specific times. A selectivity test was performed by taking 10 mg of Ni1−xMnxWO4 NC with various dyes, such as bromophenol (BP), methyl orange (MO), Congo red (CR), malachite green (MG), crystal violet (CV) and methylene blue (MB). The results are given in Figure S1, which suggest that the synthesized nanocomposite material is most sensitive towards MO, showing a photocatalytic efficiency of 94.51%. Therefore, photocatalytic experiments were conducted for MO degradation by varying the reaction parameters, such as irradiation time, pH of the MO solution, catalyst dose, and concentration of the MO solution, to optimize the photocatalytic efficiency of the synthesized material. Thus, based on the outcomes of the experiments, the kinetics and mechanism of photocatalysis was predicted.

3.5. Reusability and TOC Test

To check the stability of the synthesized material towards the treatment of MO in wastewater, a photocatalytic test was performed in cyclic mode. In cycle 1, the photocatalyst was dispersed in the MO solution under optimized reaction conditions; after the completion of the reaction, it was separated out by centrifugation, washed and dried, and then again dispersed in the MO solution for cycle 2. In this way, the same material underwent 6 cycles of the photocatalytic test to check its stability under the given conditions towards MO. Similarly, to observe the mineralization process of MO by the photocatalyst, the total organic carbon test was taken into consideration. As the irradiation time increases, the photocatalytic efficiency also increases, suggesting an increase in the mineralization process, which will reflect in a decrease in TOC value. The TOC (%) was calculated using Equation (12):
TOC   ( % ) = ( TOC 0 TOC f TOC 0 ) × 100

4. Conclusions

In the present study, we accomplished the one-pot hydrothermal synthesis of a Mn-decorated NiWO4 nanocomposite material in a Teflon-lined autoclave at 120 °C for 24 h. Various analytical and spectroscopic tests such as XRD, FTIR, UV-Vis, SEM-EDX-mapping, and TEM-SAED supported the successful incorporation of Mn in the NiWO4 lattice. The material Ni1−xMnxWO4 showed enhanced photocatalytic degradation of MO (99.06%) as compared with pristine NiWO4 (93.58%) at pH 5 using 10 mg of catalyst for 50 ppm dye concentration under 70 min of visible light irradiation. The enhanced photocatalytic efficiency belongs to improved optical properties by reducing the energy bandgap (Eg) from 3.49 eV (NiWO4) to 3.33 eV (Ni1−xMnxWO4). The SAED analysis also simulated, with XRD Miller indices data, the monoclinic wurtzite structure of the synthesized Ni1−xMnxWO4 nanocomposite material. The kinetic data was best adjusted to the L-H pseudo first-order kinetic model with R2 = 0.99. The rate of photodegradation was found to be 1.06 times higher than the pristine nanoparticles as OH radicals play the primary reactive oxidant species. The outcomes of this study suggest that the material is highly stable and can be reusable for the mineralization of MO (67.75% TOC remove) and other organic pollutants under optimized conditions for environmental remediation without producing secondary pollution.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28031140/s1, Figure S1. (a) UV–Vis spectra of various dyes after degradation with Ni1−xMnxWO4 NC and (b) bar graph representing the % degradation values for individual organic pollutant, Figure S2. Point of zero charge for both NiWO4 and Ni1−xMnxWO4 NC

Author Contributions

Conceptualization, F.A.A. and I.H.; Methodology, F.A.A. and I.H.; Software, F.A.A. and I.H.; Formal analysis, A.A.A., M.A.A. and W.S.A.-N.; Investigation, A.A.A., M.A.A. and W.S.A.-N.; Resources, A.A.A., M.A.A. and W.S.A.-N.; Data curation, I.H.; Writing – original draft, A.A.A., M.A.A. and W.S.A.-N.; Writing – review & editing, I.H.; Visualization, F.A.A. and I.H.; Supervision, F.A.A. and I.H.; Project administration, F.A.A.; Funding acquisition, F.A.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

All authors have read and agreed to the published version of the manuscript.

Data Availability Statement

Data is contained within the article.

Acknowledgments

The authors extend their appreciation to the deputyship of Research and Innovation, Ministry of Education in Saudi Arabia for funding this research work through project number (IFKSURG-2-1317).

Conflicts of Interest

The authors declare that there is no conflict of interest related to this research.

References

  1. Swathi, D.; Sabumon, P.C.; Trivedi, A. Simultaneous Decolorization and Mineralization of High Concentrations of Methyl Orange in an Anoxic Up-Flow Packed Bed Reactor in Denitrifying Conditions. J. Water Process Eng. 2021, 40, 101813. [Google Scholar] [CrossRef]
  2. You, J.; Zhang, L.; He, L.; Zhang, B. Photocatalytic Degradation of Methyl Orange on ZnO–TiO2/SO42− Heterojunction Composites. Opt. Mater. 2022, 131, 112737. [Google Scholar] [CrossRef]
  3. Bian, Y.; Zheng, G.; Ding, W.; Hu, L.; Sheng, Z. Magnetic Field Effect on the Photocatalytic Degradation of Methyl Orange by Commercial TiO2 Powder. RSC Adv. 2021, 11, 6284–6291. [Google Scholar] [CrossRef] [PubMed]
  4. Kumar, N.; Jung, U.; Jung, B.; Park, J.; Naushad, M. Zinc Hydroxystannate/Zinc-Tin Oxide Heterojunctions for the UVC-Assisted Photocatalytic Degradation of Methyl Orange and Tetracycline. Environ. Pollut. 2023, 316, 120353. [Google Scholar] [CrossRef] [PubMed]
  5. Rajapandi, S.; Pandeeswaran, M.; Kousalya, G.N. Novel Green Synthesis of N-Doped Carbon Dots from Fruits of Opuntia Ficus Indica as an Effective Catalyst for the Photocatalytic Degradation of Methyl Orange Dye and Antibacterial Studies. Inorg. Chem. Commun. 2022, 146, 110041. [Google Scholar] [CrossRef]
  6. Lu, Z.; Chen, G.; Hao, W.; Sun, G.; Li, Z. Mechanism of UV-Assisted TiO2/Reduced Graphene Oxide Composites with Variable Photodegradation of Methyl Orange. RSC Adv. 2015, 5, 72916–72922. [Google Scholar] [CrossRef]
  7. Chen, L.; Zhang, M.; Wu, J.; Zheng, X.; Liao, S.; Ou, B.; Tian, L. In Situ Embedment of CoOx on G-C3N4 as Z Scheme Heterojunction for Efficient Photocatalytic Degradation of Methyl Orange and Phenol under Visible Light. J. Alloys Compd. 2022, 927, 167047. [Google Scholar] [CrossRef]
  8. de Oliveira, R.L.; Anderson, M.A.; de Aragão Umbuzeiro, G.; Zocolo, G.J.; Zanoni, M.V.B. Assessment of By-Products of Chlorination and Photoelectrocatalytic Chlorination of an Azo Dye. J. Hazard. Mater. 2012, 205–206, 1–9. [Google Scholar] [CrossRef]
  9. Hassani, A.; Khataee, A.; Karaca, S.; Fathinia, M. Heterogeneous Photocatalytic Ozonation of Ciprofloxacin Using Synthesized Titanium Dioxide Nanoparticles on a Montmorillonite Support: Parametric Studies, Mechanistic Analysis and Intermediates Identification. RSC Adv. 2016, 6, 87569–87583. [Google Scholar] [CrossRef]
  10. Hasan, I.; Bhatia, D.; Walia, S.; Singh, P. Removal of Malachite Green by Polyacrylamide-g-Chitosan γ-Fe2O3 Nanocomposite-an Application of Central Composite Design. Groundw. Sustain. Dev. 2020, 11, 100378. [Google Scholar] [CrossRef]
  11. Hasan, I.; Ahamd, R. A Facile Synthesis of Poly (Methyl Methacrylate) Grafted Alginate@Cys-Bentonite Copolymer Hybrid Nanocomposite for Sequestration of Heavy Metals. Groundw. Sustain. Dev. 2019, 8, 82–92. [Google Scholar] [CrossRef]
  12. Hasan, I.; Alharthi, F.A. Caffeine-Alginate Immobilized CeTiO4 Bionanocomposite for Efficient Photocatalytic Degradation of Methylene Blue. J. Photochem. Photobiol. A Chem. 2022, 433, 114126. [Google Scholar] [CrossRef]
  13. Thiruvenkatachari, R.; Vigneswaran, S.; Moon, I.S. A Review on UV/TiO2 Photocatalytic Oxidation Process. Korean J. Chem. Eng. 2008, 25, 64–72. [Google Scholar] [CrossRef]
  14. Muthukumaran, M.; Gnanamoorthy, G.; Varun Prasath, P.; Abinaya, M.; Dhinagaran, G.; Sagadevan, S.; Mohammad, F.; Oh, W.C.; Venkatachalam, K. Enhanced Photocatalytic Activity of Cuprous Oxide Nanoparticles for Malachite Green Degradation under the Visible Light Radiation. Mater. Res. Express. 2020, 7, 015038. [Google Scholar] [CrossRef]
  15. Chen, X.; Wu, Z.; Liu, D.; Gao, Z. Preparation of ZnO Photocatalyst for the Efficient and Rapid Photocatalytic Degradation of Azo Dyes. Nanoscale Res. Lett. 2017, 12, 143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Sivaranjani, R.; Thayumanavan, A.; Sriram, S. Photocatalytic Activity of Zn-Doped Fe 2O 3 Nanoparticles: A Combined Experimental and Theoretical Study. Bull. Mater. Sci. 2019, 42, 1–7. [Google Scholar] [CrossRef] [Green Version]
  17. Mushtaq, F.; Chen, X.; Hoop, M.; Torlakcik, H.; Pellicer, E.; Sort, J.; Gattinoni, C.; Nelson, B.J.; Pané, S. Piezoelectrically Enhanced Photocatalysis with BiFeO3 Nanostructures for Efficient Water Remediation. iScience 2018, 4, 236–246. [Google Scholar] [CrossRef]
  18. Zhou, B.; Zhao, X.; Liu, H.; Qu, J.; Huang, C.P. Synthesis of Visible-Light Sensitive M-BiVO4 (M = Ag, Co, and Ni) for the Photocatalytic Degradation of Organic Pollutants. Sep. Purif. Technol. 2009, 77, 275–282. [Google Scholar] [CrossRef]
  19. Zhang, L.; Xu, T.; Zhao, X.; Zhu, Y. Controllable Synthesis of Bi2MoO6 and Effect of Morphology and Variation in Local Structure on Photocatalytic Activities. Appl. Catal. B 2010, 98, 138–146. [Google Scholar] [CrossRef]
  20. Babu, M.J.; Botsa, S.M.; Rani, S.J.; Venkateswararao, B.; Muralikrishna, R. Enhanced Photocatalytic Degradation of Cationic Dyes under Visible Light Irradiation by CuWO4-RGO Nanocomposite. Adv. Compos. Hybrid Mater. 2020, 3, 205–212. [Google Scholar] [CrossRef]
  21. Mohamed Jaffer Sadiq, M.; Krishna Bhat, D. Novel NiWO4-ZnO-NRGO Ternary Nanocomposites with Enhanced Photocatalytic Activity. Mater. Today Proc. 2018, 5, 22412–22420. [Google Scholar] [CrossRef]
  22. Muthuraj, V. Superior Visible Light Driven Photocatalytic Degradation of Fluoroquinolone Drug Norfloxacin over Novel NiWO4 Nanorods Anchored on G-C 3 N 4 Nanosheets. Colloids Surf. A Physicochem. Eng. Asp. 2019, 567, 43–54. [Google Scholar] [CrossRef]
  23. Rosal, F.J.O.; Gouveia, A.F.; Sczancoski, J.C.; Lemos, P.S.; Longo, E.; Zhang, B.; Cavalcante, L.S. Electronic Structure, Growth Mechanism, and Sonophotocatalytic Properties of Sphere-like Self-Assembled NiWO4 Nanocrystals. Inorg. Chem. Commun. 2018, 98, 34–40. [Google Scholar] [CrossRef]
  24. Hosseini, S.; Farsi, H.; Moghiminia, S.; Zubkov, T.; Lightcap, I.V.; Riley, A.; Peters, D.G.; Li, Z. Nickel Tungstate (NiWO4) Nanoparticles/Graphene Composites: Preparation and Photoelectrochemical Applications. Semicond. Sci. Technol. 2018, 33, 055008. [Google Scholar] [CrossRef]
  25. Pourmortazavi, S.M.; Rahimi-Nasrabadi, M.; Karimi, M.S.; Mirsadeghi, S. Evaluation of Photocatalytic and Supercapacitor Potential of Nickel Tungstate Nanoparticles Synthesized by Electrochemical Method. New J. Chem. 2018, 42, 19934–19944. [Google Scholar] [CrossRef]
  26. Liaquat, H.; Imran, M.; Latif, S.; Iqbal, S.; Hussain, N.; Bilal, M. Citric Acid-Capped NiWO4/Bi2S3 and RGO-Doped NiWO4/Bi2S3 Nanoarchitectures for Photocatalytic Decontamination of Emerging Pollutants from the Aqueous Environment. Environ. Res. 2022, 212, 113276. [Google Scholar] [CrossRef]
  27. Hao, M.; Meng, X.; Miao, Y. Synthesis of NiWO4 Powder Crystals of Polyhedron for Photocatalytic Degradation of Rhodamine. Solid State Sci. 2017, 72, 103–108. [Google Scholar] [CrossRef]
  28. Li, Z.; Shen, Y.; Yang, C.; Lei, Y.; Guan, Y.; Lin, Y.; Liu, D.; Nan, C.W. Significant Enhancement in the Visible Light Photocatalytic Properties of BiFeO3–Graphene Nanohybrids. J. Mater. Chem. A Mater. 2012, 1, 823–829. [Google Scholar] [CrossRef]
  29. Zhang, X.; Wang, B.; Wang, X.; Xiao, X.; Dai, Z.; Wu, W.; Zheng, J.; Ren, F.; Jiang, C. Preparation of M@BiFeO3 Nanocomposites (M = Ag, Au) Bowl Arrays with Enhanced Visible Light Photocatalytic Activity. J. Am. Ceram. Soc. 2015, 98, 2255–2263. [Google Scholar] [CrossRef]
  30. Irfan, S.; Li, L.; Saleemi, A.S.; Nan, C.W. Enhanced Photocatalytic Activity of La3+ and Se4+ Co-Doped Bismuth Ferrite Nanostructures. J. Mater. Chem. A Mater. 2017, 5, 11143–11151. [Google Scholar] [CrossRef]
  31. Ge, J.; Sun, Y.; Chen, W.; Song, F.; Xie, Y.; Zheng, Y.; Rao, P. Z-Scheme Heterojunction Based on NiWO4/WO3 Microspheres with Enhanced Photocatalytic Performance under Visible Light. Dalton Trans. 2021, 50, 13801–13814. [Google Scholar] [CrossRef] [PubMed]
  32. Feng, H.; Xi, Y.; Huang, Q. A Novel p–n Mn0.2Cd0.8S/NiWO4 Heterojunction for Highly Efficient Photocatalytic H2 Production. Dalton Trans. 2020, 49, 12242–12248. [Google Scholar] [CrossRef] [PubMed]
  33. Hitha, H.; John, M.; Jose, A.; Kuriakose, S.; Varghese, T. Influence of Bi3+ Doping on Structural, Optical and Photocatalytic Degradation Properties of NiWO4 Nanocrystals. J. Solid State Chem. 2021, 295, 121892. [Google Scholar] [CrossRef]
  34. Cui, Y.P.; Shang, Y.R.; Shi, R.X.; Che, Q.D.; Wang, J.P. Pt-Decorated NiWO4/WO3 Heterostructure Nanotubes for Highly Selective Sensing of Acetone. Trans. Nonferrous Met. Soc. China 2022, 32, 1981–1993. [Google Scholar] [CrossRef]
  35. Mousavi, M.; Habibi-Yangjeh, A.; Seifzadeh, D. Novel Ternary G-C3N4/Fe3O4/MnWO4 Nanocomposites: Synthesis, Characterization, and Visible-Light Photocatalytic Performance for Environmental Purposes. J. Mater. Sci. Technol. 2018, 34, 1638–1651. [Google Scholar] [CrossRef]
  36. Scherrer, P. Estimation of the Size and Internal Structure of Colloidal Particles by Means of Rontgen Rays. Nachr. Von Der Ges. Der Wiss. Zu Göttingen 1918, 26, 98–100. [Google Scholar]
  37. Dutta, D.P.; Raval, P. Effect of Transition Metal Ion (Cr3+, Mn2+ and Cu2+) Doping on the Photocatalytic Properties of ZnWO4 Nanoparticles. J. Photochem. Photobiol. A Chem. 2018, 357, 193–200. [Google Scholar] [CrossRef]
  38. Azizi, S.; Asadpour-Zeynali, K. Electrochemical Synthesis of Tungstate Bimetallic Nanoparticles with Application in Electrocatalytic Determination of Paracetamol. ChemistrySelect 2022, 7, e202104548. [Google Scholar] [CrossRef]
  39. Leeladevi, K.; Arunpandian, M.; Vinoth Kumar, J.; Chellapandi, T.; Mathumitha, G.; Lee, J.W.; Nagarajan, E.R. CoWO4 Decorated ZnO Nanocomposite: Efficient Visible-Light-Activated Photocatalyst for Mitigation of Noxious Pollutants. Phys. B Condens. Matter 2022, 626, 413493. [Google Scholar] [CrossRef]
  40. Tauc, J. Optical Properties of Solids; Ables, A., Ed.; American Elsevier: Amsterdam, The Netherlands, 1970. [Google Scholar]
  41. Modrogan, C.; Cǎprǎrescu, S.; Dǎncilǎ, A.M.; Orbuleț, O.D.; Grumezescu, A.M.; Purcar, V.; Radițoiu, V.; Fierascu, R.C. Modified Composite Based on Magnetite and Polyvinyl Alcohol: Synthesis, Characterization, and Degradation Studies of the Methyl Orange Dye from Synthetic Wastewater. Polymers 2021, 13, 3911. [Google Scholar] [CrossRef] [PubMed]
  42. Harish, S.; Navaneethan, M.; Archana, J.; Silambarasan, A.; Ponnusamy, S.; Muthamizhchelvan, C.; Hayakawa, Y. Controlled Synthesis of Organic Ligand Passivated ZnO Nanostructures and Their Photocatalytic Activity under Visible Light Irradiation. Dalton Trans. 2015, 44, 10490–10498. [Google Scholar] [CrossRef] [PubMed]
  43. Kalikeri, S.; Kamath, N.; Gadgil, D.J.; Shetty Kodialbail, V. Visible Light-Induced Photocatalytic Degradation of Reactive Blue-19 over Highly Efficient Polyaniline-TiO2 Nanocomposite: A Comparative Study with Solar and UV Photocatalysis. Environ. Sci. Pollut. Res. 2017, 25, 3731–3744. [Google Scholar] [CrossRef]
  44. Elgorban, A.M.; al Kheraif, A.A.; Syed, A. Construction of Ag2WO4 Decorated CoWO4 Nano-Heterojunction with Recombination Delay for Enhanced Visible Light Photocatalytic Performance and Its Antibacterial Applications. Colloids Surf. A Physicochem. Eng. Asp. 2021, 629, 127416. [Google Scholar] [CrossRef]
  45. Anis, S.F.; Lalia, B.S.; Palmisano, G.; Hashaikeh, R. Photoelectrochemical Activity of Electrospun WO3/NiWO4 Nanofibers under Visible Light Irradiation. J. Mater. Sci. 2018, 53, 2208–2220. [Google Scholar] [CrossRef]
  46. Zhang, Y.; Jin, Z. Accelerated Charge Transfer via a Nickel Tungstate Modulated Cadmium Sulfide p–n Heterojunction for Photocatalytic Hydrogen Evolution. Catal. Sci. Technol. 2019, 9, 1944–1960. [Google Scholar] [CrossRef]
  47. Tri, N.L.M.; Duc, D.S.; van Thuan, D.; Al Tahtamouni, T.; Pham, T.D.; Tran, D.T.; Le Chi, N.T.P.; Nguyen, V.N. Superior Photocatalytic Activity of Cu Doped NiWO4 for Efficient Degradation of Benzene in Air Even under Visible Radiation. Chem. Phys. 2019, 525, 110411. [Google Scholar] [CrossRef]
  48. Thilagavathi, T.; Venugopal, D.; Thangaraju, D.; Marnadu, R.; Palanivel, B.; Imran, M.; Shkir, M.; Ubaidullah, M.; AlFaify, S. A Facile Co-Precipitation Synthesis of Novel WO3/NiWO4 Nanocomposite with Improved Photocatalytic Activity. Mater. Sci. Semicond. Process. 2021, 133, 105970. [Google Scholar] [CrossRef]
  49. Botsa, S.M.; Jagadeesh Babu, M.; Suresh, P.; Kalyani, P.; Venkateswararao, B.; Muralikrishna, R. Spherical NiWO4-Reduced Graphene Oxide Nanocomposite for Effective Visible Light Driven Photocatalytic Activity for the Decolourisation of Organic Pollutants. Arab. J. Chem. 2020, 13, 8489–8497. [Google Scholar] [CrossRef]
  50. Habibi-Yangjeh, A.; Shekofteh-Gohari, M. Novel Magnetic Fe3O4/ZnO/NiWO4 Nanocomposites: Enhanced Visible-Light Photocatalytic Performance through p-n Heterojunctions. Sep. Purif. Technol. 2017, 184, 334–346. [Google Scholar] [CrossRef]
  51. Ikram, M.; Javed, Y.; Shad, N.A.; Sajid, M.M.; Irfan, M.; Munawar, A.; Hussain, T.; Imran, M.; Hussain, D. Facile Hydrothermal Synthesis of Nickel Tungstate (NiWO4) Nanostructures with Pronounced Supercapacitor and Electrochemical Sensing Activities. J. Alloys Compd. 2021, 878, 160314. [Google Scholar] [CrossRef]
Figure 1. XRD spectra of NiWO4 (black line) and Ni1−xMnxWO4 NC (red line) calcined at 600 °C.
Figure 1. XRD spectra of NiWO4 (black line) and Ni1−xMnxWO4 NC (red line) calcined at 600 °C.
Molecules 28 01140 g001
Figure 2. FTIR spectra of NiWO4 (black line) and Ni1−xMnxWO4 NC (red line) calcined at 600 °C.
Figure 2. FTIR spectra of NiWO4 (black line) and Ni1−xMnxWO4 NC (red line) calcined at 600 °C.
Molecules 28 01140 g002
Figure 3. UV-Vis spectra of NiWO4 (black line) and Ni1−xMnxWO4 NC (red line) with inset Tauc’s plot for calculating energy bandgap.
Figure 3. UV-Vis spectra of NiWO4 (black line) and Ni1−xMnxWO4 NC (red line) with inset Tauc’s plot for calculating energy bandgap.
Molecules 28 01140 g003
Figure 4. SEM-EDX image of (a,c) NiWO4 (b,d) Ni1–xMnxWO4 NC. (e) Selected area mapping of elements composing the synthesized nanocomposite.
Figure 4. SEM-EDX image of (a,c) NiWO4 (b,d) Ni1–xMnxWO4 NC. (e) Selected area mapping of elements composing the synthesized nanocomposite.
Molecules 28 01140 g004aMolecules 28 01140 g004b
Figure 5. (a,b) TEM images of Ni1−xMnxWO4 NC at 100 and 20 nm magnification range, (c) Gaussian average particle size distribution, and (d) SAED pattern showing the Debye–Scherer rings.
Figure 5. (a,b) TEM images of Ni1−xMnxWO4 NC at 100 and 20 nm magnification range, (c) Gaussian average particle size distribution, and (d) SAED pattern showing the Debye–Scherer rings.
Molecules 28 01140 g005
Figure 6. UV–Vis spectra of (a) NiWO4 (b) Ni1−xMnxWO4 NC and (c) Ct/C0 graph vs. pH depicting rate of MO (50 ppm) degradation for 70 min of visible light irradiation.
Figure 6. UV–Vis spectra of (a) NiWO4 (b) Ni1−xMnxWO4 NC and (c) Ct/C0 graph vs. pH depicting rate of MO (50 ppm) degradation for 70 min of visible light irradiation.
Molecules 28 01140 g006
Figure 7. UV–Vis spectra of (a) NiWO4 (b) Ni1−xMnxWO4 NC and (c) Ct/C0 graph vs. (d) pH depicting rate of MO (50 ppm) degradation for 70 min of visible light irradiation.
Figure 7. UV–Vis spectra of (a) NiWO4 (b) Ni1−xMnxWO4 NC and (c) Ct/C0 graph vs. (d) pH depicting rate of MO (50 ppm) degradation for 70 min of visible light irradiation.
Molecules 28 01140 g007
Figure 8. UV-Vis spectra for degradation of MO with variable irradiation time by (a) NiWO4 (b) Ni1−xMnxWO4 NC and (c) linear graph of −ln (Ct/C0) vs. irradiation time (t).
Figure 8. UV-Vis spectra for degradation of MO with variable irradiation time by (a) NiWO4 (b) Ni1−xMnxWO4 NC and (c) linear graph of −ln (Ct/C0) vs. irradiation time (t).
Molecules 28 01140 g008
Figure 9. UV-Vis spectra for the scavenger test for the degradation of MO by Ni1−xMnxWO4 NC.
Figure 9. UV-Vis spectra for the scavenger test for the degradation of MO by Ni1−xMnxWO4 NC.
Molecules 28 01140 g009
Figure 10. Schematic diagram showing the mechanism of photodegradation of MO by Ni1−xMnxWO4 NC under visible light source.
Figure 10. Schematic diagram showing the mechanism of photodegradation of MO by Ni1−xMnxWO4 NC under visible light source.
Molecules 28 01140 g010
Figure 11. (a) Reusability and stability test for the synthesized Ni1−xMnxWO4 NC. (b) TOC analysis during degradation of MO.
Figure 11. (a) Reusability and stability test for the synthesized Ni1−xMnxWO4 NC. (b) TOC analysis during degradation of MO.
Molecules 28 01140 g011
Figure 12. Photocurrent spectra for NiWO4 and Ni1−xMnxWO4 NC obtained at 1.3 V.
Figure 12. Photocurrent spectra for NiWO4 and Ni1−xMnxWO4 NC obtained at 1.3 V.
Molecules 28 01140 g012
Table 1. L-H first-order kinetic parameters for photocatalytic degradation of MO by NiWO4 and Ni1−xMnxWO4 NC.
Table 1. L-H first-order kinetic parameters for photocatalytic degradation of MO by NiWO4 and Ni1−xMnxWO4 NC.
Catalystkapp
(min−1)
Errort1/2
(min)
R2
NiWO40.0636.16 × 10−411.000.99
Ni1−xMnxWO4 NC0.0671.31 × 10−310.340.99
Table 2. Comparison of the literature information with the present study.
Table 2. Comparison of the literature information with the present study.
CatalystsIrradiation Time
(min)
Light SourceOrganic Pollutant% DegradationReferences
Cu-NiWO4180Visible lightBenzene96.50[47]
Bi-doped NiWO490UV IrradiationRhodamine86.71[33]
rGO-NiWO4/Bi2S340Visible lightMethyl Orange72.00[26]
WO3/NiWO480UV IrradiationMethylene blue90.63[48]
NiWO4-RGO240Visible lighto-Nitrophenol82.00[49]
Fe3O4/ZnO/NiWO4300Visible lightRhodmaine B97.90[50]
Ni1−xMnxWO470Visible lightMethyl Orange99.06%Present study
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Hasan, I.; Albaeejan, M.A.; Alshayiqi, A.A.; Al-Nafaei, W.S.; Alharthi, F.A. In Situ Hydrothermal Synthesis of Ni1−xMnxWO4 Nanoheterostructure for Enhanced Photodegradation of Methyl Orange. Molecules 2023, 28, 1140. https://doi.org/10.3390/molecules28031140

AMA Style

Hasan I, Albaeejan MA, Alshayiqi AA, Al-Nafaei WS, Alharthi FA. In Situ Hydrothermal Synthesis of Ni1−xMnxWO4 Nanoheterostructure for Enhanced Photodegradation of Methyl Orange. Molecules. 2023; 28(3):1140. https://doi.org/10.3390/molecules28031140

Chicago/Turabian Style

Hasan, Imran, Mohammed Abdullah Albaeejan, Alanoud Abdullah Alshayiqi, Wedyan Saud Al-Nafaei, and Fahad A. Alharthi. 2023. "In Situ Hydrothermal Synthesis of Ni1−xMnxWO4 Nanoheterostructure for Enhanced Photodegradation of Methyl Orange" Molecules 28, no. 3: 1140. https://doi.org/10.3390/molecules28031140

Article Metrics

Back to TopTop