Next Article in Journal
A Pyrrolidine Functionalized Poly[(ethylene glycol) Methacrylate] Resin as a Heterogeneous Catalyst for Aqueous Aldol Reactions
Next Article in Special Issue
Synthesis and Visible Light Catalytic Performance of BiOI/Carbon Nanofibers Heterojunction
Previous Article in Journal
Pt2 Dimer Anchored Vertically in Defective BN Monolayer as an Efficient Catalyst for N2 Reduction: A DFT Study
Previous Article in Special Issue
Photocatalytic Degradation of Eriochrome Black-T Using BaWO4/MoS2 Composite
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Accelerated Decoloration of Organic Dyes from Wastewater Using Ternary Metal/g-C3N4/ZnO Nanocomposites: An Investigation of Impact of g-C3N4 Concentration and Ni and Mn Doping

1
Department of Chemistry, School of Science, University of Management and Technology, Lahore 54770, Pakistan
2
Department of Physics, College of Science, Jazan University, Jazan 45142, Saudi Arabia
3
Department of Physics, Faculty of Science, Ibb University, Ibb 70270, Yemen
4
Department of Chemistry, College of Science, Jazan University, Jazan 45142, Saudi Arabia
5
UMR CNRS 7198, Institute Jean Lamour, University of Lorraine, 54011 Nancy, France
6
Department of Physics, College of Science, Al-Zulfi, Majmaah University, Al-Majmaah 11952, Saudi Arabia
7
Department of Pharmacology and Toxicology, College of Pharmacy, Jazan University, Jazan 45142, Saudi Arabia
8
Department of Applied Physics, Federal Urdu University of Arts Science and Technology, Islamabad 44000, Pakistan
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(11), 1388; https://doi.org/10.3390/catal12111388
Submission received: 19 September 2022 / Revised: 18 October 2022 / Accepted: 20 October 2022 / Published: 8 November 2022
(This article belongs to the Special Issue UV/Vis/NIR Photocatalysis and Optical Properties)

Abstract

:
Wastewater from many sectors that contains hazardous organic pollutants exacerbates environmental contamination. Consequently, outstanding photocatalytic substances that can successfully degrade hazardous substances are needed to provide pollution-free water. From this perspective, zinc oxide/g-C3N4-based composites are desirable due to their low cost, strong reactivity, and environmental friendliness. So, in the current investigation, sequences of Mn/g-C3N4/ZnO (Mn/GZ) and Ni/g-C3N4/ZnO (Ni/GZ) nanocomposites (NCs) containing different concentrations (wt.%) of g-C3N4 were made via the co-precipitation process. The chemical makeup and morphological characteristics of the produced composites were ascertained via the techniques of transmission electron microscopy (TEM), scanning electron microscopy (SEM), energy-dispersive X-ray (EDX), Fourier transform infrared (FTIR), photoluminescence (PL), and UV spectrophotometry. Methyl orange (MO) and Eriochrome Black T (EBT) dyes were used as target pollutants to assess the composite materials’ photocatalytic effectiveness. Compared to g-C3N4/ZnO and g-C3N4, the produced Mn/GZ and Ni/GZ NCs displayed better photocatalytic activity. The improved photocatalytic efficiency of the Ni/GZ and Mn/GZ NCs might be credited to synergistic interactions at the g-C3N4 and ZnO interface that result in a more efficient separation and conduction of photo-induced charges. Furthermore, the Ni/Mn atoms act as the facilitators to improve electron–hole pair separation and conduction in NCs. The nanocomposites were found to be incredibly stable, with consistently high dye decoloration efficiency over five catalytic cycles. Hence, Ni/GZ and Mn/GZ could potentially be very effective and adaptable photocatalysts for the photocatalytic decoloration of wastewater pollutants.

Graphical Abstract

1. Introduction

Our planet is currently in severe environmental jeopardy. Our world’s environmental conditions are getting worse due to the industrial pollutants being released into the ecosystem directly (water, soil, and air). Wastewater contains a variety of contaminants, including pathogenic bacteria, heavy metal ions, organic dyes, and pharmaceuticals [1,2]. Due to increased water demands and diminishing supply, these dangerous substances must be separated from wastewater and kept away from the aquatic environment. During the past two decades, great emphasis has been placed on the photocatalytic removal of organic contaminants using semiconductor materials. This method is based on the formation of reactive oxygen species (ROS) by semiconductor materials when they are exposed to sunlight [3,4]. Numerous semiconductor photocatalysts with various nanostructures have been explored, including TiO2, ZnO, CuO, NiS, SnO2, ZrO2, and WO3. Due to their large band gaps and quick electron–hole recombination, these photocatalysts are limited in visible-light photocatalysis [4,5].
ZnO is an efficient photocatalyst because of its high yield, stability, affordability, and superior optical characteristics [4,6,7,8]. Its unusual qualities make it useful in photocatalysis, solar cells, and catalytic processes. However, ZnO’s light absorption is primarily restricted to the UV area and is very weakly responsive to visible radiation because of its large Eg value (~3.37 eV) [9,10]. ZnO also has several disadvantages, including the speedy recombination of photogenerated e/h+ pairs, photo corrosion, and visible-light inactivity [11,12]. Because of the fast e/h+ pair recombination, very few electrons are transported to the pollutant’s conduction band. These issues pose a practical challenge for using ZnO in photocatalytic processes [13,14]. Many techniques have been emerge to alter its band structure, including doping metal or nonmetal and forming composites with appropriate materials. Doping helps to narrow the band gap by inserting impurities among energy levels in the form of a mid-gap state [15,16,17]. Thus, the most successful method to solve these issues is the composite creation of ZnO with semiconductors and carbon compounds [17,18].
When exposed to visible light, the g-C3N4 semiconductor demonstrated excellent photocatalytic performance. This is due to its high stability and narrow band gap, which allows it to absorb more visible light. It possesses excellent physical, thermal, optical, and electrical characteristics [19,20]. This highly stable metal-free polymer uses inexpensive starting molecules, including urea, thiourea, and melamine, through a simplistic heating procedure. Compared to inorganic conjugative materials, its softness makes it easier to coat on other materials (like CNTs and graphene) [21,22]. Furthermore, its well-crystallized lamellar structure has a higher affinity for charge transfer than other conjugative organic compounds [23,24]. As a polymeric matrix with a moderate band gap, g-C3N4 can generate hydrogen, split water, and reduce CO2 [13]. g-C3N4 is an excellent contender for a variety of uses in the visible spectrum range [25]. However, g-C3N4 is not an effective photocatalyst in pure form due to the fast recombination of photoinduced e/h+ pairs [26]. As a result, numerous efforts have been made to get around this restriction, including using vacancies, making heterojunctions, and mixing g-C3N4 with nonmetals and other metal oxides [27].
Integrating g-C3N4 and metal ions with ZnO has been one of the most successful methods for improving ZnO’s photocatalytic efficiency. It is reported that ZnO modified with g-C3N4 or doped with a metal ZnO accelerates e/h+ pair transmission and promotes the production of active species [28,29,30]. Li et al. reported that mesoporous carbon/g-C3N4 NCs made via a thermal process had improved photocatalytic efficiency against RhB as compared to g-C3N4 [31]. According to Wang et al., Mn-g-C3N4 outperformed pristine g-C3N4 in the photocatalytic diminution of Cr (VI) and degradation of RhB by nearly nine times and six times, respectively [32]. According to the findings of Wang et al., Eu-g-C3N4 demonstrated photocatalytic degradation of RhB that was six times superior to that by g-C3N4 [33]. Similarly, Tonda et al. showed that Fe-g-C3N4 displayed better photocatalytic decoloration of RhB dye in visible wavelengths, as compared to g-C3N4 [34].
In current study, we fabricated metal/g-C3N4/ZnO NCs by doping g-C3N4/ZnO with Ni and Mn metal ions individually to realize significant photocatalytic performance. The metal-doped/g-C3N4/ZnO NCs were fabricated via a surfactant (PEG) assisted precipitation method, and the resulting Ni/g-C3N4/ZnO (Ni/GZ) and Mn/g-C3N4/ZnO (Mn/GZ) NCs were used to decompose methyl orange (MO) and Eriochrome Black T (EBT) dyes. EBT and MO are water-soluble azo dyes that cause human ailments. Their removal from water is challenging since they are not destroyed by aerobic or biological oxidation [35,36]. The decoloration results of these dyes can be considered as a model for other organic contaminants/dyes. The photocatalytic decoloration study was carried out in two steps.
In step one, Mn/GZ and Ni/GZ photocatalysts of varying g-C3N4 percentages (20, 30, 40, 50, 60, 70, and 80 wt.%) were synthesized. In the current research, the effect of Ni/Mn substitution (3 wt.% for Ni and Mn metal ions, since this is the mainly cited doping amount in the literature) and g-C3N4 percentages on the photocatalytic properties of Ni/GZ and Mn/GZ NCs was observed. The Mn (4s23d5) and Ni (4s23d8) metal ions were chosen as dopants because these can be considered as preventative for transition metals (having a filling trend from half-filled d orbital towards complete filling). The Mn/50%g-C3N4/ZnO (Mn/50GZ) and Ni/60%g-C3N4/ZnO (Ni/60GZ) photocatalysts manifested the maximum decoloration efficiency under solar light. In step two, the comparative photocatalytic efficiency of Ni/60GZ and Mn/50GZ NCs was observed against EBT and MO dyes. The results showed that the photocatalytic activity of Ni/60GZ and Mn/50GZ NCs was increased because of enhanced sunlight absorption and improved separation of e−/h+ couples. The current investigation proves the usefulness of (Ni/Mn)/g-C3N4/ZnO for eliminating contaminants from our environment.

2. Materials and Methods

2.1. Materials

Nickel nitrate (Ni(NO3)2·6H2O) (99.99%), polyethylene glycol (PEG), zinc acetate dihydrate (Zn(CH3COO)2·2H2O) (98%), sodium hydroxide (NaOH) (98%), manganese (II) acetate tetrahydrate ((CH3COO)2Mn·4H2O) (99.99%), Eriochrome Black T (99.99%), methyl orange (99.99%), and urea (CO(NH2)2) (99.99%) were buyed from Merck Pakistan and which were utilized exactly as received.

2.2. Synthesis of g-C3N4

Thermal polycondensation of urea to 570 °C in a muffle furnace for 5 h at 5 °C min−1 produced a yellowish g-C3N4 powder [37]. In summary, an alumina crucible con-taining 5 grams of urea was placed in a muffle furnace. The urea was then calcined for 5 h at 5 °C min−1 at 570 °C to produce light-yellow urea. Before being processed into a finer powder, the final material was cooled to room temperature.

2.3. Synthesis of Composite Nanostructures

A range of targeted Ni/GZ and Mn/GZ NCs with varying g-C3N4 percentages (20, 30, 40, 50, 60, 70, and 80 wt.%) were produced via the PEG-assisted chemical co-precipitation method [38]. For the preparation of Ni/20GZ NCs, 0.34 gm of g-C3N4, 0.01 M Ni(NO3)2·6H2O, and 0.1 M of Zn(CH3COO)2·2H2O (2.20 g) were placed in a beaker and homogenized in 200 mL of deionized water (D. I). Then, 10 mL of PEG solution was added, and the mixture was stirred on the magnetic stirrer for 1 h. Finally, orange-brown precipitates were obtained by adding 2 M NaOH solution. After filtering off the precipitates, they were rinsed with 100% ethanol and D. I water. The precipitates were then dried at 85 °C and calcinated in an oven at 560 °C for 5 h (5 °C·min−1). The similar method was used to produce Ni/GZ and Mn/GZ NCs containing other percentages (20, 30, 40, 50, 60, 70, and 80 wt.%) of g-C3N4; the percentages for different combinations are detailed in Tables S1 and S2. For the synthesis of Mn/GZ NCs, the (CH3COO)2Mn·4H2O salt was used as the Mn precursor, and the remaining steps were the same as detailed for Ni/GZ NCs. g-C3N4/ZnO NCs and ZnO NPs were also synthesized for comparative study, as explained by Mudassar et al. [37]. The main steps involved in synthesizing NCs are given in Figure 1.

2.4. Characterization

The Structure of the Produced Materials Was Studied by XRD (Bruker AXS, D8-S4, USA) at 40 kV and 30 mA at Room Temperature with Cu K Radiation (k = 1.54056)

The structure of the produced materials was studied by XRD (Bruker AXS, D8-S4, Billerica, MA, USA) at 40 kV and 30 mA at room temperature with Cu K radiation (k = 1.54056 Å), while the elemental composition and surface morphology were determined using SEM-EDS (Hitachi, S-4800, Tokyo, Japan). A UV-vis-NIR spectrophotometer was used to analyze the reflectance and photocatalytic absorption spectra (UV-770, Jasco, Tokyo, Japan). The surface morphologies of the photocatalysts were studied using a transmission electron microscope (TEM, JEOL-JEM-1230, Billerica, MA, USA). FT-IR spectrometers with a resolution of 1 cm−1 recorded functional groups in the 4000–400 cm−1 range (Perkin 400 FT-IR, Billerica, MA, USA).

2.5. Photocatalytic Activity

The photocatalyzed dye decoloration activity of each produced photocatalyst was assessed under presence of sunlight. As organic dyes, EBT and MO aqueous solutions were used as the standard contaminants for examining the photocatalytic efficiencies. A quantity of 0.2 g of photocatalyst was diffused in 100 mL of dye solution (10 mg·L−1). Each suspension was sonicated for 15 min and then placed in the dark for 30 min to achieve adsorption–desorption equilibrium. After that, the suspension was positioned under solar light (68–73klux) (LT300, Extech, UK) in an open atmosphere, and aliquots of 5 mL were collected after every 30 min. After centrifugation, a UV-vis spectrophotometer was used to measure the sample’s photocatalytic activity. The λmax for EBT was measured at 531 nm, and that for MO was measured at 464 nm [39,40,41].

3. Results and Discussion

3.1. FTIR Analysis

Figure 2 shows a comparison of the FT-IR spectra of the photocatalysts in the 4000–450 cm−1 wavenumber band. The observed band at 879 cm−1 in the FT-IR spectrum of ZnO confirms the presence of the metal–oxygen bond and the absence of all other functional groups [11]. Stretching of the O-H bond causes the peak at around 3300–3450 cm−1, while bending of the O-H bond causes the peak at about 1696 cm−1 [11]. The bending at 1504 cm−1 is caused by H-O-H vibrations, which suggests that the ZnO NPs absorbed H2O. A wide-ranging band at 3178 cm−1 in the FT-IR spectrum of g-C3N4 appeared due to O-H bond stretching; the peaks in the range 1500–2000 cm−1 are due to the stretching vibrations of C=N, and peaks in the range 1500–1000 cm−1 are due to C-N bond stretching [42]. A peak at 805 cm−1 indicates the triazine unit. Also, a band at 2355 cm−1 can be accredited to adsorbed carbon dioxide in all the samples [43]. The FT-IR spectra of the Ni/GZ and Mn/GZ composites contain the peaks corresponding to both ZnO and g-C3N4, signifying that the composites were formed successfully [43,44,45].

3.2. XRD Analysis

X-ray diffractograms were used to characterize the phase and purity of the synthesized samples, as exhibited in Figure 3a,b. The XRD pattern of zinc oxide showed peaks (2θ values) with conforming crystal planes at 31.82° (100), 34.73° (002), 36.34° (101), 47.61° (102), 56.48° (110), 62.79° (103), 68.29° (112), and 69.41° (201), correspondingly, and these planes are recognized as the hexagonal wurtzite structure of ZnO (JCPDS # 043-0002) [46,47,48,49]. The g-C3N4 XRD pattern contained two distinct peaks at 13.02° and 27.3°, which correspond to the (100) and (002) facets of the crystals and confirmed its nanosheet morphology (JCPDS # 00-087-1526) [50,51]. The principal peaks ((100), (002), and (101)) in the XRD patterns of the Ni/GZ and Mn/GZ composites showed a minor decrease in intensity, supporting the successful incorporation of Ni/Mn metal ions into ZnO and its heterostructure development with g-C3N4 (Figure 3b) [52,53]. The successful synthesis of Ni/GZ and Mn/GZ composites is supported by the existence of characteristic g-C3N4 and ZnO diffraction peaks in the XRD patterns [17].

3.3. Compositional and Morphological Study

To check the morphology and structural composition of the generated photocatalysts, TEM and SEM micrographs were acquired. The sphere-shaped morphology of ZnO NPs with 10–15 nm diameter is exhibited in Figure 4a. A lamellar sheet-like structure was observed for pure g-C3N4 in the TEM image (Figure 4b). The (Ni/Mn)-ZnO NPs were consistently integrated with thin sheets of g-C3N4, as exhibited in the TEM images of Mn/GZ (Figure 4c) and Ni/GZ (Figure 4d) NCs. The photocatalytic efficiency of the Ni/GZ and Mn/GZ NCs is improved by the correct integration of NPs into g-C3N4 layers, which aids quicker separation of e/h+ pairs. The samples’ exterior architecture was characterized by SEM, and the subsequent findings are shown in Figure 5a–d. The SEM image of ZnO NPs shows sphere-shaped particles (Figure 5a). Figure 5b shows the free g-C3N4 in the form of stacked thin sheets. The sheets are built up in piles to produce a bulk of nanosheets. The dispersal of spheroidal and rod-shaped particles on the g-C3N4 NSs was revealed in the scanning electron micrograph (SEM) picture of Ni/GZ NCs (Figure 5c), while the Mn/GZ NCs exhibited the fluffier sheet morphology, as shown in Figure 5d. It is widely accepted that this kind of composite shape considerably improves the active sites and surface area, and, correspondingly, the catalytic proficiency of the heterojunction composites.
An EDX study was performed to determine the chemical and elemental makeup of the produced composites. The Zn and O peaks were seen in the ZnO EDX spectra, confirming the purity of NPs (Figure 6a). Only peaks for the elements carbon and nitrogen were seen in the EDX spectrum of g-C3N4 (Figure 6b). The EDX spectra of Ni/GZ (Figure 6c) and Mn/GZ (Figure 6d) NCs showed elemental peaks for carbon, nitrogen, zinc, and oxygen, in addition to peaks for nickel and manganese, respectively. The elemental data of each fabricated sample has been given by Tables S3–S6.

3.4. Optical Properties

UV-vis DRS experiments were employed to investigate the optical characteristics of the synthesized samples, as shown in Figure 7a. The corresponding Eg values of the synthesized samples were assessed by plotting the Kubelka-Munk function, as shown in Figure 7b [54]. The ZnO sample’s reflectance band edge was about 375 nm, and it exhibited the greatest band gap value of 3.1 eV. In contrast, Mn/GZ NC had the smallest band gap energy of 1.9 eV. When g-C3N4 was incorporated into ZnO, the UV-vis reflectance decreased (or the absorption increased) throughout the entire wavelength range. These results are in accord with earlier results [55]. The band edge of GZ photocatalysts was somewhat expanded to a longer wavelength after adding g-C3N4. The same trend was also observed for Mn/GZ and Ni/GZ samples. Further, the integration of Ni/Mn ions into ZnO along with the addition of g-C3N4 broadened the light absorption of Mn/GZ and Ni/GZ photocatalysts into the visible region, as compared to that of pure ZnO [56]. The estimated band gap energies for GZ and Ni/GZ NC were 2.8 and 2.6 eV, respectively. These results indicate that the rate of e/h+ pair recombination is reduced in Mn/GZ and Ni/GZ NCs.

3.5. Photocatalytic Study

The photocatalytic efficiency of the produced materials was observed under sunlight in two phases. In the first phase, the photocatalytic activity of the fabricated set of (Mn/Ni)/(20, 30, 40, 50, 60, 70, & 80) GZ NCs was investigated using EBT and MO solutions in sunlight (Figure 8a–d). The rates of dye decoloration were monitored using a UV-Vis spectrophotometer. To achieve adsorption-desorption equilibrium, each dye-composite solution was maintained in the dark for one hour before being exposed to the sun and realize the created materials’ adsorption capacities. The Mn/GZ NCs in the adsorption experiment absorbed various amounts of EBT (Figure 8a) and MO dye (Figure 8b). The maximum dye adsorption (20% of EBT and 19% of MO) was observed for the Mn/50 GZ NCs as compared to other Mn/GZ NCs in the synthesized series. Likewise, in the adsorption experiment, the Ni/60 GZ NC revealed the best adsorption as compared to other Ni/GZ composites in the produced series (23% of EBT and 24% of MO).
The photocatalytic decoloration efficiency levels of Mn/g-C3N4/ZnO NCs against EBT and MO dyes are given by Figure 8a,b, respectively. The photocatalytic activity of Mn/GZ NCs increased with increasing concentration of g-C3N4 up to 50 wt.%. Thus, 50 wt.% g-C3N4 is the optimal concentration, and increasing the g-C3N4 concentration beyond this (<50 wt.%.) causes a reduction in the photocatalytic efficiency of Mn/GZ NCs (Figure 8a,b). Under the impact of sunlight, the Mn/50GZ NC totally decolorized MO and EBT in 70 and 50 min, correspondingly. Likewise, Ni/60GZ NC showed the best photocatalytic efficiency against EBT and MO dyes as compared to other Ni/GZ NCs (Figure 8c,d). The observed optimal concentration of g-C3N4 for Ni/GZ NCs was 60 wt.%. The Ni/60GZ NC completely decolorized EBT and MO in 60 and 90 min, respectively.
There is a possibility that the greater charge separation is responsible for the better photocatalytic efficiency exhibited by the Ni/60GZ and Mn/50GZ NCs [57,58,59]. The Ni/60GZ and Mn/50GZ NCs would exhibit rapid charge transfer and higher absorption of visible radiation as a result of g-C3N4 coupling in ZnO, as well as doping with Mn and Ni. Intrinsically, for Ni/60GZ and Mn/50GZ NCs, the observed optimal weight proportions of g-C3N4 were 60% and 50%, respectively. A decreasing trend in photocatalytic proficiency was observed for NCs when the g-C3N4 concentration increased beyond the optimal limits. This might be because, if we increase the g-C3N4 contents, lengthier transport paths are generated for photoinduced charges, and extra charge recombination centers are created in the NCs [57,58,59]. A preliminary inquiry is needed to analyze this rationalization further.
In the second phase, the comparative photocatalytic efficiency of Mn/50GZ and Ni/60GZ NCs was observed against EBT and MO dyes. From Figure 9a–d, it is evident that Mn/50GZ NC had better photocatalytic efficiency than Ni/60GZ NCs against both dyes. According to recent research, the optical, electrical, and catalytic characteristics of g-C3N4 and ZnO are significantly improved by the addition of Mn and Ni. So, this could be a reason for the highest photocatalytic efficiency of the Mn/50GZ and Ni/60GZ NCs compared to g-C3N4 and ZnO. Additionally, Mn ions have a greater effect on the optical, electrical, and photocatalytic capabilities of g-C3N4 and ZnO than do Ni and other metal ions [60]. This could be the cause for the Mn/50GZ NC’s superior photocatalytic activities when compared to Ni/60GZ NC and other synthesized composites.
The kinetics of EBT dye decolorization are explained by the Langmuir-Hinshelwood model for ZnO, g-C3N4, Ni/60GZ, Mn/50GZ, 60% g-C3N4/ZnO, and 50% g-C3N4/ZnO NCs, as shown in Figure 10a,b [61]. Figure 10a makes it clear that the EBT dye decoloration by the manufactured samples when exposed to sunlight fits to pseudo-first-order dynamics. The determined values of each NC’s decolorization rate constant (k) are displayed as a bar graph in Figure 10b. The highest and lowest “k” values were found in Mn/50GZ (0.027 min−1) and ZnO (0.008 min−1). The Mn/50GZ NCs totally decolored the EBT in 50 min, and their “k” value is 3.3 and 2.58 times higher than those of ZnO (0.008 min−1) and g-C3N4 (0.0105 min−1), which both decolored the EBT in less time. The maximum “k” value for Mn/50GZ supports the maximal synergic activity of g-C3N4, Mn, and ZnO in Mn/50GZ NC, which increased the composite’s active sites and photocatalytic efficiency against EBT dye. Similarly, the observed kinetics of MO dye decoloration by ZnO, g-C3N4, Ni/60GZ, Mn/50GZ, 60% g-C3N4/ZnO, and 50% g-C3N4/ZnO NCs are given in Figure 10c,d. The photocatalytic efficiency of Mn/50GZ and Ni/60GZ NCs was significantly higher than those in various prior reported research, as given in Table 1 [62,63,64,65,66,67]. The most effective photocatalyst was utilized in the recycling study, which was Mn/50GZ NC.

3.6. Photoluminescence Spectra

An important method frequently used to assess the photo-excited (e/h+ pairs) charge separation capability in photocatalytic systems is photoluminescence (PL) spectroscopy [68]. The PL emission spectra of the produced photocatalysts at 365 nm excitation wavelength are shown in Figure S1. The strong emission peak for g-C3N4 can be seen at about 440 nm, which indicates a high rate of recombination for the photo-excited e–h pair of graphitic carbon nitride. When compared to ZnO, Mn/50GZ and Ni/60GZ exhibit a significant decrease in PL intensity. This decrease may be attributed to zinc defects, surface oxygen vacancies, and the growth of Mn/50GZ and Ni/60GZ heterostructures [69]. Mn/50GZ exhibited the smallest PL peak, demonstrating that the strong junction between the Mn, ZnO, and g-C3N4 sheets strongly inhibits e–h pair recombination as compared to the Ni/60GZ heterostructure. This finding demonstrates that the photo-excited electrons used in the generation of reactive oxygen species (ROS) are productively conducted by the Mn/50GZ and Ni/60GZ nanocomposites. As a result, the rate of charge recombination is decreased, and Mn/50GZ and Ni/60GZ NC photocatalytic competence is increased.

3.7. Recyclability and Constancy of the Mn/50GZ Composite

A photocatalyst must sustain its stability and endurance under repeated photocatalytic activities in order to be useful in practical applications. Five consecutive photocatalytic recycling experiments were performed to monitor the stability of the Mn/50GZ NC. In the recyclability test, the Mn/50GZ NC kept up its rate of dye degradation. The composite’s dye decoloration efficiency did not decrease significantly even after the fifth recycle, as shown in Figure 11a. Interestingly, following the fifth cycle, there were no discernible variations in the XRD profile of the Mn/50GZ NC (Figure 11b). Possible explanations for the small reduction in catalytic efficiency of the nanocomposite involve partial blockage of active sites. The Mn/50GZ NC might thus function as a reliable, reusable, and efficient photocatalytic material.
It is crucial to identify the fundamental ROS photoinduced by the Mn/GZ and Ni/GZ NCs responsible for the photocatalytic decomposition of dyes. In order to investigate the potential mechanism of photocatalytic dye decoloration, three chemical traps were employed to capture the ROS generated during the dye decoloration reaction. The influence of the trapping compounds on the dye decoloration reaction is shown in Figure 12a,b. The results show that OH and O2 are the principal reactive species involved in the photocatalytic dye degradation, with 91% and 60% inhibition, respectively. On the other hand, h+, accounting for 14% of the inhibition, is the lesser oxidative species engaging in dye photodegradation.

3.8. Mechanism

The enhanced decoloration of EBT/MO dyes by Ni/GZ and Mn/GZ NCs may be accredited to the production of e/h+ pairs in the produced photocatalysts, as shownvia schematic sketch (Figure 13). When sunlight irradiates the NCs, both ZnO and g-C3N4 are excited, and e/h+ couples are formed on their respective conduction bands (CB) and valence bands (VB) [18]. ZnO and g-C3N4 valence bond electrons are promoted to their conduction bands.The excited electrons from the CB of g-C3N4 (−1.12 eV) are transported to the CB of ZnO (−0.5 eV). On the basis of the CB/VB edge potentials, Photo-induced electrons may easily migrate from the g-C3N4 conduction band (CB) to the CB of ZnO. Simultaneously, the h+ in the VB of ZnO might shift to g-C3N4, because ZnO has higher VB potential than g-C3N4 [70]. To enhance the electron transportation and reduce e/h+ pair recombination, the Mn/Ni metal ions in the Mn/GZ and Ni/GZ NCs not only function as a facilitator at the g-C3N4 and ZnO interface but also decrease the Eg value of ZnO [62]. The produced e and h+ react with the water and oxygen adsorbed on the photocatalyst surface to form radicals (OH and O2) [71]. These radicals are used to decolorize dye molecules by converting them into low-molecular-weight intermediates, which are then oxidized into H2O, CO2, and inorganic ions.

4. Conclusions

In this study, extremely effective (Mn/Ni)/g-C3N4/ZnO (Mn/GZ and Ni/GZ) photocatalysts were made via the co-precipitation method. The assembly and purity of the samples were studied using the XRD, TEM, SEM, EDX, and FTIR methods. Photocatalytic dye decoloration testing of the synthesized (Mn/Ni)/(20, 30, 40, 50, 60, 70, and 80) GZ samples was carried out against MO and EBT, and it was observed that the Mn/50GZ NCs had a very high decolorizing efficiency under sunlight. The Mn/50GZ NC completely decolorized EBT and MO in 50 and 70 min, respectively, while Ni/60GZ NC decolorized the same dyes in 60 and 90 min, respectively. Further, Mn/GZ and Ni/GZ NCs exhibited pseudo-first-order kinetics during the dye-reduction reaction. The number of active sites, surface defects, and reduction in e/h+ pairs in nanocomposites were improved by the successful integration of metal ions, ZnO, on the fragile sheets of g-C3N4, as exhibited by TEM images and PL study. In general, the findings of this work contribute to a better understanding of the processes involved in synthesizing photocatalysts and their uses for the decoloration of organic pollutants from wastewater.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12111388/s1, Figure S1: Photoluminescence spectra of ZnO, Ni/60GZ, Mn/50GZ, and g-C3N4 samples; Table S1: Composition of the synthesized Ni/g-C3N4/ZnO composites; Table S2: Composition of the synthesized Mn/g-C3N4/ZnO composites; Table S3: Elemental data of ZnO nanoparticles; Table S4: Elemental data of g-C3N4 nanosheets; Table S5: Elemental data of Mn/g-C3N4/ZnO nanocomposites; Table S6: Elemental data of Ni/g-C3N4/ZnO nanocomposites.

Author Contributions

Formal analysis, M.S. and O.M.; methodology, M.J. and S.K.A.; resources, M.M.F. and M.S.K.; software, M.Y.A.; supervision, S.S. and M.J.; validation, M.H.A.N.; visualization, I.S.S. and D.A.H.; writing – original draft, M.A.Q.; writing—review & editing, S.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research work was funded by the Deanship of Scientific Research at Majmaah University, under Project Number R-2022-335, Al Majmaah, Saudi Arabia.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to thank the Deanship of Scientific Research at Majmaah University for supporting this work under Project Number R-2022-335.

Conflicts of Interest

The authors disclose no possible conflict of interest.

References

  1. Finke, E.-J.; Beyer, W.; Loderstädt, U.; Frickmann, H. The risk of contracting anthrax from spore-contaminated soil—A military medical perspective. Eur. J. Microbiol. Immunol. 2020, 10, 29–63. [Google Scholar] [CrossRef] [PubMed]
  2. Zou, L.; Wang, H.; Jiang, X.; Yuan, G.; Wang, X. Enhanced photocatalytic efficiency in degrading organic dyes by coupling CdS nanowires with ZnFe2O4 nanoparticles. Sol. Energy 2020, 195, 271–277. [Google Scholar] [CrossRef]
  3. Basaleh, A. Construction of mesoporous ZnFe2O4-g-C3N4 nanocomposites for enhanced photocatalytic degradation of acridine orange dye under visible light illumination adopting soft-and hard-template-assisted routines. J. Mater. Res. Technol. 2021, 11, 1260–1271. [Google Scholar] [CrossRef]
  4. Dhasmana, A.; Uniyal, S.; Kumar, V.; Gupta, S.; Kesari, K.K.; Haque, S.; Lohani, M.; Pandey, J. Scope of nanoparticles in environmental toxicant remediation. In Environmental Biotechnology: For Sustainable Future; Springer: Singapore, 2019; pp. 31–44. [Google Scholar]
  5. Sharma, S.; Dutta, V.; Raizada, P.; Hosseini-Bandegharaei, A.; Thakur, V.; Nguyen, V.-H.; VanLe, Q.; Singh, P. An overview of heterojunctioned ZnFe2O4 photocatalyst for enhanced oxidative water purification. J. Environ. Chem. Eng. 2021, 9, 105812. [Google Scholar]
  6. Alshorifi, F.T.; Alswat, A.A.; Mannaa, M.A.; Alotaibi, M.T.; El-Bahy, S.M.; Salama, R.S. Facile and Green Synthesis of Silver Quantum Dots Immobilized onto a Polymeric CTS–PEO Blend for the Photocatalytic Degradation of p-Nitrophenol. ACS Omega 2021, 6, 30432–30441. [Google Scholar] [CrossRef]
  7. Pirhashemi, M.; Habibi-Yangjeh, A.; Pouran, S.R. Review on the criteria anticipated for the fabrication of highly efficient ZnO-based visible-light-driven photocatalysts. J. Ind. Eng. Chem. 2018, 62, 1–25. [Google Scholar] [CrossRef]
  8. Jilani, A.; Ansari, M.O.; Ur Rehman, G.; Shakoor, M.B.; Hussain, S.Z.; Othman, M.H.D.; Ahmad, S.R.; Dustgeer, M.R.; Alshahrie, A. Phenol removal and hydrogen production from water: Silver nanoparticles decorated on polyaniline wrapped zinc oxide nanorods. J. Ind. Eng. Chem. 2022, 109, 347–358. [Google Scholar] [CrossRef]
  9. Shekofteh-Gohari, M.; Habibi-Yangjeh, A.; Abitorabi, M.; Rouhi, A. Magnetically separable nanocomposites based on ZnO and their applications in photocatalytic processes: A review. Crit. Rev. Environ. Sci. Technol. 2018, 48, 806–857. [Google Scholar] [CrossRef]
  10. Jin, Y.; Long, J.; Ma, X.; Zhou, T.; Zhang, Z.; Lin, H.; Long, J.; Wang, X. Synthesis of caged iodine-modified ZnO nanomaterials and study on their visible light photocatalytic antibacterial properties. Appl. Catal. B Environ. 2019, 256, 117873. [Google Scholar] [CrossRef]
  11. Kuang, M.; Zhang, J.; Wang, W.; Chen, J.; Liu, R.; Xie, S.; Wang, J.; Ji, Z. Synthesis of octahedral-like ZnO/ZnFe2O4 heterojunction photocatalysts with superior photocatalytic activity. Solid State Sci. 2019, 96, 105901. [Google Scholar] [CrossRef]
  12. Umar, K.; Adnan, R. Biomass Mediated Green Synthesis of ZnO/TiO2 Nanocomposite and Enhanced Photocatalytic Activity for the Decolorization of Rhodamine B under Visible Light. In Key Engineering Materials; Trans Tech Publications Ltd.: Zurich, Switzerland, 2022; pp. 36–42. [Google Scholar]
  13. Hong, J.; Xia, X.; Wang, Y.; Xu, R. Mesoporous carbon nitride with in situ sulfur doping for enhanced photocatalytic hydrogen evolution from water under visible light. J. Mater. Chem. 2012, 22, 15006–15012. [Google Scholar] [CrossRef]
  14. Djerdj, I.; Jagličić, Z.; Arčon, D.; Niederberger, M. Co-doped ZnO nanoparticles: Minireview. Nanoscale 2010, 2, 1096–1104. [Google Scholar] [CrossRef] [PubMed]
  15. Chauhan, N.; Singh, V.; Kumar, S.; Dhiman, R. Influence of nickel, silver, and sulphur doping on the photocatalytic efficiency of mesoporous ZnO nanoparticles. Arab. J. Sci. Eng. 2020, 45, 249–259. [Google Scholar] [CrossRef]
  16. Ahmad, I. Comparative study of metal (Al, Mg, Ni, Cu and Ag) doped ZnO/g-C3N4 composites: Efficient photocatalysts for the degradation of organic pollutants. Sep. Purif. Technol. 2020, 251, 117372. [Google Scholar] [CrossRef]
  17. Zhang, S.; Su, C.; Ren, H.; Li, M.; Zhu, L.; Ge, S.; Wang, M.; Zhang, Z.; Li, L.; Cao, X. In-situ fabrication of g-C3N4/ZnO nanocomposites for photocatalytic degradation of methylene blue: Synthesis procedure does matter. Nanomaterials 2019, 9, 215. [Google Scholar] [CrossRef] [Green Version]
  18. Zhang, H.; Zhu, C.; Zhang, G.; Li, M.; Tang, Q.; Cao, J. Palladium modified ZnFe2O4/g-C3N4 nanocomposite as an efficiently magnetic recycling photocatalyst. J. Solid State Chem. 2020, 288, 121389. [Google Scholar] [CrossRef]
  19. Chen, Z.; Zhang, S.; Liu, Y.; Alharbi, N.S.; Rabah, S.O.; Wang, S.; Wang, X. Synthesis and fabrication of g-C3N4-based materials and their application in elimination of pollutants. Sci. Total Environ. 2020, 731, 139054. [Google Scholar] [CrossRef]
  20. Bai, X.; Wang, L.; Zong, R.; Zhu, Y. Photocatalytic activity enhanced via g-C3N4 nanoplates to nanorods. J. Phys. Chem. C 2013, 117, 9952–9961. [Google Scholar] [CrossRef]
  21. Wang, X.; Chen, X.; Thomas, A.; Fu, X.; Antonietti, M. Metal-containing carbon nitride compounds: A new functional organic–metal hybrid material. Adv. Mater. 2009, 21, 1609–1612. [Google Scholar] [CrossRef]
  22. Dong, F.; Sun, Y.; Wu, L.; Fu, M.; Wu, Z. Facile transformation of low cost thiourea into nitrogen-rich graphitic carbon nitride nanocatalyst with high visible light photocatalytic performance. Catal. Sci. Technol. 2012, 2, 1332–1335. [Google Scholar] [CrossRef]
  23. Shen, B.; Hong, Z.; Chen, Y.; Lin, B.; Gao, B. Template-free synthesis of a novel porous g-C3N4 with 3D hierarchical structure for enhanced photocatalytic H2 evolution. Mater. Lett. 2014, 118, 208–211. [Google Scholar] [CrossRef]
  24. Chang, F.; Xie, Y.; Li, C.; Chen, J.; Luo, J.; Hu, X.; Shen, J. A facile modification of g-C3N4 with enhanced photocatalytic activity for degradation of methylene blue. Appl. Surf. Sci. 2013, 280, 967–974. [Google Scholar] [CrossRef]
  25. Kuriki, R.; Sekizawa, K.; Ishitani, O.; Maeda, K. Visible-light-driven CO2 reduction with carbon nitride: Enhancing the activity of ruthenium catalysts. Angew. Chem. Int. Ed. 2015, 54, 2406–2409. [Google Scholar] [CrossRef] [PubMed]
  26. Mao, L.; Jin, H.; Wang, M.; Hu, Y.; Chen, S.; He, Q.; Chang, J.; Hong, C.; Zhou, Y.; Wang, D. Neurologic manifestations of hospitalized patients with coronavirus disease 2019 in Wuhan, China. JAMA Neurol. 2020, 77, 683–690. [Google Scholar] [CrossRef] [Green Version]
  27. Guo, Y.; Sun, X.; Chen, Q.; Liu, Y.; Lou, X.; Zhang, L.; Zhang, X.; Li, Y.; Guan, J. Photocatalytic Applications of g-C3N4 Based on Bibliometric Analysis. Catalysts 2022, 12, 1017. [Google Scholar] [CrossRef]
  28. Suzuki, V.Y.; Amorin, L.H.; Fabris, G.S.; Dey, S.; Sambrano, J.R.; Cohen, H.; Oron, D.; La Porta, F.A. Enhanced Photocatalytic and Photoluminescence Properties Resulting from Type-I Band Alignment in the Zn2GeO4/g-C3N4 Nanocomposites. Catalysts 2022, 12, 692. [Google Scholar] [CrossRef]
  29. Machín, A.; Fontánez, K.; Duconge, J.; Cotto, M.C.; Petrescu, F.I.; Morant, C.; Márquez, F. Photocatalytic degradation of fluoroquinolone antibiotics in solution by Au@ ZnO-rGO-gC3N4 composites. Catalysts 2022, 12, 166. [Google Scholar] [CrossRef]
  30. Yaqoob, A.; Mohd Noor, N.H.B.; Umar, K.; Adnan, R.; Ibrahim, M.N.M.; Rashid, M. Graphene oxide-ZnO nanocomposite: An efficient visible light photocatalyst for degradation of rhodamine B. Appl. Nanosci. 2021, 11, 1291–1302. [Google Scholar] [CrossRef]
  31. Li, L.; Sun, S.-Q.; Wang, Y.-X.; Wang, C.-Y. Facile synthesis of ZnO/g-C3N4 composites with honeycomb-like structure by H2 bubble templates and their enhanced visible light photocatalytic performance. J. Photochem. Photobiol. A Chem. 2018, 355, 16–24. [Google Scholar] [CrossRef]
  32. Wang, J.-C.; Cui, C.-X.; Li, Y.; Liu, L.; Zhang, Y.-P.; Shi, W. Porous Mn doped g-C3N4 photocatalysts for enhanced synergetic degradation under visible-light illumination. J. Hazard. Mater. 2017, 339, 43–53. [Google Scholar] [CrossRef]
  33. Wang, M.; Guo, P.; Zhang, Y.; Lv, C.; Liu, T.; Chai, T.; Xie, Y.; Wang, Y.; Zhu, T. Synthesis of hollow lantern-like Eu (III)-doped g-C3N4 with enhanced visible light photocatalytic perfomance for organic degradation. J. Hazard. Mater. 2018, 349, 224–233. [Google Scholar] [CrossRef] [PubMed]
  34. Tonda, S.; Kumar, S.; Kandula, S.; Shanker, V. Fe-doped and-mediated graphitic carbon nitride nanosheets for enhanced photocatalytic performance under natural sunlight. J. Mater. Chem. A 2014, 2, 6772–6780. [Google Scholar] [CrossRef]
  35. Wu, L.; Liu, X.; Lv, G.; Zhu, R.; Tian, L.; Liu, M.; Li, Y.; Rao, W.; Liu, T.; Liao, L. Study on the adsorption properties of methyl orange by natural one-dimensional nano-mineral materials with different structures. Sci. Rep. 2021, 11, 10640. [Google Scholar] [CrossRef] [PubMed]
  36. Boudouaia, N.; Bengharez, Z.; Jellali, S. Preparation and characterization of chitosan extracted from shrimp shells waste and chitosan film: Application for Eriochrome black T removal from aqueous solutions. Appl. Water Sci. 2019, 9, 91. [Google Scholar] [CrossRef] [Green Version]
  37. Shi, L.; Liang, L.; Ma, J.; Wang, F.; Sun, J. Remarkably enhanced photocatalytic activity of ordered mesoporous carbon/g-C3N4 composite photocatalysts under visible light. Dalton Trans. 2014, 43, 7236–7244. [Google Scholar] [CrossRef]
  38. Sher, M.; Shahid, S.; Javed, M. Synthesis of a novel ternary (g-C3N4 nanosheets loaded with Mo doped ZnOnanoparticles) nanocomposite for superior photocatalytic and antibacterial applications. J. Photochem. Photobiol. B Biol. 2021, 219, 112202. [Google Scholar] [CrossRef]
  39. Slobodan, D.; Nada, T.; Zhiyu, P. Absorption spectroscopy of EBT model GAFCHROMIC™ film. Med. Phys. 2007, 34, 2207. [Google Scholar]
  40. Jethave, G.; Fegade, U.; Attarde, S.; Ingle, S. Decontamination study of Eriochrome Black-T from waste water by using AlTiPbO nanoparticles (ATPO-NPs) for sustainable clean environment. J. Water Environ. Nanotechnol. 2019, 4, 263–274. [Google Scholar]
  41. Jamshidi, E.; Manteghi, F. Methyl Orange Adsorption by Fe2O3@ Co-Al-Layered Double Hydroxide. Multidiscip. Digit. Publ. Inst. Proc. 2020, 41, 64. [Google Scholar]
  42. Qamar, M.A.; Shahid, S.; Javed, M. Synthesis of dynamic g-C3N4/Fe@ ZnO nanocomposites for environmental remediation applications. Ceram. Int. 2020, 46, 22171–22180. [Google Scholar] [CrossRef]
  43. Palanivel, B.; Maiyalagan, T.; Jayarman, V.; Ayyappan, C.; Alagiri, M. Rational design of ZnFe2O4/g-C3N4 nanocomposite for enhanced photo-Fenton reaction and supercapacitor performance. Appl. Surf. Sci. 2019, 498, 143807. [Google Scholar] [CrossRef]
  44. Saeed, H.; Nadeem, N.; Zahid, M.; Yaseen, M.; Noreen, S.; Jilani, A.; Shahid, I. Mixed metal ferrite (Mn0. 6Zn0. 4Fe2O4) intercalated g-C3N4 nanocomposite: Efficient sunlight driven photocatalyst for methylene blue degradation. Nanotechnology 2021, 32, 505714. [Google Scholar] [CrossRef] [PubMed]
  45. Xiao, Z.; Liming, W.; Yong, S.; Lihui, X.; Mingrui, X. Preparation and photocatalytic properties of magnetic g-C3N4/MnZnFe2O4 composite. J. Phys. Conf. Ser. 2021, 1790, 012075. [Google Scholar] [CrossRef]
  46. Shanmugam, N.; Suthakaran, S.; Kannadasan, N.; Sathishkumar, K. Synthesis and characterization of Te doped ZnO nanosheets for photocatalytic application. JO Heterocycl. 2015, 105, 15–20. [Google Scholar] [CrossRef]
  47. Ahmad, A.; Wei, Y.; Syed, F.; Tahir, K.; Rehman, A.U.; Khan, A.; Ullah, S.; Yuan, Q. The effects of bacteria-nanoparticles interface on the antibacterial activity of green synthesized silver nanoparticles. Microb. Pathog. 2017, 102, 133–142. [Google Scholar] [CrossRef]
  48. Kotresh, M.; Patil, M.; Inamdar, S. Reaction temperature based synthesis of ZnO nanoparticles using co-precipitation method: Detailed structural and optical characterization. Optik 2021, 243, 167506. [Google Scholar] [CrossRef]
  49. Sher, M.; Khan, S.A.; Shahid, S.; Javed, M.; Qamar, M.A.; Chinnathambi, A.; Almoallim, H.S. Synthesis of novel ternary hybrid g-C3N4@ Ag-ZnO nanocomposite with Z-scheme enhanced solar light-driven methylene blue degradation and antibacterial activities. J. Environ. Chem. Eng. 2021, 9, 105366. [Google Scholar] [CrossRef]
  50. Jourshabani, M.; Shariatinia, Z.; Badiei, A. Facile one-pot synthesis of cerium oxide/sulfur-doped graphitic carbon nitride (g-C3N4) as efficient nanophotocatalysts under visible light irradiation. J. Colloid Interface Sci. 2017, 507, 59–73. [Google Scholar] [CrossRef]
  51. Kalisamy, P.; Lallimathi, M.; Suryamathi, M.; Palanivel, B.; Venkatachalam, M. ZnO-embedded S-doped g-C3N4 heterojunction: Mediator-free Z-scheme mechanism for enhanced charge separation and photocatalytic degradation. RSC Adv. 2020, 10, 28365–28375. [Google Scholar] [CrossRef]
  52. Norouzzadeh, P.; Golzan, M.; Mabhouti, K.; Naderali, R. Effect of Mn-substitution on impedance spectroscopy and magnetic properties of Al-doped ZnO nanoparticles. Nanotechnology 2020, 31, 325704. [Google Scholar] [CrossRef]
  53. Rana, A.K.; Kumar, Y.; Rajput, P.; Jha, S.N.; Bhattacharyya, D.; Shirage, P.M. Search for origin of room temperature ferromagnetism properties in Ni-doped ZnO nanostructure. ACS Appl. Mater. Interfaces 2017, 9, 7691–7700. [Google Scholar] [CrossRef] [Green Version]
  54. Simmons, E. Reflectance spectroscopy: Application of the Kubelka-Munk theory to the rates of photoprocesses of powders. Appl. Opt. 1976, 15, 951–954. [Google Scholar] [CrossRef] [PubMed]
  55. Luu Thi, L.A.; Neto, M.M.; Van, T.P.; Nguyen Ngoc, T.; Nguyen Thi, T.M.; Nguyen, X.S.; Nguyen, C.T. In Situ g-C3N4@ Zno Nanocomposite: One-Pot Hydrothermal Synthesis and Photocatalytic Performance under Visible Light Irradiation. Adv. Mater. Sci. Eng. 2021, 2021, 6651633. [Google Scholar] [CrossRef]
  56. Rosli, N.I.M.; Lam, S.-M.; Sin, J.-C.; Satoshi, I.; Mohamed, A.R. Photocatalytic performance of ZnO/g-C3N4 for removal of phenol under simulated sunlight irradiation. J. Environ. Eng. 2018, 144, 04017091. [Google Scholar] [CrossRef]
  57. Mohamed, R.; Shawky, A. CNT supported Mn-doped ZnO nanoparticles: Simple synthesis and improved photocatalytic activity for degradation of malachite green dye under visible light. Appl. Nanosci. 2018, 8, 1179–1188. [Google Scholar] [CrossRef]
  58. Labhane, P.; Patle, L.; Sonawane, G.; Sonawane, S. Fabrication of ternary Mn doped ZnO nanoparticles grafted on reduced graphene oxide (RGO) sheet as an efficient solar light driven photocatalyst. Chem. Phys. Lett. 2018, 710, 70–77. [Google Scholar] [CrossRef]
  59. Paul, D.R.; Sharma, R.; Panchal, P.; Malik, R.; Sharma, A.; Tomer, V.K.; Nehra, S. Silver doped graphitic carbon nitride for the enhanced photocatalytic activity towards organic dyes. J. Nanosci. Nanotechnol. 2019, 19, 5241–5248. [Google Scholar] [CrossRef]
  60. Sun, J.-X.; Yuan, Y.-P.; Qiu, L.-G.; Jiang, X.; Xie, A.-J.; Shen, Y.-H.; Zhu, J.-F. Fabrication of composite photocatalyst g-C3N4–ZnO and enhancement of photocatalytic activity under visible light. Dalton Trans. 2012, 41, 6756–6763. [Google Scholar] [CrossRef]
  61. Farouq, R. Coupling Adsorption-Photocatalytic Degradation of Methylene Blue and Maxilon Red. J. Fluoresc. 2022, 32, 1381–1388. [Google Scholar] [CrossRef]
  62. Iqbala, S.; Iqbala, M.; Sibtaina, A.; Iqbalb, A.; Farooqic, Z.H.; Ahmadd, S.; Mustafaa, K.; Musaddiqa, S. Solar driven photocatalytic degradation of organic pollutants via Bi. Desalin. Water Treat. 2021, 216, 140–150. [Google Scholar] [CrossRef]
  63. Sobahi, T.R.; Amin, M. Synthesis of ZnO/ZnFe2O4/Pt nanoparticles heterojunction photocatalysts with superior photocatalytic activity. Ceram. Int. 2020, 46, 3558–3564. [Google Scholar] [CrossRef]
  64. Falak, P.; Hassanzadeh-Tabrizi, S.; Saffar-Teluri, A. Synthesis, characterization, and magnetic properties of ZnO-ZnFe2O4 nanoparticles with high photocatalytic activity. J. Magn. Magn. Mater. 2017, 441, 98–104. [Google Scholar] [CrossRef]
  65. Jaffari, Z.H.; Lam, S.-M.; Sin, J.-C.; Zeng, H.; Mohamed, A.R. Magnetically recoverable Pd-loaded BiFeO3 microcomposite with enhanced visible light photocatalytic performance for pollutant, bacterial and fungal elimination. Sep. Purif. Technol. 2020, 236, 116195. [Google Scholar] [CrossRef]
  66. Wu, Z.; Chen, X.; Liu, X.; Yang, X.; Yang, Y. A ternary magnetic recyclable ZnO/Fe3O4/g-C3N4 composite photocatalyst for efficient photodegradation of monoazo dye. Nanoscale Res. Lett. 2019, 14, 147. [Google Scholar] [CrossRef]
  67. Kulkarni, S.D.; Kumbar, S.; Menon, S.G.; Choudhari, K.; Santhosh, C. Magnetically separable core–shell ZnFe2O4@ ZnO nanoparticles for visible light photodegradation of methyl orange. Mater. Res. Bull. 2016, 77, 70–77. [Google Scholar] [CrossRef]
  68. Kumaresan, N.; Sinthiya, M.M.A.; Kumar, M.P.; Ravichandran, S.; Babu, R.R.; Sethurman, K.; Ramamurthi, K. Investigation on the g-C3N4 encapsulated ZnO nanorods heterojunction coupled with GO for effective photocatalytic activity under visible light irradiation. Arab. J. Chem. 2020, 13, 2826–2843. [Google Scholar] [CrossRef]
  69. Anitha, S.; Muthukumaran, S. Microstructure, crystallographic and photoluminescence examination of Ni doped ZnO nanoparticles co-doped with Co by sol–gel method. J. Mater. Sci. Mater. Electron. 2017, 28, 12995–13005. [Google Scholar] [CrossRef]
  70. Shi, Y.; Li, L.; Xu, Z.; Sun, H.; Amin, S.; Guo, F.; Shi, W.; Li, Y. Engineering of 2D/3D architectures type II heterojunction with high-crystalline g-C3N4 nanosheets on yolk-shell ZnFe2O4 for enhanced photocatalytic tetracycline degradation. Mater. Res. Bull. 2022, 150, 111789. [Google Scholar] [CrossRef]
  71. Wu, Y.; Wang, Y.; Di, A.; Yang, X.; Chen, G. Enhanced photocatalytic performance of hierarchical ZnFe2O4/g-C3N4 heterojunction composite microspheres. Catal. Lett. 2018, 148, 2179–2189. [Google Scholar] [CrossRef]
Figure 1. Schematic description of the synthesis of metal/g-C3N4/ZnO nanocomposites.
Figure 1. Schematic description of the synthesis of metal/g-C3N4/ZnO nanocomposites.
Catalysts 12 01388 g001
Figure 2. FTIR spectra of ZnO, g-C3N4, and (Ni/Mn)/g-C3N4/ZnO NCs.
Figure 2. FTIR spectra of ZnO, g-C3N4, and (Ni/Mn)/g-C3N4/ZnO NCs.
Catalysts 12 01388 g002
Figure 3. XRD patterns of ZnO, g-C3N4, and (Ni/Mn)/g-C3N4/ZnO NCs (a) Magnified XRD patterns of ZnO and (Ni/Mn)/g-C3N4/ZnO NCs (b).
Figure 3. XRD patterns of ZnO, g-C3N4, and (Ni/Mn)/g-C3N4/ZnO NCs (a) Magnified XRD patterns of ZnO and (Ni/Mn)/g-C3N4/ZnO NCs (b).
Catalysts 12 01388 g003
Figure 4. Morphological analysis of ZnO (a), g-C3N4 (b), Ni/g-C3N4/ZnO (c), and Mn/g-C3N4/ZnO (d) by TEM.
Figure 4. Morphological analysis of ZnO (a), g-C3N4 (b), Ni/g-C3N4/ZnO (c), and Mn/g-C3N4/ZnO (d) by TEM.
Catalysts 12 01388 g004
Figure 5. Morphological analysis of ZnO (a), g-C3N4 (b), Ni/g-C3N4/ZnO (c), and Mn/g-C3N4/ZnO (d) by SEM.
Figure 5. Morphological analysis of ZnO (a), g-C3N4 (b), Ni/g-C3N4/ZnO (c), and Mn/g-C3N4/ZnO (d) by SEM.
Catalysts 12 01388 g005
Figure 6. EDX analysis of ZnO (a), g-C3N4 (b), g-C3N4/ZnO (c), Ni/g-C3N4/ZnO (c), and Mn/g-C3N4/ZnO (d).
Figure 6. EDX analysis of ZnO (a), g-C3N4 (b), g-C3N4/ZnO (c), Ni/g-C3N4/ZnO (c), and Mn/g-C3N4/ZnO (d).
Catalysts 12 01388 g006
Figure 7. UV-vis DRS spectra (a) and band gap energy of ZnO, GZ, Mn/GZ, and Ni/GZ NCs (b).
Figure 7. UV-vis DRS spectra (a) and band gap energy of ZnO, GZ, Mn/GZ, and Ni/GZ NCs (b).
Catalysts 12 01388 g007
Figure 8. Photocatalytic decoloration rates of EBT (a) and MO dye (b) by Mn/50GZ NC, and decoloration rates of EBT (c) and MO dye (d) by Mn/50GZ NC.
Figure 8. Photocatalytic decoloration rates of EBT (a) and MO dye (b) by Mn/50GZ NC, and decoloration rates of EBT (c) and MO dye (d) by Mn/50GZ NC.
Catalysts 12 01388 g008
Figure 9. Photocatalytic dye decoloration rates of EBT (a) and MO (b) and bar graphs of the percentage of decoloration of EBT (c) and MO (d) by Ni/60GZ, Mn/50GZ, 50-GZ, 60-GZ, and g-C3N4.
Figure 9. Photocatalytic dye decoloration rates of EBT (a) and MO (b) and bar graphs of the percentage of decoloration of EBT (c) and MO (d) by Ni/60GZ, Mn/50GZ, 50-GZ, 60-GZ, and g-C3N4.
Catalysts 12 01388 g009
Figure 10. EBT decoloration kinetic characteristics (a) and k (min−1) values (b) and MO decoloration kinetic characteristics (c) and k (min−1) values (d) by Ni/60GZ, Mn/50GZ, 50-GZ, 60-GZ, and g-C3N4.
Figure 10. EBT decoloration kinetic characteristics (a) and k (min−1) values (b) and MO decoloration kinetic characteristics (c) and k (min−1) values (d) by Ni/60GZ, Mn/50GZ, 50-GZ, 60-GZ, and g-C3N4.
Catalysts 12 01388 g010
Figure 11. XRD analysis of Mn/50GZ NCs (a) and the recyclability test (b) after the fifth cycle.
Figure 11. XRD analysis of Mn/50GZ NCs (a) and the recyclability test (b) after the fifth cycle.
Catalysts 12 01388 g011
Figure 12. Scavenger effects (a) and the capture of reactive species with percentage inhibition (b).
Figure 12. Scavenger effects (a) and the capture of reactive species with percentage inhibition (b).
Catalysts 12 01388 g012
Figure 13. Mechanism of MO/EBT dye photocatalytic decoloration using Mn/GZ and Ni/GZ NCs as photocatalysts.
Figure 13. Mechanism of MO/EBT dye photocatalytic decoloration using Mn/GZ and Ni/GZ NCs as photocatalysts.
Catalysts 12 01388 g013
Table 1. Comparison of decoloration capabilities of Mn/50GZ with some earlier works.
Table 1. Comparison of decoloration capabilities of Mn/50GZ with some earlier works.
Sr. NoPhotocatalystContaminantLight SourceDye (mg/L)Catalyst (mg/L)Radiation Time (min.)Decoloration%Ref.
1Mn-ZnO/CSACBGSunlight3020017596[62]
2ZnO/Fe2O4/PtCiprofloxacinXe lamp1015075100[63]
3Mn-ZnO/RGORhBVisible202014099[58]
4ZnO/ZnFe2O4MBVisible105010098[64]
5Pt-BiFeO3MGVisible1010024096[65]
6ZnO/Fe3O4/g-C3N4MOVisible301015097.87[66]
7Mn-ZnO/CNTMGVisible50100150>95[57]
8ZnFe2O4@ZnOMOVisible5666024099[67]
9Mn/50% g-C3N4/ZnOEBTSunlight1020050100Present work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Qamar, M.A.; Shahid, S.; Javed, M.; Shariq, M.; Fadhali, M.M.; Madkhali, O.; Ali, S.K.; Syed, I.S.; Awaji, M.Y.; Shakir Khan, M.; et al. Accelerated Decoloration of Organic Dyes from Wastewater Using Ternary Metal/g-C3N4/ZnO Nanocomposites: An Investigation of Impact of g-C3N4 Concentration and Ni and Mn Doping. Catalysts 2022, 12, 1388. https://doi.org/10.3390/catal12111388

AMA Style

Qamar MA, Shahid S, Javed M, Shariq M, Fadhali MM, Madkhali O, Ali SK, Syed IS, Awaji MY, Shakir Khan M, et al. Accelerated Decoloration of Organic Dyes from Wastewater Using Ternary Metal/g-C3N4/ZnO Nanocomposites: An Investigation of Impact of g-C3N4 Concentration and Ni and Mn Doping. Catalysts. 2022; 12(11):1388. https://doi.org/10.3390/catal12111388

Chicago/Turabian Style

Qamar, Muhammad Azam, Sammia Shahid, Mohsin Javed, Mohammad Shariq, Mohammed M. Fadhali, Osama Madkhali, Syed Kashif Ali, Imam Saheb Syed, Majed Yusef Awaji, Mohd. Shakir Khan, and et al. 2022. "Accelerated Decoloration of Organic Dyes from Wastewater Using Ternary Metal/g-C3N4/ZnO Nanocomposites: An Investigation of Impact of g-C3N4 Concentration and Ni and Mn Doping" Catalysts 12, no. 11: 1388. https://doi.org/10.3390/catal12111388

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop