Next Article in Journal
Bio-Adipic Acid Production from Muconic Acid Hydrogenation on Palladium-Transition Metal (Ni and Zn) Bimetallic Catalysts
Next Article in Special Issue
Application of BiOX Photocatalyst to Activate Peroxydisulfate Ion-Investigation of a Combined Process for the Removal of Organic Pollutants from Water
Previous Article in Journal
Influence of the Valence of Iron on the NO Reduction by CO over Cu-Fe-Mordenite
Previous Article in Special Issue
Photocatalytic Decomposition of Azo Dyes and Phenols Using Polymer Composites Containing Nanostructured Poly(Titanium Oxide) Doped with Gold or Silver Nanoparticles under Light Irradiation in a Wide Wavelength Range
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation of S-Scheme g-C3N4/ZnO Heterojunction Composite for Highly Efficient Photocatalytic Destruction of Refractory Organic Pollutant

1
Department of Engineering Fundamental Sciences, Faculty of Engineering, Tarsus University, 33400 Tarsus, Turkey
2
Department of Environmental Engineering, Mersin University, 33343 Mersin, Turkey
3
Department of Environmental Engineering, Faculty of Engineering, Gebze Technical University, 41400 Gebze, Turkey
4
Research Laboratory of Advanced Water and Wastewater Treatment Processes, Department of Applied Chemistry, Faculty of Chemistry, University of Tabriz, Tabriz 51666-16471, Iran
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(3), 485; https://doi.org/10.3390/catal13030485
Submission received: 6 December 2022 / Revised: 15 February 2023 / Accepted: 22 February 2023 / Published: 27 February 2023

Abstract

:
In this study, graphitic carbon nitride (g-C3N4)-based ZnO heterostructure was synthesized using a facile calcination method with urea and zinc nitrate hexahydrate as the initiators. According to the scanning electron microscopic (SEM) images, spherical ZnO particles can be seen along the g-C3N4 nanosheets. Additionally, the X-ray diffraction (XRD) analysis reveals the successful synthesis of the g-C3N4/ZnO. The photocatalytic activity of the synthesized catalyst was tested for the decolorization of crystal violet (CV) as an organic refractory contaminant. The impacts of ZnO molar ratio, catalyst amount, CV concentration, and H2O2 concentration on CV degradation efficiency were investigated. The obtained outcomes conveyed that the ZnO molar ratio in the g-C3N4 played a prominent role in the degradation efficiency, in which the degradation efficiency reached 95.9% in the presence of 0.05 mmol of ZnO and 0.10 g/L of the catalyst in 10 mg/L of CV through 120 min under UV irradiation. Bare g-C3N4 was also tested for dye decolorization, and a 76.4% dye removal efficiency was obtained. The g-C3N4/ZnO was also tested for adsorption, and a 32.3% adsorption efficiency was obtained. Photocatalysis, in comparison to adsorption, had a dominant role in the decolorization of CV. Lastly, the results depicted no significant decrement in the CV degradation efficiency in the presence of the g-C3N4/ZnO photocatalyst after five consecutive runs.

1. Introduction

Wastewaters from the textile industry have adverse impacts on aquatic life due to the high concentration of hazardous chemicals released into the environment [1]. Diverse salts, chlorinated chemicals, and surfactants are commonly found in such wastewaters [2]. Since most dyes own high reactivity, even a trace amount of dye in water affects the gas solubility in water bodies [3,4]. Hence, researchers have devised several treatment techniques for the disintegration of hazardous contaminants [5]. Amongst these techniques, the photocatalytic process as a commonly used method for wastewater treatment has gained high interest [6].
The photocatalytic process is considered an eco-friendly, low-cost, and effective alternative method for decreasing water pollution when compared to other chemical, physical, and biological approaches. [7,8]. By using light irradiation, electrons in the valence band (VB) of a semiconductor are stimulated to the conduction band (CB), generating electron–hole pairs [9,10]. Afterward, the formed electron–hole pairs are separated to trigger the desired redox reaction on the active catalyst surfaces [11]. Metal oxides, especially titanium dioxide (TiO2) and zinc oxide (ZnO), are frequently utilized in photocatalytic investigations owing to their physical stability and strong catalytic activity [12,13]. Moreover, ZnO has unique features, such as high abundance, low toxicity, ecofriendliness and cost-effectiveness [14]. ZnO is an n-type semiconductor oxide comparable to TiO2. As ZnO has the same band gap energy as TiO2 but exhibits greater absorption efficiency across a significant portion of the solar spectrum, ZnO has been suggested as a potential replacement photocatalyst [15,16,17]. Mirzaeifard et al. [18] reported that in the presence of 0.1 g of S-doped ZnO and 5 mg/L of RhB at a pH = 5, a degradation efficiency of 100% was achieved through 90 min. In addition, Ahmad et al. [19] prepared Au-ZnO which could degrade RhB (95%) within 60 min under the basic pH of RhB. Nevertheless, a high recombination ratio of electron hole pairs and photo corrosion have constrained the usage of ZnO as a photocatalyst [20].
Direct Z-scheme heterojunction and S-scheme heterojunction both use the same basic carrier separation mechanism. The first two generations of direct Z-scheme heterojunctions, conventional and all-solid-state, both have significant drawbacks. As a result, S-scheme heterojunction with a clear identification and a practical charge-transfer channel can successfully prevent needless misunderstandings (Figure 1). In addition, due to their distinctive structure and photoelectronic properties, such as greater specific surface area and more active sites, 2D-layered heterojunction photocatalysts have drawn significant interest and exhibit considerable potential in photocatalysis. Theoretically, by combining the S-scheme heterojunction and gradual 2D hybrid interfaces, one can maximize each one’s benefits and improve charge separation and utilization in photocatalysis. The development of sophisticated S-scheme heterojunctions with gradual 2D coupling interfaces for significantly accelerated photocatalytic destruction of refractory organic pollutants, however, has received very little attention to date. ZnO-based heterostructures have been identified as an effective way of increasing photocatalytic efficiency [21,22,23,24]. Recently, graphitic carbon nitride (g-C3N4)-based ZnO nanocomposites have attracted great attention due to their low band gap, high physicochemical/photochemical stability, and electron–hole separation ability [25,26]. In general, semiconductor-mediated photocatalysis begins with the absorption of proper photon energy equal to or greater than the energy gap of the desired photocatalyst to produce e/h+ pairs, followed by the production of reactive species that leads to various redox reactions on the heterogeneous photocatalysts’ surface [27,28].
The latest achievements in the photocatalytic synthesis of organic chemicals using g-C3N4-based composites have been described by Akhundi et al. [29]. Owing to the intriguing properties of g-C3N4, including an appropriate bandgap, biocompatibility, cost-effectiveness, and proper stability, g-C3N4 has attracted widespread attention [30,31]. As a polymeric semiconductor, g-C3N4 is used in photovoltaic solar cells, chemical sensors, and in separation, detection, bioimaging, photocatalytic, and membrane applications [32]. Accordingly, electron–hole pairs can be generated by light irradiation, which is responsible for initiating the photocatalytic reactions. However, electron–hole pairs can rapidly recombine, which can hamper the degradation efficiency. Therefore, attempts to enhance the photocatalytic activity of g-C3N4-based photocatalysts have propelled them to the frontline of investigations [33]. In this way, various adjustments, such as doping with efficient elements and preparing nanocomposites, can be performed to decrease the electron and hole recombination by capturing the electrons [34]. Various nanocomposites have been designed by combining g-C3N4 with inorganic materials, including TiO2, CdS, and Fe2O3 [35,36,37,38]. Recently, modified g-C3N4 by ZnO has demonstrated remarkable catalytic properties [39]. As an instance, Liu et al. [40] reported a g-C3N4/ZnO nanocomposite structure derived from melamine and zinc chloride with improved photoactivity. As a result, the nanocomposite demonstrated better photocatalytic results compared to bare g-C3N4 or ZnO. In addition, Sun et al. [41] deduced that a g-C3N4/ZnO composite obtained by heating melamine and zinc acetate had a remarkable impact on the photocatalytic degradation of methyl orange and p-nitrophenol under visible light in comparison to the single-phase g-C3N4.
Figure 2 provides the energy band structure of the g-C3N4/ZnO nanocomposite, which demonstrates the lower CB potential energy of g-C3N4 rather than ZnO (ECB = −0.85 eV vs. −0.40 eV). Therefore, by exposing it to visible light, the excited electrons from the CB of g-C3N4 can readily inject into the CB of ZnO. The potential energy match in the g-C3N4/ZnO is a driving force for transferring electrons, which effectively reduces the recombination of e/h+ pairs. Afterward, the electrons on the CB of ZnO can react with O2 to form superoxide radicals (O2•−), which are able to generate hydroxyl radicals (·OH) through further reactions. On the other hand, OH radicals are also produced by the oxidation of water molecules by h+ in the VB of g-C3N4 [42].
In this study, g-C3N4/ZnO nanocomposites with different ZnO molar ratios were synthesized using a simple method that is different from the literature. However, the present work is focused on the photocatalytic activity of the g-C3N4/ZnO nanocomposites with different ZnO molar ratios under UV light, which can help understand their performance under real ambient test conditions. To our knowledge, there are no reports on the application of g-C3N4/ZnO nanocomposites for the photocatalytic degradation of crystal violet under UV light. It is observed that the photodegradation of crystal violet is faster and more efficient when the g-C3N4/ZnO nanocomposite is used as a photocatalyst. This is because the recombination rate of photogenerated electron–hole pairs decreases, and the g-C3N4/ZnO nanocomposite increases visible light absorption compared to individual ZnO and g-C3N4.

2. Results and Discussion

2.1. Characterization of g-C3N4/ZnO Heterogeneous Photocatalyst

The powder X-ray diffraction patterns of the as-prepared bare g-C3N4 and different molar ratios of ZnO-decorated g-C3N4 nanocomposites are shown in Figure 3A. The measured powder XRD patterns of the g-C3N4/ZnO nanocomposites replicate Joint Committee on Powder Diffraction Standards (JCPDS) card no. 36-1451 and show peaks that are linked to the hexagonal wurtzite structure of ZnO. The XRD patterns of the bare g-C3N4 (JCPDS card no. 87-1526), ZnO, g-C3N4/ZnO (0.25), g-C3N4/ZnO (0.50), and g-C3N4/ZnO (1.00) demonstrate a strong peak at 27.5°, which corresponds to the (002) plane, revealing graphitic stacking of g-C3N4 in the bare g-C3N4 and g-C3N4/ZnO nanocomposites (Figure 3A). By further increasing the molar ratio of ZnO from 0.50 mmol to 1.00 mmol, it can be observed that the crystal growth of the bare g-C3N4 is limited [41]. As the amount of ZnO increases in the samples, the peaks also increase. In addition, the g-C3N4 peaks decrease, which is just opposite of ZnO. This is because the density of g-C3N4 is lower than ZnO (Figure 3A).
Figure 3B displays the Fourier-transform infrared spectra of the g-C3N4 and g-C3N4/ZnO samples. The N–H and hydroxyl groups of adsorbed H2O molecules give g-C3N4 a broad peak at 3180 cm−1. Aromatic C–N stretching modes generated by the g-C3N4 peaks can be assigned to several peaks ranging from 1285 to 1367 cm−1. The out-of-plane bending vibration of triazine rings is revealed by the intense band at 806 cm−1. The ZnO sample shows characteristic bands in the region of 593–458 cm−1, which correspond to the ZnO bond vibration, and another band of about 3160 cm−1, which is related to the hydroxyl group stretching vibrations. For the g-C3N4/ZnO hybrid structure, all the characteristic peaks related to g-C3N4 emerge, and the structure is proven to be successfully synthesized. The typical bands of the ZnO bond vibration appear at about 458, 473, and 455 cm−1 for the g-C3N4/ZnO (0.25 mmol), g-C3N4/ZnO (0.50 mmol), and g-C3N4/ZnO (1.00 mmol) samples, respectively. Through ZnO crystallization, the intensities of the peaks at 1626–1206 cm−1 drop, and the peaks are merged into a broad absorption band with increasing ZnO content. Hence, the creation of a conjugated material consisting of g-C3N4 and ZnO is vivid in the FT-IR spectra of the g-C3N4/ZnO composites, offering a channel for charge carrier transfer, resulting in photocatalytic activity enhancement [39].
SEM analysis was used to investigate the surface morphology and microstructure of the prepared samples. The SEM images of the bare g-C3N4 (Figure 4A) are in good consistency with the morphology of g-C3N4, which has been previously reported in the literature [42,43]. As illustrated in Figure 4A, the bare g-C3N4 appears as agglomerated pieces containing irregular smaller crystals. Then, a comparison of the surface morphologies of the g-C3N4/ZnO composite structure was performed (Figure 4B–D). The SEM images shown in Figure 4 reveal the formation of ZnO nanoparticles on the g-C3N4 nanosheets. The presented ZnO nanoparticles are not in a globular shape due to the nearness of g-C3N4 sheets. Moreover, Figure 4A shows that the uneven and projecting boundaries of the overlapping graphitic carbon nitride layers are what have created the g-C3N4 multilayer roughness. In Figure 4B–D, we can see the morphological changes of the coating around the g-C3N4 sample as the amount of ZnO increases.
The optical properties of the samples were analyzed using a UV–Vis absorption spectroscopy. The normalized absorption spectra of the prepared g-C3N4 and g-C3N4/ZnO nanocomposites are depicted in Figure 5A. The bare g-C3N4 sample has a characteristic absorption peak at 410 nm. We also observe that the absorbance of the g-C3N4/ZnO nanocomposite sample is slightly increased compared to the g-C3N4 sample in the UV–Vis wavelength region. The absorption nature of the g-C3N4/ZnO nanocomposites increases over the entire visible spectrum compared to the bare g-C3N4 and ZnO samples. This improved light absorption intensity in the visible range will be beneficial for the photocatalytic degradation process. The energy bandgap was characterized using the Kubelka–Munk equation for direct band-gap semiconductors, which was derived from a Tauc plot of (αhʋ)2 vs. hʋ, as shown in Figure 5B. According to Figure 5B, the intersection of the line extrapolation and the x-axis reveals the energy band gaps. The equation is α = A(hʋ-Eg)n/hʋ, where α is the absorption coefficient, hʋ is the energy of the incident photon, Eg is the band gap of the material, A is the constant, and n is a factor that depends on the nature of the electron transition [44]. The band gap energies of the ZnO, bare g-C3N4, g-C3N4/ZnO (0.25), g-C3N4/ZnO (0.50), and g-C3N4/ZnO (1.00) samples are determined to be 3.10, 2.70, 2.73, 2.66, and 2.61 eV, respectively, as demonstrated in Figure 5B.
Figure 5C displays the adsorption–desorption N2 isotherms for the g-C3N4/ZnO hybrid structures. The N2 adsorption–desorption isotherms produced at 77 K were used to determine the surface area and pore volumes of the g-C3N4/ZnO (0.25 mmol), g-C3N4/ZnO (0.50 mmol) and g-C3N4/ZnO (1.00 mmol) hybrid structures (Table 1). The surface areas obtained for the g-C3N4, g-C3N4/ZnO (0.25 mmol), g-C3N4/ZnO (0.50 mmol), and g-C3N4/ZnO (1.00 mmol) nanocomposites are, respectively, 47.01 m2/g, 67.15 m2/g, 80.77 m2/g, and 49.97 m2/g. The highest surface area and pore volume are obtained for the g-C3N4/ZnO (0.50 mmol) sample as 80.77 m2/g and 0.38 cm3/g, respectively. The low surface area and pore volume of the hybrid structures can be predicted owing to the changes in the bond structures.

2.2. The Effect of Different Amounts of ZnO

To determine the impact of diverse amounts of ZnO on CV degradation efficiency, different amounts of ZnO in the g-C3N4/ZnO photocatalyst were provided. The removal of CV was also investigated for the bare g-C3N4 catalyst to observe the photocatalytic performances of the designed nanocomposites. Initially, the amounts of ZnO selected to explore the degradation efficiency were 0.25, 0.50, and 1.00 mmol. Afterward, 0.10 g/L of the prepared photocatalyst was inserted into the CV aqueous solution, and the photocatalytic decolorization was investigated under UV light irradiation (Figure 6A). The results depict that 0.50 mmol of ZnO can degrade the CV solution up to 95.9% within 120 min. Lower removal efficiencies are obtained for 0.25 mmol and 1.00 mmol of ZnO. The reason could be explained by the fact that lower amounts of ZnO are insufficient to degrade the dye molecules. On the other hand, the active centers of the catalyst might have been inactivated because 1.00 mmol of ZnO was agglomerated due to overdose. Therefore, the 0.50 mmol ZnO-decorated g-C3N4 catalyst was chosen for further evaluation.

2.3. The Effect of g-C3N4/ZnO Catalyst Concentration

The effect of g-C3N4/ZnO catalyst concentration on dye removal efficiency was examined. Three different catalyst amounts (50, 100, and 200 mg/L) were used to examine the impact on CV removal efficiency (Figure 6B). The outcomes reveal that by increasing the amount of the catalyst from 0.05 g/L to 0.10 g/L, the degradation efficiency can be increased from 82.2% to 95.9%, respectively. By increasing the catalyst amount, the provided surface area as a platform for the reaction and generation of diverse reactive oxygen species (ROSs) is incremented. However, enhancing the catalyst amount up to 0.20 g/L can decrease the removal efficiency to 94.2%, which might be due to the light penetration prevention.

2.4. The Effect of CV Concentration

The influence of initial CV concentration (10, 20, and 30 mg/L) on the removal efficiency is presented in Figure 6C. Increasing the dye concentration from 10 to 30 mg/L causes competition as there are more dye molecules on the constant active sites of the catalyst and the generated ROS [45]. Moreover, high contaminant concentrations cause more intense color and adversely affect dye removal efficiency due to decreasing UV light penetration into the pollutant solution. It seems that most of the UV light might be absorbed by the dye molecules rather than the catalyst [46].

2.5. The Effect of H2O2 Concentration

The effect of H2O2 concentration on CV removal efficiency is demonstrated in Figure 6D. As it is well known, the addition of H2O2 promotes the formation of more OH• radicals by reacting with UV light irradiation, and these OH• radicals are able to improve the photocatalytic oxidation [47]. A complete decolorization (100%) of the CV dye was obtained at 0.65 mM of H2O2 within 120 min. However, an increase in H2O2 concentration cannot increase the dye removal efficiency due to the prevention of OH• formation by excessive H2O2 molecules [48].
The efficiency of catalyst reuse was studied by recycling the same catalyst slurry for the degradation of 10 mg/L of CV under similar batch conditions. The reuse experiments were performed for three cycles and by maintaining the catalyst slurry concentration at constant values for each run.

2.6. Kinetic Study

The pseudo-first- and second-order reaction kinetics were calculated for the present study. The kinetic constants of the reaction k1 (1/min) and k2 (L/mg.min) were calculated from the slope of the line of the kinetic models of pseudo-first order (ln C0/C = k1t) and second order (1/C − 1/ C0 = k2t). The kinetic constants k1 and k2 and the regression coefficients (R2) are compared in Table 2. The results show that the kinetic constants decrease with an increase in the initial concentration of CV from 10 to 30 mg/L. The pseudo-second-order kinetic model supplies the best fit to the experimental data for photocatalytic decolorization of CV using the g-C3N4/ZnO catalyst (Figure S1).

2.7. The Effect of Different Processes

Adsorption experiments were also performed to compare the efficiency of adsorption and the photocatalytic potential of the g-C3N4/ZnO catalyst on the removal of the CV dye solution using 0.1 g/L of catalyst loading. The results in Figure 7A indicate that photocatalysis is the dominant mechanism for the decolorization of the CV dye. It is reported in the literature that the pH of zero-point charge (pHPZC) of ZnO nanoparticles is around 9.5–10, which means that for pH values lower than 10, the surface of ZnO nanoparticles is positively charged [49,50]. Additionally, the g-C3N4 sample has positive zeta potential values [51]. In an aqueous solution, the methyl groups of CV dissociate and then CV molecules are converted to cationic dye ions. Therefore, the surface of the g-C3N4 and ZnO nanoparticles is positively charged at the original pH (6.8) of the solution, which is disadvantageous for the reach of positively charged CV dye ions due to electrostatic repulsive force. This causes less adsorption of the CV dye on the g-C3N4/ZnO catalyst. Moreover, the bare g-C3N4 was also tested for dye decolorization, and a 76.4% dye removal efficiency was obtained. Moreover, a comparative study with the literature is presented in Table 3. The results depict that our heterogeneous catalyst supplies effective decolorization with a lower catalyst amount compared to other studies.
The TOC of the solution was monitored during the degradation reactions to ensure the mineralization of the pollutants. It can be clearly seen that the degradation reaches steady-state conditions after 45 min of the process, with 80.12% efficiency (Figure 7B), which indicates that the g-C3N4/ZnO nanocomposite can effectively mineralize CV. After 45 min, degradation is almost not observed. The result show that when the catalyst is new (virgin), the rate of TOC removal is faster, and later its performance remains constant.

2.8. The Comparison of Removal Efficiency in the Presence of an Anionic Dye

Additionally, the photocatalytic degradation efficiency of a negatively charged reactive red 180 (RR180) dye was compared with the positively charged CV dye (Figure 7C). Removal efficiencies of 40.8% and 95.9% were obtained for the RR180 and CV dyes within 120 min, respectively. This could be attributed to the molecular structure of the RR180 dye (linear formula: C29H19N3Na4O17S5 with molecular weight: 933.76) being more complex than the CV dye (linear formula: C25H30N3Cl with molecular weight: 407.98).

3. Materials and Methods

3.1. Materials

Zinc nitrate hexahydrate (Zn(NO3)2·6H2O, >99%), urea (CH4N2O >99.5%), and ammonium hydroxide solution (NH4OH, 26%) were purchased from Sigma-Aldrich (USA). All reagents used in this work were of analytical grade and were used without further purifications. Crystal violet (CV), which was used as a model pollutant, was procured from Sigma-Aldrich (USA).

3.2. Synthesis of g-C3N4

Graphite-like g-C3N4 nanosheets were synthesized according to the procedure reported previously in the literature [39]. A total of 50 g of urea was first calcinated at 580 °C for 3 h at a heating rate of 5 °C/min under argon gas flow using a one-step synthesis technique. The material was cooled to room temperature, washed with HNO3 solution (0.1 mol/L) and water, and dried at 70 °C overnight.

3.3. The Synthesis of Different Molar Ratios of ZnO-Decorated g-C3N4

The ZnO-decorated g-C3N4 (named g-C3N4/ZnO) nanohybrid was obtained in the presence of urea and different proportions of Zn(NO3)2·6H2O. Firstly, 10 mL of distilled water was adjusted to a pH of 10 with the ammonium solution. A total of 166 mmol of urea was added to the distilled water/ammonia solution mixture. The urea was completely dissolved, and different proportions of [Zn(NO3)2·6H2O]x (x; 0.25 mmol, 0.50 mmol, and 1.00 mmol) were added to form the hybrid structure. Then, the mixture was vigorously stirred at 70 °C for 2 h, and the solvent was evaporated at 80 °C. The obtained precipitation was heated at 550 °C for 3 h in a furnace (heating rate of 5 °C/min) and then allowed to cool down to room temperature. A schematic illustration of the g-C3N4/ZnO synthesis procedure is shown in Figure 8.

3.4. Photocatalytic Degradation Experiments

The experimental studies were performed in a batch Pyrex photo reactor (column-shaped) with a 500 mL capacity. A total of 100 mL of dye solution was prepared. The photocatalytic reactor was blown from the bottom at a flow rate of 150 mL/min to accelerate the mixing of the catalyst. Cooled air was used to keep a constant temperature of 25 ± 1 °C. Six UVA lamps (Philips TL8 W Actinic BL, emitting wavelengths at 365 nm) were placed in a hexagonal position around the reactor, which was surrounded by an aluminum-coated film, to obtain a uniform reflection. The samples were taken from the photocatalytic reactor at predetermined times (5, 15, 30, 45, 60, and 120 min) during the 120 min reaction and centrifuged at 6000 rpm for 5 min to separate the catalyst from the aqueous solution. Maximum absorbance was measured by using a Hach DR3900 UV–Vis spectrophotometer (Düsseldorf, Germany) to calculate dye removal efficiency. The dye removal efficiency was calculated by subtracting the final concentration from the initial concentration and dividing it by the initial concentration.
The reusability behavior of the catalyst was examined through five consecutive runs. After each run, the treated solution was discharged from the reactor after the catalyst was precipitated. Then, a new solution was filled into the reactor and the catalyst was utilized again. The catalyst was washed with 50 mL of deionized water, centrifuged, and dried in an oven.

3.5. Characterizations Techniques

The crystal structures of the as-prepared samples were characterized using an X-ray diffraction (XRD, Bruker AXS D8 Advance, Billerica, Massachusetts, USA) diffractometer. The absorption spectra of the g-C3N4/ZnO nanocomposites were recorded using a UV–Vis spectrophotometer (Thermo scientific-Genesys 150, Menlo Park, CA, USA). A Quanta 650 FEG field emission scanning electron microscope manufactured by FEI (USA) was used to evaluate the surface and section morphology of the nanocomposites. The Fourier-transform infrared spectroscopy (FTIR) spectra of the nanocomposites were recorded using a PerkinElmer FTIR spectrometer (Waltham, MA, USA). The surface area and total pore volume were measured by using the Brunauer Emmett-Teller analysis (BET, MicroActive for TriStar II Plus 2.00, Norcross, GA, USA).

4. Conclusions

In summary, the ZnO-decorated g-C3N4 heterogeneous catalyst was successfully synthesized via a one-step calcination method. It was characterized using FTIR, XRD, UV–Vis, SEM, and BET analyses. The XRD and FTIR measurements demonstrated the strong coordination of ZnO with bare g-C3N4 and the partial breakdown of the g-C3N4 catalyst’s crystalline structure upon the addition of ZnO. Furthermore, the ZnO-decorated g-C3N4 composite exhibited high photocatalytic activity for the degradation of CV dye under UV light. The results displayed no decrease in the CV dye removal efficiency up to five cycles for the g-C3N4 /ZnO photocatalyst. The synergistic interactions between the ZnO and g-C3N4 provide a forceful method for combining two visible-light photocatalysts to expand their UV light absorption. This study provides some new insights into the fabrication of advanced catalysts for highly efficient photocatalytic applications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal13030485/s1, Figure S1: Kinetic study for CV decoloriaztion.

Author Contributions

Conceptualization, B.S. and Z.B.; methodology, T.S.R.; investigation, N.D., K.O. and A.K.; writing—original draft preparation, K.O. and A.K.; writing—review and editing, N.D., K.O. and AK. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hachem, F.C.; Bocquillon, O.; Zahraa, M. Decolourization of textile industry wastewater by the photocatalytic degradation process. Dye. Pigment. 2001, 49, 117–125. [Google Scholar] [CrossRef]
  2. Natarajan, S.; Bajaj, H.C.; Tayade, R.J. Recent advances based on the synergetic effect of adsorption for removal of dyes from wastewater using photocatalytic process. J. Environ. Sci. 2017, 65, 201–222. [Google Scholar] [CrossRef] [PubMed]
  3. Anisuzzaman, S.M.; Joseph, C.G.; Affandi, N.A.; Maruja, S.N.; Vijayan, V. Current Trends in the Utilization of Photolysis and Photocatalysis Treatment Processes for the Remediation of Dye Wastewater: A Short Review. Chem. Eng. 2022, 6, 58. [Google Scholar] [CrossRef]
  4. Ambaye, T.G.; Hagos, K. Photocatalytic and biological oxidation treatment of real textile wastewater. Nanotechnol. Environ. Eng. 2020, 5, 28. [Google Scholar] [CrossRef]
  5. Souza, R.P.; Freitas, T.K.F.S.; Domingues, F.S.; Pezoti, O.; Ambrosio, E.; Ferrari, L.A.M.; Garcia, J.C. Photocatalytic activity of TiO2, ZnO and Nb2O5 applied to degradation of textile wastewater. J. Photochem. Photobiol. A Chem 2016, 329, 9–17. [Google Scholar] [CrossRef]
  6. Eskikaya, O.; Ozdemir, S.; Tollu, G.; Dizge, N.; Ramaraj, R.; Manivannan, A.; Balakrishnan, D. Synthesis of two different zinc oxide nanoflowers and comparison of antioxidant and photocatalytic activity. Chemosphere 2022, 306, 153389. [Google Scholar] [CrossRef] [PubMed]
  7. Lv, Y.; Lin, J.; Peng, S.; Zhang, L.; Yu, L. Effective ways to enhance the photocatalytic activity of ZnO nanopowders: High crystalline degree, more oxygen vacancies, and preferential growth. New J. Chem. 2019, 43, 19223–19231. [Google Scholar] [CrossRef]
  8. Vattikuti, S.V.P.; Reddy, P.A.K.; Shim, J.; Chan, B. Visible-Light-Driven Photocatalytic Activity of SnO2−ZnO Quantum Dots Anchored on g-C3N4 Nanosheets for Photocatalytic Pollutant Degradation and H2 Production. ACS Omega 2018, 3, 7587–7602. [Google Scholar] [CrossRef] [Green Version]
  9. 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. [Google Scholar] [CrossRef]
  10. Mohammed, I. The photocatalytic performance of the g-C3N4/ZnOcomposite photocatalyst toward degradation of organic pollutants and its inactivity toward hydrogen evolution: The influence of light irradiation and charge transfer. Chem. Phys. Lett. 2019, 739, 136992. [Google Scholar] [CrossRef]
  11. Li, J.; Zhou, M.; Ye, Z.; Ma, C.; Wang, H.; Huo, P.; Yan, Y. Enhanced photocatalytic activity of g-C3N4-ZnO/HNTs composited heterostructure photocatalysts for degradation of tetracycline under visible light irradiation. RSC Adv. 2015, 5, 91177–91189. [Google Scholar] [CrossRef]
  12. Manikanika; Chopra, L. Photocatalytic activity of zinc oxide for dye and drug degradation: A review. Mater. Today Proc. 2022, 52, 1653–1656. [Google Scholar] [CrossRef]
  13. Nam, Y.; Lim, J.H.; Ko, K.C.; Lee, J.Y. Photocatalytic activities of TiO2 nanoparticles: A theoretical aspect. J. Mater. Chem. A 2019, 7, 13833–13859. [Google Scholar] [CrossRef]
  14. Jadoun, S.; Yáñez, J.; Mansilla, H.D.; Riaz, U.; Chauhan, N.P.S. Conducting polymers/zinc oxide-based photocatalysts for environmental remediation: A review. Environ. Chem. Lett. 2022, 20, 2063–2083. [Google Scholar] [CrossRef]
  15. Nagaraju, G.; Shivaraju, G.C.; Banuprakash, G.; Rangappa, D. Photocatalytic Activity of ZnO Nanoparticles: Synthesis via Solution Combustion Method. Mater. Today Proc. 2017, 4, 11700–11705. [Google Scholar] [CrossRef]
  16. Forouzandeh, M.M.; Ganjali, F.; Zamiri, E.; Zarei, S.S.; Jalali, F.; Padervand, M.; Taheri, L.R.; Maleki, A. Efficient Photodegradation of Eriochrome Black-T by a Trimetallic Magnetic Self-Synthesized Nanophotocatalyst Based on Zn/Au/Fe-Embedded Poly(vinyl alcohol). Langmuir 2022, 38, 13728–13743. [Google Scholar] [CrossRef] [PubMed]
  17. Mclaren, A.; Valdes-Solis, T.; Li, G.; Tsang, S.C. Shape and Size Effects of ZnO Nanocrystals on Photocatalytic Activity. J. Am. Chem. Soc. 2009, 131, 12540–12541. [Google Scholar] [CrossRef]
  18. Mirzaeifard, Z.; Shariatinia, Z.; Jourshabani, M.; Darvishi, S.M.R. ZnO Photocatalyst Revisited: Effective Photocatalytic Degradation of Emerging Contaminants Using S-Doped ZnO Nanoparticles under Visible Light Radiation. Ind. Eng. Chem. Res. 2020, 59, 15894–15911. [Google Scholar] [CrossRef]
  19. Ahmad, M.R.; Khan, W.; Qureshi, M.M.; Gul, M.T.; Haq, A.; Ullah, S.R.; Rab, A.; Menaa, F. Phytogenic fabrication of ZnO and gold decorated ZnO nanoparticles for photocatalytic degradation of Rhodamine B. J. Environ. Chem. Eng. 2021, 9, 104725. [Google Scholar] [CrossRef]
  20. Chen, D.; Wang, K.; Xiang, D.; Zong, R.; Yao, W.; Zhu, Y. Significantly enhancement of photocatalytic performances via core–shell structure of ZnO@mpg-C3N4. Appl. Catal. B Environ. 2014, 147, 554–561. [Google Scholar] [CrossRef]
  21. Safaei, J.; Mohamed, N.A.; Mohamad, N.; Mohamad, F.; Soh, M.F.; Ludin, N.A.; Ibrahim, M.A.; Roslam, W.I.; Wan, N.; Mat, T.; et al. Graphitic carbon nitride (g-C3N4) electrodes for energy conversion and storage: A review on photoelectrochemical water splitting, solar cells and supercapacitors. J. Mater. Chem. A 2018, 6, 22346–22380. [Google Scholar] [CrossRef]
  22. Padervand, M.; Heidarpour, H.; Goshadehzehn, M.; Hajiahmadi, S. Photocatalytic degradation of 3-methyl-4-nitrophenol over Ag/AgCl-decorated/[MOYI]-coated/ZnO nanostructures: Material characterization, photocatalytic performance, and in-vivo toxicity assessment of the photoproducts. Environ. Technol. Innov. 2021, 21, 101212. [Google Scholar] [CrossRef]
  23. Deng, X.; Wang, D.; Li, H.; Jiang, W.; Zhou, T.; Wen, Y.; Yu, B.; Che, G.; Wang, L. Boosting interfacial charge separation and photocatalytic activity of 2D/2D g-C3N4/ZnIn2S4 S-scheme heterojunction under visible light irradiation. J. Alloy. Compd. 2020, 894, 162209. [Google Scholar] [CrossRef]
  24. Huu, H.T.; Thi, M.D.N.; Nguyen, V.P.; Thi, L.N.; Phan, T.T.T.; Hoang, Q.D.; Vo, V. One-pot synthesis of S-scheme MoS2/g-C3N4 heterojunction as effective visible light photocatalyst. Sci. Rep. 2021, 11, 14787. [Google Scholar] [CrossRef]
  25. Mohamed, R.M.; Mkhalid, I.A.; Alhaddad, M.; Basaleh, A.; Alzahrani, K.A.; Ismail, A.A. Enhanced CO2 photocatalytic conversion into CH3OH over visible-light-driven Pt nanoparticle-decorated mesoporous ZnO–ZnS S-scheme heterostructures. Ceram. Int. 2021, 47, 26779–26788. [Google Scholar] [CrossRef]
  26. Kang, S.; Li, Z.; Xu, Z.; Zhang, Z.; Sun, J.; Bian, J.; Bai, L.; Qu, Y.; Jing, L. Synthesis of mixed-valence Cu phthalocyanine/graphene/ g-C3N4 ultrathin heterojunctions as efficient photocatalysts for CO2 reduction. Catal. Sci. Technol. 2022, 12, 4817–4825. [Google Scholar] [CrossRef]
  27. Swamy, C.K.; Hezam, A.; Ramesh, A.M.; Ramakrishnegowda, D.H.; Purushothama, D.K.; Krishnegowda, J.; Shivanna, S. Microwave hydrothermal synthesis of copper induced ZnO/g-C3N4 heterostructure with efficient photocatalytic degradation through S-scheme mechanism. J. Photochem. Photobiol. A Chem. 2021, 418, 113394. [Google Scholar] [CrossRef]
  28. Chao, Y.; Liu, B.; Zhou, B.; Tian, S.; Zhang, L. Structure and properties of graphene oxide during the synthesis process at fixed temperatures. Ceram. Int. 2021, 47, 1748–17493. [Google Scholar] [CrossRef]
  29. Akhundi, A.; Habibi-Yangjeh, A. High performance magnetically recoverable g-C3N4/Fe3O4/Ag/Ag2SO3 plasmonic photocatalyst for enhanced photocatalytic degradation of water pollutants. Adv. Powder Technol. 2017, 28, 565–574. [Google Scholar] [CrossRef]
  30. Fu, J.; Yu, J.; Jiang, C.; Cheng, B. g-C3N4-Based Heterostructured Photocatalysts. Energy Mater. 2018, 8, 1701503. [Google Scholar] [CrossRef]
  31. Padervand, M.; Rhimi, B.; Wang, C. One-pot synthesis of novel ternary Fe3N/Fe2O3/C3N4 photocatalyst for efficient removal of rhodamine B and CO2 reduction. J. Alloy. Compd. 2021, 852, 156955. [Google Scholar] [CrossRef]
  32. Pomilla, F.R.; Cortes, M.A.L.R.M.; Hamilton, J.W.J.; Molinari, R.; Barbieri, G.; Marci, G.; Palmisano, L.; Sharma, P.K.; Brown, A.; Byrne, J.A. An Investigation into the Stability of Graphitic CN as a Photocatalyst for CO2 Reduction. J. Phys. Chem. C 2018, 122, 28727–28738. [Google Scholar] [CrossRef]
  33. Dong, G.; Zhang, L. Porous structure dependent photoreactivity of graphitic carbon nitride under visible light. J. Mater. Chem. 2012, 22, 1160–1166. [Google Scholar] [CrossRef]
  34. Akhundi, A.; Badiei, A.; Ziarani, G.M.; Habibi-Yangjeh, A.; Muñoz-Batista, M.J.; Luque, R. Graphitic carbon nitride-based photocatalysts: Toward efficient organic transformation for value-added chemicals production. Mol. Catal. 2020, 488, 110902. [Google Scholar] [CrossRef]
  35. Wu, Y.; Tao, L.; Zhao, J.; Yue, X.; Deng, W.; Li, Y.; Wang, C. TiO2/g-C3N4 nanosheets hybrid photocatalyst with enhanced photocatalytic activity under visible light irradiation. Res. Chem. Intermed. 2016, 42, 3609–3624. [Google Scholar] [CrossRef]
  36. Liu, R.; Yang, W.; He, G.; Zheng, W.; Li, M.; Tao, W.; Tian, M. Ag-Modified g-C3N4 Prepared by a One-Step Calcination Method for Enhanced Catalytic Efficiency and Stability. ACS Omega 2020, 5, 19615–19624. [Google Scholar] [CrossRef]
  37. Xu, Y.; Zhang, W.D. CdS/ g-C3N4 Hybrids with Improved Photostability and Visible Light Photocatalytic Activity. Eur. J. Inorg. Chem. 2015, 10, 1744–1751. [Google Scholar] [CrossRef]
  38. Li, Y.; Zhu, S.; Liang, Y.; Li, Z.; Wu, S.; Chang, C.; Luo, S.; Cui, Z. Synthesis of α-Fe2O3/g-C3N4 Photocatalyst for high-efficiency water splitting under full light. Mater. Des. 2020, 196, 109191. [Google Scholar] [CrossRef]
  39. Paul, D.R.; Gautam, S.; Panchal, P.; Nehra, S.P.; Choudhary, P.; Sharma, A. ZnO-Modified g-C3N4: A Potential Photocatalyst for Environmental Application. ACS Omega 2020, 5, 3828–3838. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Liu, W.; Wang, M.; Xu, C.; Chen, S. Facile synthesis of g-C3N4/ZnO composite with enhanced visible light photooxidation and photoreduction properties. Chem. Eng. J. 2012, 209, 386–393. [Google Scholar] [CrossRef]
  41. 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] [PubMed]
  42. Fageria, P.; Nazir, R.; Gangopadhyay, S.; Barshilia, H.C.; Pande, S. Graphitic-carbon nitride support for the synthesis of shape-dependent ZnO and their application in visible light photocatalysts. RSC Adv. 2015, 5, 80397–80409. [Google Scholar] [CrossRef]
  43. Tauc, J.; Grigorovici, R.; Vancu, A. Optical Properties and Electronic Structure of Amorphous Germanium. Phys. Stat. Sol. 1966, 15, 627–637. [Google Scholar] [CrossRef]
  44. 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. 2018, 13, 2826–2843. [Google Scholar] [CrossRef]
  45. Ngullie, R.C.; Alaswad, S.O.; Bhuvaneswari, K.; Shanmugam, P.; Pazhanivel, T.; Arunachalam, P. Synthesis and Characterization of Efficient g-C3N4/ZnONanocomposites Photocatalyst for Photocatalytic Degradation of Methylene Blue. Coatings 2020, 10, 500. [Google Scholar] [CrossRef]
  46. Wahab, H.S.; Hussain, A.A. Photocatalytic oxidation of phenol red onto nanocrystalline TiO2 particles. J. Nanostruct. Chem. 2016, 6, 261–274. [Google Scholar] [CrossRef] [Green Version]
  47. Bilici, Z.; Bouchareb, R.; Sacak, T.; Yatmaz, H.C.; Dizge, N. Recycling of TiO2-containing waste and utilization by photocatalytic degradation of a reactive dye solution. Water Sci. Technol. 2021, 83, 1242–1249. [Google Scholar] [CrossRef]
  48. Vahid, V.; Somayeh, F.; Ramazan, K.; Alireza, K. Enhancing the permeability and antifouling properties of cellulose acetate ultrafiltration membrane by incorporation of ZnO@graphitic carbon nitride nanocomposite. Carbohydr. Polym. 2021, 256, 117413. [Google Scholar] [CrossRef]
  49. Daneshvar, N.; Rasoulifard, M.H.; Khataee, A.R.; Hosseinzadeh, F. Removal of C.I. Acid Orange 7 from aqueous solution by UV irradiation in the presence of ZnO nanopowder. J. Hazard. Mater. 2007, 143, 95–101. [Google Scholar] [CrossRef]
  50. Miao, J.; Jia, Z.; Lu, H.B.; Habibi, D.; Zhang, J.C. Heterogeneous photocatalytic degradation of mordant black 11 with ZnO nanoparticles under UV–Vis light. J. Taiwan Inst. Chem. Eng. 2014, 45, 1636–1641. [Google Scholar] [CrossRef]
  51. Kumar, R.; Barakat, M.A.; Alseroury, F.A. Oxidized g-C3N4/polyaniline nanofiber composite for the selective removal of hexavalent chromium. Sci. Rep. 2017, 7, 12850. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Ye, L.; Liu, J.; Jiang, Z.; Peng, T.; Zan, L. Facets coupling of BiOBr-g-C3N4 composite photocatalyst for enhanced visible-light-driven photocatalytic activity. Appl. Catal. B Environ. 2013, 142–143, 1–7. [Google Scholar] [CrossRef]
  53. Liu, Y.; Ma, Z. TiOF2/ g-C3N4 composite for visible-light driven photocatalysis. Colloids Surf. A Physicochem. Eng. Asp. 2021, 618, 126471. [Google Scholar] [CrossRef]
  54. Tian, J.; Chen, Z.; Jing, J.; Feng, C.; Sun, M.; Li, W. Enhanced photocatalytic performance of the MoS2/g-C3N4 heterojunction composite prepared by vacuum freeze drying method. J. Photochem. Photobiol. A Chem. 2020, 390, 112260. [Google Scholar] [CrossRef]
  55. Qian, L.; Jie, J.; Changhai, L.; Song, X.; Zhongyu, L. Constructing a novel p-n heterojunction photocatalyst LaFeO3/g-C3N4 with enhanced visible-light-driven photocatalytic activity. J. Alloy. Compd. 2017, 709, 542–548. [Google Scholar] [CrossRef]
  56. Yongchao, B.; Kezheng, C. A novel Z-scheme visible light driven Cu2O/Cu/g-C3N4 photocatalyst using metallic copper as a charge transfer mediator. Mol. Catal. 2017, 432, 187–195. [Google Scholar] [CrossRef]
Figure 1. A representative diagram for charge-transfer processes in an S-scheme heterojunction under light irradiation.
Figure 1. A representative diagram for charge-transfer processes in an S-scheme heterojunction under light irradiation.
Catalysts 13 00485 g001
Figure 2. A representative diagram for the energy band structure of g-C3N4/ZnO heterostructure.
Figure 2. A representative diagram for the energy band structure of g-C3N4/ZnO heterostructure.
Catalysts 13 00485 g002
Figure 3. (A) XRD patterns of the bare g-C3N4, ZnO, and synthesized different molar ratios g-C3N4/ZnO samples, and (B) FTIR spectra of the g-C3N4/ZnO samples.
Figure 3. (A) XRD patterns of the bare g-C3N4, ZnO, and synthesized different molar ratios g-C3N4/ZnO samples, and (B) FTIR spectra of the g-C3N4/ZnO samples.
Catalysts 13 00485 g003
Figure 4. SEM images of (A) bare g-C3N4, (B) g-C3N4/ZnO (0.25), (C) g-C3N4/ZnO (0.50), and (D) g-C3N4/ZnO (1.00) samples.
Figure 4. SEM images of (A) bare g-C3N4, (B) g-C3N4/ZnO (0.25), (C) g-C3N4/ZnO (0.50), and (D) g-C3N4/ZnO (1.00) samples.
Catalysts 13 00485 g004
Figure 5. (A) UV–Vis absorption spectra, (B) Tauc plot of the g-C3N4, ZnO, and different molar ratios g-C3N4/ZnO nanocomposites, and (C) nitrogen adsorption−desorption isotherms of the g-C3N4/ZnO nanocomposites.
Figure 5. (A) UV–Vis absorption spectra, (B) Tauc plot of the g-C3N4, ZnO, and different molar ratios g-C3N4/ZnO nanocomposites, and (C) nitrogen adsorption−desorption isotherms of the g-C3N4/ZnO nanocomposites.
Catalysts 13 00485 g005
Figure 6. (A) The effect of ZnO on dye removal efficiency (experimental conditions: [catalyst] = 0.10 g/L). (B) The effect of catalyst amount on dye removal efficiency (experimental conditions: ZnO ratio = 0.5 mmol and [dye] = 10 mg/L). (C) The effect of dye concentration on dye removal efficiency (experimental conditions: ZnO ratio = 0.5 mmol and [catalyst] = 0.10 g/L). (D) The effect of H2O2 concentration on dye removal efficiency (experimental conditions: ZnO ratio = 0.5 mmol, [catalyst] = 0.10 g/L, and [dye] = 10 mg/L).
Figure 6. (A) The effect of ZnO on dye removal efficiency (experimental conditions: [catalyst] = 0.10 g/L). (B) The effect of catalyst amount on dye removal efficiency (experimental conditions: ZnO ratio = 0.5 mmol and [dye] = 10 mg/L). (C) The effect of dye concentration on dye removal efficiency (experimental conditions: ZnO ratio = 0.5 mmol and [catalyst] = 0.10 g/L). (D) The effect of H2O2 concentration on dye removal efficiency (experimental conditions: ZnO ratio = 0.5 mmol, [catalyst] = 0.10 g/L, and [dye] = 10 mg/L).
Catalysts 13 00485 g006
Figure 7. (A) The effect of different processes on dye removal efficiency, (B) TOC alterations during photocatalysis of CV, and (C) comparison of the degradation efficiency of RR180 and CV during photocatalysis. (Experimental conditions: ZnO ratio = 0.5 mmol, [catalyst] = 0.10 g/L, and [dye] = 10 mg/L).
Figure 7. (A) The effect of different processes on dye removal efficiency, (B) TOC alterations during photocatalysis of CV, and (C) comparison of the degradation efficiency of RR180 and CV during photocatalysis. (Experimental conditions: ZnO ratio = 0.5 mmol, [catalyst] = 0.10 g/L, and [dye] = 10 mg/L).
Catalysts 13 00485 g007
Figure 8. The synthesis procedure of different molar ratios of ZnO-decorated g-C3N4.
Figure 8. The synthesis procedure of different molar ratios of ZnO-decorated g-C3N4.
Catalysts 13 00485 g008
Table 1. The surface area and pore volume of the prepared samples.
Table 1. The surface area and pore volume of the prepared samples.
SampleSurface Area (m2/g)Pore Volume (cm3/g)
g-C3N4/ZnO (0.25 mmol)67.150.41
g-C3N4/ZnO (0.50 mmol)80.770.38
g-C3N4/ZnO (1.00 mmol)49.970.12
g-C3N447.010.17
Table 2. The pseudo-first- and second-order kinetic rate constants with regression coefficients for decolorization of CV at different dye concentration conditions (experimental conditions: solution pH = 6.8, and [catalyst] = 0.10 g/L).
Table 2. The pseudo-first- and second-order kinetic rate constants with regression coefficients for decolorization of CV at different dye concentration conditions (experimental conditions: solution pH = 6.8, and [catalyst] = 0.10 g/L).
CCV (mg/L)First OrderSecond Order
k1 (1/min)R2k2 (L/mg min)R2
100.02350.75940.02020.9497
200.01270.82970.00200.9780
300.0090.92900.00060.9805
Table 3. Comparison of g-C3N4-based heterogeneous catalysts for dye decolorization.
Table 3. Comparison of g-C3N4-based heterogeneous catalysts for dye decolorization.
CatalystsDye Concentration (mg/L)PollutantsDegradation EfficiencyReferences
g-C3N4/ZnO10Crystal Violet95.9% for 120 minThis study
g-C3N4/Ag10Methyl Orange98.7% for 120 min[36]
g-C3N4/CdS5Rhodamine B95% for 120 min[37]
g-C3N4/BiOBr10Rhodamine B95% for 30 min[52]
g-C3N4/TiOF25Rhodamine B78.3% for 80 min[53]
g-C3N4/MoS210Rhodamine B97.6% for 50 min[54]
g-C3N4/LaFeO315Rhodamine B70% for 160 min[55]
g-C3N4/Cu2O/Cu10Methyl Orange99% for 40 min[56]
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

Sert, B.; Bilici, Z.; Ocakoglu, K.; Dizge, N.; Rad, T.S.; Khataee, A. Preparation of S-Scheme g-C3N4/ZnO Heterojunction Composite for Highly Efficient Photocatalytic Destruction of Refractory Organic Pollutant. Catalysts 2023, 13, 485. https://doi.org/10.3390/catal13030485

AMA Style

Sert B, Bilici Z, Ocakoglu K, Dizge N, Rad TS, Khataee A. Preparation of S-Scheme g-C3N4/ZnO Heterojunction Composite for Highly Efficient Photocatalytic Destruction of Refractory Organic Pollutant. Catalysts. 2023; 13(3):485. https://doi.org/10.3390/catal13030485

Chicago/Turabian Style

Sert, Buse, Zeynep Bilici, Kasim Ocakoglu, Nadir Dizge, Tannaz Sadeghi Rad, and Alireza Khataee. 2023. "Preparation of S-Scheme g-C3N4/ZnO Heterojunction Composite for Highly Efficient Photocatalytic Destruction of Refractory Organic Pollutant" Catalysts 13, no. 3: 485. https://doi.org/10.3390/catal13030485

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