Next Article in Journal
The Impact of Sediment–Water Ratio and Hydraulic Residence Time on the Release of Inorganic Nitrogen from Sediments in the Pearl River Delta
Next Article in Special Issue
Taguchi L9 (34) Orthogonal Array Design for Photocatalytic Degradation of Methylene Blue Dye by Green ZnO Particles Biosynthesized by Chrysanthemum spp. Flower Extract
Previous Article in Journal
Oxidation of an Azo-Dye via the Photo-Fenton Process under Heterogeneous and Homogeneous Conditions
Previous Article in Special Issue
Impregnation of Modified Magnetic Nanoparticles on Low-Cost Agro-Waste-Derived Biochar for Enhanced Removal of Pharmaceutically Active Compounds: Performance Evaluation and Optimization Using Response Surface Methodology
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

TiO2/Activated Carbon/2D Selenides Composite Photocatalysts for Industrial Wastewater Treatment

1
Institute of Chemistry, Khwaja Fareed University of Engineering and Information Technology, Rahim Yar Khan 64200, Pakistan
2
Institute of Chemical and Environmental Engineering, Khwaja Fareed University of Engineering and Information Technology, Rahim Yar Khan 64200, Pakistan
3
Department of Chemical Engineering, University of Gujrat, Gujrat 50700, Pakistan
4
Institute of Health Sciences, Khwaja Fareed University of Engineering and Information Technology, Rahim Yar Khan 64200, Pakistan
5
Optics Valley Laboratory, Wuhan 430074, China
6
State Key Laboratory of Magnetic Resonance and Atomic Molecular Physics, Key Laboratory of Magnetic Resonance in Biological Systems, National Center for Magnetic Resonance in Wuhan, Wuhan Institute of Physics and Mathematics, Innovation Academy for Precision Measurement Science and Technology, Chinese Academy of Sciences, Wuhan 430074, China
7
Wuhan National Laboratory for Optoelectronics, Huazhong University of Science and Technology, Wuhan 430071, China
8
Nawaz Sharif Medical College, University of Gujrat, Gujrat 50700, Pakistan
9
Institute of Physics, Khwaja Fareed University of Engineering and Information Technology, Rahim Yar Khan 64200, Pakistan
*
Authors to whom correspondence should be addressed.
Water 2023, 15(9), 1788; https://doi.org/10.3390/w15091788
Submission received: 27 March 2023 / Revised: 24 April 2023 / Accepted: 3 May 2023 / Published: 6 May 2023
(This article belongs to the Special Issue Recent Advances in Nanomaterials for Water Treatment)

Abstract

:
Even in the 21st century, water contamination has been a big problem and industrial processes are to be blamed for polluted water supplies. The use of sunlight in the process of photocatalysis is an efficient way to purify wastewater. Composites of TiO2/activated carbon/two-dimensional selenides performed better than either of the individual material or binary composites for this application. A straightforward hydrothermal technique was employed in the synthesis of photocatalysts. The synthesized photocatalytic composites were verified with the help of UV-Visible spectroscopy, FTIR, XRD, and SEM. The heterostructures absorbed nearly all of the sun’s UV and visible light. These photons are then converted into usable reducing electrons and oxidizing species such as O2 and OH to decompose organic pollutants from industrial wastewater. Since there were additional pathways available for charge transfer along with several active edge sites, the composite photocatalysts are proven more active than individual TiO2 and 2D MoSe2 components. With the help of a cascade-driven mechanism of electrons, these channels can transmit more charges than single-component heterojunctions. The results provided a realistic method for developing photocatalyst composites powered by solar light for use in industrial wastewater treatment. Results of degradation of methylene blue suggest that the synthesized composites possess better photocatalytic activity. This enhanced photocatalytic activity is not limited to organic dyes. Other hazardous organic pollutants present in industrial wastewater can be decomposed by using this approach.

1. Introduction

Water is essential for every individual to sustain life on earth [1]. Rapidly growing populations, urbanization, extensive agricultural practices and industrialization are the main reasons for water pollution, a threatening issue of the modern world [2,3,4]. In the previous decade, a rapid increase in industries and industrial processes have created a huge impact in water pollution [5]. Industrial waste materials are the most frightening environmental issues. They impose serious health effects on humans via surface wastewater bodies [6,7]. About 2.2 million people die annually due to waterborne diseases in developing countries [8]. The consumption of contaminated water causes diarrhea and approximately 1.8 million children lose their lives every year from this disease [9,10]. There is an urgent need to minimize the problem of water pollution and to save human beings and aquatic life from these chronic diseases.
Water treatment facilities are preventive measures against waterborne diseases [11]. A huge amount of water is used in textile industries for different processes. Water is used for washing purposes in the industry. If this water is released as such without treatment, it imposes a serious threat on ecosystem [12]. This water is contaminated with animal oils, fiber lint, organic dyes and other hazardous chemicals. These chemicals are used to improve the quality of raw material. The liquid wastes along with several contaminants are discarded from these industries, causing environmental pollution [13,14]. The colored wastewater of these industries contains an unpleasant odor, chemical oxygen demand, biochemical oxygen demand, high pH and suspended solids. Suspended solids are composed of different inorganic salts, heavy metals and hazardous chemicals [15,16]. A high concentration of carcinogenic organic dyes is also present in industrial effluents [17]. Synthetic dyes can cause genetic mutations and cancer in human beings. These organic dyes enter the food chain from wastewater of the textile industry. Dyes are extensively used in the textile industry in different processes [18]. Textile industries are responsible for one-fifth of industrial water pollution worldwide [19]. Methylene blue (MB) is the most commonly used dye in textile industry. A large amount of MB is released in wastewater from textile industries [20]. Above a certain concentration it is harmful for human beings, microbes and the environment. MB is responsible for blindness, abdominal disorders, digestive and destructive disorders [21,22]. It also causes diarrhea, cyanosis, gastritis, jaundice, methemoglobinemia, shock, skin and eye irritation, tissue necrosis and vomiting [20].
Many conventional techniques such as biological treatments, adsorption, filtration, sedimentation, osmosis, reverse osmosis and chemical treatments are applied for water purification. Conventional techniques have been applied for the removal of contaminants from wastewater, thereby enhancing water quality. These techniques are not adequately effective to clean wastewater containing diverse contaminants [23]. The urgent need of the time is to develop alternative wastewater techniques. The techniques which are capable of completely eliminating these hazardous contaminants from wastewater are always preferred [24].
Advanced oxidation processes are recently revolutionized as wastewater treatment technique. These produce highly reactive free radicals (OH) by degrading inorganic and organic pollutants. OH free radicals oxidize carbonaceous species into inorganic ions and CO2 gas, a product due to their strong oxidizing nature [25]. Oxidants such as ozone and hydrogen peroxide can trigger the formation of OH. This can be completed by using energy resources such as heat, ultrasounds and ultraviolet light or by using homogeneous or heterogeneous photocatalysts such as Fenton’s reagent, TiO2 or ZnO [26]. Among these advanced oxidation processes, a photocatalytic oxidation process is an attractive technique. In this process, impurities and pollutants can be eliminated with the help of oxidation activated by free radicals at a normal temperature and pressure. This results in non-selective oxidation of contaminants in water such as carbon dioxide, anions and cations. OH, free radicals possess very high oxidation potential (OH/H2O) = 2.80 eV/SHE. It is the second most proficient oxidation potential. Maximum oxidation potential carries fluorine (E° = 3.0 eV) [27]. Free radicals have a very short lifespan in water and vanish rapidly from reaction medium [28]. The prime drawbacks of traditional advanced oxidation processes include the following: tedious instrumentation, inadequate mineralization of pollutants/contaminants, high-cost processing, ultraviolet light activity and half-life time of O3 [29,30,31].
Semiconductor-material-based photocatalysts have been proven as the best solution to cover these drawbacks. Semiconductor photocatalysts do not form any kind of secondary pollutants in water. Because of their chemically stable nature, photocatalysts can be reprocessed in aqueous medium [32]. TiO2 has been an extensively studied semiconducting material, in the previous decade. Its wide applications are due to its easy availability, low price and non-poisonous character, high stability, porous structure and greater surface area [33,34,35].
Major limitations of TiO2 include suppression of photocatalytic activity, slow photocatalytic degradation rates, higher band energy, and aggregation of TiO2 nanoparticles due to unstable nanosized particles [36,37,38,39]. Photocatalytic properties of TiO2 can be increased in a visible region by minimizing a quick recombination of the photogenerated charge carrier, by depositing metal nanoparticles on its surface [40,41,42].
A very effective strategy to expand absorption in the visible range from UV solar light and to enhance photo-induced charge separation is the construction of heterostructure [43,44]. For this purpose, transition metal semiconductors can increase the photo-induced charge separation of TiO2 and their Mn+/M(n−1)+ (M = transition metal) cycle directly activates persulphates to reduce pollutants in wastewater [45]. In addition to the limitations of semiconductors, major hindrances in the practical applications of composites include ion diffusion, slow charge transfer and deficient electromagnetically active sites [46].
In recent research, two-dimensional transition metal dichalcogenides have grasped the attention of researchers due to their unique electronic and optical properties. These materials possess excellent light absorbance and fast electron migrating properties [47]. A large number of active sites on their crystal edges provide more surface area for photocatalytic reaction. Therefore, transition metal dichalcogenides are predictably suitable co-catalysts for TiO2, to reduce its carrier recombination and improve its light absorbance capacity [48,49,50]. For example, MoSe2, has been proven to be a promising semiconductor because of its narrow band gap (1.7 to 1.9 eV), which absorbs a broader range of solar light [51]. It contains better antiphotocorrosive stability and an exceptional two-dimensional (2D) layer structure for surface reaction [52,53,54]. In 2D MoSe2, Mo atoms are entrapped between the layers of Se atoms via van der Waals cohesive forces [55]. The layered structure of 2D MoSe2 can be referenced from a recent study which proved that interlayer spacing of MoS2 can be expanded in different phases [56]. The lower cost of MoSe2 and easy availability proves it is a better cocatalyst than the graphene and noble metals [48,57].
Carbon materials are used to modify TiO2 because of their excellent electrical conductivity and strong visible light absorbing property. These properties enhance the efficiency of photo-induced carrier separation [58,59]. Activated carbon (AC) is the most promising adsorbent. The large surface area of AC, high surface activity and highly porous structure enhance its absorption character. Different varieties of activated carbon are used as adsorbents, catalysts and catalyst supporters [60]. Literature survey regarding publications on relevant topic has been given in Table 1.
TiO2/Activated Carbon/2D-Selenide photocatalysts were synthesized using the simple hydrothermal method. These composites absorb broad spectrum solar light in the ultraviolet, visible and NIR region due to the presence of 2D MoSe2. The 2D MoSe2 reduces the band gap of TiO2 and AC increases the light absorption tendency of TiO2. Maximum light absorption power and the reduced band gap proves TiO2/AC/2DMoSe2 a better photocatalyst for degradation of organic hazardous pollutants in wastewater.

2. Materials and Methods

2.1. Materials

Charcoal (C), polyethylene glycol (C2nH4n+2On+1), ethylene diamine tetra acetic acid ([CH2N(CH2CO2H)2]2), sodium selenite (Na2O3Se), hydrazine (N2H4), ammonium paramolybdate tetrahydrate ((NH4)6Mo7O24.4H2O), and other reagents were imported from Sigma-Aldrich. All the chemicals were of analytical grade so there was no need to further purify them.

2.2. Activation of Charcoal

Sequence wise procedure for the activation of carbon has been shown in Figure 1. A total of 3 g of charcoal was dipped completely into 1 M hydrochloric acid and left for 24 h. After that the mixture was stirred for 2 h at 25 °C, filtered, and washed with distilled water, 0.1 M sodium hydroxide (NaOH) and again with distilled water. The activated carbon was then dried and saved for further use.

2.3. Synthesis of TiO2

Schematic representation of the synthesis of TiO2 has been shown in Figure 2. TiO2 was synthesized using the hydrothermal method [53]. Then, 20 mL of titanium tetra butoxide (C16H36O4Ti) was added in excess of 0.1 M NaOH by continuous stirring. The pH of this solution was decreased from 8–9 to 2–3 by adding 0.1 M HCl. The mixture was added into a Teflon-lined autoclave, screwed tightly and kept in an oven at 170 °C for 24 h. Precipitates of TiO2 were obtained, filtered, washed with methanol (CH3OH) and distilled water, and dried. Annealing of TiO2 crystals was conducted at 300 °C for 3 h and then saved for further use.

2.4. Synthesis of TiO2/xAC/2DMoSe2

Schematic representation of synthesis of TiO2/xAC/2D MoSe2 has been shown in Figure 3 TiO2/xAC/2DMoSe2 heterostructures were prepared as follows. In the first step, 2D MoSe2 was prepared by adding 15 mL of polyethylene glycol (PEG) in 15 mL of distilled water while stirring continued. The solution was divided into two equal halves into beaker A and B. Then, 1–2 drops of ethylene diamine along with 0.359 g of sodium selenite and 5 mL of hydrazine along with 0.176 g ammonium paramolybdate tetrahydrate were added in beaker A and B, respectively, by constant stirring. Both the mixtures were stirred for 30 min and mixed into beaker C by continuous stirring for 10 min. In the second step, TiO2/xAC/MoSe2 was prepared by adding 1% (by mass) of TiO2 and activated carbon into the above mixture. The mixture was magnetically stirred for 20 min and then transferred into a 100 mL autoclave. After, the airtight sealed autoclave was placed in an oven for 24 h at 200 °C. The precipitates obtained were washed with methanol, distilled water, and dried at 60 °C overnight. The resulting composites were given the name of TiO2/xAC/MoSe2.
The samples of 0.25%, 0.5%, 0.75% and 100% by mass ratios of activated carbon were denoted as TiO2/0.25AC/2DMoSe2, TiO2/0.5AC/2DMoSe2, TiO2/0.75AC/2DMoSe2 and TiO2/1AC/2DMoSe2, respectively.

2.5. Characterization

A PerkinElmer 100 FT-IR spectrometer (Waltham, MA, USA) was used to record FTIR spectra of synthesized composites. The range of the FTIR spectrometer was set at 400–4500 cm−1 in transmittance mode in the Hi-Tech Laboratory of KFUEIT Rahim Yar Khan, Pakistan. The XRD studies were performed at the University of Peshawar, Pakistan by using the JEOL X-ray diffractometer (Model: JDX-3532, JEOL, Tokyo Japan). X-rays of CuKα with Wavelength = 1.5418 Å and 2Theta-Range: 0 to 160°. The SEM characterization was performed in the Central Resource Laboratory of the University of Peshawar Pakistan by using SEM (Model: JSM 5910, JEOL, Tokyo Japan). UV-Visible spectroscopy was performed in the laboratory of the Institute of Chemical & Environmental Engineering, KFUEIT Rahim Yar Khan Pakistan. Quartz cuvettes were used as a sample holder. Distilled water was used as a reference.

3. Results and Discussion

3.1. FTIR Study of Composites

In FTIR spectra, peaks at 830 cm−1, 1090 cm−1, 2941 cm−1 and 3650 cm−1 were representing O-Mo-O, Se-O bond, C-H stretching and OH stretching, respectively (Figure 4) [61,62]. Bands in the range of 550–750 cm−1 ascribe a stretching vibration of the O-Ti bond [63]. The absorption peak at 3410.51 cm−1 shows -OH stretching vibration. This stretching peak shows absorption of water molecules, alcohols and phenolic compounds on the surface of photocatalyst [64,65]. The peak at 1600.89 cm−1 is of C=C aromatic ring stretching vibration, shifted from 1634 cm−1 to 1600.89 cm−1. Peaks that appeared at 832 cm−1 and 600 cm−1 show the formation of oxygen metal bonding. Therefore, these peaks represent the titanium dioxide loaded with activated carbon [66]. The band between 550 and 650 cm−1 shows a stretching vibration of Ti-O. At 1060.43 cm−1 a peak indicates activated carbon. The peak intensity at 1060 cm−1 increases with an increase in activated carbon. An absorption peak at 949 cm−1 and 550 cm−1 represents O-Ti-O-C bonds [67]. It was ascribed to the titania mixture described by Zhang et al. A peak present at 1060.43 cm−1 shows the C-O-Ti bond. A minute conjugation of Ti-O and a bulk of activated carbon is due to the electron affinity difference [68]. At 1060 cm−1 the peak shows the Se-O bond [14].

3.2. Structural Properties

The XRD spectra of synthesized composites are shown (Figure 5). These spectra were recorded at the National Center of Excellence in Physical Chemistry Labs, University of Peshawar Pakistan, by using a diffractometer (X-ray diffractometer, model. JDX-3532, JEOL, Tokyo Japan) using X-rays; CuKα CuKα (λ = 1.5418 Å), 2θ = 0 to 160°.
The crystal structure of synthesized photocatalysts were characterized by XRD. It is clear from the spectrum that there is no prominent difference in the diffraction peaks of pure TiO2 and synthesized photocatalyst nano composites (Figure 5). It clarifies that the incorporation of activated carbon and 2DMoSe2 has a negligible effect on the crystal structure of TiO2. Due to a small amount of MoSe2 and its high dispersion power on the surface of activated carbon and TiO2, no prominent peak of MoSe2 appeared in the XRD spectrum of the composites [69,70]. Due to the amorphous nature of AC and less diffracting 2D material, no signals of activated carbon were found in the spectrum of photocatalyst [71]. Ammonium ions from ammonium molybdate easily enter into the layers of MoSe2. These ammonium ions between the layers of MoSe2 result in shifting of the peak to 9.25° [69,72]. The anatase structure of TiO2 is indicated by (101) plane. The (110) plane represents the rutile TiO2. As the percentage of AC increases the crystallinity of composites decreases successively. As a result, the intensity of peaks decreases gradually.

3.3. Morphological Properties

SEM photographs of synthesized photocatalysts are shown in Figure 6. Unfortunately, the SEM that we had in access, did not have high enough resolution to accurately characterize such small materials, but the pictures we captured were not a total failure, as demonstrated in Figure 6. Figure 6 show evidence of pristine MoSe2 nanostructures on the surfaces of TiO2 particles, and agglomerates are round by seeming, suggesting that they are made up of skinny, non-linear units. Pure activated carbon does not exhibit any crystalline shape as shown in Figure 6a. The spherical shape of pure TiO2 nanocrystals appears in Figure 6b. This spherical shape in transformed into semi-spherical shape due to addition of activated carbon. Activated carbon surrounds a large surface area over TiO2 nanocrystals [73,74]. Decrease in particle size of TiO2 is an indication of even distribution of activated carbon in TiO2 nanocrystals [75]. In the 2D MoSe2 lattice structure, two layers of Se atoms sandwich the Mo atom. Weak van der Waals interaction between these layers develops a few-layered structures of 2D MoSe2, known as the monolayer structure [76]. These monolayers of 2D MoSe2 affect the crystalline structure of TiO2 to some extent. From Figure 6c–f, an increasing percentage of activated carbon reduces the spherical shape of TiO2 and the monolayer structure of 2D MoSe2.

3.4. Optical Properties

The results of the degradation of methylene blue dye are shown in Figure 7, Figure 8, Figure 9, Figure 10 and Figure 11. 10 mg/L Methylene blue (MB) solution was used by catalyst loading of 0.1 g/dm3 to evaluate efficiency of composites [77] 0.01 M NaBH4 solution. The organic dye was selected as it shows color change as a result of degradation. This color change is the visual evidence of completion of photocatalytic chemical reaction. Being a member of the class cationic thiazine dye MB is widely known by its biochemical and chemical applications. MB shows a deep blue color in the aqueous solution in its oxidized form and is colorless in its reduced form [78]. MB shows absorption at 290 nm because of π to π* transitions while at 664 nm because of n to π* electronic shifting. These wavelengths help to monitor dye reduction [79]. NaBH4 is well known due to its strong reducing property. It was used in dye solution, in such an extent that its low redox potential cannot complete the reaction and kinetically show forbidden behavior [80].
In the first step, the photocatalytic activity of synthesized composites was examined by degrading methylene blue in aqueous solution. In the absence of photocatalysts in blank solution, a very small change appeared in the color and intensity after 150 min. This shows that NaBH4, individually has no tendency to degrade dye. However, in the presence of a small amount of photocatalysts, a significant increase in dye degradation appeared.
The degradation of methylene blue was achieved by adding 1 mL of NaBH4 (0.001 M) in 50 mL of 10 µg/L dye solution by constant stirring. A sample from this solution was taken as the standard. After 5 min, 15 mg of photocatalyst was added into solution. The prepared mixture was kept under dark for 150 min. Samples were taken out from the reaction mixture at 30 min intervals to study the effect of time on degradation of MB. After that, Philips light emitting Diode (LED) of 24 W was used to irradiate the solution keeping approximately 10 cm away from it to examine the photocatalytic degradation efficiency of synthesized composites. The LED lamp emits radiations of λ > 400 nm. The stirring of the solution continued until the completion of the reaction. A small volume of the sample was taken out from the reaction mixture at 30 min intervals for the kinetic studies of the MB degradation using UV-visible spectroscopy.
The process is repeated for all the synthesized composites in order to study their photocatalytic behavior. Photocatalytic degradation of MB was repeated three times under the same conditions to perform the re-usability test of the synthesized photocatalysts. All the experiments were performed in 150 min. The photocatalytic efficiency of synthesized composites was calculated by applying the following formula:
P e r c e n t a g e   e f f i c i e n c y = C 0 C C 0 × 100
where C0 = concentration of MB solution before photocatalytic reaction; C = concentration of MB solution after photo irradiation for given time t.
Figure 7a shows the photocatalytic degradation of pure TiO2 while Figure 8a, Figure 9a, Figure 10a and Figure 11a shows the photocatalytic degradation of synthesized composites having MB with the passage of time. From these results, it is clear that the photocatalyst with the higher percentage of AC in composite shows the maximum absorption of light and maximum degradation efficiency.
The effect of light irradiation time on the MB concentration ratio (C/C0) is shown in Figure 7b, Figure 8b, Figure 9b, Figure 10b and Figure 11b. C and C0 are the initial concentrations of MB in and concentrations at time t, in the aqueous phase. The C/C0 ratio makes a convenient visual comparison of the photocatalytic effect of the composite.
On the base of the Langmuir–Hinshelwood model, photocatalytic reactions show pseudo first order kinetics. The first order rate constant k can be calculated by applying formula:
ln C/C0 = −kt
Figure 7c, Figure 8c, Figure 9c, Figure 10c and Figure 11c show that pure TiO2 and synthesized photocatalytic composites show pseudo-first order kinetics.
The percentage of degradation efficiency of pure TiO2 and synthesized composites is shown in Figure 7d, Figure 8d, Figure 9d, Figure 10d and Figure 11d. From these graphical representations it is clear that increased amounts of AC in composites increases the light absorption capability of the photocatalyst which in turn results in increased photocatalytic activity.
Pure TiO2 shows the absorption maximum at 390 nm. It shows that TiO2 works only in the ultraviolet region. Pure MoSe2 shows remarkable absorption in the ultraviolet, visible and infrared region [79]. TiO2/xAC/2D MoSe2 show absorption in the visible to infrared region. A large distribution of pore size structures in the photocatalyst, due to the presence of activated carbon and 2D MoSe2 will make it able to absorb all the ultraviolet, visible and near infrared light.
Table 2 shows the comparative % efficiency of AC, TiO2 alone, combinations of TiO2 with AC and CNTs and our synthesized TiO2/AC/2D MoSe2 to degrade MB. From the literature, it was found that the degradation efficiency of TiO2 and activated carbon is not high when we use these materials individually to degrade organic pollutants in wastewater [81,82]. The degradation efficiency of activated carbon is only 55%. The degradation property of activated carbon is due to the presence of active sites on its surface. The light absorption power of pure TiO2 decreases with the passage of time, so it showed 73% degradation at 30 minutes’ interval. Incorporation of activated carbon with TiO2 increased light absorption behavior of photocatalyst [68]. In another study, TiO2@AC showed 65% degradation of MB in 180 min. The Xenon lamp was used as a light source for the degradation of MB in that study [75]. Askari et al., (2017) decorated TiO2 with multi-walled carbon nano tubes (MWCNT) and used synthesized photocatalyst for the degradation of MB dye in water. The UV Philip lamp was used as a radiation source in that study. The photocatalyst showed 61.1% degradation of MB at the rate of 30 min [83]. Yoon, C.-J., et al., (2021) developed the TiO2@carbon nanotube and evaluated the photocatalytic activity of the TiO2@carbon nanotube photocatalyst by the degradation of MB in water. The photocatalyst showed 85% degradation of MB when the UV radiation source was used [84]. Alghamdi, Y.G., et al., (2022) synthesized biomass based activated carbon loaded with TiO2 and used it for the photocatalytic degradation of oflxacin and reactive red 120. The photocatalyst showed 82% degradation in the UV light source [85]. Justh et al., (2019) studied the photocatalytic properties of TiO2@carbon aerogel composites. Composites were prepared using the atomic layer deposition method. Photocatalysts showed 55% degradation at 240 min interval [86]. All the above studies show that TiO2 along with AC absorbs light in the ultraviolet region. Although carbon nano-tubes and activated carbon increased the photocatalytic activity of TiO2 up to 85%, these composites required ultraviolet light for photodegradation. A main portion of sunlight is composed of the visible light spectrum.
In the present study, 2D MoSe2 was incorporated in TiO2/AC to reduce the band gap of TiO2 and to increase the absorption spectrum of composites in the visible and NIR region. From the values given in Table 2, it is also clear that the increased amount of AC proved effective in increasing the efficiency of the photocatalyst as it increases the light absorption capacity of the composite. The results show that maximum efficiency of synthesized nanocomposite was observed with the increase in ratio of AC. The maximum degradation efficiency (83%) of TiO2/AC/2D MoSe2 was observed at maximum ratio of AC i.e., 1:1:1. The increase in light absorbance decreases dye intensity in water with the passage of time. As time increases, the absorbance of light decreases which indicates that dye molecules are dissociating into their fragments. The photocatalytic activity of the synthesized composites is due to the reduced band gap of TiO2 because of the presence of 2DMoSe2 and the enhanced light absorption property due to the presence of elemental activated carbon.

4. Parameters Affecting Photodegradation of Methylene Blue

4.1. Effect of Irradiation Time

The effect of time of irradiation on dye degradation was studied by successive increase in time intervals irradiation on the MB samples in the presence of photocatalysts as shown in Table 3. As the reaction time increases, a decrease in the absorption peak of MB appears. At the same time, the color changes from blue to colorless. The reason for the decrease in the absorption spectra is reduction of MB chromophore.

4.2. Effect of Initial Dye Concentration

To study the effect of initial concentration of MB on degradation rate, aqueous solution of MB was prepared with concentration 5 mg/L, 10 mg/L, 15 mg/L, 20 mg/L and 25 mg/L. The maximum percentage degradation shown by the photocatalyst, “TiO2/AC/2DMoSe2”, is given in the Table 4.
The nature of dye, its initial concentration and presence of foreign species in the dye solution greatly affect the rate of photodegradation [88]. At a lower concentration of MB, its absorption capacity is high. It is because of the availability of more active sites on the surface of the photocatalyst for low MB concentration [89]. Active sites of the photocatalyst are covered by higher adsorption of dye molecules which decreases the photodegradation rate at a higher concentration of MB. A higher concentration of dye molecule increases the screening effect of light and minimizes the production of OH active radicals [90]. In the present study, the degradation rate of MB was found to be high at a lower concentration of MB.

4.3. Effect of pH

To determine the impact of pH, MB solutions with different pH values were used. A little change in degradation was observed when pH is changed from 7.0 to 4.0 (Table 5). Degradation efficiency further decreased when pH decreased further from 4.0 to 2.0. While an increase in degradation was observed when the pH of the MB solution increased from 7 to 9.5. The maximum degradation efficiency was achieved after 120 min of irradiation.
MB absorbs on highly negative charged photocatalysts, as it is a cationic dye [91]. At a high pH (in basic medium), photocatalysts try to gain a negative charge which results in increased adsorption of positively charged dyes. This increased adsorption of dye on the surface of the photocatalyst is due to attraction of oppositely charged ions [92]. In acidic medium (at lower pH), positive ions of dye compete with H+ of the medium, that results in a decrease in adsorption on the surface of photocatalyst and a decrease in photodegradation. Similar is the case of MB dye. Reduction in the adsorption of MB on photocatalysts’ surface reduces reaction between MB and OH. There is no repulsion between MB and OH and repulsion between the negative surface of the photocatalyst and OH. As a result, OH will remain in reaction medium and play its vital role in the photodegradation process [93,94]. Jia et al., (2017), in a photocatalytic study of TiO2, revealed that maximum absorption of MB on its surface takes place in basic medium. That was because of the electrostatic attraction between cationic dye on the negatively charged active surface of TiO2 [95].

4.4. Effect of Catalyst Loading

It is necessary to find out the optimum quantity of the photocatalyst for optimum photodegradation of MB. The effect of catalyst loading on the photodegradation of MB was checked by irradiating UV/Visible radiations and keeping other parameters constant. The amount of photocatalyst varied from 0.01 g/dm3 to 0.25 g/dm3 for 5 ppm MB solution. Photodegradation efficiency is shown in Table 6. It can be seen from the data that degradation efficiency increases as the amount of catalyst increases and then decreases.
After the optimum amount of photocatalyst, further increase in the amount results in agglomeration. At higher levels of concentration turbidity of solution also increases. These two factors inhibit the absorption of photons on the surface of the photocatalyst and decrease its efficiency [96].

5. Proposed Photocatalytic Mechanism

From the Figure 12, it is clear that the energy band gap of MoSe2 was found to be 1.80 eV, so it can cover a broad solar energy region. The flat-band potential of MoSe2 was found to be 2.27 V on NHE. The photocatalytic mechanism of the synthesized composite is shown in the figure schematically. When light falls on the photocatalyst, valence band electrons of TiO2 are transferred to the conduction band leaving holes in the valence band. These valence electrons of TiO2 can be transferred to the conduction band of MoSe2 as it lies nearer to the VB of TiO2 as compared to its conduction band. 2D MoSe2 sheets act as active sites to carry out reactions. Separation of photogenerated electrons and reduction of charge recombination becomes easy in such heterostructures, which leads to efficient and enhanced photocatalytic activity. The adjustable band gap improves the light absorption range in the visible region of the composite. As the photocatalytic activity of the sample is linked with visible light absorption, 2D MoSe2 improves the activity of TiO2 in the visible region and near the IR region. In addition to this, activated carbon enhances the absorption capacity of the composite.
Photochemical reactions involved in degradation of dye are given below.
Photoexcitation: TiO2/AC/2D MoSe2 + hv (TiO2 surface) → eCBTiO2+h+VBTiO2
Entrapment of free electrons: eCB (TiO2) → eCB (MoSe2)
Photoexcited electron scavenging (Reduction): (O2) ads + eCB(MoSe2) → O2
Decomposition reaction: MB+ O2 → CO2 + H2O
Entrapment of holes: h+VB(TiO2) → h+VB (MoSe2)
Oxidation of hydroxyls: OH + h+ VB (MoSe2) → OH
Photodegradation by OH• radicals: MB + OH →CO2 + H2O
The above reactions describe details of the photocatalytic degradation process. Pollutants transfer from the bulk of wastewater to the surface of the photocatalyst. Adsorption of pollutants on the surface of the photocatalyst is activated by photons of light. Photons of light (UV-Vis-NIR region) excite electrons from the valence band of TiO2 to its conduction band and produce holes on the valence band of TiO2. Photo-excited electrons migrate from the conduction band of TiO2 to the conduction band of MoSe2. These electrons reduce molecular oxygen into oxygen free radicals. Being very reactive, these oxygen free radicals decompose organic pollutants such as MB into carbon dioxide gas and water molecules. Holes generated as a result of photo-excitation move from the valence band of TiO2 to the valence band of MoSe2. These holes oxidize water molecules into OH. These OH are involved in the decomposition of organic pollutants. Decomposition products of this reaction are CO2 and H2O. After these reactions, O2 and OH could be generated by entrapped UV-Vis-NIR light. These free radicals are strong oxidizing agents and can easily decompose organic pollutants present in wastewater.

6. Conclusions

Composites of TiO2 with 2D MoSe2 were synthesized in two steps and then AC was incorporated in these composites. The simple hydrothermal method was used for this purpose. TiO2 nanocrystals were synthesized from titanium tetrabutoxide in NaOH solution. Nanocrystals were calcinated at a high 300 °C for 3 h to obtain a stable phase of TiO2 nanoparticles. 2DMoSe2 was prepared by using sodium selenite and ammonium paramolybdate tetrahydrate. TiO2/xAC/2DMoSe2 composites were synthesized with the addition of AC and TiO2 in the reaction mixture of sodium selenite and ammonium paramolybdate tetrahydrate. The synthesis and morphology of TiO2/xAC/2DMoSe2 composites were established by FTIR, XRD and SEM. UV—vis spectroscopy and FT-IR analysis indicated the successful stabilization of composites. The studies revealed that composites are very useful as active photocatalysts to degrade MB and are expected to be equally useful to degrade other organic pollutants present in wastewater in the visible light spectrum. The extended activity of synthesized composites credited to the anatase-phase of TiO2 facilitated with the narrow-band gap 2D MoSe2. Moreover, the presence of AC in the composite helps to absorb extended solar radiation which resulted in an increase in photocatalytic activity of composites. As the percentage of AC increases in composites, degradation power increased directly. The environment-friendly synthesis of TiO2/xAC/2DMoSe2 using simple aqueous medium is an encouraging proposal to prepare composites of other metals as well. Photocatalytic composites can be used efficiently to fight with different environmental pitfalls including toxic organic dyes and health hazardous organic pollutants present in industrial wastewater. Synthesized composites showed maximum (83%) degradation efficiency that is greater than the degradation efficiency of any individual semiconductor i.e., TiO2 and 2D MoSe2 and previously reported TiO2 based photocatalysts. This indicates a very useful application of synthesized nano composites in industrial water treatments.

Author Contributions

Conceptualization, M.S.T.; Methodology, M.B.T.; Validation, M.S.; Formal analysis, S.A.; Investigation, M.B.T.; Resources, G.M.K., X.Z. and B.J.; Data curation, M.S.; Writing—original draft, S.A.; Writing—review & editing, S.N.; Visualization, S.N.; Supervision, M.S.T. and G.M.K.; Funding acquisition, X.Z. and B.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China at the Chinese Academy of Science and the APC was funded by Grant numbers 21974149 and 22174152.

Data Availability Statement

All data included in this study are available upon request by contacting the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Krishna, R.; Mishra, J.; Ighalo, J.O. Rising demand for rain water harvesting system in the world: A case study of Joda Town, India. World Sci. News 2020, 146, 47–59. [Google Scholar]
  2. Olvera, R.C.; Silva, S.L.; Robles-Belmont, E.; Lau, E.Z. Review of nanotechnology value chain for water treatment applications in Mexico. Resour.-Effic. Technol. 2017, 3, 1–11. [Google Scholar] [CrossRef]
  3. Singha, I.; Mishrab, P.K. Nano-membrane filtration a novel application of nanotechnology for waste water treatment. Mater. Today Proc. 2020, 29, 327–332. [Google Scholar] [CrossRef]
  4. Kamali, M.; Persson, K.M.; Costa, M.E.; Capela, I. Sustainability criteria for assessing nanotechnology applicability in industrial wastewater treatment: Current status and future outlook. Environ. Int. 2019, 125, 261–276. [Google Scholar] [CrossRef]
  5. Das, A.; Dey, A. P-Nitrophenol-Bioremediation using potent Pseudomonas strain from the textile dye industry effluent. J. Environ. Chem. Eng. 2020, 8, 103830. [Google Scholar] [CrossRef]
  6. Anas, M.; Han, D.S.; Mahmoud, K.; Park, H.; Abdel-Wahab, A. Photocatalytic degradation of organic dye using titanium dioxide modified with metal and non-metal deposition. Mater. Sci. Semicond. Process. 2016, 41, 209–218. [Google Scholar] [CrossRef]
  7. de Vidales, M.J.M.; Nieto-Márquez, A.; Morcuende, D.; Atanes, E.; Blaya, F.; Soriano, E.; Fernández-Martínez, F. 3D printed floating photocatalysts for wastewater treatment. Catal. Today 2019, 328, 157–163. [Google Scholar] [CrossRef]
  8. Ighalo, J.O.; Adeniyi, A.G.; Adeniran, J.A.; Ogunniyi, S. A systematic literature analysis of the nature and regional distribution of water pollution sources in Nigeria. J. Clean. Prod. 2021, 283, 124566. [Google Scholar] [CrossRef]
  9. Kumar, S.; Ahlawat, W.; Bhanjana, G.; Heydarifard, S.; Nazhad, M.M.; Dilbaghi, N. Nanotechnology-based water treatment strategies. J. Nanosci. Nanotechnol. 2014, 14, 1838–1858. [Google Scholar] [CrossRef]
  10. Qu, X.; Brame, J.; Li, Q.; Alvarez, P.J. Nanotechnology for a safe and sustainable water supply: Enabling integrated water treatment and reuse. Acc. Chem. Res. 2013, 46, 834–843. [Google Scholar] [CrossRef]
  11. Nwabor, O.F.; Nnamonu, E.; Martins, P.; Ani, O. Water and waterborne diseases: A review. Int. J. Trop. Dis. Health 2016, 12, 1–14. [Google Scholar]
  12. Munir, H.M.S.; Feroze, N.; Ramzan, N.; Sagir, M.; Babar, M.; Tahir, M.S.; Shamshad, J.; Mubashir, M.; Khoo, K.S. Fe-zeolite catalyst for ozonation of pulp and paper wastewater for sustainable water resources. Chemosphere 2022, 297, 134031. [Google Scholar] [CrossRef] [PubMed]
  13. Chen, L.; Caro, F.; Corbett, C.J.; Ding, X. Estimating the environmental and economic impacts of widespread adoption of potential technology solutions to reduce water use and pollution: Application to China’s textile industry. Environ. Impact Assess. Rev. 2019, 79, 106293. [Google Scholar] [CrossRef]
  14. Mittal, H.; Kumar, A.; Khanuja, M. In-situ oxidative polymerization of aniline on hydrothermally synthesized MoSe2 for enhanced photocatalytic degradation of organic dyes. J. Saudi Chem. Soc. 2019, 23, 836–845. [Google Scholar] [CrossRef]
  15. Khan, S.; Malik, A. Environmental and health effects of textile industry wastewater. In Environmental Deterioration and Human Health; Springer: Berlin/Heidelberg, Germany, 2014; pp. 55–71. [Google Scholar]
  16. Tounsadi, H.; Metarfi, Y.; Taleb, M.; El Rhazi, K.; Rais, Z. Impact of chemical substances used in textile industry on the employee's health: Epidemiological study. Ecotoxicol. Environ. Saf. 2020, 197, 110594. [Google Scholar] [CrossRef]
  17. Bener, S.; Atalay, S.; Ersöz, G. The hybrid process with eco-friendly materials for the treatment of the real textile industry wastewater. Ecol. Eng. 2020, 148, 105789. [Google Scholar] [CrossRef]
  18. Gupta, R.; Pandit, C.; Pandit, S.; Gupta, P.K.; Lahiri, D.; Agarwal, D.; Pandey, S. Potential and future prospects of biochar-based materials and their applications in removal of organic contaminants from industrial wastewater. J. Mater. Cycles Waste Manag. 2022, 24, 852–876. [Google Scholar] [CrossRef]
  19. Jaganathan, V.; Cherurveettil, P.; Chellasamy, A.; Premapriya, M. Environmental pollution risk analysis and management in textile industry: A preventive mechanism. In Proceedings of the 1st International Conference on Social Sciences and Humanities, ICSH 2014, Gaborone, Botswana, 4–7 June 2014. [Google Scholar]
  20. Khan, I.; Saeed, K.; Zekker, I.; Zhang, B.; Hendi, A.H.; Ahmad, A.; Ahmad, S.; Zada, N.; Ahmad, H.; Shah, L.A. Review on methylene blue: Its properties, uses, toxicity and photodegradation. Water 2022, 14, 242. [Google Scholar] [CrossRef]
  21. Santoso, E.; Ediati, R.; Kusumawati, Y.; Bahruji, H.; Sulistiono, D.; Prasetyoko, D. Review on recent advances of carbon based adsorbent for methylene blue removal from waste water. Mater. Today Chem. 2020, 16, 100233. [Google Scholar] [CrossRef]
  22. Abdelrahman, E.A.; Hegazey, R.; El-Azabawy, R.E. Efficient removal of methylene blue dye from aqueous media using Fe/Si, Cr/Si, Ni/Si, and Zn/Si amorphous novel adsorbents. J. Mater. Res. Technol. 2019, 8, 5301–5313. [Google Scholar] [CrossRef]
  23. Shandilya, P.; Mittal, D.; Soni, M.; Raizada, P.; Hosseini-Bandegharaei, A.; Saini, A.K.; Singh, P. Fabrication of fluorine doped graphene and SmVO4 based dispersed and adsorptive photocatalyst for abatement of phenolic compounds from water and bacterial disinfection. J. Clean. Prod. 2018, 203, 386–399. [Google Scholar] [CrossRef]
  24. Raizada, P.; Sudhaik, A.; Singh, P.; Shandilya, P.; Saini, A.K.; Gupta, V.K.; Lim, J.-H.; Jung, H.; Hosseini-Bandegharaei, A. Fabrication of Ag3VO4 decorated phosphorus and sulphur co-doped graphitic carbon nitride as a high-dispersed photocatalyst for phenol mineralization and E. coli disinfection. Sep. Purif. Technol. 2019, 212, 887–900. [Google Scholar] [CrossRef]
  25. Pare, B.; Singh, P.; Jonnalgadda, S. Degradation and Mineralization of Victoria Blue B Dye in a Slurry Photoreactor Using Advanced Oxidation Process; CSIR: New Delhi, India, 2009. [Google Scholar]
  26. Pare, B.; Singh, P.; Jonnalagadda, S. Visible Light Induced Heterogeneous Advanced Oxidation Process to Degrade Pararosanilin Dye in Aqueous Suspension of ZnO; CSIR: New Delhi, India, 2008. [Google Scholar]
  27. Watanabe, N.; Nakamura, S.; Liu, B.; Wang, N. Utilization of Structure from Motion for processing CORONA satellite images: Application to mapping and interpretation of archaeological features in Liangzhu Culture, China. Archaeol. Res. Asia 2017, 11, 38–50. [Google Scholar] [CrossRef]
  28. Sharma, K.; Dutta, V.; Sharma, S.; Raizada, P.; Hosseini-Bandegharaei, A.; Thakur, P.; Singh, P. Recent advances in enhanced photocatalytic activity of bismuth oxyhalides for efficient photocatalysis of organic pollutants in water: A review. J. Ind. Eng. Chem. 2019, 78, 1–20. [Google Scholar] [CrossRef]
  29. Priya, B.; Raizada, P.; Singh, N.; Thakur, P.; Singh, P. Adsorptional photocatalytic mineralization of oxytetracycline and ampicillin antibiotics using Bi2O3/BiOCl supported on graphene sand composite and chitosan. J. Colloid Interface Sci. 2016, 479, 271–283. [Google Scholar] [CrossRef]
  30. Raizada, P.; Singh, P.; Kumar, A.; Pare, B.; Jonnalagadda, S.B. Zero valent iron-brick grain nanocomposite for enhanced solar-Fenton removal of malachite green. Sep. Purif. Technol. 2014, 133, 429–437. [Google Scholar] [CrossRef]
  31. Sudhaik, A.; Raizada, P.; Shandilya, P.; Singh, P. Magnetically recoverable graphitic carbon nitride and NiFe2O4 based magnetic photocatalyst for degradation of oxytetracycline antibiotic in simulated wastewater under solar light. J. Environ. Chem. Eng. 2018, 6, 3874–3883. [Google Scholar] [CrossRef]
  32. Raizada, P.; Kumari, J.; Shandilya, P.; Dhiman, R.; Singh, V.P.; Singh, P. Magnetically retrievable Bi2WO6/Fe3O4 immobilized on graphene sand composite for investigation of photocatalytic mineralization of oxytetracycline and ampicillin. Process Saf. Environ. Prot. 2017, 106, 104–116. [Google Scholar] [CrossRef]
  33. Chen, D.; Cheng, Y.; Zhou, N.; Chen, P.; Wang, Y.; Li, K.; Huo, S.; Cheng, P.; Peng, P.; Zhang, R. Photocatalytic degradation of organic pollutants using TiO2-based photocatalysts: A review. J. Clean. Prod. 2020, 268, 121725. [Google Scholar] [CrossRef]
  34. Jafari, S.; Zhao, F.; Zhao, D.; Lahtinen, M.; Bhatnagar, A.; Sillanpää, M. A comparative study for the removal of methylene blue dye by N and S modified TiO2 adsorbents. J. Mol. Liq. 2015, 207, 90–98. [Google Scholar] [CrossRef]
  35. Beiki, H.; Esfahany, M.N.; Etesami, N. Turbulent mass transfer of Al2O3 and TiO2 electrolyte nanofluids in circular tube. Microfluid. Nanofluid. 2013, 15, 501–508. [Google Scholar] [CrossRef]
  36. Kusiak-Nejman, E.; Morawski, A.W. TiO2/graphene-based nanocomposites for water treatment: A brief overview of charge carrier transfer, antimicrobial and photocatalytic performance. Appl. Catal. B Environ. 2019, 253, 179–186. [Google Scholar] [CrossRef]
  37. Lee, C.-G.; Javed, H.; Zhang, D.; Kim, J.-H.; Westerhoff, P.; Li, Q.; Alvarez, P.J. Porous electrospun fibers embedding TiO2 for adsorption and photocatalytic degradation of water pollutants. Environ. Sci. Technol. 2018, 52, 4285–4293. [Google Scholar] [CrossRef] [PubMed]
  38. Mallakpour, S.; Nikkhoo, E. Surface modification of nano-TiO2 with trimellitylimido-amino acid-based diacids for preventing aggregation of nanoparticles. Adv. Powder Technol. 2014, 25, 348–353. [Google Scholar] [CrossRef]
  39. Raza, N.; Raza, W.; Gul, H.; Azam, M.; Lee, J.; Vikrant, K.; Kim, K.-H. Solar-light-active silver phosphate/titanium dioxide/silica heterostructures for photocatalytic removal of organic dye. J. Clean. Prod. 2020, 254, 120031. [Google Scholar] [CrossRef]
  40. Alamelu, K.; Ali, B.J. Au nanoparticles decorated sulfonated graphene-TiO2 nanocomposite for sunlight driven photocatalytic degradation of recalcitrant compound. Sol. Energy 2020, 211, 1194–1205. [Google Scholar] [CrossRef]
  41. Moradi, M.; Khorasheh, F.; Larimi, A. Pt nanoparticles decorated Bi-doped TiO2 as an efficient photocatalyst for CO2 photo-reduction into CH4. Sol. Energy 2020, 211, 100–110. [Google Scholar] [CrossRef]
  42. Xie, Z.; Zhao, F.; Zou, S.; Zhu, F.; Zhang, Z.; Wang, W. TiO2 nanorod arrays decorated with Au nanoparticles as sensitive and recyclable SERS substrates. J. Alloy. Compd. 2021, 861, 157999. [Google Scholar] [CrossRef]
  43. Xiao, W.-Z.; Xu, L.; Rong, Q.-Y.; Dai, X.-Y.; Cheng, C.-P.; Wang, L.-L. Two-dimensional H-TiO2/MoS2 (WS2) van der Waals heterostructures for visible-light photocatalysis and energy conversion. Appl. Surf. Sci. 2020, 504, 144425. [Google Scholar] [CrossRef]
  44. Wang, S.; Yun, J.-H.; Luo, B.; Butburee, T.; Peerakiatkhajohn, P.; Thaweesak, S.; Xiao, M.; Wang, L. Recent progress on visible light responsive heterojunctions for photocatalytic applications. J. Mater. Sci. Technol. 2017, 33, 1–22. [Google Scholar] [CrossRef]
  45. Yang, L.; Hao, X.; Yu, D.; Zhou, P.; Peng, Y.; Jia, Y.; Zhao, C.; He, J.; Zhan, C.; Lai, B. High visible-light catalytic activity of Bis-PDI-T@ TiO2 for activating persulfate toward efficient degradation of carbamazepine. Sep. Purif. Technol. 2021, 263, 118384. [Google Scholar] [CrossRef]
  46. Kang, L.; Zhang, M.; Zhang, J.; Liu, S.; Zhang, N.; Yao, W.; Ye, Y.; Luo, C.; Gong, Z.; Wang, C. Dual-defect surface engineering of bimetallic sulfide nanotubes towards flexible asymmetric solid-state supercapacitors. J. Mater. Chem. A 2020, 8, 24053–24064. [Google Scholar] [CrossRef]
  47. Wang, S.; Li, Y.; Wang, X.; Zi, G.; Zhou, C.; Liu, B.; Liu, G.; Wang, L.; Huang, W. One-step supramolecular preorganization constructed crinkly graphitic carbon nitride nanosheets with enhanced photocatalytic activity. J. Mater. Sci. Technol. 2022, 104, 155–162. [Google Scholar] [CrossRef]
  48. Wu, L.; Shi, S.; Li, Q.; Zhang, X.; Cui, X. TiO2 nanoparticles modified with 2D MoSe2 for enhanced photocatalytic activity on hydrogen evolution. Int. J. Hydrog. Energy 2019, 44, 720–728. [Google Scholar] [CrossRef]
  49. Hu, Z.; Wu, Z.; Han, C.; He, J.; Ni, Z.; Chen, W. Two-dimensional transition metal dichalcogenides: Interface and defect engineering. Chem. Soc. Rev. 2018, 47, 3100–3128. [Google Scholar] [CrossRef]
  50. Wang, D.; Xu, Y.; Sun, F.; Zhang, Q.; Wang, P.; Wang, X. Enhanced photocatalytic activity of TiO2 under sunlight by MoS2 nanodots modification. Appl. Surf. Sci. 2016, 377, 221–227. [Google Scholar] [CrossRef]
  51. Liang, Z.; Bai, X.; Hao, P.; Guo, Y.; Xue, Y.; Tian, J.; Cui, H. Full solar spectrum photocatalytic oxygen evolution by carbon-coated TiO2 hierarchical nanotubes. Appl. Catal. B Environ. 2019, 243, 711–720. [Google Scholar] [CrossRef]
  52. Coehoorn, R.; Haas, C.; Dijkstra, J.; Flipse, C.d.; De Groot, R.; Wold, A. Electronic structure of MoSe2, MoS2, and WSe2. I. Band-structure calculations and photoelectron spectroscopy. Phys. Rev. B 1987, 35, 6195. [Google Scholar] [CrossRef]
  53. Boyjoo, Y.; Sun, H.; Liu, J.; Pareek, V.K.; Wang, S. A review on photocatalysis for air treatment: From catalyst development to reactor design. Chem. Eng. J. 2017, 310, 537–559. [Google Scholar] [CrossRef]
  54. Ren, Z.; Liu, X.; Zhuge, Z.; Gong, Y.; Sun, C.Q. MoSe2/ZnO/ZnSe hybrids for efficient Cr (VI) reduction under visible light irradiation. Chin. J. Catal. 2020, 41, 180–187. [Google Scholar] [CrossRef]
  55. Rivera, P.; Schaibley, J.R.; Jones, A.M.; Ross, J.S.; Wu, S.; Aivazian, G.; Klement, P.; Seyler, K.; Clark, G.; Ghimire, N.J. Observation of long-lived interlayer excitons in monolayer MoSe2–WSe2 heterostructures. Nat. Commun. 2015, 6, 6242. [Google Scholar] [CrossRef] [PubMed]
  56. Chen, J.; Liu, B.; Cai, H.; Liu, S.; Yamauchi, Y.; Jun, S.C. Covalently Interlayer-Confined Organic–Inorganic Heterostructures for Aqueous Potassium Ion Supercapacitors. Small 2023, 19, 2204275. [Google Scholar] [CrossRef] [PubMed]
  57. Xie, T.; Liu, Y.; Wang, H.; Wu, Z. Layered MoSe2/Bi2WO6 composite with PN heterojunctions as a promising visible-light induced photocatalyst. Appl. Surf. Sci. 2018, 444, 320–329. [Google Scholar] [CrossRef]
  58. Leary, R.; Westwood, A. Carbonaceous nanomaterials for the enhancement of TiO2 photocatalysis. Carbon 2011, 49, 741–772. [Google Scholar] [CrossRef]
  59. Yan, Y.; Kuang, W.; Shi, L.; Ye, X.; Yang, Y.; Xie, X.; Shi, Q.; Tan, S. Carbon quantum dot-decorated TiO2 for fast and sustainable antibacterial properties under visible-light. J. Alloy. Compd. 2019, 777, 234–243. [Google Scholar] [CrossRef]
  60. Moosavi, S.; Lai, C.W.; Gan, S.; Zamiri, G.; Akbarzadeh Pivehzhani, O.; Johan, M.R. Application of efficient magnetic particles and activated carbon for dye removal from wastewater. ACS Omega 2020, 5, 20684–20697. [Google Scholar] [CrossRef]
  61. Siddiqui, I.; Mittal, H.; Kohli, V.K.; Gautam, P.; Ali, M.; Khanuja, M. Hydrothermally synthesized micron sized, broom-shaped MoSe2 nanostructures for superior photocatalytic water purification. Mater. Res. Express 2018, 5, 125020. [Google Scholar] [CrossRef]
  62. Pan, J.; Zhu, X.; Chen, X.; Zhao, Y.; Liu, J. Gd3+-Doped MoSe2 nanosheets used as a theranostic agent for bimodal imaging and highly efficient photothermal cancer therapy. Biomater. Sci. 2018, 6, 372–387. [Google Scholar] [CrossRef]
  63. Subramani, A.; Byrappa, K.; Kumaraswamy, G.; Ravikumar, H.; Ranganathaiah, C.; Lokanatha Rai, K.; Ananda, S.; Yoshimura, M. Hydrothermal preparation and characterization of TiO2: AC composites. Mater. Lett. 2007, 61, 4828–4831. [Google Scholar] [CrossRef]
  64. Alkorbi, A.S.; Javed, H.M.A.; Hussain, S.; Latif, S.; Mahr, M.S.; Mustafa, M.S.; Alsaiari, R.; Alhemiary, N.A. Solar light-driven photocatalytic degradation of methyl blue by carbon-doped TiO2 nanoparticles. Opt. Mater. 2022, 127, 112259. [Google Scholar] [CrossRef]
  65. Thinkohkaew, K.; Piroonpan, T.; Jiraborvornpongsa, N.; Potiyaraj, P. Development of multifunctional polypropylene nonwoven fabric by radiation induced grafting of TiO2 nanoparticles and trifluoroethyl methacrylate for protective textile applications. Materialia 2022, 21, 101355. [Google Scholar] [CrossRef]
  66. Jimoh, A.; Akpeji, B.; Azeez, S.; Ayipo, Y.; Abdulsalam, Z.; Adebayo, Z.; Ajao, A.; Zakariyah, A.; Elemike, E. Biosynthesis of Ag and TiO2 nanoparticles and the evaluation of their antibacterial activities. Inorg. Chem. Commun. 2022, 141, 109503. [Google Scholar] [CrossRef]
  67. Dessai, S.; Ayyanar, M.; Amalraj, S.; Khanal, P.; Vijayakumar, S.; Gurav, N.; Rarokar, N.; Kalaskar, M.; Nadaf, S.; Gurav, S. Bioflavonoid mediated synthesis of TiO2 nanoparticles: Characterization and their biomedical applications. Mater. Lett. 2022, 311, 131639. [Google Scholar] [CrossRef]
  68. Liu, S.; Chen, X.; Chen, X. A TiO2/AC composite photocatalyst with high activity and easy separation prepared by a hydrothermal method. J. Hazard. Mater. 2007, 143, 257–263. [Google Scholar] [CrossRef] [PubMed]
  69. Cai, Q.; Wang, F.; Xiang, J.; Dan, M.; Yu, S.; Zhou, Y. Layered MoS2 Grown on Anatase TiO2 {001} Promoting Interfacial Electron Transfer to Enhance Photocatalytic Evolution of H2 From H2S. Front. Environ. Chem. 2020, 1, 591645. [Google Scholar] [CrossRef]
  70. Yuan, Y.-J.; Ye, Z.-J.; Lu, H.-W.; Hu, B.; Li, Y.-H.; Chen, D.-Q.; Zhong, J.-S.; Yu, Z.-T.; Zou, Z.-G. Constructing anatase TiO2 nanosheets with exposed (001) facets/layered MoS2 two-dimensional nanojunctions for enhanced solar hydrogen generation. ACS Catal. 2016, 6, 532–541. [Google Scholar] [CrossRef]
  71. Wang, S.; Yi, L.; Halpert, J.E.; Lai, X.; Liu, Y.; Cao, H.; Yu, R.; Wang, D.; Li, Y. A novel and highly efficient photocatalyst based on P25–graphdiyne nanocomposite. Small 2012, 8, 265–271. [Google Scholar] [CrossRef]
  72. Dashora, A.; Ahuja, U.; Venugopalan, K. Electronic and optical properties of MoS2 (0 0 0 1) thin films: Feasibility for solar cells. Comput. Mater. Sci. 2013, 69, 216–221. [Google Scholar] [CrossRef]
  73. Mohammad, N.H.; EL-Sherbiny, G.M.; Hammad, A.A.; Askar, A.A.; El-Nour, S.A.A. Gamma-ray and sunlight-induced synthesis of silver nanoparticles using bacterial cellulose and cell-free filtrate produced by Komagataeibacter rhaeticus N1 MW322708 strain. Cellulose 2022, 29, 1791–1805. [Google Scholar] [CrossRef]
  74. Shah, I.H.; Ashraf, M.; Sabir, I.A.; Manzoor, M.A.; Malik, M.S.; Gulzar, S.; Ashraf, F.; Iqbal, J.; Niu, Q.; Zhang, Y. Green synthesis and Characterization of Copper oxide nanoparticles using Calotropis procera leaf extract and their different biological potentials. J. Mol. Struct. 2022, 1259, 132696. [Google Scholar] [CrossRef]
  75. Parvathiraja, C.; Katheria, S.; Siddiqui, M.R.; Wabaidur, S.M.; Islam, M.A.; Lai, W.-C. Activated Carbon-Loaded Titanium Dioxide Nanoparticles and Their Photocatalytic and Antibacterial Investigations. Catalysts 2022, 12, 834. [Google Scholar] [CrossRef]
  76. Dai, C.; Zhou, Z.; Tian, C.; Li, Y.; Yang, C.; Gao, X.; Tian, X. Large-scale synthesis of graphene-like MoSe2 nanosheets for efficient hydrogen evolution reaction. J. Phys. Chem. C 2017, 121, 1974–1981. [Google Scholar] [CrossRef]
  77. Tichapondwa, S.M.; Newman, J.; Kubheka, O. Effect of TiO2 phase on the photocatalytic degradation of methylene blue dye. Phys. Chem. Earth Parts A/B/C 2020, 118, 102900. [Google Scholar] [CrossRef]
  78. Small, J.M.; Hintelmann, H. Methylene blue derivatization then LC–MS analysis for measurement of trace levels of sulfide in aquatic samples. Anal. Bioanal. Chem. 2007, 387, 2881–2886. [Google Scholar] [CrossRef]
  79. Joseph, S.; Mathew, B. Microwave-assisted green synthesis of silver nanoparticles and the study on catalytic activity in the degradation of dyes. J. Mol. Liq. 2015, 204, 184–191. [Google Scholar] [CrossRef]
  80. Al-Hamoud, K.; Shaik, M.R.; Khan, M.; Alkhathlan, H.Z.; Adil, S.F.; Kuniyil, M.; Assal, M.E.; Al-Warthan, A.; Siddiqui, M.R.H.; Tahir, M.N. Pulicaria undulata Extract-Mediated Eco-Friendly Preparation of TiO2 Nanoparticles for Photocatalytic Degradation of Methylene Blue and Methyl Orange. ACS Omega 2022, 7, 4812–4820. [Google Scholar] [CrossRef]
  81. Rahman, Q.I.; Ali, A.; Ahmad, N.; Lohani, M.B.; Mehta, S.; Muddassir, M. Synthesis and characterization of CuO rods for enhanced visible light driven dye degradation. J. Nanosci. Nanotechnol. 2020, 20, 7716–7723. [Google Scholar] [CrossRef]
  82. Rahman, Q.I.; Hasan, S.; Ali, A.; Mehta, S.; Raja, M.A.; Ahmad, N.; Khan, A.R.; Muddassir, M. Synthesis and Characterizations of Nitrogen (N) Doped Strontium Titanate (SrTiO3) Nanoparticles for Enhanced Visible Light Driven Photocatalytic Degradation. J. Nanosci. Nanotechnol. 2020, 20, 6475–6481. [Google Scholar] [CrossRef]
  83. Askari, M.B.; Banizi, Z.T.; Seifi, M.; Dehaghi, S.B.; Veisi, P. Synthesis of TiO2 nanoparticles and decorated multi-wall carbon nanotube (MWCNT) with anatase TiO2 nanoparticles and study of optical properties and structural characterization of TiO2/MWCNT nanocomposite. Optik 2017, 149, 447–454. [Google Scholar] [CrossRef]
  84. Yoon, C.-J.; Lee, S.-H.; Kwon, Y.-B.; Kim, K.; Lee, K.-H.; Kim, S.M.; Kim, Y.-K. Fabrication of sustainable and multifunctional TiO2@ carbon nanotube nanocomposite fibers. Appl. Surf. Sci. 2021, 541, 148332. [Google Scholar] [CrossRef]
  85. Alghamdi, Y.G.; Krishnakumar, B.; Malik, M.A.; Alhayyani, S. Design and preparation of biomass-derived activated carbon loaded TiO2 photocatalyst for photocatalytic degradation of reactive red 120 and ofloxacin. Polymers 2022, 14, 880. [Google Scholar] [CrossRef] [PubMed]
  86. Justh, N.; Mikula, G.J.; Bakos, L.P.; Nagy, B.; László, K.; Parditka, B.; Erdélyi, Z.; Takáts, V.; Mizsei, J.; Szilágyi, I.M. Photocatalytic properties of TiO2@ polymer and TiO2@ carbon aerogel composites prepared by atomic layer deposition. Carbon 2019, 147, 476–482. [Google Scholar] [CrossRef]
  87. Saravanan, L.; Patil, R.A.; Gultom, P.; Kumar, B.; Manikandan, A.; Fu, Y.-P.; Chueh, Y.-L.; Cheng, C.-L.; Yeh, W.-C.; Ma, Y.-R. Rutile-phase TiO2@ carbon core-shell nanowires and their photoactivation in visible light region. Carbon 2021, 181, 280–289. [Google Scholar] [CrossRef]
  88. Salama, A.; Mohamed, A.; Aboamera, N.M.; Osman, T.; Khattab, A. Photocatalytic degradation of organic dyes using composite nanofibers under UV irradiation. Appl. Nanosci. 2018, 8, 155–161. [Google Scholar] [CrossRef]
  89. Elsayed, E.M.; Elnouby, M.S.; Gouda, M.; Elessawy, N.A.; Santos, D. Effect of the morphology of tungsten oxide embedded in sodium alginate/polyvinylpyrrolidone composite beads on the photocatalytic degradation of methylene blue dye solution. Materials 2020, 13, 1905. [Google Scholar] [CrossRef]
  90. Abdellah, M.; Nosier, S.; El-Shazly, A.; Mubarak, A. Photocatalytic decolorization of methylene blue using TiO2/UV system enhanced by air sparging. Alex. Eng. J. 2018, 57, 3727–3735. [Google Scholar] [CrossRef]
  91. Azeez, F.; Al-Hetlani, E.; Arafa, M.; Abdelmonem, Y.; Nazeer, A.A.; Amin, M.O.; Madkour, M. The effect of surface charge on photocatalytic degradation of methylene blue dye using chargeable titania nanoparticles. Sci. Rep. 2018, 8, 7104. [Google Scholar] [CrossRef]
  92. Nguyen Thi Thu, T.; Nguyen Thi, N.; Tran Quang, V.; Nguyen Hong, K.; Nguyen Minh, T.; Le Thi Hoai, N. Synthesis, characterisation, and effect of pH on degradation of dyes of copper-doped TiO2. J. Exp. Nanosci. 2016, 11, 226–238. [Google Scholar] [CrossRef]
  93. Hejazi, R.; Mahjoub, A.R.; Khavar, A.H.C.; Khazaee, Z. Fabrication of novel type visible-light-driven TiO2@ MIL-100 (Fe) microspheres with high photocatalytic performance for removal of organic pollutants. J. Photochem. Photobiol. A Chem. 2020, 400, 112644. [Google Scholar] [CrossRef]
  94. Hendekhale, N.R.; Mohammad-Khah, A. A novel synthesis of Co2ZrO5 and m-ZrO2 nanoparticles by sono-precipitation and hydrothermal methods and their application in UV/Visible-photocatalytic studies. J. Environ. Chem. Eng. 2020, 8, 104065. [Google Scholar] [CrossRef]
  95. Jia, Z.; La, L.; Zhang, W.; Liang, S.; Jiang, B.; Xie, S.; Habibi, D.; Zhang, L. Strong enhancement on dye photocatalytic degradation by ball-milled TiO2: A study of cationic and anionic dyes. J. Mater. Sci. Technol. 2017, 33, 856–863. [Google Scholar] [CrossRef]
  96. Khade, G.; Gavade, N.; Suwarnkar, M.; Dhanavade, M.; Sonawane, K.; Garadkar, K. Enhanced photocatalytic activity of europium doped TiO2 under sunlight for the degradation of methyl orange. J. Mater. Sci. Mater. Electron. 2017, 28, 11002–11011. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of activation of carbon.
Figure 1. Schematic representation of activation of carbon.
Water 15 01788 g001
Figure 2. Schematic representation of stepwise synthesis of TiO2.
Figure 2. Schematic representation of stepwise synthesis of TiO2.
Water 15 01788 g002
Figure 3. Schematic representation of stepwise synthesis of TiO2/ AC/2DMoSe2.
Figure 3. Schematic representation of stepwise synthesis of TiO2/ AC/2DMoSe2.
Water 15 01788 g003
Figure 4. FTIR patterns: Pure TiO2, TiO2/AC and TiO2/xAC/2DMoSe2 composite.
Figure 4. FTIR patterns: Pure TiO2, TiO2/AC and TiO2/xAC/2DMoSe2 composite.
Water 15 01788 g004
Figure 5. XRD patterns with different composition of AC in TiO2/xAC/2DMoSe2.
Figure 5. XRD patterns with different composition of AC in TiO2/xAC/2DMoSe2.
Water 15 01788 g005
Figure 6. SEM images, (a) Pure AC; (b) Pure TiO2 and (cf) TiO2/xAC/2DMoSe2 composites.
Figure 6. SEM images, (a) Pure AC; (b) Pure TiO2 and (cf) TiO2/xAC/2DMoSe2 composites.
Water 15 01788 g006aWater 15 01788 g006b
Figure 7. Photocatalytic degradation of Pure TiO2: (a) Degradation of MB at different time intervals; (b) C/C0 against time; (c) ln C/C0 against time; (d) %age Efficiency.
Figure 7. Photocatalytic degradation of Pure TiO2: (a) Degradation of MB at different time intervals; (b) C/C0 against time; (c) ln C/C0 against time; (d) %age Efficiency.
Water 15 01788 g007
Figure 8. Photocatalytic degradation of TiO2/0.25AC/2DMoSe2: (a) Degradation of MB at different time intervals; (b) C/C0 against time; (c) ln C/C0 against time; (d) %age Efficiency.
Figure 8. Photocatalytic degradation of TiO2/0.25AC/2DMoSe2: (a) Degradation of MB at different time intervals; (b) C/C0 against time; (c) ln C/C0 against time; (d) %age Efficiency.
Water 15 01788 g008
Figure 9. Photocatalytic degradation of TiO2/0.5AC/2DMoSe2: (a) Degradation of MB at different time intervals; (b) C/C0 against time; (c) ln C/C0 against time; (d) %age Efficiency.
Figure 9. Photocatalytic degradation of TiO2/0.5AC/2DMoSe2: (a) Degradation of MB at different time intervals; (b) C/C0 against time; (c) ln C/C0 against time; (d) %age Efficiency.
Water 15 01788 g009aWater 15 01788 g009b
Figure 10. Photocatalytic degradation of TiO2/0.75AC/2DMoSe2: (a) Degradation of MB at different time intervals; (b) C/C0 against time; (c) ln C/C0 against time; (d) %age Efficiency.
Figure 10. Photocatalytic degradation of TiO2/0.75AC/2DMoSe2: (a) Degradation of MB at different time intervals; (b) C/C0 against time; (c) ln C/C0 against time; (d) %age Efficiency.
Water 15 01788 g010
Figure 11. Photocatalytic degradation of TiO2/1AC/2DMoSe2: (a) Degradation of MB at different time intervals; (b) C/C0 against time; (c) ln C/C0 against time; (d) %age efficiency.
Figure 11. Photocatalytic degradation of TiO2/1AC/2DMoSe2: (a) Degradation of MB at different time intervals; (b) C/C0 against time; (c) ln C/C0 against time; (d) %age efficiency.
Water 15 01788 g011
Figure 12. Proposed mechanism of action of TiO2/AC/2DMoSe2 photocatalyst.
Figure 12. Proposed mechanism of action of TiO2/AC/2DMoSe2 photocatalyst.
Water 15 01788 g012
Table 1. No. of publications on TiO2/MoSe2 from 2019 to 2023 at different sources.
Table 1. No. of publications on TiO2/MoSe2 from 2019 to 2023 at different sources.
Sr. No.SourceNo. of Publications in Years
201920202021202220232019–2023
1Google Scholar54578395115806684527
2Science Direct 6620151057
Table 2. Comparison of photodegradation efficiencies of previously studied and synthesized composites.
Table 2. Comparison of photodegradation efficiencies of previously studied and synthesized composites.
Sr. No.CompositeRatioPhotodegradation %Age Efficiency
1ACpure55% [75]
2TiO2Pure71–73% [75]
3TiO2/ACNA55% [86]
4TiO2@ACNA65% [87]
5TiO2/MWCNTNA61.1% [83]
6TiO2@carbon nanotubeNA85% [84]
7AC-TiO2/(OFL)NA82% [85]
8TiO2/AC/2D MoSe21:0.25:176% present work
9TiO2/AC/2D MoSe21:0.5:180% present work
10TiO2/AC/2D MoSe21:0.75:181% present work
11TiO2/AC/2D MoSe21:1:183% present work
Table 3. Effect of irradiation time on MB degradation using photocatalysts.
Table 3. Effect of irradiation time on MB degradation using photocatalysts.
PhotocatalystTime (min)0306090120150
Pure TiO2MB mg/L1097.35.14.22.9
TiO2/0.25AC/2DMoSe2106.25.34.12.92.4
TiO2/0.50AC/2DMoSe2107.66.64.83.32.0
TiO2/0.75AC/2DMoSe2107.06.55.53.51.9
TiO2/AC/2DMoSe2106.96.45.33.61.7
Table 4. Initial concentration of MB versus percentage degradation.
Table 4. Initial concentration of MB versus percentage degradation.
Sr. No.MB (mg/L)% Degradation
1587
21083
31574
42069
62565
Table 5. Effect of change of pH on photodegradation.
Table 5. Effect of change of pH on photodegradation.
Sr. No.pHPercentage Degradation
17.072
27.574
38.079
48.580
59.082
69.583
710.081
Table 6. Effect of catalyst load on photodegradation efficiency.
Table 6. Effect of catalyst load on photodegradation efficiency.
Sr. No.Catalyst Loading (g/dm3)Percentage Degradation
10.0179
20.1083
30.1573
40.2066
50.2562
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

Ahmad, S.; Tahir, M.S.; Kamal, G.M.; Zhang, X.; Nazir, S.; Tahir, M.B.; Jiang, B.; Safdar, M. TiO2/Activated Carbon/2D Selenides Composite Photocatalysts for Industrial Wastewater Treatment. Water 2023, 15, 1788. https://doi.org/10.3390/w15091788

AMA Style

Ahmad S, Tahir MS, Kamal GM, Zhang X, Nazir S, Tahir MB, Jiang B, Safdar M. TiO2/Activated Carbon/2D Selenides Composite Photocatalysts for Industrial Wastewater Treatment. Water. 2023; 15(9):1788. https://doi.org/10.3390/w15091788

Chicago/Turabian Style

Ahmad, Shehzad, Muhammad Suleman Tahir, Ghulam Mustafa Kamal, Xu Zhang, Saima Nazir, Muhammad Bilal Tahir, Bin Jiang, and Muhammad Safdar. 2023. "TiO2/Activated Carbon/2D Selenides Composite Photocatalysts for Industrial Wastewater Treatment" Water 15, no. 9: 1788. https://doi.org/10.3390/w15091788

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