Next Article in Journal
Quasi-2D SnO2 Thin Films for Gas Sensors: Chemoresistive Response and Temperature Effect on Adsorption of Analytes
Next Article in Special Issue
Sustainable Packaging Design for Molded Expanded Polystyrene Cushion
Previous Article in Journal
Gas Atomization of Duplex Stainless Steel Powder for Laser Powder Bed Fusion
Previous Article in Special Issue
Effectiveness of Dimple Microtextured Copper Substrate on Performance of Sn-0.7Cu Solder Alloy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Near-Infrared (NIR) Silver Sulfide (Ag2S) Semiconductor Photocatalyst Film for Degradation of Methylene Blue Solution

by
Zahrah Ramadlan Mubarokah
1,
Norsuria Mahmed
1,2,*,
Mohd Natashah Norizan
2,3,
Ili Salwani Mohamad
2,3,
Mohd Mustafa Al Bakri Abdullah
1,2,
Katarzyna Błoch
4,
Marcin Nabiałek
4,
Madalina Simona Baltatu
5,*,
Andrei Victor Sandu
5,6,7 and
Petrica Vizureanu
5,8
1
Faculty of Chemical Engineering & Technology, Universiti Malaysia Perlis (UniMAP), Arau 01000, Malaysia
2
Centre of Excellence Geopolymer and Green Technology (CEGeoGTech), Universiti Malaysia Perlis (UniMAP), Arau 01000, Malaysia
3
Faculty of Electronic Engineering & Technology, Universiti Malaysia Perlis (UniMAP), Arau 02600, Malaysia
4
Faculty of Mechanical Engineering and Computer Science, Częstochowa University of Technology, 42-201 Częstochowa, Poland
5
Department of Technologies and Equipments for Materials Processing, Faculty of Materials Science and Engineering, Gheorghe Asachi Technical University of Iaşi, Blvd. Mangeron, No. 51, 700050 Iasi, Romania
6
National Institute for Research and Development in Environmental Protection INCDPM, Splaiul Independentei 294, 060031 Bucharest, Romania
7
Romanian Inventors Forum, Str. Sf. P. Movila 3, 700089 Iasi, Romania
8
Technical Sciences Academy of Romania, Dacia Blvd 26, 030167 Bucharest, Romania
*
Authors to whom correspondence should be addressed.
Submission received: 26 November 2022 / Revised: 24 December 2022 / Accepted: 27 December 2022 / Published: 3 January 2023
(This article belongs to the Special Issue Design and Applications of Functional Materials, Volume II)

Abstract

:
A silver sulfide (Ag2S) semiconductor photocatalyst film has been successfully synthesized using a solution casting method. To produce the photocatalyst films, two types of Ag2S powder were used: a commercialized and synthesized powder. For the commercialized powder (CF/comAg2S), the Ag2S underwent a rarefaction process to reduce its crystallite size from 52 nm to 10 nm, followed by incorporation into microcrystalline cellulose using a solution casting method under the presence of an alkaline/urea solution. A similar process was applied to the synthesized Ag2S powder (CF/syntAg2S), resulting from the co-precipitation process of silver nitrate (AgNO3) and thiourea. The prepared photocatalyst films and their photocatalytic efficiency were characterized by Fourier transform infrared spectroscopy (FTIR), X-ray diffraction (XRD), and UV-visible spectroscopy (UV-Vis). The results showed that the incorporation of the Ag2S powder into the cellulose films could reduce the peak intensity of the oxygen-containing functional group, which indicated the formation of a composite film. The study of the crystal structure confirmed that all of the as-prepared samples featured a monoclinic acanthite Ag2S structure with space group P21/C. It was found that the degradation rate of the methylene blue dye reached 100% within 2 h under sunlight exposure when using CF/comAg2S and 98.6% for the CF/syntAg2S photocatalyst film, and only 48.1% for the bare Ag2S powder. For the non-exposure sunlight samples, the degradation rate of only 33–35% indicated the importance of the semiconductor near-infrared (NIR) Ag2S photocatalyst used.

1. Introduction

Since the prehistoric age, humans have used dyes from natural resources for coloring purposes. The lengthy process, poor colorfastness, excessive cost of producing natural dyes, and the increasing demand for textiles led to the discovery of synthetic dyes from petroleum compounds that outperformed the properties of natural dyes [1,2,3]. Statistically, more than 7 × 105 tons of synthetic dyes are produced worldwide, and about 1 × 104 of them are used in industry [4]. However, these synthetic dyes have a negative impact on the environment, such as water pollution. Textile industries discharge around 35% of dye wastewater into water bodies as effluent annually [5]. These dyes can contaminate the aquatic habitat, which may enter the food chain [6,7]. Methylene blue (MB), or C16H18N3SCl, is one of the synthetic dye materials used in textile industries. When dissolved in water, it results in the formation of a blue-colored solution. The discharging of MB dye into the environment is considered as a significant threat for aesthetical and toxicological reasons. In terms of aquatic life, its existence in the water could reduce sunlight transmittance and decrease oxygen solubility due to the high molar absorption coefficient (~8.4 × 104 L·mol−1·cm−1 at 664 nm) of MB dyes [8,9,10]. At a certain concentration, it can cause serious threats to human health, including respiratory distress, abdominal disorders, blindness, and digestive and mental disorders [11].
In recent years, tremendous attempts have been made to improve industrial wastewater treatment methods. These methods can be divided into physical [12,13,14,15,16], biological [17,18,19,20,21], and chemical methods. For the chemical methods, there are several conventional chemical dye-removal processes that have been used, e.g., the coagulant flocculation, electrochemical destruction, advanced oxidation process (AOP), etc. [22,23,24,25]. Amongst these, the advanced oxidation processes (AOPs) have received the most attention due to their significant competence in acting toward a wide range of organic or inorganic dye pollutants in the aqueous phase, by converting those pollutants into stable inorganic compounds such as water, carbon dioxide, and salt without any footprint and sludge production [26,27]. The AOPs also rely on the in situ production of greatly reactive hydroxyl radicals (OH•) [28].
In the midst of the AOPs, heterogeneous photocatalysis has been considered a leading method and desired breakthrough to cope with organic contaminants. Photocatalysis is a reaction that involves light (photoreaction) and accelerates the reaction due to the presence of a catalyst that absorbs light energy to form the reducing and oxidizing (electron–hole pairs) ions on the surface of the catalyst. The electron’s handover occurs during the oxidation–reduction process [29]. The reduction of an acceptor occurs when the electrons (e) combine with oxygen in the water to generate an anion (O2), which oxidizes the hydroxyl radical (OH•), while the hole (h+) will oxidize the dissolved hydroxyl and convert it into a radical with great energy [30]. These processes work simultaneously. The photocatalytic properties of various metal oxides (e.g., titania, TiO2, and zinc oxide, ZnO) have been extensively studied. The results of these studies show that the metal oxides become active potential photocatalysts in wastewater treatment [31,32,33,34]. However, due to a wide bandgap, they can only absorb the ultraviolet (UV) light and show poor performance in the visible–near-infrared (NIR) range of the solar spectrum [29,35,36,37]. In addition, the solar spectrum itself is dominated by the visible (46%) and NIR spectral (49%) rather than the UV spectral (5%) [38]. In addition, the manufacturing cost is relatively high for many of the metal oxide materials [39,40]. Therefore, metal sulfides have become an option.
Among all of the metal sulfide photocatalyst materials, silver sulfide (Ag2S) shows more efficient charge separation, attributed to the remarkable synergistic effects of strong NIR light absorption and excellent surface properties [41,42]. Furthermore, compared to the other metal sulfide compounds such as cadmium selenide, CdS, and lead selenide, PbS, Ag2S is found to be acceptable for wastewater treatment due to its non-toxic properties [43,44,45]. Silver sulfide has an eminent performance in the degradation of pollutants, solar energy conversion, and production of hydrogen with various approaches that have been well established for synthesizing a variety of visible–NIR-light driven Ag2S photocatalysts. For example, Ag2S synthesized using the facile ion-exchange method at room temperature can be used as an effective photocatalyst for the decomposition of methylene orange (MO). The results of a study confirmed the excellent photo-oxidation performance of Ag2S since MO can be completely photodegraded with Ag2S in just 30 min, and 70 min under visible light and NIR light irradiation. This performance is due to the narrow band gap of Ag2S, 1.078 eV, and the lower recombination efficiency and photogenerated electron–hole pairs of Ag2S during the photocatalytic process [46]. The Ag2S photocatalyst has usually been used in the powder form [43,47,48,49]. This is because the photocatalyst powder can be well dispersed in suspensions [50].
However, this photocatalyst powder exhibits certain drawbacks especially for nano-sized powders/particles. During the degradation process, the photocatalyst may undergo coagulation due to the instability of the nano-sized particles, which will hamper the light incidence on the active centers, consequently reducing its catalytic activity [51]. Furthermore, for the slurry system, the main challenge is to recover the nano-sized photocatalyst particles from the treated water [52,53,54]. This condition leads to the requirements of high filtration costs of catalyst removal [55], hindering its industrial application and impractical. In addition, the Ag2S nanoparticles are susceptible to photo-corrosion, which means that the sulfide ions have the potential to be oxidized into sulfur by photogenerated holes when they are exposed to irradiation [56]. There have been a lot of efforts towards increasing the photocatalytic efficiency and preventing the photo-corrosion of Ag2S [57,58,59,60]. A photocatalyst film is one of the plausible material technologies to diminish the impacts of photocatalyst powder on the environment [61]. A photocatalyst film has promising performance in terms of adsorption and efficient mass transport. The adsorption ability is maxed out by the larger surface area that allows more catalyst deposit in the film, resulting in higher photocatalytic activity. Moreover, the mass transport of reactants, intermediates, and products is enhanced by the compact substrate that allows better contact with the catalyst, and prevents the leakage of the powder of the photocatalyst material into clean water [62,63]. Table 1 shows the cost comparison for some types of photocatalyst, especially titania (TiO2), which is a common photocatalyst used for industrial purposes. Those types of photocatalyst are used under a UV lamp-assisted reactor for the photocatalytic processes and consume a certain amount of energy (power). For an industrial batch process, it will consume a lot of energy with a higher cost of operation. Thus, our study introduces a composite film (cellulose/Ag2S) that can effectively degrade methylene blue concentrations using only sunlight, which is cost-effective. Furthermore, the composite film can also be easily removed from the water after the treatment process without a costly separation method.
To date, bio-based products such as cellulose film have become attractive compounds due to their excellent properties that are harmless to the environment. A cellulose film has distinctive characteristics such as transparency, robustness, low water content, etc. These properties lead to the wide usage of cellulose film as a renewable alternative to petroleum-based materials [68]. Furthermore, it is hypothesized that cellulose may become one of the solutions to improve the efficiency of electron donors and to establish the hole (h+) transporter during the light irradiation process. To the best of our knowledge, the incorporation of Ag2S powder into cellulose film has not been studied, although the integration between semiconductor catalyst and cellulose film is a promising innovation for photocatalytic applications. Therefore, the integration of Ag2S in cellulose-derived film is hoped to enhance the photocatalytic efficiency. Thus, this study aimed to synthesize cellulose/Ag2S film in the presence of an alkali/urea solution using a simple solution casting method to describe the transformation of the functional group and the structure of the cellulose thin film due to the deposition of Ag2S. We also compared the photocatalytic efficiency of the commercial Ag2S and synthesized Ag2S doped into the cellulose film via the degradation of the MB solution.

2. Materials and Methods

The current study successfully synthesized films that were stated as CFs for cellulose film, cellulose films with commercial Ag2S (CF/comAg2S), and cellulose films with synthesized Ag2S (CF/syntAg2S). The materials and methods used for synthesizing the Ag2S photocatalyst film are described in the section below.

2.1. Materials and Reagents

The materials and reagents used consist of microcrystalline cellulose, MCC (≤100%), commercial silver sulfide, Ag2S, silver nitrate, AgNO3, thiourea, (NH2)2CS, and polyvinyl alcohol, PVA. All of these materials were bought from Sigma-Aldrich. The PVA was dissolved with distilled water at the temperature of 90 °C until the transparent PVA solution with a concentration of 5.0% (w/w) occurred. Chemicals including sodium hydroxide (NaOH), glycerol (C3H8O3), and acetone (C3H6O) were supplied by HmbG chemicals. The acetone was diluted in distilled water with a ratio of 2:1 as the agent of regeneration for cellulose films, called an acetone bath. Urea (CO(NH2)2) was obtained from Bendosen Laboratory Chemicals.

2.2. Preparation of Sodium Hydroxide/Urea (NaOH/urea) Aqueous Solution

To obtain the NaOH/urea solution, the NaOH pellets and urea powder were weighed to obtain 7 g and 12 g of mass, respectively [47,69]. Subsequently, each sample of NaOH and urea was diluted in a separate beaker glass before being mixed in a 100 mL volumetric flask. The balanced chemical reaction worked in the mixture can be written as:
2NaOH(aq) + CO(NH2)2(aq) → 2NH3(g) + Na2CO3(aq)
The solution was agitated to obtain a homogenous solution. Once the homogeneity was obtained, the solution was stored for the next process.

2.3. Liquid Phase Rarefaction of Commercial Ag2S by Magnetic Stirring Technique

Prior to the incorporation of the commercial Ag2S into the cellulose film, it was essential to break down the bulk Ag2S into smaller particle sizes. In this step, the commercial Ag2S powder was weighed according to the mass variation that is displayed in Table 2.
After the desired amount of Ag2S was obtained, the powder was placed into a beaker containing 6 mL of distilled water and vigorously stirred using a magnetic stirrer under ambient temperature for 24 h. The crystallite size of rarefaction Ag2S was then investigated using X-ray diffraction (XRD) to confirm its size reduction.

2.4. Synthesis of the Cellulose Film

The cellulose films were prepared using a solution casting method. The cellulose with NaOH/urea aqueous solution was used as the main reagent to synthesize the cellulose films. Both commercial Ag2S and synthesized Ag2S were incorporated into the cellulose system to study the different behavior of the cellulose film with commercial and synthesized Ag2S. The composition of NaOH, urea, and water was different for each process as shown in Table 3. Figure 1 shows the overall schematic procedure to synthesize the CFs, CF/comAg2S, and CF/syntAg2S samples.

2.4.1. Synthesis Cellulose Film (CF) and Cellulose Film/Commercial Ag2S (CF/comAg2S)

At the beginning of the process, 15 mL of NaOH/urea aqueous solution was precooled to a temperature of ~1 °C in the ice bath. After that, 3 wt% of microcrystalline cellulose (MCC) was added into the precooled solvent, and the mixture was rapidly agitated for 20 min. Then, the cellulose solution was taken out from the ice bath and 2.5 wt% of glycerol [70] was wisely dropped into the solution. Afterwards, the solution was spun for another hour to achieve the homogeneity. Once the dope and viscous solution was obtained, it was cast onto an 8.5 cm × 6 cm × 0.2 cm glass mold. Subsequently, the casted sample was regenerated in an acetone coagulant bath [71] until opaque-sheet-like cellulose hydrogel was formed. The hydrogel was soaked in distilled water for neutralization. Before drying, the hydrogel was submerged in a 5 wt% polyvinyl alcohol (PVA) bath for 2 h [72]. Finally, a transparent cellulose film was obtained. The same steps were also conducted in order to synthesize the CF/comAg2S specimens, except the 6mL of refracted Ag2S produced from Section 2.3 was first poured into the NaOH/urea solution prior to the addition of the desired amount of MCC into the precooled solution.

2.4.2. Synthesis of Cellulose Film/Synthesized Ag2S (CF/syntAg2S)

The co-precipitation method was used to synthesize the Ag2S particles. Silver nitrate and thiourea were the main precursors to produce the Ag2S slurry. The presence of silver and sulfide ions in the solution was required for the chemical deposition process to occur in order to complete the formation of Ag2S. The chemical reaction involved during the deposition process is shown in the following formulas [73,74]:
(NH2)2CS + H2O ↔ (NH2)2CO + H2S
2AgNO3 + H2S → Ag2S + 2HNO3
Table 3 shows the final composition that was utilized in this section for the whole synthesis process:
Table 3. Sample code and sample composition to synthesize samples CF/syntAg2S.
Table 3. Sample code and sample composition to synthesize samples CF/syntAg2S.
Sample CodeMolarity of AgNO3AgNO3 SolutionThiourea Solution
AgNO3WaterNaOHThioureaUreaWater
CF/syntAg2S10.1 M0.084 g5.0 mL1.6 g1.3 g1.6 g10.5 mL
CF/syntAg2S20.3 M0.254 g5.0 mL1.6 g1.3 g1.6 g10.5 mL
CF/syntAg2S30.5 M0.422 g5.0 mL1.6 g1.3 g1.6 g10.5 mL
According to Table 3, the total volume of solution that was used in the synthesis of CF/syntAg2S samples was 20 mL (5 mL AgNO3 solution and 15 mL thiourea solution). The solutions were prepared in two different beakers. At the beginning of the process, 15 mL of thiourea solution was poured into a 50 mL glass beaker while stirring on a magnetic stirrer. After that, the 5 mL of AgNO3 solution was added dropwise into the stirred thiourea solution. At this stage, a black precipitate was obtained, which indicated the formation of Ag2S particles. The solution was left to stir for another 15 min to ensure all of the substances completely reacted. Then, the precipitated solution was precooled in the ice bath for 20 min. Subsequently, 3% w/v of MCC powder was added into the black cold-precipitated solution and continuously stirred until the MCC was well dispersed in the solution. The mixed solution was then removed from the ice bath, dripped with glycerol, and spun for an entire hour. The cellulose/synthesized AgNO3 solution was transferred into the glass mold and then regenerated by using an acetone bath and PVA bath, in sequence. Lastly, the cellulose film obtained after the regenerated film was open-air dried for 2 days. All of the procedures above were repeated for different molarities of synthesized AgNO3.

2.5. Photocatalytic Activity Test

For photocatalytic testing, a concentration of 10 ppm of methylene blue (MB) solution was prepared. An amount of 0.1 g of MB powder was diluted with distilled water in a 100 mL volumetric flask. Afterwards, 0.7 g of each sample was submerged into a glass beaker consisting of 70 mL of 10 ppm MB and immediately exposed to sunlight for a total of 300 min. Every 30 min, about 40 mL of the irradiated solution was transferred into a glass vial with a dropper to investigate its degradation behavior using an ultraviolet-visible (UV-Vis) spectrophotometer, (Lambda 25, Perkin Elmer, Waltham, MA, USA) with 650 nm of wavelength. The degradation ratio of MB was determined using Equation (4); where At is the degradation ratio, I0 is base absorbance, and It is absorbency after time t [33,73,75].
A t = I 0 I t I 0 × 100

2.6. Characterizations

To identify chemical bonds and to locate the functional group of the cellulose films, the samples were characterized using a Fourier-transform infrared (FTIR) analysis. The analysis was carried out using an ATR-Perkin Elmer Spectrometer 2000 FT-IR, based on the attenuated total reflection (ATR) approach to ensure non-destructive analysis. The measurement was conducted under 650 cm−1–4000 cm−1 and 4 cm−1 resolution. The phase determination and crystallite size measurement were conducted using a Rigaku RINT 2000 X-ray diffraction (XRD). The reflection mode with monochromator-filtered Cu K α radiation (λ = 0.15418 nm) at 30 kV and 10 mA was applied. Samples were prepared in two different ways depending on the type of tested sample. For the powder sample, it was loaded onto the XRD sample holder with a small circle cavity in the middle of the holder. For film samples, the cellulose film was cut into a size of 25 mm, thickness ≤ 8 mm, and stuck on the holder. The prepared samples were subsequently scanned with 2theta in the range between 15° and 80°. The Debye-Scherrer Equation (5) was generated to compute the crystallite size and the degree of crystallinity was obtained from Equation (6) as follows:
D = K λ β   cos θ
Degree   of   Crystallinity = Area   of   all   crystalline   peaks Total   area   under   XRD   peaks × 100
where, D = crystallite size (nm); K = Scherrer’s constant (K = 0.94); λ = the wavelength of X-ray (1.54178 Å); β = full width half maximum (FWHM); θ = angle of diffraction (rad) [76].

3. Results and Discussion

3.1. The Formation of Cellulose-Photocatalyst Film

In nature, the structure of cellulose includes both axial C–H bonds and equatorial hydroxyl groups. Hence, it has amphiphilic properties with the inter-layer region being hydrophobic and the intra-layer region being hydrophilic. The intra- and intermolecular hydrogen bonds should be broken to a large extent when the cellulose is dissolved in the solvent [77]. Thus, when cellulose is dissolved in the NaOH/urea solution, it is expected that the NaOH interacts with the hydrophilic hydroxyl groups of cellulose, breaking the hydrogen bond between the chains, and the urea reacts with the hydrophobic portions, resulting in a solution that dissolves cellulose [78].
In this study, it was also discovered that the semi-crystalline properties of the cellulose can impact the end-product of the prepared films. During the drying process, the cellulose films will experience severe shrinkage [72,79,80]. Nevertheless, when 2.5% w/v glycerol and 5% PVA bath was applied, the shrinkage of the film samples became less. This is because the existence of PVA as a plasticizer in the film can improve the toughness of the cellulose glycerol composite films [73]. The commercial Ag2S powder was micron- sized and might affect the distribution of particles when incorporated in the MCC film. Thus, the rarefaction process was conducted as shown in Figure 2.
The particle size transformations were involved during the process. The stirring action from the magnetic stirrer was deployed in a clockwise direction at a certain speed in the liquid while applying fluidic shear forces among the involved substances [81]. The existence of opposing forces between the applied forces and the friction forces within the system (Figure 2) caused the occurrence of shear stress in the Ag2S powders, which can reduce the Ag2S particle size. The longer the magnetic bar rotates, the more forces work, resulting in more energy provided to wear down the Ag2S powder. Thus, smaller crystal Ag2S were formed. According to the peak analysis and crystallite size calculation, it was confirmed that before stirring, the Ag2S particle has a crystallite size of 52 nm and this size was reduced to 10 nm after the rarefaction process. Thus, the purpose of the rarefaction in this study was succeeded.
The distribution of both the rarefacted and synthesized Ag2S powder in the cellulose film can be observed in Figure 3. According to Figure 3d–f, the particles from the synthesized Ag2S particles evenly spread in the whole area of the cellulose matrix. Meanwhile, the samples that were loaded with the commercial Ag2S had diverse particle distributions (Figure 3a–c). It was shown that part of the cellulose film was not affected by the Ag2S. The degree of crystallinity and the crystallite size of the synthesized Ag2S can be categorized as one of the factors that affected this particle distribution. Therefore, further discussion about the structure and chemical function of the samples will be explained in the following section.

3.2. Crystalline Phase Investigation of the Synthesized Ag2S powder

Figure 4 shows the XRD spectra of the synthesized Ag2S particles that were obtained from the reaction of 0.3 M AgNO3 and thiourea. The diffraction peaks of Ag2S that were observed at 2theta = 22.4°, 25.9°, 28.9°, 31.5°, 34.4°, 36.8°, 37.7°, 40.7°, 43.6°, and 46.2° showed similar patterns with the commercial Ag2S and confirmed the formation of Ag2S phase (ICDD No. 00-014-0072) with monoclinic structure.

3.3. Crystal Structure and Phase Identification of Cellulose Film/Synthesized Ag2S (CF/syntAg2S)

To ensure the formation of the synthesized silver sulfide in the cellulose film samples, it was necessary to delve into the characterization of their crystal structure and phase identification. Furthermore, the influence of the AgNO3 concentration on the crystal structure of the sample was also studied. The crystallite is crucial because the crystallinity of the catalyst in the film will have a direct impact on the photocatalytic properties.
The study of the crystal structure of the CF/syntAg2S sample confirmed that all as-prepared samples have diffraction peaks of a monoclinic Acanthine Ag2S structure with space group P21/C. The specific lattice parameter of these samples can be indexed as a = 9.5200 Å, b = 6.9300 Å, and c = 8.2900 Å. Moreover, the peak of (002) due to the reflection of the cellulose plane also existed in all of the samples that had undergone different AgNO3/thiourea loadings during the synthesis. The acquired peaks of the Ag2S and cellulose were matched well with the references ICDD No. 00-009-0422 and ICDD No. 00-050-2241, respectively. The average grain size and degree of crystallinity of the samples are shown in Table 4.
In terms of peak intensity, it was observed that the higher the concentration, the lower and broader the peak (see Figure 5), which is attributed to the low crystallite size of the sample [82]. Table 4 illustrates that the smallest yield of crystallite size was found from the highest concentration of the substituent (CF/syntAg2S). Furthermore, Table 4 also depicts that the substituent was responsible for the shifted peak that occurred in the sample with concentrated loadings. The shifted peak towards the lower 2theta indicated that the lattice parameter of the pristine cellulose had been incorporated by the synthesized Ag2S [83,84].

3.4. Functional Group Analysis of CF/comAg2S and CF/syntAg2S

According to Figure 6, the commercial (Figure 6a–d) and synthesized (Figure 6g–j) samples had comparable FTIR patterns along the observation scanning range. However, in terms of peak intensity, the sample with the synthesized Ag2S had a sharp and narrow peak compared to the commercial sample. The spectra illustrated the same pattern among the thin films that were either impregnated by the commercial or synthesized Ag2S. The broad bands that were found from wavenumber 3336.5 cm−1 to 3419.8 cm−1 were designated to the hydroxyl (O-H) group stretching vibration. The shifted peak among these bands was attributed to the interaction of the hydrogen bond of the pristine cellulose with glycerol and PVA. These displacements were also found by [71,85], in which the high content of glycerol and PVA in the cellulose matrix impacted the sharpening and shifting to the higher wavenumber. The spectral band located from 2916.5 cm−1 to 2912.59 cm−1, as well as a weak peak at 897.00 cm−1 were attributed to the characteristic of symmetrical stretching of C-H from the alkyl group [32] and glycosidic CH deformation [72], respectively.
Thereafter, the transmittance band at the area of 1753.70 cm−1 to 1752.13 cm−1 was associated with the stretching vibration of the C=O ester carbonyl group and the medium peak at the range of 1238 cm−1 was denoted as C-O-C asymmetric stretching [86]. The authentic peak of cellulose was found at the region around 1059 cm−1 and remarked as C-O-C stretching vibrations of aliphatic primary and secondary alcohols in cellulose [87,88]. The peak of C-H bending was observed at the regions of 1374 cm−1 to 1373 cm−1. The study by Cazón et al. announced that the chemical content of glycerol and PVA in the cellulose matrix may contribute to the displacement of the observed peak. Nevertheless, Figure 6e,f,k,l indicates that the loadings of metal silver sulfide into the cellulose can reduce the intensity of the peak with an oxygen-containing functional group. The more concentrated the Ag2S, the lower the intensity of the peak-contained oxygen functional group. According to Kumar et al., it was confirmed that the strong electrostatic linkages between Ag2S and the functional groups of cellulose were responsible for the lower intensity of the transmittance peak into the oxygen-containing functional group [89]. This means that the different intensity of the peak, particularly in the range of oxygen-containing group regions, was a characteristic that showed the success of Ag2S deposited into the cellulose matrix.

3.5. Photocatalytic Degradation Mechanism of Methylene Blue Using the Cellulose/Ag2S Films

The degradation of the methylene blue (MB) dye solution with sunlight exposure and without exposure was investigated. The results are shown in Table 5. For the non-exposure samples, the degradation rate of both synthesized and commercial Ag2S incorporated in the cellulose film was only 35% and 33% (due to the adsorption by cellulose film), respectively, even after 5 h of processing. Compared to the same samples that were exposed to sunlight, the degradation rate of MB reached 98–100% within 2 h. This result showed that solar energy seemed to have a great influence on the photocatalytic activity of the Ag2S-containing samples, since this semiconductor catalyst absorbed in the NIR region. Nevertheless, according to the direct observation of photocatalytic activity under sunlight exposure, there were two kinds of reactions that might have occurred: firstly, the reaction between the cellulose surface and the MB solution; secondly, the reaction among deposited Ag2S in the cellulose film and the MB solution. Figure 7 shows the reaction mechanism that might have occurred between the cellulose and MB solution during the process.
The cellulose interacts with MB through several bonds, such as the electrostatic bond between the ion N+ in MB and ion O in cellulose. The electrostatic attraction that was involved in the adsorption mechanism of MB on cellulose led to the enhancement of the MB molecules to quickly fill the adsorption sites on the surface of the cellulose film, resulting in a high rate of MB adsorption [90]. The Van der Waals force (C=C) and hydrogen bond also formed during this process [91]. These bonds made it possible for the cellulose to experience the photolysis reaction when exposed to the solar source.
On the other hand, under solar illumination, the existence of the catalyst Ag2S in the cellulose film established the photocatalytic activity due to the configuration of the reducing and oxidizing (electron–hole pairs) ions on the surface of the catalyst, which may degrade the concentration of the MB dyes solution until it reached 100% discoloration (see Table 5). During the process of solar irradiation, photons of solar light with an energy that was either equal to or higher than the band gap of Ag2S (~1.06 eV) [92] were degraded by Ag2S. Upon the process of photon absorption, an electron (e-) from the valence band (VB) was stimulated up to the conduction band (CB). In addition, it was linked to the development of a hole (h+) in the valence band, which led to the formation of electron–hole pairs that took part in the process of reduction and oxidation. In this process, the electrons in the surface reacted with the oxygen (O2) that was dissolved in the aqueous solution, which then resulted in the production of anionic superoxide radicals (•O2) [93]. Simultaneously, the photogenerated holes also reacted with water to create hydroxyl radicals, which further oxidized the dye molecules. The produced anionic superoxide radicals (•O2) and hydroxyl radicals oxidized the dye molecules that can decompose the MB dyes into CO2 or H2O and their intermediates [94]. Overall, this phenomenon can be expressed through the following chemical reactions:
Ag2S + hve + h+
e + O2 → •O2
h+ + H2O → •OH + H+
Methylene   Blue     O 2 OH   free   radicals     intermediates   +   CO 2   +   H 2 O
However, without light energy, the processes involving reduction–oxidation reactions in the photocatalysis process will not be formed. As a result, the MB dyes wastewater will only interact with the cellulose film as described above.
The incorporation of Ag2S into cellulose films served primarily to increase the rate of dye degradation by creating a massive contact surface. The catalyst, Ag2S, was distributed well toward the cellulose films, leading to the films outperforming the powder counterparts in terms of activity because they effectively absorbed light and perhaps underwent internal scattering within the cellulose/Ag2S film, inducing higher charge carrier formation and hence better photocatalytic efficiency. Table 5 shows the photocatalytic efficiency for all samples based on the MB degradation rate (%). From the table, when Ag2S powder was used, the degradation rate of MB only reached about 75% even after 5 h of sunlight exposure. For pristine cellulose, the degradation of MB reached 100% after 4 h exposure. The high rate of MB adsorption by pristine cellulose might be due to the electrostatic attraction that was involved in the adsorption mechanism of the MB molecules to quickly fill the adsorption sites on the surface of the cellulose [90]. For CF/com Ag2S, 100% of the MB was degraded in 2 h of exposure time compared to the CF/syntAg2S sample with 98.6% degradation. The deterioration in the treatment that applied the catalyst was compared to the degradation in the photolysis experiment that was also conducted without the presence of a catalyst in the aqueous solution of methylene blue. By observing the direct photolysis of MB, it was clear that there were no appreciable color changes over the 5 h of exposure time. The degradation rate also reached its lowest value, ranging from 0% to 1.66%, as shown in Figure 8a,c.
Based on Figure 8b,d, the concentration of the dye solution was elevated to more than 90% in just 60 min for the pristine cellulose, CF/comAg2S and CF/syntAg2S samples, while for the Ag2S powder, less than 40% of the MB was degraded. This showed that the distribution of Ag2S powders in regenerated cellulose matrices can enhance the photocatalytic activity up to 100% efficiency due to higher surface area, and the affinity of the photocatalyst films that react with the dye molecules during the sunlight irradiation [95].
In addition, the surface of the cellulose film acted as a host during the process of dye adsorption, which in turn enabled the donor and acceptor molecules to interact with one another. This process could delay the recombination of charge carriers under direct sunlight. Charge recombination must be avoided during this process since it can bring down the efficiency of the photocatalyst [96,97,98]. In terms of the degradation ratio, it can be seen that there were differences in the photodegradation characterization of the two types of samples that were synthesized. The samples with commercial Ag2S doping will have an increasing ability with increasing Ag2S loading, where the higher content that can be accepted by cellulose film is 0.1 g. Meanwhile, the other types of samples (CF/syntAg2S) showed the opposite nature. This may occur due to overload on the sample (Figure 3d–f) with loading that might exceed the limit, resulting in poor interfacial charge carrier migration and an increase in the recombination rate of photoexcited electron–hole pairs due to an increase in the concentration of Ag sources [73,74]. Thus, it was found that the CF/comAg2S samples had the efficiency to degrade 100% of the MB content compared to the CF/syntAg2S (98.6%) in 2 h.

4. Conclusions

It has been successfully demonstrated that silver sulfide (Ag2S) particles, which were incorporated into regenerated cellulose, outperformed its properties in the powder state. This was because the distribution of Ag2S particles on the cellulose matrix could increase its photocatalytic activity as the contact surface between the catalyst and dyes became larger. Furthermore, the molecules contained in the cellulose were found to have the ability to interact with molecules in the dyes, especially the methylene blue (MB) dyes. The samples with the commercial Ag2S showed an increase in photocatalytic activity up to 100% degradation after 120 min of sunlight exposure with the maximum of 0.1 g Ag2S loading onto the cellulose film, while other types of samples (CF/syntAg2S) showed 98.6%. This situation might be due to the overload of synthesized Ag2S particles on the sample that exceeded the photocatalyst limit on the film. As a result, poor interfacial charge carrier migration and an increase in the recombination rate of photoexcited electron–hole pairs occurred. Thus, it can be concluded that the photocatalytic efficiency of the photocatalyst film originating from the commercial Ag2S particles was the highest, with 100% of the degradation rates in 2 h. For the non-exposure sunlight samples, the degradation rates were only 33–35%, showing the importance of the NIR semiconductor Ag2S catalyst used.

Author Contributions

Conceptualization, Z.R.M., N.M. and M.N.N.; methodology, I.S.M., M.M.A.B.A. and K.B.; software, A.V.S.; validation, Z.R.M., N.M. and P.V.; formal analysis, M.N.N.; investigation, I.S.M., M.M.A.B.A. and K.B.; data curation, K.B., M.N. and M.S.B.; writing—original draft preparation, Z.R.M., N.M., M.N.N. and I.S.M.; writing—review and editing, K.B., M.N. and A.V.S.; visualization, M.M.A.B.A.; supervision, P.V.; project administration, M.M.A.B.A.; funding acquisition, M.S.B. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Fundamental Research Grant Scheme FRGS/1/2021/TK0/UNIMAP/02/33 from the Ministry of Education Malaysia (MOE) and by Gheorghe Asachi Technical University of Iasi—TUIASI Romania, Scientific Research Funds, FCSU-2022.

Conflicts of Interest

The authors declare that they have no known competing financial interest or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Ziarani, G.M.; Moradi, R.; Lashgari, N.; Kruger, H.G. Introduction and Importance of Synthetic Organic Dyes. In Metal-Free Synthetic Organic Dyes; Elsevier: Amsterdam, The Netherlands, 2018; pp. 1–7. ISBN 978-0-12-815647-6. [Google Scholar]
  2. Fleck, B.; Cabral, I.; Souto, A.P. Eco Printing of Linen and Tencel Substrates with Onion Skins and Red Cabbage. Mater. Circ. Econ. 2022, 4, 2. [Google Scholar] [CrossRef]
  3. Kusumastuti, A.; Anis, S.; Fardhyanti, D.S. Production of natural dyes powder based on chemo-physical technology for textile application. IOP Conf. Ser. Earth Environ. Sci. 2019, 258, 012028. [Google Scholar] [CrossRef]
  4. Elsahida, K.; Fauzi, A.M.; Sailah, I.; Siregar, I.Z. Sustainability of the use of natural dyes in the textile industry. IOP Conf. Ser. Earth Environ. Sci. 2019, 399, 012065. [Google Scholar] [CrossRef]
  5. Hasan, K.M.F.; Wang, H.; Mahmud, S.; Islam, A.; Habib, M.A.; Genyang, C. Enhancing mechanical and antibacterial performances of organic cotton materials with greenly synthesized colored silver nanoparticles. IJCST, 2022; ahead-of-print. [Google Scholar] [CrossRef]
  6. Bahuguna, A.; Singh, S.; Bahuguna, A.; Sharma, S.; Dadarwal, B. Physical method of Wastewater treatment-A review. J. Res. Environ. Earth Sci. 2021, 7, 2348–2532. [Google Scholar]
  7. Berradi, M.; Hsissou, R.; Khudhair, M.; Assouag, M.; Cherkaoui, O.; El Bachiri, A.; El Harfi, A. Textile finishing dyes and their impact on aquatic environs. Heliyon 2019, 5, e02711. [Google Scholar] [CrossRef]
  8. Kosswattaarachchi, A.M.; Cook, T.R. Repurposing the Industrial Dye Methylene Blue as an Active Component for Redox Flow Batteries. ChemElectroChem 2018, 5, 3437–3442. [Google Scholar] [CrossRef]
  9. Lawagon, C.P.; Amon, R.E.C. Magnetic rice husk ash “cleanser” as efficient methylene blue adsorbent. Environ. Eng. Res. 2019, 25, 685–692. [Google Scholar] [CrossRef]
  10. Zhou, M.; Chen, J.; Hou, C.; Liu, Y.; Xu, S.; Yao, C.; Li, Z. Organic-free synthesis of porous CdS sheets with controlled windows size on bacterial cellulose for photocatalytic degradation and H2 production. Appl. Surf. Sci. 2019, 470, 908–916. [Google Scholar] [CrossRef]
  11. Wang, H.; Li, G.; Fakhri, A. Fabrication and structural of the Ag2S-MgO/graphene oxide nanocomposites with high photocatalysis and antimicrobial activities. J. Photochem. Photobiol. B Biol. 2020, 207, 111882. [Google Scholar] [CrossRef]
  12. Hammood, Z.A.; Chyad, T.F.; Al-Saedi, R. Adsorption Performance of Dyes Over Zeolite for Textile Wastewater Treatment. Ecol. Chem. Eng. S 2021, 28, 329–337. [Google Scholar] [CrossRef]
  13. Januário, E.F.D.; Vidovix, T.B.; Calsavara, M.A.; Bergamasco, R.; Vieira, A.M.S. Membrane surface functionalization by the deposition of polyvinyl alcohol and graphene oxide for dyes removal and treatment of a simulated wastewater. Chem. Eng. Process.—Process Intensif. 2022, 170, 108725. [Google Scholar] [CrossRef]
  14. Khan, M.M.; Pradhan, D.; Sohn, Y. (Eds.) Nanocomposites for Visible Light-induced Photocatalysis. In Springer Series on Polymer and Composite Materials; Springer International Publishing: Cham, Germany, 2017; ISBN 978-3-319-62445-7. [Google Scholar]
  15. Popa, N.; Visa, M. New hydrothermal charcoal TiO2 composite for sustainable treatment of wastewater with dyes and cadmium cations load. Mater. Chem. Phys. 2021, 258, 123927. [Google Scholar] [CrossRef]
  16. Singh, R.; Munya, V.; Are, V.N.; Nayak, D.; Chattopadhyay, S. A Biocompatible, pH-Sensitive, and Magnetically Separable Superparamagnetic Hydrogel Nanocomposite as an Efficient Platform for the Removal of Cationic Dyes in Wastewater Treatment. ACS Omega 2021, 6, 23139–23154. [Google Scholar] [CrossRef]
  17. Azimi, B.; Abdollahzadeh-Sharghi, E.; Bonakdarpour, B. Anaerobic-aerobic processes for the treatment of textile dyeing wastewater containing three commercial reactive azo dyes: Effect of number of stages and bioreactor type. Chin. J. Chem. Eng. 2021, 39, 228–239. [Google Scholar] [CrossRef]
  18. Haddad, M.; Abid, S.; Hamdi, M.; Bouallagui, H. Reduction of adsorbed dyes content in the discharged sludge coming from an industrial textile wastewater treatment plant using aerobic activated sludge process. J. Environ. Manag. 2018, 223, 936–946. [Google Scholar] [CrossRef]
  19. Idris, I.; Rahmadhani, I.; Sudiana, I.M. Feasibility of Aspergillus keratitidis InaCC1016 for synthetic dyes removal in dyes wastewater treatment. IOP Conf. Ser. Earth Environ. Sci. 2020, 439, 012027. [Google Scholar] [CrossRef]
  20. Ikram, M.; Zahoor, M.; El-Saber Batiha, G. Biodegradation and decolorization of textile dyes by bacterial strains: A biological approach for wastewater treatment. Z. Für Phys. Chem. 2021, 235, 1381–1393. [Google Scholar] [CrossRef]
  21. Shen, L.; Jin, Z.; Xu, W.; Jiang, X.; Shen, Y.; Wang, Y.; Lu, Y. Enhanced Treatment of Anionic and Cationic Dyes in Wastewater through Live Bacteria Encapsulation Using Graphene Hydrogel. Ind. Eng. Chem. Res. 2019, 58, 7817–7824. [Google Scholar] [CrossRef]
  22. Ben Ayed, S.; Azam, M.; Al-Resayes, S.I.; Ayari, F.; Rizzo, L. Cationic Dye Degradation and Real Textile Wastewater Treatment by Heterogeneous Photo-Fenton, Using a Novel Natural Catalyst. Catalysts 2021, 11, 1358. [Google Scholar] [CrossRef]
  23. Dong, Y.-Y.; Zhu, Y.-H.; Ma, M.-G.; Liu, Q.; He, W.-Q. Synthesis and characterization of Ag@AgCl-reinforced cellulose composites with enhanced antibacterial and photocatalytic degradation properties. Sci. Rep. 2021, 11, 3366. [Google Scholar] [CrossRef]
  24. Hussin, N.A.M.; Abidin, C.Z.A.; Fahmi; Ibrahim, A.H.; Ahmad, R.; Singa, P.K. Optimization of Anthraquinone Dye Wastewater Treatment using Ozone in the Presence of Persulfate Ion in a Semi-batch Reactor. IOP Conf. Ser. Earth Environ. Sci. 2021, 920, 012019. [Google Scholar] [CrossRef]
  25. Saeed, M.; Adeel, S.; Muneer, M.; ul Haq, A. Photo Catalysis: An Effective Tool for Treatment of Dyes Contaminated Wastewater. In Bioremediation and Biotechnology, Vol 2; Bhat, R.A., Hakeem, K.R., Dervash, M.A., Eds.; Springer International Publishing: Cham, Germany, 2020; pp. 175–187. ISBN 978-3-030-40332-4. [Google Scholar]
  26. Brienza, M.; Katsoyiannis, I. Sulfate Radical Technologies as Tertiary Treatment for the Removal of Emerging Contaminants from Wastewater. Sustainability 2017, 9, 1604. [Google Scholar] [CrossRef] [Green Version]
  27. Crini, G.; Lichtfouse, E. Advantages and disadvantages of techniques used for wastewater treatment. Environ. Chem. Lett. 2019, 17, 145–155. [Google Scholar] [CrossRef]
  28. Sengupta, S.; Pal, C.K. Chemistry in Wastewater Treatment. In Advanced Materials and Technologies for Wastewater Treatment; CRC Press: Boca Raton, FL, USA, 2021; ISBN 978-1-00-313830-3. [Google Scholar]
  29. Riente, P.; Noël, T. Application of metal oxide semiconductors in light-driven organic transformations. Catal. Sci. Technol. 2019, 9, 5186–5232. [Google Scholar] [CrossRef] [Green Version]
  30. Gisbertz, S.; Pieber, B. Heterogeneous Photocatalysis in Organic Synthesis. ChemPhotoChem 2020, 4, 456–475. [Google Scholar] [CrossRef] [Green Version]
  31. Al-Mayyahi, A.; Al-Asadi, A.A. Advanced oxidation processes (aops) for wastewater treatment and reuse: A brief review. Asian J. Appl. Sci. Technol 2018, 2, 13. [Google Scholar]
  32. Kumar, N.; Mittal, H.; Reddy, L.; Nair, P.; Ngila, J.C.; Parashar, V. Morphogenesis of ZnO nanostructures: Role of acetate (COOH) and nitrate (NO3) ligand donors from zinc salt precursors in synthesis and morphology dependent photocatalytic properties. RSC Adv. 2015, 5, 38801–38809. [Google Scholar] [CrossRef]
  33. Zailan, S.N.; Bouaissi, A.; Mahmed, N.; Abdullah, M.M.A.B. Influence of ZnO Nanoparticles on Mechanical Properties and Photocatalytic Activity of Self-cleaning ZnO-Based Geopolymer Paste. J. Inorg. Organomet. Polym. 2020, 30, 2007–2016. [Google Scholar] [CrossRef]
  34. Subramaniam, M.N.; Goh, P.-S.; Lau, W.-J.; Ng, B.-C.; Ismail, A.F. Chapter 3—development of nanomaterial-based photocatalytic membrane for organic pollutants removal. In Advanced Nanomaterials for Membrane Synthesis and its Applications; Lau, W.-J., Ismail, A.F., Isloor, A., Al-Ahmed, A., Eds.; Micro and Nano Technologies; Elsevier: Amsterdam, The Netherlands, 2019; pp. 45–67. ISBN 978-0-12-814503-6. [Google Scholar]
  35. Ishchenko, O.M.; Rogé, V.; Lamblin, G.; Lenoble, D. Tio2- and zno-based materials for photocatalysis: Material properties, device architecture and emerging concepts. In Semiconductor Photocatalysis—Materials, Mechanisms and Applications; Cao, W., Ed.; InTech: London, UK, 2016; ISBN 978-953-51-2484-9. [Google Scholar]
  36. Kanakkillam, S.S.; Krishnan, B.; Guzman, S.S.; Martinez, J.A.A.; Avellaneda, D.A.; Shaji, S. Defects rich nanostructured black zinc oxide formed by nanosecond pulsed laser irradiation in liquid. Appl. Surf. Sci. 2021, 567, 150858. [Google Scholar] [CrossRef]
  37. Haounati, R.; El Guerdaoui, A.; Ouachtak, H.; El Haouti, R.; Bouddouch, A.; Hafid, N.; Bakiz, B.; Santos, D.M.F.; Labd Taha, M.; Jada, A.; et al. Design of direct Z-scheme superb magnetic nanocomposite photocatalyst Fe3O4/Ag3PO4@Sep for hazardous dye degradation. Sep. Purif. Technol. 2021, 277, 119399. [Google Scholar] [CrossRef]
  38. Hu, X.; Li, Y.; Tian, J.; Yang, H.; Cui, H. Highly efficient full solar spectrum (UV-vis-NIR) photocatalytic performance of Ag2S quantum dot/TiO2 nanobelt heterostructures. J. Ind. Eng. Chem. 2017, 45, 189–196. [Google Scholar] [CrossRef]
  39. Dias, P.; Mendes, A. Hydrogen Production from Photoelectrochemical Water Splitting. In Encyclopedia of Sustainability Science and Technology; Meyers, R.A., Ed.; Springer: New York, NY, USA, 2018; pp. 1–52. ISBN 978-1-4939-2493-6. [Google Scholar]
  40. Danish, M.S.S.; Bhattacharya, A.; Stepanova, D.; Mikhaylov, A.; Grilli, M.L.; Khosravy, M.; Senjyu, T. A Systematic Review of Metal Oxide Applications for Energy and Environmental Sustainability. Metals 2020, 10, 1604. [Google Scholar] [CrossRef]
  41. Sadovnikov, S.I.; Gusev, A.I. Recent progress in nanostructured silver sulfide: From synthesis and nonstoichiometry to properties. J. Mater. Chem. A 2017, 5, 17676–17704. [Google Scholar] [CrossRef] [Green Version]
  42. Shafi, A.; Ahmad, N.; Sultana, S.; Sabir, S.; Khan, M.Z. Ag2S-sensitized NiO–ZnO heterostructures with enhanced visible light photocatalytic activity and acetone sensing property. ACS Omega 2019, 4, 12905–12918. [Google Scholar] [CrossRef] [Green Version]
  43. Yu, W.; Yin, J.; Li, Y.; Lai, B.; Jiang, T.; Li, Y.; Liu, H.; Liu, J.; Zhao, C.; Singh, S.C.; et al. Ag2S Quantum Dots as an Infrared Excited Photocatalyst for Hydrogen Production. ACS Appl. Energy Mater. 2019, 2, 2751–2759. [Google Scholar] [CrossRef]
  44. Badawi, A. Effect of the non-toxic Ag2S quantum dots size on their optical properties for environment-friendly applications. Phys. E Low-Dimens. Syst. Nanostructures 2019, 109, 107–113. [Google Scholar] [CrossRef]
  45. Shahri, N.N.M.; Taha, H.; Hamid, M.H.S.A.; Kusrini, E.; Lim, J.-W.; Hobley, J.; Usman, A. Antimicrobial activity of silver sulfide quantum dots functionalized with highly conjugated Schiff bases in a one-step synthesis. RSC Adv. 2022, 12, 3136–3146. [Google Scholar] [CrossRef]
  46. Jiang, W.; Wu, Z.; Yue, X.; Yuan, S.; Lu, H.; Liang, B. Photocatalytic performance of Ag2S under irradiation with visible and near-infrared light and its mechanism of degradation. RSC Adv. 2015, 5, 24064–24071. [Google Scholar] [CrossRef]
  47. Yuan, L.; Lu, S.; Yang, F.; Wang, Y.; Jia, Y.; Kadhim, M.; Yu, Y.; Zhang, Y.; Zhao, Y. A facile room-temperature synthesis of three-dimensional coral-like Ag2S nanostructure with enhanced photocatalytic activity. J. Mater. Sci. 2019, 54, 3174–3186. [Google Scholar] [CrossRef]
  48. Al-Shehri, B.M.; Shkir, M.; Bawazeer, T.M.; AlFaify, S.; Hamdy, M.S. A rapid microwave synthesis of Ag2S nanoparticles and their photocatalytic performance under UV and visible light illumination for water treatment applications. Phys. E Low-Dimens. Syst. Nanostructures 2020, 121, 114060. [Google Scholar] [CrossRef]
  49. Cui, C.; Li, X.; Liu, J.; Hou, Y.; Zhao, Y.; Zhong, G. Synthesis and Functions of Ag2S Nanostructures. Nanoscale Res. Lett. 2015, 10, 431. [Google Scholar] [CrossRef] [PubMed]
  50. Khan, M.M. Metal oxide powder photocatalysts. In Multifunctional Photocatalytic Materials for Energy; Elsevier: Amsterdam, The Netherlands, 2018; pp. 5–18. ISBN 978-0-08-101977-1. [Google Scholar]
  51. Nalajala, N.; Patra, K.K.; Bharad, P.A.; Gopinath, C.S. Why the thin film form of a photocatalyst is better than the particulate form for direct solar-to-hydrogen conversion: A poor man’s approach. RSC Adv. 2019, 9, 6094–6100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Garusinghe, U.M.; Raghuwanshi, V.S.; Batchelor, W.; Garnier, G. Water Resistant Cellulose—Titanium Dioxide Composites for Photocatalysis. Sci. Rep. 2018, 8, 2306. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Katheresan, V.; Kansedo, J.; Lau, S.Y. Efficiency of various recent wastewater dye removal methods: A review. J. Environ. Chem. Eng. 2018, 6, 4676–4697. [Google Scholar] [CrossRef]
  54. Samsami, S.; Mohamadizaniani, M.; Sarrafzadeh, M.-H.; Rene, E.R.; Firoozbahr, M. Recent advances in the treatment of dye-containing wastewater from textile industries: Overview and perspectives. Process Saf. Environ. Prot. 2020, 143, 138–163. [Google Scholar] [CrossRef]
  55. Wandiga, S.O.; Masese, F.; Mbugua, S.N.; Macharia, J.W.; Otieno, M.A. Challenges and solutions to water problems in Africa. In Handbook of Water Purity and Quality; Elsevier: Amsterdam, The Netherlands, 2021; pp. 35–56. ISBN 978-0-12-821057-4. [Google Scholar]
  56. Zhu, C.; Zhang, X.; Zhang, Y.; Li, Y.; Wang, P.; Jia, Y.; Liu, J. Ultrasonic-Assisted Synthesis of CdS/Microcrystalline Cellulose Nanocomposites with Enhanced Visible-Light-Driven Photocatalytic Degradation of MB and the Corresponding Mechanism Study. Front. Chem. 2022, 10, 892680. [Google Scholar] [CrossRef]
  57. Liu, X.; Sayed, M.; Bie, C.; Cheng, B.; Hu, B.; Yu, J.; Zhang, L. Hollow CdS-based photocatalysts. J. Mater. 2021, 7, 419–439. [Google Scholar] [CrossRef]
  58. Nawrot, K.C.; Wawrzyńczyk, D.; Bezkrovnyi, O.; Kępiński, L.; Cichy, B.; Samoć, M.; Nyk, M. Functional CdS-Au Nanocomposite for Efficient Photocatalytic, Photosensitizing, and Two-Photon Applications. Nanomaterials 2020, 10, 715. [Google Scholar] [CrossRef]
  59. Haghighatzadeh, A.; Kiani, M.; Mazinani, B.; Dutta, J. Facile synthesis of ZnS–Ag2S core–shell nanospheres with enhanced nonlinear refraction. J. Mater. Sci. Mater. Electron. 2020, 31, 1283–1292. [Google Scholar] [CrossRef] [Green Version]
  60. Zhang, Y.; Zhou, W.; Jia, L.; Tan, X.; Chen, Y.; Huang, Q.; Shao, B.; Yu, T. Visible light driven hydrogen evolution using external and confined CdS: Effect of chitosan on carriers separation. Appl. Catal. B Environ. 2020, 277, 119152. [Google Scholar] [CrossRef]
  61. Nalajala, N.; Salgaonkar, K.N.; Chauhan, I.; Mekala, S.P.; Gopinath, C.S. Aqueous Methanol to Formaldehyde and Hydrogen on Pd/TiO2 by Photocatalysis in Direct Sunlight: Structure Dependent Activity of Nano-Pd and Atomic Pt-Coated Counterparts. ACS Appl. Energy Mater. 2021, 4, 13347–13360. [Google Scholar] [CrossRef]
  62. Thamer, N.H.; Hatem, O.A. Synthesis and Characterization of TiO2—Ag-Cellulose Nanocomposite and evaluation of photo catalytic degradation efficiency for Congo Red. IOP Conf. Ser. Earth Environ. Sci. 2022, 1029, 012026. [Google Scholar] [CrossRef]
  63. Mohamed, M.A.; Salleh, W.N.W.; Jaafar, J.; Ismail, A.F.; Mutalib, M.A.; Sani, N.A.A.; Asri, S.E.A.M.; Ong, C.S. Physicochemical characteristic of regenerated cellulose/N-doped TiO2 nanocomposite membrane fabricated from recycled newspaper with photocatalytic activity under UV and visible light irradiation. Chem. Eng. J. 2016, 284, 202–215. [Google Scholar] [CrossRef]
  64. Verbruggen, S.; Tytgat, T.; Passel, S.; Martens, J.; Lenaerts, S. Cost-effectiveness analysis to assess commercial TiO2 photocatalysts for acetaldehyde degradation in air. Chem. Pap. 2014, 68, 1273–1278. [Google Scholar] [CrossRef] [Green Version]
  65. Asha, R.C.; Vishnuganth, M.A.; Remya, N.; Selvaraju, N.; Kumar, M. Livestock Wastewater Treatment in Batch and Continuous Photocatalytic Systems: Performance and Economic Analyses. Water Air Soil Pollut. 2015, 226, 132. [Google Scholar] [CrossRef]
  66. Akbarzadeh, R.; Ibrahim, Q.; Jen, T.-C. Photocatalysis and Energy Cost Analysis of Vanadia Titania Thin Films Synthesis. In Proceedings of the 2019 International Conference on Power Generation Systems and Renewable Energy Technologies (PGSRET), Istanbul, Turkey, 26–27 August 2019; pp. 1–3. [Google Scholar]
  67. Mehrjouei, M.; Müller, S.; Möller, D. Catalytic and photocatalytic ozonation of tert-butyl alcohol in water by means of falling film reactor: Kinetic and cost–effectiveness study. Chem. Eng. J. 2014, 248, 184–190. [Google Scholar] [CrossRef]
  68. Ibrahim, N.A.; Salleh, K.M.; Fudholi, A.; Zakaria, S. Drying Regimes on Regenerated Cellulose Films Characteristics and Properties. Membranes 2022, 12, 445. [Google Scholar] [CrossRef]
  69. Yang, Q.; Fukuzumi, H.; Saito, T.; Isogai, A.; Zhang, L. Transparent Cellulose Films with High Gas Barrier Properties Fabricated from Aqueous Alkali/Urea Solutions. Biomacromolecules 2011, 12, 2766–2771. [Google Scholar] [CrossRef]
  70. Cazón, P.; Vázquez, M.; Velazquez, G. Cellulose-glycerol-polyvinyl alcohol composite films for food packaging: Evaluation of water adsorption, mechanical properties, light-barrier properties and transparency. Carbohydr. Polym. 2018, 195, 432–443. [Google Scholar] [CrossRef]
  71. Geng, H.; Yuan, Z.; Fan, Q.; Dai, X.; Zhao, Y.; Wang, Z.; Qin, M. Characterisation of cellulose films regenerated from acetone/water coagulants. Carbohydr. Polym. 2014, 102, 438–444. [Google Scholar] [CrossRef]
  72. Cazón, P.; Velazquez, G.; Vázquez, M. Novel composite films from regenerated cellulose-glycerol-polyvinyl alcohol: Mechanical and barrier properties. Food Hydrocoll. 2019, 89, 481–491. [Google Scholar] [CrossRef]
  73. Holi, A.M.; Zainal, Z.; Ayal, A.K.; Chang, S.-K.; Lim, H.N.; Talib, Z.A.; Yap, C.-C. Ag2S/ZnO Nanorods Composite Photoelectrode Prepared by Hydrothermal Method: Influence of Growth Temperature. Optik 2019, 184, 473–479. [Google Scholar] [CrossRef]
  74. Holi, A.M.; Zainal, Z.; Al-Zahrani, A.A.; Ayal, A.K.; Najm, A.S. Effect of Varying AgNO3 and CS(NH2)2 Concentrations on Performance of Ag2S/ZnO NRs/ITO Photoanode. Energies 2022, 15, 2950. [Google Scholar] [CrossRef]
  75. Haounati, R.; Alakhras, F.; Ouachtak, H.; Saleh, T.A.; Al-Mazaideh, G.; Alhajri, E.; Jada, A.; Hafid, N.; Addi, A.A. Synthesized of Zeolite@Ag2O Nanocomposite as Superb Stability Photocatalysis Toward Hazardous Rhodamine B Dye from Water. Arab. J. Sci. Eng. 2022, 207, 112157. [Google Scholar] [CrossRef]
  76. Hajizadeh, Z.; Taheri-Ledari, R.; Asl, F.R. Identification and analytical methods. In Heterogeneous Micro and Nanoscale Composites for the Catalysis of Organic Reactions; Elsevier: Amsterdam, The Netherlands, 2022; pp. 33–51. ISBN 978-0-12-824527-9. [Google Scholar]
  77. Medronho, B.; Romano, A.; Miguel, M.G.; Stigsson, L.; Lindman, B. Rationalizing cellulose (in) solubility: Reviewing basic physicochemical aspects and role of hydrophobic interactions. Cellulose 2012, 19, 581–587. [Google Scholar] [CrossRef]
  78. Xiong, B.; Zhao, P.; Cai, P.; Zhang, L.; Hu, K.; Cheng, G. NMR spectroscopic studies on the mechanism of cellulose dissolution in alkali solutions. Cellulose 2013, 20, 613–621. [Google Scholar] [CrossRef]
  79. Mohamed, M.A.; Salleh, W.N.W.; Jaafar, J.; Ismail, A.F.; Abd Mutalib, M.; Jamil, S.M. Incorporation of N-doped TiO2 nanorods in regenerated cellulose thin films fabricated from recycled newspaper as a green portable photocatalyst. Carbohydr. Polym. 2015, 133, 429–437. [Google Scholar] [CrossRef]
  80. Zeng, J.; Liu, S.; Cai, J.; Zhang, L. TiO2 Immobilized in Cellulose Matrix for Photocatalytic Degradation of Phenol under Weak UV Light Irradiation. J. Phys. Chem. C 2010, 114, 7806–7811. [Google Scholar] [CrossRef]
  81. Desai, J.A.; Adhikari, N.; Kaul, A.B. Chemical exfoliation efficacy of semiconducting WS2 and its use in an additively manufactured heterostructure graphene–WS2–graphene photodiode. RSC Adv. 2019, 9, 25805–25816. [Google Scholar] [CrossRef] [Green Version]
  82. Ma, Y.; Zhao, Z.; Xian, Y.; Wan, H.; Ye, Y.; Chen, L.; Zhou, H.; Chen, J. Highly Dispersed Ag2S Nanoparticles: In Situ Synthesis, Size Control, and Modification to Mechanical and Tribological Properties towards Nanocomposite Coatings. Nanomaterials 2019, 9, 1308. [Google Scholar] [CrossRef] [Green Version]
  83. Wang, Q.; Cai, J.; Zhang, L. In situ synthesis of Ag3PO4/cellulose nanocomposites with photocatalytic activities under sunlight. Cellulose 2014, 21, 3371–3382. [Google Scholar] [CrossRef]
  84. Wu, J.; Zhao, N.; Zhang, X.; Xu, J. Cellulose/silver nanoparticles composite microspheres: Eco-friendly synthesis and catalytic application. Cellulose 2012, 19, 1239–1249. [Google Scholar] [CrossRef]
  85. Rouhi, M.; Razavi, S.H.; Mousavi, S.M. Optimization of crosslinked poly (vinyl alcohol) nanocomposite films for mechanical properties. Mater. Sci. Eng. C 2017, 71, 1052–1063. [Google Scholar] [CrossRef] [PubMed]
  86. Abderrahim, B.; Abderrahman, E.; Aqil, M. Ouassini Krim Kinetic Thermal Degradation of Cellulose, Polybutylene Succinate and a Green Composite: Comparative Study. Sci. Educ. Publ. 2015, 3, 95–110. [Google Scholar] [CrossRef]
  87. Dutta, D.; Hazarika, R.; Dutta, P.D.; Goswami, T.; Sengupta, P.; Dutta, D.K. Synthesis of Ag–Ag2S Janus nanoparticles supported on an environmentally benign cellulose template and their catalytic applications. RSC Adv. 2016, 6, 85173–85181. [Google Scholar] [CrossRef]
  88. Fujii, Y.; Imagawa, K.; Omura, T.; Suzuki, T.; Minami, H. Preparation of Cellulose/Silver Composite Particles Having a Recyclable Catalytic Property. ACS Omega 2020, 5, 1919–1926. [Google Scholar] [CrossRef] [Green Version]
  89. Kumar, P.; Agnihotri, R.; Wasewar, K.L.; Uslu, H.; Yoo, C. Status of adsorptive removal of dye from textile industry effluent. Desalination Water Treat. 2012, 50, 226–244. [Google Scholar] [CrossRef]
  90. Tan, K.B.; Abdullah, A.Z.; Horri, B.A.; Salamatinia, B. Adsorption Mechanism of Microcrystalline Cellulose as Green Adsorbent for the Removal of Cationic Methylene Blue Dye. J. Chem. Soc. Pak. 2016, 38, 651–664. [Google Scholar]
  91. Manna, S.; Roy, D.; Saha, P.; Gopakumar, D.; Thomas, S. Rapid methylene blue adsorption using modified lignocellulosic materials. Process Saf. Environ. Prot. 2017, 107, 346–356. [Google Scholar] [CrossRef]
  92. Zamiri, R.; Abbastabar Ahangar, H.; Zakaria, A.; Zamiri, G.; Shabani, M.; Singh, B.; Ferreira, J.M.F. The structural and optical constants of Ag2S semiconductor nanostructure in the Far-Infrared. Chem. Cent. J. 2015, 9, 28. [Google Scholar] [CrossRef] [Green Version]
  93. Andrade-Guel, M.; Díaz-Jiménez, L.; Cortés-Hernández, D.; Cabello-Alvarado, C.; Ávila-Orta, C.; Bartolo-Pérez, P.; Gamero-Melo, P. Microwave assisted sol–gel synthesis of titanium dioxide using hydrochloric and acetic acid as catalysts. Boletín De La Soc. Española De Cerámica Y Vidr. 2019, 58, 171–177. [Google Scholar] [CrossRef]
  94. Isac, L.; Cazan, C.; Enesca, A.; Andronic, L. Copper Sulfide Based Heterojunctions as Photocatalysts for Dyes Photodegradation. Front. Chem. 2019, 7, 694. [Google Scholar] [CrossRef] [PubMed]
  95. Fan, J.; Yu, D.; Wang, W.; Liu, B. The self-assembly and formation mechanism of regenerated cellulose films for photocatalytic degradation of C.I. Reactive Blue 19. Cellulose 2019, 26, 3955–3972. [Google Scholar] [CrossRef]
  96. Fan, G.-Z.; Wang, Y.-X.; Song, G.-S.; Yan, J.-T.; Li, J.-F. Preparation of microcrystalline cellulose from rice straw under microwave irradiation. J. Appl. Polym. Sci. 2017, 134, 44901. [Google Scholar] [CrossRef]
  97. Ahmaruzzaman, M. Biochar based nanocomposites for photocatalytic degradation of emerging organic pollutants from water and wastewater. Mater. Res. Bull. 2021, 140, 111262. [Google Scholar] [CrossRef]
  98. Tavker, N.; Gaur, U.K.; Sharma, M. Agro-waste extracted cellulose supported silver phosphate nanostructures as a green photocatalyst for improved photodegradation of RhB dye and industrial fertilizer effluents. Nanoscale Adv. 2020, 2, 2870–2884. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of sequence process to synthesize the samples.
Figure 1. Schematic illustration of sequence process to synthesize the samples.
Materials 16 00437 g001
Figure 2. Illustrated scheme of the rarefaction process of commercial semiconductor Ag2S.
Figure 2. Illustrated scheme of the rarefaction process of commercial semiconductor Ag2S.
Materials 16 00437 g002
Figure 3. Prepared cellulose films: (a) CF/comAg2S1, (b) CF/comAg2S2, (c) CF/comAg2S3, (d) CF/syntAg2S1, (e) CF/syntAg2S2, and (f) CF/syntAg2S3.
Figure 3. Prepared cellulose films: (a) CF/comAg2S1, (b) CF/comAg2S2, (c) CF/comAg2S3, (d) CF/syntAg2S1, (e) CF/syntAg2S2, and (f) CF/syntAg2S3.
Materials 16 00437 g003
Figure 4. X-ray diffraction spectrum of synthesized Ag2S.
Figure 4. X-ray diffraction spectrum of synthesized Ag2S.
Materials 16 00437 g004
Figure 5. X-ray diffraction spectra of: (a) database of Ag2S (ICDD No. 00-009-0422), (b) database of cellulose (ICDD No. 00-050-2241), (c) CF/syntAg2S1, (d) CF/syntAg2S2, and (e) CF/syntAg2S3.
Figure 5. X-ray diffraction spectra of: (a) database of Ag2S (ICDD No. 00-009-0422), (b) database of cellulose (ICDD No. 00-050-2241), (c) CF/syntAg2S1, (d) CF/syntAg2S2, and (e) CF/syntAg2S3.
Materials 16 00437 g005
Figure 6. ATR—FTIR spectra of: (a,g) cellulose, (b) CF/comAg2S1, (c) CF/comAg2S2, (d) CF/comAg2S3, (e,f) inset CF/comAg2S graph, (h) CF/syntAg2S1, (i) CF/syntAg2S2, (j) CF/syntAg2S3, and (k,l) inset CF/syntAg2S graph.
Figure 6. ATR—FTIR spectra of: (a,g) cellulose, (b) CF/comAg2S1, (c) CF/comAg2S2, (d) CF/comAg2S3, (e,f) inset CF/comAg2S graph, (h) CF/syntAg2S1, (i) CF/syntAg2S2, (j) CF/syntAg2S3, and (k,l) inset CF/syntAg2S graph.
Materials 16 00437 g006
Figure 7. Predicted mechanism reaction between methylene blue and cellulose.
Figure 7. Predicted mechanism reaction between methylene blue and cellulose.
Materials 16 00437 g007
Figure 8. (a) Degradation ratio of MB with Ag2S commercial as a catalyst; (b) The concentration of MB after photocatalytic activity using CF/comAg2S samples; (c) Degradation ratio of MB with synthesized Ag2S as a catalyst; (d) The concentration of MB after photocatalytic activity employing CF/syntAg2S samples.
Figure 8. (a) Degradation ratio of MB with Ag2S commercial as a catalyst; (b) The concentration of MB after photocatalytic activity using CF/comAg2S samples; (c) Degradation ratio of MB with synthesized Ag2S as a catalyst; (d) The concentration of MB after photocatalytic activity employing CF/syntAg2S samples.
Materials 16 00437 g008
Table 1. Comparison of the cost-effectiveness assessment of several photocatalysts.
Table 1. Comparison of the cost-effectiveness assessment of several photocatalysts.
Type of PhotocatalystTypes of WastewaterEnergy SourcesCatalyst Dosage (g/L)Maximum Photodegradation EfficiencyTime to Obtain Maximum Photodegradation (min.)Total Power Consumed (kWh)Operational Cost (USD/kg) bRefs.
Rutile TiO2AcetaldehydeUV lamps
(56 W)
5.0180100not available127.13[64]
GAC a–TiO2Livestock wastewaterUV lamps
(56 W)
610060.5240.68[65]
V2O5–TiO2Methylene blueUV lamps4.759212018.7205[66]
TiO2/O3Tert-butyl alcoholUVA lamps
(15W)
not available751037not available[67]
Cellulose/Ag2SMethylene blueSolar energy710012000Current study
a Granular activated carbon. b Energy costs (USD) per kg of wastewater removal.
Table 2. Sample code and composition of refracted commercial Ag2S.
Table 2. Sample code and composition of refracted commercial Ag2S.
Sample CodeAg2S Mass (Gram)Volume of Distilled Water (mL)
CF/comAg2S10.0756.0
CF/comAg2S20.1006.0
CF/comAg2S30.3006.0
Table 4. Phase identification and crystallinity of the CF/syntAg2S sample.
Table 4. Phase identification and crystallinity of the CF/syntAg2S sample.
Sample
Code
Index Miller CelluloseIndex Miller Ag2SCrystallite Size (Å)Degree of Crystallinity (%)
( 002 ) ( 3 ¯ 11 ) ( 220 ) ( 022 )
2θ Stnd.2θ Ob.2θ Stnd.2θ Ob.2θ Stnd.2θ Ob.2θ Stnd.2θ Ob.
CF/syntAg2S122.7822.7631.56-34.5034.4436.9237.1822147.93
CF/syntAg2S222.7822.3031.5631.1834.5033.9936.9236.4718876.76
CF/syntAg2S322.7822.7831.5631.4834.5034.4936.9236.9917437.75
Table 5. Comparison of the photocatalytic efficiency of the prepared samples.
Table 5. Comparison of the photocatalytic efficiency of the prepared samples.
Exposure Time (Min)Degradation Ratio (%)
Non-ExposureAg2S PowderPristine CelluloseWith Exposure
CF/comAg2SCF/syntAg2SCF/comAg2S2CF/syntAg2S1
0000000
53.590.6619.7844.1430.9931.49
105.971.6023.3753.0940.7744.42
157.843.0926.8563.4352.1071.71
2010.444.3130.2271.1058.5683.54
2512.215.6432.8275.9164.9286.69
3014.426.4634.5381.6669.7889.45
6022.7612.4339.3990.8898.9593.98
9027.4016.2445.4194.9299.0197.29
12029.7821.6048.1896.1910098.56
15031.3325.0352.9899.5610099.78
18031.9328.1854.9299.5610099.78
21032.3230.1759.2899.83100100
24032.4933.8162.4999.97100100
27032.7134.7566.74100100100
30033.0435.0875.03100100100
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

Mubarokah, Z.R.; Mahmed, N.; Norizan, M.N.; Mohamad, I.S.; Abdullah, M.M.A.B.; Błoch, K.; Nabiałek, M.; Baltatu, M.S.; Sandu, A.V.; Vizureanu, P. Near-Infrared (NIR) Silver Sulfide (Ag2S) Semiconductor Photocatalyst Film for Degradation of Methylene Blue Solution. Materials 2023, 16, 437. https://0-doi-org.brum.beds.ac.uk/10.3390/ma16010437

AMA Style

Mubarokah ZR, Mahmed N, Norizan MN, Mohamad IS, Abdullah MMAB, Błoch K, Nabiałek M, Baltatu MS, Sandu AV, Vizureanu P. Near-Infrared (NIR) Silver Sulfide (Ag2S) Semiconductor Photocatalyst Film for Degradation of Methylene Blue Solution. Materials. 2023; 16(1):437. https://0-doi-org.brum.beds.ac.uk/10.3390/ma16010437

Chicago/Turabian Style

Mubarokah, Zahrah Ramadlan, Norsuria Mahmed, Mohd Natashah Norizan, Ili Salwani Mohamad, Mohd Mustafa Al Bakri Abdullah, Katarzyna Błoch, Marcin Nabiałek, Madalina Simona Baltatu, Andrei Victor Sandu, and Petrica Vizureanu. 2023. "Near-Infrared (NIR) Silver Sulfide (Ag2S) Semiconductor Photocatalyst Film for Degradation of Methylene Blue Solution" Materials 16, no. 1: 437. https://0-doi-org.brum.beds.ac.uk/10.3390/ma16010437

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