Next Article in Journal
Study on Imagery Modeling of Electric Recliner Chair: Based on Combined GRA and Kansei Engineering
Previous Article in Journal
Predicting Temperature and Humidity in Roadway with Water Trickling Using Principal Component Analysis-Long Short-Term Memory-Genetic Algorithm Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

One-Pot Facile Synthesis of ZrO2-CdWO4: A Novel Nanocomposite for Hydrogen Production via Photocatalytic Water Splitting

by
Ahmed Hussain Jawhari
Department of Chemistry, College of Science, Jazan University, P.O. Box 114, Jazan 45142, Saudi Arabia
Appl. Sci. 2023, 13(24), 13344; https://doi.org/10.3390/app132413344
Submission received: 22 October 2023 / Revised: 10 December 2023 / Accepted: 15 December 2023 / Published: 18 December 2023
(This article belongs to the Section Nanotechnology and Applied Nanosciences)

Abstract

:
ZrO2-based nanocomposites are highly versatile materials with huge potential for photocatalysis. In this study, ZrO2-CdWO4 nanocomposites (NC) were prepared via the green route using aqueous Brassica rapa leaf extract, and its photocatalytic water-splitting application was evaluated. Brassica rapa leaf extract acts as a reducing agent and abundant phytochemicals are adsorbed onto the nanoparticle surfaces, improving the properties of ZrO2-CdWO4 nanocomposites. As-prepared samples were characterized by using various spectroscopic and microscopic techniques. The energy of the direct band gap (Eg) of ZrO2-CdWO4 was determined as 2.66 eV. FTIR analysis revealed the various functional groups present in the prepared material. XRD analysis showed that the average crystallite size of ZrO2 and CdWO4 in ZrO2-CdWO4 was approximately 8 nm and 26 nm, respectively. SEM and TEM images suggested ZrO2 deposition over CdWO4 nanorods, which increases the roughness of the surface. The prepared sample was also suggested to be porous. BET surface area, pore volume, and half pore width of ZrO2-CdWO4 were estimated to be 19.6 m2/g. 0.0254 cc/g, and 9.457 Å, respectively. PL analysis suggested the conjugation between the ZrO2 and CdWO4 by lowering the PL graph on ZrO2 deposition over CdWO4. The valence and conduction band edge positions were also determined for ZrO2-CdWO4. These band positions suggested the formation of a type I heterojunction between ZrO2 and CdWO4. ZrO2-CdWO4 was used as a photocatalyst for hydrogen production via water splitting. Water-splitting results confirmed the ability of the ZrO2-CdWO4 system for enhanced hydrogen production. The effect of various parameters such as photocatalyst amount, reaction time, temperature, water pH, and concentration of sacrificial agent was also optimized. The results suggested that 250 mg of ZrO2-CdWO4 could produce 1574 µmol/g after 5 h at 27 °C, pH 7, using 30 vol. % of methanol. ZrO2-CdWO4 was reused for up to seven cycles with a high hydrogen production efficiency. This may prove to be useful research on the use of heterojunction materials for photocatalytic hydrogen production.

1. Introduction

Increasing populations and modernization, human dependence on energy is increasing continuously, and this is becoming a burden on the global environmental system, as most of the energy is generated from fossil fuels [1]. Various nations have been raising issues of pollution and energy crisis, which is why alternative means of developing renewable and clean energy systems are constantly being explored [1,2].
Hydrogen (H2) gas, which acts as a secondary energy source, holds a significant position in this aspect [3]. H2 has advantages over other energy resources, because of its lightweight, high calorific value, and pollution-free clean combustion [4]. Since hydrogen gas can be a green source of energy, new and easier methods of production are constantly being discovered; however, currently with the existing technologies, the H2 production method is quite expensive [5].
Various methods have been adopted for H2 production such as petroleum-based petrochemical processes [6], steam reforming of methane or propane [7], thermal decomposition of water by nuclear energy [8], solid electrolysis [9], biomass gasification [10] and splitting of water [11].
H2 production via water electrolysis or water splitting holds a clear edge over other contemporary methods since it does not emit secondary pollution [9,11,12]. Production of H2 by the water-splitting process on the surface of the photocatalyst in the presence of light is being recognized as the most advanced and resourceful technology [12]. The most commonly used materials to streamline this process are nanosized semiconductors that act as photocatalysts and prove to be promising candidates due to their unique physical and optical properties [13]. Despite having many properties, the drawback of these nanomaterials is their inability to work better in the presence of visible or natural solar light [14,15].
Generally, a photocatalyst directly absorbs a major portion of sunlight for efficient H2 production. This requires a suitable band gap. A larger-band-gap photocatalyst absorbs sunlight mainly in the UV region. For example, TiO2 was determined to be a stable and efficient photocatalyst for producing H2 but its larger band gap (>3 eV) limits light absorption [14,15]. It is known that photocatalysts produce electrons and holes in the presence of light, which are responsible for the photolysis process; however, due to the large band gap in the photocatalyst or due to reduced charge separation or high electron-hole recombination, their efficiency is nullified [15]. These shortcomings ultimately reduce the yield potential and photocatalytic properties of such materials under natural solar light. This may not prove useful considering H2 production requirements [15].
In view of the above deficiency, many upgradations have been made, the most useful is the development of heterojunction composite materials [16,17,18]. Heterojunction composites have better optical properties due to new crystallographic phases than the original oxides, the creation of defect energy levels in the band gap region, and changes in the surface characteristics of the individual oxides due to the formation of new sites at the interface of the components, increasing the stability of the photoactive crystalline phase [16,17,18].
From a photocatalytic point of view, coupling two materials leads to the formation of a heterojunction with different positions of energy bands. Previous studies have suggested that there can be three types of heterojunctions depending on the position of the band edge such as straddling gap (type I), staggered gap (type II) or broken gap (type III) [19]. Currently, a type I heterojunction is considered attractive and advanced, responsible for many benefits such as a narrow energy band gap, prevention of recombination of photogenerated electron holes, and charge separation [19]. This further allows them to capture solar/visible light efficiently and smooth and long-term circulation of photogenerated electrons and holes over the surface, and ultimately increases the photocatalytic efficiency of materials [16,17,18].
To practically verify this subject and to increase the importance of the H2 production method, this study was conducted, in which two bare semiconductors (CdWO4 and ZrO2) with a large energy band gap were coupled to the developed heterojunction.
Notably, monoclinic wolframite CdWO4 is known for good photocatalytic performance due to its layered structure, self-trapped excitons, optical, chemical, and structural properties, and thermal stability. But modified (doped or coupled) CdWO4 may have better photocatalytic performance than bare CdWO4 due to a wider band gap in bare CdWO4 (~3.5–4.0 eV) [16,17,18]. For instance, Yogesh and others reported nanostructured CdS-sensitized CdWO4 nanorods that exhibited highly efficient H2 generation than bare CdWO4 [20].
In addition to CdWO4, ZrO2, a n-type semiconductor, received much attention due to its substantial amount of oxygen vacancies, strong ion exchangeability, and physicochemical stability. However, it is a broad-band-gap semiconductor (5.5 eV) that requires UV-C light (<280 nm) for electron-hole generation [21]. One method to overcome this problem is by doping ZrO2 with different transition metal ions or coupling with other metal oxides, creating composites, bringing about heterojunction formation, and incorporating co-catalysts. For instance, a binder-free WS2/ZrO2 hybrid nanocomposite (NC) was reported as a photocatalyst for H2 production via water splitting [22]. The coupling of ZrO2 with other semiconductors like SiO2, TiO2, ZnO, WO3, and NiO have also been reported [23,24].
Considering the advantages of both these materials as bare photocatalysts, along with their composites, this study assessed the photocatalytic activity of the heterojunction material developed by coupling ZrO2 and CdWO4.
Many methods like coprecipitation [25], hydrothermal [26], sol–gel [27], microwave [28], and bioprecipitation [16] have been used to develop heterojunction material. However, among all these methods, bioprecipitation is the cheapest, easiest, sustainable, and green [17,21], hence the bioprecipitation method has been chosen as the synthesis method in the present study.
In the bioprecipitation method, plant extracts were used, which act as reducing, capping, and stabilizing agents [16,17,18,29,30,31]. The plant extracts are usually a rich source of phytochemicals that adsorb onto the material surface during preparation and enhance the properties of materials such as a stabilized particle, reduced particle size, enhanced surface functional sites, and large surface area [16,17,18,21,32]. The presence of functional groups on the surface of the material also enhances the photocatalytic activity of the material by easily absorbing sunlight [16,17,18].
Keeping all these features in mind, the green method was adopted for the present (ZrO2-CdWO4) preparation for which leaf extracts of Brassica rapa plants were employed. Brassica rapa leaf extracts contain abundant phytochemicals like kaempferol, isorhamnetin, quercetin glycosides, phenylpropanoid derivatives, indole alkaloids, sterol glucosides, and isorhamnetin glycosylated acylated with different hydroxycinnamic [17,18]. This plant extract has also been used for a number of nanoparticles (NPs)/(NCs) [16,17,18].
The objective of this study is to prepare a ZrO2-CdWO4 heterojunction (type I) NC using a Brassica rapa leaf extract. In this study, water-splitting photocatalytic activity has been explored by developing a heterojunction between ZrO2 and CdWO4.
It is worth noting here that Cd is a toxic element, but its toxicity can be greatly reduced in the form of NC such as CdWO4. Notably, ZrO2 is a ceramic material that may further provide stability and strength to the composed NC, due to which the toxicity of Cd can be completely reduced. That is why ZrO2 is combined with CdWO4 in this study (as a type I heterojunction) to demonstrate the novelty of this material.

2. Materials and Methods

2.1. Materials

Leaves of Brassica rapa were collected from the general supermarket store, Jazan, KSA. In terms of precursor chemicals, cadmium iodide, A.R (Mol. wt. = 366.21), was procured from HI Media laboratory limited, Mumbai, India; purified sodium tungstate dihydrate (Mol. wt. = 329.86 g/mol) and zirconium dioxide 97% (Mol wt. = 123.22 g/mol) were purchased from Merck Ltd., Germany. All these chemicals were of analytical grade and used without any kind of purification.

2.2. Methods

Among various preparation methods, the bioprecipitation method is considered to be a sustainable process. Plant extracts may be used in the bioprecipitation method [17]. Plant extracts contain phytochemicals, which act as reducing agents, capping agents, and functionalizing agents in NP formation [16,17]. Plant extracts are cheap, natural, and zero toxicity reagents that are easily available. As such, the NPs made through this process are less toxic, cheaper, and environmentally benign. That is why this process was given priority in this study and it was used in preparing the current photocatalyst.

2.2.1. Preparation of Brassica rapa Leaf Extract

The Brassica rapa leaves were washed several times with distilled water, and then ground to a thick paste using a mixer grinder (Blender, Model 24CB10C, Waring Commercial, Stamford, CT, USA). An amount of 10 g of the fresh leaf paste was dispersed in 100 mL of distilled water and heated at 70 °C for 40 min. The prepared extract was then filtered by Whatman paper No. 1, and the obtained filtrate was placed into the refrigerator for further use.

2.2.2. Synthesis of the ZrO2-CdWO4 NC

A previous study by Fatima et al. [17] was followed to synthesize a ZrO2-CdWO4 NC. The ZrO2-CdWO4 NC was prepared by taking 150 mL of 0.2 M (9.8958 g) sodium tungstate dihydrate solution and 50 mL of 0.2 M (3.6621 g) cadmium iodide solutions separately, which were later brought together, homogeneously. Thereafter, 30 mL of Brassica rapa leaf extract was added to the above solution and mixed thoroughly. A volume of 50 mL was taken out separately from a stock solution and 50 mL of 0.02 M (0.1232 g) of zirconium dioxide (ZrO2) solution (which had been previously sonicated) was added under continuous stirring on a magnetic stirrer at 650 rpm to obtain the dispersed solution. The solution mixture was further stirred for 30 min, then heated at 100 °C for 10 h. The sample was then separated out by centrifuging it for 15 min at 8000 rpm, washed several times with distilled water, followed by ethanol, and then dried in an oven at 80 °C for 6 h. Thereafter, the dried sample was calcined in a muffle furnace at 400 °C for 2 h [17] as shown in Scheme 1. The yield of the sample was estimated to be approximately 2.5 g.

2.3. Characterization

The UV–Vis spectrum was recorded on a UV–Vis–NIR spectrophotometer (UV-3600-DRS spectrophotometer, Shimadzu Corporation, Kyoto, Japan). Band gap energy was calculated from a Tauc plot from the data obtained from the UV–Visible absorption spectrum of the solid sample using a Shimadzu, Kyoto, Japan UV/Vis-3600, spectrophotometer in the range of 200–800 nm. The FTIR spectrum was recorded with a Perkin Elmer-USA in the wavenumber range of 4000–400 cm−1. The powder X–ray diffraction patterns were recorded on Philips PW-3710 diffractometer using Cu-Kα radiation (λ = 1.54 Å), a Cu-filter, 35 kV generator voltage, 30 mA current, and a proportional counter detector. The SEM micrographs of the samples were obtained using a scanning electron microscope (Model: JEOL/EO-JSM-6510, Tokyo, Japan). Transmission electron microscopy (TEM) images were obtained using a JEOL HRTEM, JEM-2100F-Tokyo, Japan transmission electron microscope and were analyzed at 200 kV. Diluted samples were placed on copper grids and analyzed at 200 kV of tension. The N2 physisorption analysis of the sample was performed using a Micromeretics Tristar 2000 instrument. The sample was degassed under helium gas at 300 °C for 2 h prior to analysis. The photoluminescence (PL) spectra of the samples were measured using a fluorescence spectrophotometer (F-4500, Hitachi, Tokyo, Japan).

2.4. H2 Production Experiment

The photocatalytic H2 production experiments were performed in a photochemical reactor connected with a GC instrument. The reactor was a tightly closed glass reactor (total volume of 1000 mL) equipped with a cold-water circulation pump to reduce the heat generated during the photocatalytic reaction process. For each photocatalytic experiment, 100 mg of the synthesized catalyst powder was dispersed in a mixed solution of 400 mL DI water and sacrificial reagent (20 vol. % of methanol). Then, the solution was transferred into the reactor and purged with argon gas for 30 min in order to flush all the atmospheric gases. The empty headspace was kept constant at 500 mL for all reactions. When all gases were completely purged, the reactor was irradiated using a 300 W Xenon arc lamp (provided a flux of ~125 mW cm−2 for 2 h) without any UV cutoff filter. The amount of H2 from the reaction was determined by flowing argon as a gas carrier through the reactor in order to carry the generated gases for gas chromatography analysis (GC-1000). The measurement was performed every 30 min at a 5 h span of photoreaction.

3. Results

3.1. UV–Visible Spectroscopy

For observing the formation of NPs, optical absorption measurements are the first step. Figure 1a shows the UV-DRS spectra of ZrO2, CdWO4, and ZrO2-CdWO4. Pure ZrO2 and pure CdWO4 have an absorption edge at 228 nm nanoparticles and 255 nm nanorods, respectively, while the ZrO2-CdWO4 NC occurs at approximately 240 nm. The absorption band corresponding to these nanorods gets blue-shifted on ZrO2 doping. The shift of the absorption band towards a shorter wavelength indicates a decrease in particle diameter and the narrowing of the band gap. The narrowing of the band gap could also be due to many interactions on the ZrO2 deposition and the making of a heterojunction between ZrO2 and CdWO4. Here, the optical absorption changes mainly owing to the coupled semiconductor system following the deposition of ZrO2 on CdWO4 nanorods.
The UV absorption spectrum (Figure 1a) of the sample was recorded for measuring the direct and indirect bandgap of the ZrO2, CdWO4, and the ZrO2-CdWO4 NC using the Tauc relation, Equation (1) [16,17,18]:
α h v 1 / n = A ( h ν E g )
where A is a proportionality constant, hv is the incident photon energy, α is the molar extinction coefficient, Eg is the bandgap of the sample, and n is the exponent to evaluate the type of electronic transition causing the absorption, which has values of ½ and 2 for direct and indirect band gap transitions, respectively. The bandgap was determined from the Tauc plot [16,17,18,33]. In the Tauc plot, the intercept of the linear portion of the (ɑhν)2 vs. on the x-axis is shown in Figure 1b and (ɑhν)1/2 vs. hν is shown in Figure 1c. The bandgap values for ZrO2, CdWO4 and the ZrO2-CdWO4 NC are given in Table 1. The decrease in the band gap suggests the successful coupling of ZrO2 to CdWO4.
The band gap of the present ZrO2-CdWO4 NC was determined as narrower than that of previously reported CdWO4-based NCs such as ZnO-CdWO4, CuO-CdWO4, and ZnO-CuO-CdWO4, as shown in Table 1.
Apart from the preparation method, the band gap also depends on many other factors, such as plant extract, calcination temperature, and coupled semiconductors.
Using estimated band gaps (Eg) and electronegativity (X) of conjugated semiconductors (ZrO2 and CdWO4), the bands (valence band, VB) and the conduction band, CB) edge position (EVB and ECB) of the current photocatalyst were determined by applying Equation (2) [16], the result of which is given in Table 2.
EVB = X − Ee + 0.5Eg, ECB = EVB − Eg
Here, Ee is the energy of free electrons, which is ~4.5 eV.
Using the edge positions of these bands (VB and CB), the heterojunction design was proposed as type I [19]. Type I heterojunctions aid charge separation and prevention of electron-hole recombination.

3.2. FT-IR Analysis

The infrared spectrum of the ZrO2-CdWO4 NC (Figure 2) showed a broad-spectrum peak at approximately 3429.5 cm−1, contributing to the stretching vibrations of the O–H bond [35]. The peaks that appeared at approximately 2919 and 2854.1 cm−1 are for the stretching vibration of the C-H bond coming from the Brassica rapa leaf extract [36]. A sharp peak at 1633.6 cm−1 was assigned for the bending vibration of Zr-OH and adsorbed moisture content [33]. The absorption peaks at 1020.7 and 829.3 cm−1 in the sample are assigned for the –OH group and for the vibration of the Zr-O bond, confirming that ZrO2 is present in the sample [37]. The sharp peaks at 890.0 cm−1, 717.6 cm−1, 587.96 cm−1, 412.3 cm−1 and 412 cm−1 correspond to Cd–O stretching vibrations of Cd–O–W, stretching vibrations of the W-O-W bond in the W O 4 2 group and in-plane deformation of the W O 4 2 group [16,17,18,37,38,39,40].
The FT-IR spectrum of the current ZrO2-CdWO4 NC can be verified with the FT-IR spectrum of the previously reported NC based on CdWO4 prepared using plant extracts (Table 3). The current nanocomposite showed an almost identical FTIR spectrum to the previous one, confirming the validation of the present analysis. The presence of typical peaks corresponding to the plant extracts (Brassica rapa leaf) as well as the nanomaterial (ZrO2-CdWO4), which is consistent with the previous study, suggest the preparation of the ZrO2-CdWO4 NC that was chemically functionalized with phytochemicals from the plant extract. These functional groups can aid light absorption as suggested in the previous study.

3.3. XRD Analysis

Figure 3 shows the XRD pattern of the ZrO2-CdWO4 NC within the 2θ range of 10–80°. XRD analysis was conducted using the Crystal Impact Match program. Results obtained confirmed that CdWO4 has monoclinic symmetry with cell dimensions of a = 4.9750 Å, b = 5.7650 Å, c = 5.0370 Å and β = 91.840° (space group P 1 2/c 1) and ZrO2 has tetragonal symmetry with cell dimensions a= 3.6120 Å and c= 5.2120 Å (space group P 42/n m c). The 2θ values at 23.26°, 28.83°, 35.49°, 47.44°, and 51.93° corresponded to (110), (111), (002), (022), and (221) lattice planes, indicating a wolframite structure of CdWO4 (entry number 969013916] [17,39,40]. The 2θ values at 29.10°, 50.55°, 60.21°, and 63.14° correspond to the tetragonal phase of ZrO2 (entry number 961526428) [41,42]. The amount of CdWO4 and ZrO2 in samples is 90.2% and 9.8%, respectively. The XRD pattern of ZrO2-CdWO4 can also be verified from the XRD spectrum of the previously reported CdWO4-based NC such as ZnO-CdWO4, CuO-CdWO4, and CuO-ZnO-CdWO4. The X-ray diffraction (XRD) pattern of the ZrO2-CdWO4 nanocomposite showed distinct peaks that corresponded to the crystal facets of the monoclinic phase of ZrO2, specifically (111), (200), (220), (311), and (400). These results were consistent with the crystal structure identified by JCPDS no. 00–037-1484. Furthermore, the ZrO2-CdWO4 NC exhibited diffraction peaks that were identified as the crystal facets (110), (−112), (211), (−203), (311), and (−231) of the tetragonal phase of ZrO2 (JCPDS # 00–017-0923). The X-ray diffraction (XRD) pattern of ZrO2-CdWO4 nanocrystals indicates the presence of ZrO2 in both the monoclinic and tetragonal phases.
The crystallite size was calculated using Scherrer’s Equation (3) [17]:
D = K λ / β   C o s   θ
where D is the crystallite size, constant K = 0.94, λ is the X-ray wavelength of Cu Kα radiation (1.54 A), and β is the full width at half maximum (FWHM). The average crystallite size corresponding to the highest peaks of ZrO2 and CdWO4 was determined as ~8 nm and ~26 nm, respectively.

3.4. SEM Analysis

Particles were in an aggregated state. Aggregation of particles occurred due to van der Waal forces of attraction between them (Figure 4a). The average diameter of the ZrO2-CdWO4 NC increased due to the presence of ZrO2. The EDAX pattern of the ZrO2-CdWO4 NC is shown. The composition of the sample shows the presence of the zirconium, oxygen, cadmium, and tungsten elements. The contents of Zr, O, Cd, and W are 5.74 wt%, 21.99 wt%, 27.65 wt%, and 44.62 wt% (Figure 4b). Also, the spectral peaks reveal the presence of Zr and O at 2.05 and 0.5Kev, respectively, which confirms the presence of Zr and O in CdWO4 nanorods.

3.5. TEM Analysis

The particle size, pattern, morphology, and distribution of crystallite were confirmed by transmission electron microscopy. TEM images of the biogenically synthesized ZrO2-CdWO4 NC at 400 °C of calcination are given in Figure 5. The presence of ZrO2 on the rod-shaped CdWO4 nanorods increases the roughness. Further, the particles were well crystallized without defects and self-aggregated. Self-aggregation is caused by the high generation rate of CdWO4 nanorods. Due to fewer electron density differences between the particles, the ZrO2 particles and CdWO4 nanorods appeared to be slightly different in color. The particle size diameter is nearly 16–30 nm. The nanorods CdWO4 in NC can also be verified from the TEM images of previously reported CdWO4-based NCs such as ZnO-CdWO4, CuO-CdWO4, and CuO-ZnO-CdWO4.

3.6. Nitrogen Adsorption–Desorption Isotherm Analysis

In the present study, nitrogen (N2) absorption and desorption analysis of ZrO2-CdWO4 were also studied, the result of which is seen in Figure 6. This analysis gives the information about the surface area and pore size of the prepared sample. Figure 6 reveals the type III adsorption isotherm, which gave a surface area of ZrO2-CdWO4 of 19.6 m2/g. The pore volume and half pore width of the prepared sample were estimated as 0.0254 cc/g and 9.457 Å, respectively.

3.7. Photoluminescence Analysis

To elucidate the photogenerated charge carrier separation at the interface of ZrO2-CdWO4, photoluminescence (PL) spectra were analyzed. The PL spectrum shows the comparative e-h+ recombination rate in the bare semiconductors and heterojunction [16]. Figure 7 shows the PL spectra of ZrO2 and CdWO4, and also the heterojunction (ZrO2-CdWO4). As can be seen in the PL spectra, the PL intensity was significantly reduced when the heterojunction was formed as compared to ZrO2 and CdWO4 [15,16]. The decrease in the PL intensity upon ZrO2 loading might be due to the decrease in the rate of e-h+ recombination upon loading ZrO2 on CdWO4 [15,16]. Therefore, the heterojunction photocatalyst shows even better photolytic activity than the virgin ZrO2 and CdWO4.

3.8. Photocatalytic Performance Results

As reported in previous studies, these catalysts exhibit better H2 production activity in the presence of a sacrificial agent than a virgin photocatalyst. It is well known that the amount of protons, i.e., H+ ions, for H2 production in the presence of a sacrificial agent increases. The sacrificial agent provides H+ ions, which are responsible for increased photocatalyst activity and water splitting. In this study, methanol was used as a sacrificial agent and it was found that this sacrificial agent increases the rate of H2 production. The increased H2 production rates with sacrificial agents can be easily seen in Figure 8a.
A systematic study of the photocatalytic activity of the ZrO2-CdWO4 NC was performed for H2 production via water splitting with the sacrificial agent (methanol). H2 production via water splitting was performed under various conditions such as different catalyst amounts, time, temperatures, and sacrificial agents.
The time effect has also been monitored for this study, and is shown in Figure 8a. Figure 8a confirms that H2 production increases as time increases. It can be seen from Figure 8b that the H2 production rate increases as the photocatalyst is added. The point is that the number of surface sites increases with the increase in photocatalyst concentration (facilitates visible light absorption), due to which the catalytic activity of the catalyst also increases. This is because the rate of H2 production showed a marked increase when increasing the concentration of the photocatalyst prepared in this study. As can be seen, the concentration of the photocatalyst ranged from 25 mg to 250 mg, and H2 production ranged from 165 µmol/g to 1574 µmol/g.
Figure 8c shows the temperature effect on water splitting. The increased temperature facilitates the easy excitation of electrons from VB to CB, charge separation, and prevention of e–h+ recombination, which is responsible for enhanced water splitting. In addition, the temperature increase reduces the viscosity of water, which is also responsible for water splitting. Different pH effects were also measured for H2 production efficiency, as shown in Figure 8d, and pH 7.0 was observed as the optimum pH for good H2 production.
The effect of MeOH (sacrificial agent) concentration on H2 production via water splitting was also observed. As shown in Figure 8e, H2 production initially increases with an increase in the concentration of MeOH, and the H2 production rate decreases. The reason behind the decrease in the H2 production rate might be the blocking of the surface site on the high loading of MeOH. The blocking of surface sites may eventually decrease the visible light absorption for photocatalytic activity.

3.9. The Mechanism Involved

A heterojunction photocatalyst was used in this study. According to previous studies, heterojunction photocatalysts can be arranged in a variety of ways depending on the band edge positions of their constituent semiconductors, including dominant straddling gap (type I), staggered gap (type II) or broken gap (type III) heterojunction photocatalysts [19]. The present heterojunction photocatalyst was made up of two separate semiconductor units (ZrO2 and CdWO4) whose band (VB and CB) positions are different to each other. According to their band edge positions, the ZrO2-CdWO4 semiconductor interface can be arranged as a straddling gap (type I) heterojunction. A Schottky barrier is formed due to different band positions or type I arrangement of the heterojunction (suggested by UV-DRS and PL spectra). Therefore, the possibility of migration of e to the catalyst surface is increased via charge separation, and prevention of e–h+ recombination, as shown in Scheme 2 [16,18].
When visible light falls on such a heterojunction photocatalyst, a valence band (VB) e can be excited from one of the units (either ZrO2 or CdWO4) and move to the conduction band (CB) of that unit, leading to the formation of h+ in VB. Due to the difference in the positions of the bands (VB and CB) in such units (ZrO2 and CdWO4) of the developed heterojunction, the possibility of migration of e and h+ from one unit to the other unit may increase. This way, the excited e and h+ do not go back and can rapidly and continuously transfer from one unit (ZrO2) to the other unit (CdWO4) of the semiconductor, thus causing charge separation, and also preventing charge recombination.
Due to this charge separation and prevention of charge recombination, the electronic charge becomes distributed over the entire surface of the heterojunction photocatalyst. In such a situation, the e and h+ reach the surface of the catalyst and react with water, split the water, and produce H2 gas. Because, in this study, a sacrificial agent such as MeOH has been used, the MeOH acts as a proton source and therefore increases the photocatalytic activity occurring in the presence of these extra protons. Based on the theory presented in the present study, the speculated mechanism has been shown through Scheme 2 [16]. Our group [43,44] has previously reported the use of Pt-Ag/Ag3PO4-WO3 NCs for photocatalytic H2 production from bioethanol.

3.10. The Reuasabilty Study

It is very important to assess the stability and life cycle of the photocatalyst for cheap and smooth production of H2 gas. In view of this, the photocatalyst used in this study was also used for several cycles. To investigate the reuse activity of the catalyst, an experiment was conducted with 250 mg of ZrO2-CdWO4 at room temperature (27 °C) and 30 vol% methanol. The catalyst used in each cycle is separated by centrifuging and then washed with deionized water and dried, which was reused. The results of this experiment are shown in Figure 9, which makes it clear that even after 7 cycles, the present catalyst showed excellent H2 production activity. The present catalyst retained 75% of its activity even after 7 cycles. The 25% drop in activity may be due to a reduction in catalyst weight and a reduction in surface activity. It is also worth noting here that this catalyst displayed almost no activity in the absence of a sacrificial agent.

3.11. Comparative Analysis

A comparative study was conducted for this study. First, the photocatalyst was compared with its virgin version of NPs. The versions of NPs here are ZrO2 and CdWO4. The photocatalytic activity of both these NPs for water splitting was observed and it was found that both these versions of the nano-catalysts showed lower H2 production rates as compared to the heterojunction photocatalyst (ZrO2-CdWO4). The low efficiency of these virgin nano-photocatalysts may be due to the low charge separation and fast electron-hole recombination between VB and CB, whereas a heterojunction photocatalyst is a junction that is formed when zirconium oxide and cadmium tungstate are mixed and it can prevent electron-hole recombination; therefore, this photocatalyst is responsible for the increased H2 production rate (Figure 8a).
The present photocatalyst was also compared with several previous photocatalysts and it can be seen from Table 4 that the present photocatalyst has shown comparable performance to the previous photocatalyst and has also shown improved results. This suggests that the existing photocatalyst is a better catalyst for H2 production.
However, comparing photocatalysts in terms of their H2 production capacity would be a biased analysis as the performance parameters and conditions differed for each reported catalyst. Many photocatalysts are given in Table 4; most of these catalysts were made through the chemical method; in comparison, the present catalyst has been made through the precipitation method, which makes the present catalyst cheap, low cost, and environmentally friendly. Additionally, the current photocatalyst is a type I heterojunction, which performs the water-splitting process more smoothly than other catalysts for seven reused cycles. This gives it an advantage over the other reported catalyst. Therefore, the present catalyst is better than other reported photocatalysts. Although, to validate these inferences, more concrete fundamental mechanistic studies are required.

4. Conclusions

This study describes using plant extracts to successfully prepare ZrO2-CdWO4, a heterojunction material. A heterojunction was made by depositing tetragonal zirconium dioxide over the monoclinic cadmium tungstate. The 2.66 eV energy band gap was reported for this photocatalyst, which was lower than the bare oxide (ZrO2 and CdWO4) as well as various reported CdWO4-based nanocomposites (ZnO-CdWO4 and CuO-ZnO-CdWO4). A ZrO2 heterojunction with CdWO4 may lower the band gap energy and aid visible light photon absorption. Critically, it was reported that heterojunction formation forms a Schottky barrier, which increases charge separation and prevents electron-hole combinations. This mechanism exhibits a complete distribution of electrons on the semiconductor surface, which is important for photocatalytic applications. In this study, hydrogen production was enabled via water splitting to understand photocatalytic application. The heterojunction material resulted in improved hydrogen production. After a 5 h reaction at 27 °C, pH 7, and using 30 vol. % of methanol, 1574 µmol/g hydrogen gas was produced using 250 mg of ZrO2-CdWO4. This is a useful study on heterojunction materials being used for hydrogen production.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the article.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Macedo, D.P.; Marques, A.C. Is the energy transition ready for declining budgets in RD&D for fossil fuels? Evidence from a panel of European countries. J. Clean. Prod. 2023, 417, 138102. [Google Scholar]
  2. Rennuit-Mortensen, A.W.; Rasmussen, K.D.; Grahn, M. How replacing fossil fuels with electrofuels could influence the demand for renewable energy and land area. Smart Energy 2023, 10, 100107. [Google Scholar] [CrossRef]
  3. Baquero, J.E.G.; Monsalve, D.B. A Proposal for the Transformation of Fossil Fuel Energy Economies to Hydrogen Economies through Social Entrepreneurship. In Entrepreneurial Innovation for Securing Long-Term Growth in a Short-Term Economy; IGI Global: Hershey, PA, USA, 2021; pp. 48–70. [Google Scholar]
  4. Zhang, L.; Jia, C.; Bai, F.; Wang, W.; An, S.; Zhao, K.; Li, Z.; Li, J.; Sun, H. A comprehensive review of the promising clean energy carrier: Hydrogen production, transportation, storage, and utilization (HPTSU) technologies. Fuel 2024, 355, 129455. [Google Scholar] [CrossRef]
  5. Song, H.; Luo, S.; Huang, H.; Deng, B.; Ye, J. Solar-Driven Hydrogen Production: Recent Advances, Challenges, and Future Perspectives. ACS Energy Lett. 2022, 7, 1043–1065. [Google Scholar] [CrossRef]
  6. Okere, C.J.; Sheng, J.J. Review on clean hydrogen generation from petroleum reservoirs: Fundamentals, mechanisms, and field applications. Int. J. Hydrogen Energy 2023, 48, 38188–38222. [Google Scholar] [CrossRef]
  7. Tamilselvan, R.; Selwynraj, A.I. Model development for biogas generation, purification and hydrogen production via steam methane reforming. Int. J. Hydrogen Energy 2024, 50, 211–225. [Google Scholar] [CrossRef]
  8. Ji, M.; Shi, M.; Wang, J. Life cycle assessment of nuclear hydrogen production processes based on high temperature gas-cooled reactor. Int. J. Hydrogen Energy 2023, 48, 22302–22318. [Google Scholar] [CrossRef]
  9. Zhang, Y.; Wang, Z.; Du, Z.; Li, Y.; Qian, M.; Van herle, J.; Wang, L. Techno-economic analysis of solar hydrogen production via PV power/concentrated solar heat driven solid oxide electrolysis with electrical/thermal energy storage. J. Energy Storage 2023, 72, 107986. [Google Scholar] [CrossRef]
  10. Song, H.; Yang, G.; Xue, P.; Li, Y.; Zou, J.; Wang, S.; Yang, H.; Chen, H. Recent development of biomass gasification for H2 rich gas production. Appl. Energy Combust. Sci. 2022, 10, 100059. [Google Scholar] [CrossRef]
  11. Idriss, H. Hydrogen production from water: Past and present. Curr. Opin. Chem. Eng. 2020, 29, 74–82. [Google Scholar] [CrossRef]
  12. Chung, K.H.; Jung, H.H.; Kim, S.J.; Park, Y.K.; Kim, S.C.; Jung, S.C. Hydrogen Production through Catalytic Water Splitting Using Liquid-Phase Plasma over Bismuth Ferrite Catalyst. Int. J. Mol. Sci. 2021, 22, 13591. [Google Scholar] [CrossRef] [PubMed]
  13. Tahir, M.B.; Sohaib, M.; Sagir, M.; Rafique, M. Role of Nanotechnology in Photocatalysis. Encycl. Smart Mater. 2022, 2, 578–589. [Google Scholar]
  14. Majeed, I.; Arif, A.; Faizan, M.; Khan, M.A.; Imran, M.; Ali, H.; Nadeem, M.A. CdS nanorods supported copper-nickel hydroxide for hydrogen production under direct sunlight irradiation. J. Environ. Chem. Eng. 2021, 9, 105670. [Google Scholar] [CrossRef]
  15. Althabaiti, S.A.; Khan, Z.; Malik, M.A.; Bawaked, S.M.; Al-Sheheri, S.Z.; Mokhtar, M.; Siddiqui, S.I.; Narasimharao, K. Biomass-derived carbon deposited TiO2 nanotube photocatalysts for enhanced hydrogen production. Nanoscale Adv. 2023, 5, 3671–3683. [Google Scholar] [CrossRef] [PubMed]
  16. Althabaiti, S.A.; Malik, M.A.; Kumar Khanna, M.; Bawaked, S.M.; Narasimharao, K.; Al-Sheheri, S.Z.; Fatima, B.; Siddiqui, S.I. One-Pot Facile Synthesis of CuO–CdWO4 Nanocomposite for Photocatalytic Hydrogen Production. Nanomaterials 2022, 12, 4472. [Google Scholar] [CrossRef] [PubMed]
  17. Fatima, B.; Siddiqui, S.I.; Rajor, H.K.; Malik, M.A.; Narasimharao, K.; Ahmad, R.; Vikrant, K.; Kim, T.; Kim, K.H. Photocatalytic removal of organic dye using green synthesized zinc oxide coupled cadmium tungstate nanocomposite under natural solar light irradiation. Environ. Res. 2023, 216, 114534. [Google Scholar] [CrossRef] [PubMed]
  18. Fatima, B.; Siddiqui, S.I.; Ahmad, R.; Linh, N.T.T.; Thai, V.N. CuO-ZnO-CdWO4: A sustainable and environmentally benign photocatalytic system for water cleansing. Environ. Sci. Pollut. Res. 2021, 28, 53793–53803. [Google Scholar] [CrossRef] [PubMed]
  19. Sharma, K.; Raizada, P.; Hasija, V.; Singh, P.; Bajpai, A.; Nguyen, V.H.; Rangabhashiyam, S.; Kumar, P.; Nadda, A.K.; Kim, S.Y.; et al. ZnS-based quantum dots as photocatalysts for water purification. J. Water Process Eng. 2021, 43, 102217. [Google Scholar] [CrossRef]
  20. Sethi, Y.A.; Panmand, R.P.; Kadam, S.R.; Kulkarni, A.K.; Apte, S.K.; Naik, S.D.; Kale, B.B. Nanostructured CdS sensitized CdWO4 nanorods for hydrogen generation from hydrogen sulfide and dye degradation under sunlight. J. Colloid Interface Sci. 2017, 487, 504–512. [Google Scholar] [CrossRef]
  21. Alagarsamy, A.; Chandrasekaran, S.; Manikandan, A. Green synthesis and characterization studies of biogenic zirconium oxide (ZrO2) nanoparticles for adsorptive removal of methylene blue dye. J. Mol. Struct. 2022, 1247, 131275. [Google Scholar] [CrossRef]
  22. Vattikuti, S.P.; Devarayapalli, K.C.; Nagajyothi, P.C.; Shim, J. Binder-free WS2/ZrO2 hybrid as a photocatalyst for organic pollutant degradation under UV/simulated sunlight and tests for H2 evolution. J. Alloys Compd. 2019, 809, 151805. [Google Scholar] [CrossRef]
  23. Rani, V.; Sharma, A.; Kumar, A.; Singh, P.; Thakur, S.; Singh, A.; Raizada, P. ZrO2-Based Photocatalysts for Wastewater Treatment: From Novel Modification Strategies to Mechanistic Insights. Catalysts 2022, 12, 1418. [Google Scholar] [CrossRef]
  24. López, M.C.U.; Lemus, M.A.A.; Hidalgo, M.C.; González, R.L.; Owen, P.Q.; Oros-Ruiz, S.; López, S.A.U.; Acosta, J. Synthesis and characterization of ZnO-ZrO2 nanocomposites for photocatalytic degradation and mineralization of phenol. J. Nanomater. 2019, 2019, 1015876. [Google Scholar]
  25. Bai, j.; Huang, Y.; Wei, D.; Fan, Z.; Seo, H.J. Synthesis and characterization of semiconductor heterojunctions based on Zr6Nb2O17 nanoparticles. Mater. Sci. Semicond. Process. 2020, 112, 105010. [Google Scholar] [CrossRef]
  26. Ayappan, C.; Palanivel, B.; Jayaraman, V.; Maiyalagan, T.; Mani, A. One-step hydrothermal synthesis of CaWO4/α-Ag2WO4 heterojunction: An efficient photocatalyst for removal of organic contaminants. Mater. Sci. Semicond. Process. 2019, 104, 104693. [Google Scholar] [CrossRef]
  27. Chen, C.L.; Pu, H.B.; Wang, M.; Wang, X.; Zang, Y.; Gao, C.Y. A facile way for fabrication of nano-sized p-NiO/n-SiC heterojunction using sol–gel technique. Superlattices Microstruct. 2019, 136, 106251. [Google Scholar] [CrossRef]
  28. Kubiak, A.; Bielan, Z.; Kubacka, M.; Gabała, E.; Zgoła-Grześkowiak, A.; Janczarek, M.; Zalas, M.; Zielińska-Jurek, A.; Siwińska-Ciesielczyk, K.; Jesionowski, T. Microwave-assisted synthesis of a TiO2-CuO heterojunction with enhanced photocatalytic activity against tetracycline. Appl. Surf. Sci. 2020, 520, 146344. [Google Scholar] [CrossRef]
  29. Radini, I.A.; Hasan, N.; Malik, M.A.; Khan, Z. Biosynthesis of iron nanoparticles using Trigonella foenum-graecum seed extract for photocatalytic methyl orange dye degradation and antibacterial applications. J. Photochem. Photobiol. B Biol. 2018, 183, 154–163. [Google Scholar] [CrossRef]
  30. Jawhari, A.H.; Hasan, N. Nanocomposite Electrocatalysts for Hydrogen Evolution Reactions (HERs) for Sustainable and Efficient Hydrogen Energy—Future Prospects. Materials 2023, 16, 3760. [Google Scholar] [CrossRef]
  31. Alamier, W.M.; Hasan, N.; Syed, I.S.; Bakry, A.M.; Ismail, K.S.; Gedda, G.; Girma, W.M. Silver Nanoparticles’ Biogenic Synthesis Using Caralluma subulata Aqueous Extract and Application for Dye Degradation and Antimicrobials Activities. Catalysts 2023, 13, 1290. [Google Scholar] [CrossRef]
  32. Alamier, W.M.; Hasan, N.; Nawaz, M.D.S.; Ismail, K.S.; Shkir, M.; Malik, M.A.; Oteef, M.D.Y. Biosynthesis of NiFe2O4 nanoparticles using Murayya koenigii for photocatalytic dye degradation and antibacterial application. J. Mat. Res. Technol. 2023, 22, 1331–1348. [Google Scholar] [CrossRef]
  33. Tabassum, N.; Kumar, D.; Verma, D.; Bohara, R.A.; Singh, M.P. Zirconium oxide (ZrO2) nanoparticles from antibacterial activity to cytotoxicity: A next-generation of multifunctional nanoparticles. Mater. Today Commun. 2021, 26, 102156. [Google Scholar] [CrossRef]
  34. Murali, G.; Gopalakrishnan, S.; Lakhera, S.K.; Neppolian, B.; Ponnusamy, S.; Harish, S.; Navaneethan, M. Enhanced solar light-driven hydrogen evolution of activated carbon sphere supported TiO2 hybrid nanocomposites. Diam. Relat. Mater. 2022, 128, 109226. [Google Scholar] [CrossRef]
  35. Fathima, J.B.; Pugazhendhi, A.; Venis, R. Synthesis and characterization of ZrO2 nanoparticles-antimicrobial activity and their prospective role in dental care. Microb. Pathog. 2017, 2017, 245–251. [Google Scholar] [CrossRef] [PubMed]
  36. Abid, R.; Islam, M.; Saeed, H.; Ahmad, A.; Imtiaz, F.; Yasmeen, A.; Rathore, H.A. Antihypertensive potential of Brassica rapa leaves: An in vitro and in silico approach. Front. Pharmacol. 2022, 13, 996755. [Google Scholar] [CrossRef] [PubMed]
  37. Jia, R.; Zhang, G.; Wu, Q.; Ding, Y. Preparation, structures and photoluminescent enhancement of CdWO4–TiO2 composite nanofilms. Appl. Surf. Sci. 2006, 253, 2038–2042. [Google Scholar] [CrossRef]
  38. Tian, N.; Huang, H.; Zhang, Y. Mixed-calcination synthesis of CdWO4/g-C3N4 heterojunction with enhanced visible-light-driven photocatalytic activity. Appl. Surf. Sci. 2015, 358, 343–349. [Google Scholar] [CrossRef]
  39. Priya, A.M.; Selvan, R.K.; Senthilkumar, B.; Satheeshkumar, M.K.; Sanjeeviraja, C. Synthesis and characterization of CdWO4 nanocrystals. Ceram. Int. 2011, 37, 2485–2488. [Google Scholar] [CrossRef]
  40. Xing, D.; Liu, Y.; Zhou, P.; Wang, Z.; Wang, P.; Zheng, Z.; Huang, B. Enhanced photocatalytic hydrogen evolution of CdWO4 through polar organic molecule modification. Int. J. Hydrogen Energy 2019, 44, 4754–4763. [Google Scholar] [CrossRef]
  41. Igawa, N.; Ishii, Y. Crystal Structure of Metastable Tetragonal Zirconia up to 1473 K. J. Am. Ceram. Soc. 2004, 84, 1169–1171. [Google Scholar] [CrossRef]
  42. Siddiqui, M.R.H.; Al-Wassil, A.I.; Al-Otaibi, A.M.; Mahfouz, R.M. Effects of precursor on the morphology and size of ZrO2 nanoparticles, synthesized by sol-gel method in non-aqueous medium. Mater. Res. 2012, 15, 986–989. [Google Scholar] [CrossRef]
  43. Jawhari, A.H.; Hasan, N.; Radini, I.A.; Malik, M.A.; Narasimharao, K. Pt-Ag/Ag3PO4-WO3 nanocomposites for photocatalytic H2 production from bioethanol. Fuel 2023, 344, 127998. [Google Scholar] [CrossRef]
  44. Jawhari, A.H.; Hasan, N.; Radini, I.A.; Narasimharao, K.; Malik, M.A. Noble Metals Deposited LaMnO3 Nanocomposites for Photocatalytic H2 Production. Nanomaterials 2022, 12, 2985. [Google Scholar] [CrossRef]
  45. Wang, H.; Qi, H.; Sun, X.; Jia, S.; Li, X.; Miao, T.J.; Xiong, L.; Wang, S.; Zhang, X.; Liu, X.; et al. High quantum efficiency of hydrogen production from methanol aqueous solution with PtCu-TiO2 photocatalysts. Nat. Mater. 2023, 22, 619–626. [Google Scholar] [CrossRef] [PubMed]
  46. Velázquez, J.J.; Fernández-González, R.; Díaz, L.; Pulido Melián, E.; Rodríguez, V.D.; Núñez, P. Effect of reaction temperature and sacrificial agent on the photocatalytic H2-production of Pt-TiO2. J. Alloys Compd. 2017, 721, 405–410. [Google Scholar] [CrossRef]
Scheme 1. Flow chart of the synthesis of the ZrO2-CdWO4 NC.
Scheme 1. Flow chart of the synthesis of the ZrO2-CdWO4 NC.
Applsci 13 13344 sch001
Figure 1. (a) UV–Visible diffuse absorbance spectrum of (a) ZrO2, CdWO4, ZrO2-CdWO4; (b) a Tauc plot of ZrO2, CdWO4, and ZrO2-CdWO4 as an indirect band gap; and (c) a Tauc plot of ZrO2, CdWO4, and ZrO2-CdWO4 as a direct band gap.
Figure 1. (a) UV–Visible diffuse absorbance spectrum of (a) ZrO2, CdWO4, ZrO2-CdWO4; (b) a Tauc plot of ZrO2, CdWO4, and ZrO2-CdWO4 as an indirect band gap; and (c) a Tauc plot of ZrO2, CdWO4, and ZrO2-CdWO4 as a direct band gap.
Applsci 13 13344 g001
Figure 2. FT-IR spectra of the ZrO2-CdWO4 NC.
Figure 2. FT-IR spectra of the ZrO2-CdWO4 NC.
Applsci 13 13344 g002
Figure 3. X-ray diffraction pattern for the ZrO2-CdWO4 NCs.
Figure 3. X-ray diffraction pattern for the ZrO2-CdWO4 NCs.
Applsci 13 13344 g003
Figure 4. (a) SEM images and (b) EDX of ZrO2-CdWO4 NC.
Figure 4. (a) SEM images and (b) EDX of ZrO2-CdWO4 NC.
Applsci 13 13344 g004
Figure 5. TEM images of the ZrO2-CdWO4 NC.
Figure 5. TEM images of the ZrO2-CdWO4 NC.
Applsci 13 13344 g005
Figure 6. Nitrogen adsorption–desorption isotherm (inset pore diameter distribution) of the ZrO2-CdWO4 NC.
Figure 6. Nitrogen adsorption–desorption isotherm (inset pore diameter distribution) of the ZrO2-CdWO4 NC.
Applsci 13 13344 g006
Figure 7. PL spectra of ZrO2, CdWO4, and the ZrO2-CdWO4 NC.
Figure 7. PL spectra of ZrO2, CdWO4, and the ZrO2-CdWO4 NC.
Applsci 13 13344 g007
Figure 8. Optimization results of H2 production via water splitting using ZrO2-CdWO4: (a) time effect, (b) catalyst amount effect, (c) temperature effect, (d) pH effect, and (e) MeOH vol% effect.
Figure 8. Optimization results of H2 production via water splitting using ZrO2-CdWO4: (a) time effect, (b) catalyst amount effect, (c) temperature effect, (d) pH effect, and (e) MeOH vol% effect.
Applsci 13 13344 g008
Scheme 2. Plausible reaction mechanism for H2 production over ZrO2-CdWO4.
Scheme 2. Plausible reaction mechanism for H2 production over ZrO2-CdWO4.
Applsci 13 13344 sch002
Figure 9. Reutilization of ZrO2-CdWO4 for H2 production.
Figure 9. Reutilization of ZrO2-CdWO4 for H2 production.
Applsci 13 13344 g009
Table 1. Comparative table for different NP/NC sizes and band gap.
Table 1. Comparative table for different NP/NC sizes and band gap.
Order NPs/NCsPreparation MethodSize of Particles (nm)Band Gap (eV)Ref.
1.CdWO4Green synthesis using Brassica rapa leave extract26–324.0[34]
2.CuO-CdWO4Green synthesis using Brassica rapa leave extract-2.56[16]
3.ZnO-CdWO4Green synthesis using Lemon extract 10–503.12[17]
4.CuO-ZnO- CdWO4Green synthesis using Brassica rapa leave extract28–623.13[18]
5.ZrO2Green synthesis using Sapindus mukorossi plant extract 5–105.5[21]
6.CdWO4Green synthesis using Brassica rapa leave extract-3.06[This study]
7.ZrO2-CdWO4Green synthesis using Brassica rapa leave extract16–292.66
Table 2. Band edge position in ZrO2-CdWO4 heterojunction.
Table 2. Band edge position in ZrO2-CdWO4 heterojunction.
SemiconductorsEg (eV)X (eV)EVB (eV)ECB (eV)
ZrO24.165.913.49−0.67
CdWO43.066.283.310.25
Table 3. Comparative FTIR analysis of various CdWO4 NP-based composites prepared using plant extracts.
Table 3. Comparative FTIR analysis of various CdWO4 NP-based composites prepared using plant extracts.
OrderNCsPlant ExtractFTIR PeaksRef.
Vibrational Frequencies of Functional Groups due to Plant ExtractVibrational Frequencies of M-O Due to NPs
1.CdWO4Brassica rapa leaf3419 cm−1 -O-H; 2923 cm−1 and 2853 cm−1 C-H of CH3, and CH2; 1460–1020 cm−1 aromatic C–C and C–O, respectively900–500 cm−1 Intrinsic vibrations of CdWO4[34]
2.ZnO-CdWO4Lemon leaf3459 cm−1 -OH; 1671 cm−1 -O-H/amide; 1469 cm−1 -C-H; 1084 cm−1 C–N844 cm−1 W–O–W bond in the WO4; 586 cm−1 Cd–O; 447 cm−1 and 417 cm−1 Zn–O[17]
3.CuO-CdWO4Brassica rapa leaf3481–3400 cm−1 -O-H; 1670–1590 cm−1 C = O, -NH, and aromatic C = C; 1460–1020 cm−1 aromatic C–C and C–O, respectively1000–400 cm−1 CuO and CdWO4[16]
4.CuO-ZnO-CdWO4Brassica rapa leaf3664–3060 cm−1 O-H; 2924 and 2854 cm−1 C-H of CH3 and CH2, 1746 and 1634 cm−1 -C = O and -O-H836 cm−1 Cd-O-W; 708 cm−1 W-O; 450 cm−1 Cd-O; 531 cm−1 and 595 cm−1 Zn-O and Cu-O, respectively[18]
5.ZrO2-CdWO4Brassica rapa leaf3429 cm−1 O-H; 2919 and 2854 cm−1 C-H of CH3 and CH2, 1633 cm−1 Zr-OH829 cm−1 Zr-O; 890.0 cm−1, 717.6 cm−1 and (587.96 cm−1 and 412.3 cm−1 and 412 cm−1) Cd–O–W, W-O-W bond in WO42− group and in-plane deformation of the WO42− group[This study]
Table 4. Comparative table of different nanomaterials for H2 production.
Table 4. Comparative table of different nanomaterials for H2 production.
OrderMaterialSacrificial AgentLight SourcesH2 Production
(µ mol/g)
Ref.
1.CuO-CdWO4MeOHSunlight~1600[16]
2.TiO2/ACSGlycerine Sunlight 203[34]
3.Pt-Cu-TiO2MeOHLED irradiation ~2384[45]
4.Pt/TiO2MeOHSunlight195.5[46]
5.ZrO2-CdWO4MeOHSunlight1574This study
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Jawhari, A.H. One-Pot Facile Synthesis of ZrO2-CdWO4: A Novel Nanocomposite for Hydrogen Production via Photocatalytic Water Splitting. Appl. Sci. 2023, 13, 13344. https://0-doi-org.brum.beds.ac.uk/10.3390/app132413344

AMA Style

Jawhari AH. One-Pot Facile Synthesis of ZrO2-CdWO4: A Novel Nanocomposite for Hydrogen Production via Photocatalytic Water Splitting. Applied Sciences. 2023; 13(24):13344. https://0-doi-org.brum.beds.ac.uk/10.3390/app132413344

Chicago/Turabian Style

Jawhari, Ahmed Hussain. 2023. "One-Pot Facile Synthesis of ZrO2-CdWO4: A Novel Nanocomposite for Hydrogen Production via Photocatalytic Water Splitting" Applied Sciences 13, no. 24: 13344. https://0-doi-org.brum.beds.ac.uk/10.3390/app132413344

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