Next Article in Journal
Study of the Electrical Properties of Aluminate Cement Adhesives for Porcelain Insulators
Next Article in Special Issue
Enhanced Photodegradation of p-Nitrobenzoic Acid by Binary Mixtures with Ba2+/TiO2 and MCM-41
Previous Article in Journal
The Design and Research of a New Hybrid Surface Plasmonic Waveguide Nanolaser
Previous Article in Special Issue
Enhanced Photocatalytic Activity of Spherical Nd3+ Substituted ZnFe2O4 Nanoparticles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Tuned S-Scheme Cu2S_TiO2_WO3 Heterostructure Photocatalyst toward S-Metolachlor (S-MCh) Herbicide Removal

1
Product Design, Mechatronics and Environmental Department, Transilvania University of Brasov, 35000 Brasov, Romania
2
Renewable Energy Systems and Recycling Research Center, Transilvania University of Brasov, Eroilor 29 Street, 35000 Brasov, Romania
*
Author to whom correspondence should be addressed.
Materials 2021, 14(9), 2231; https://doi.org/10.3390/ma14092231
Submission received: 11 March 2021 / Revised: 19 April 2021 / Accepted: 23 April 2021 / Published: 26 April 2021

Abstract

:
A dual S-scheme Cu2S_TiO2_WO3 heterostructure was constructed by sol–gel method using a two-step procedure. Due to the synthesis parameters and annealing treatment the heterostructure is characterized by sulfur deficit and oxygen excess allowing the passivation of oxygen vacancies. The photocatalytic activity was evaluated under UV and UV–Vis irradiation scenarios using S-MCh as reference pollutant. The heterostructure is composed on orthorhombic Cu2S, anatase TiO2 and monoclinic WO3 with crystallite sizes varying from 65.2 Å for Cu2S to 97.1 Å for WO3. The heterostructure exhibit a dense morphology with pellets and particle-like morphology closely combined in a relatively compact assembly. The surface elemental composition indicate that the heterostructure maintain a similar atomic ratio as established during the synthesis with a slight sulfur deficit due to the annealing treatments. The results indicate that the three-component heterostructure have higher photocatalytic efficiency (61%) comparing with two-component heterostructure or bare components. Moreover, Cu2S_TiO2_WO3 exhibit a superior constant rate (0.114 s−1) due to the high concentration of photogenerated charge carriers, efficient charge separation and migration.

1. Introduction

Aquatic contamination with recalcitrant organic compounds released by industrial, agricultural, and transportation activities represent a serious threat to human health and life quality. Pesticide compounds including herbicides, insecticides, and fungicides have large contamination potential due to their continuous use in agricultural area and long term persistency in the ecological system [1,2,3]. The agricultural accumulation of organic compounds, such as S-metolachlor, cyprodinil, iprodione, tebuconazole, etc., in sediment and biota, are sources of serious environmental pollution [4,5]. Finding new and sustainable pathways for water decontamination represent an important target for the future of agriculture and wastewater plants. Photocatalytic technologies use the light energy as driving force to induce organic pollutant oxidation. The use of mono-component photocatalysts such as TiO2 [6], WO3 [7], SnO2 [8], Cu2S [9], CuO [10], or NiO [11] presents several disadvantages: (i) limited light absorption range, (ii) fast charge carriers recombination, and (iii) reduced chemical stability.
The photocatalytic process is based on three major steps: (1) the electron/hole pairs formation during irradiation due to the semiconductors absorption of photons with energy (hv) equal to or above their bandgap (Eg); (2) the separation and diffusion of photogenerated carriers under the action of the internal electric field; (3) formation of oxidative and super-oxidative species due to the surfaces reactions induced by the photogenerated electrons and holes [12,13,14]. TiO2 and WO3 have band gap values of 3.2 eV and 2.8 eV, which means that they are mostly active in the UV spectra. The charge carriers are photo-generated when the photons energy are equal or higher than the band gap value. Cu2S is photoactivated in Vis spectra due to his band gap value which may vary between 1.2 eV and 2.0 eV, depending on the stoichiometric composition. When coupled, the band gap values shift in order to generate a build in electric field required for charge carriers mobility through the heterostructure. Traditional heterojunctions (such as type-II heterojunctions, p-n junction) typically produce unfavorable losses of the photogenerated charges, while the Z-scheme heterostructure represent the starting point on the development of S-scheme heterostructure [15,16]. If compared with Z-scheme heterostructure, the S-scheme efficiently use the build-in electric field in order to reduce the migration distance of photogenerated electrons and holes based on the synergetic interface between the semiconductor components [17,18].
The S-scheme heterostructure consists of two or more semiconductors with suitable position of the energy bands. The driving force responsible of the charge migration is represented by the internal electric field established between the heterostructure components [19,20]. The photogenerated charges are separated in space, based on the semiconductors potential: the holes are located in the conduction band (CB) of the reduction photocatalyst and the electrons are in the valence band (VB) of the oxidation photocatalyst, respectively [21,22]. The photocatalytic reactions generating the (super)oxidative (•OH, •O2) species are initiated by the photogenerated holes and electrons. Therefore, efficient S-scheme heterostructure is highly desirable, in order to promote the charge migration and to sustain favorable charge potentials [23,24]. Until now, several S-scheme heterostructures such as SnFe2O4/ZnFe2O4 [25], TiO2/W18O49 [26], NiO/BiOI [27], BiVO4@MoS2 [28], WO3/CdIn2S4 [29], CdS/UiO-66 [30] were tested for dyes or pharmaceutical active compounds removal. The SnFe2O4/ZnFe2O4 show enhanced Vis light absorbance and direct S-scheme path of charge separation and transfer. After 60 min of 300 W Vis light irradiation the NiO/BiOI exhibited 90% Rhodamine B photocatalytic removal.
Herein, we report for the first time the photocatalytic removal of S-MCh herbicide using a dual S-scheme heterostructure and two irradiation scenarios. The composition and morphology was investigated and correlated with the photocatalytic properties and energy band diagram. The dual S-scheme mechanism facilitates the photogenerated charge carriers migration and the efficient use of redox potential to induce pollutant degradation. Compared with Z-scheme mechanism, the S-scheme heterostructure use the build-in electric field to mitigate the charge transmission distance due to the synergic semiconductors interface. The S-scheme mechanism is able to use more efficiently the photogenerated charge carriers based on the charge density difference between the heterostructure components. The kinetic evaluation indicates a higher oxidation reaction rate for the three-component heterostructure compared with two-components or single component photocatalysts.

2. Materials and Methods

2.1. Photocatalyst Materials

2.1.1. Preparation of Mono-Component Photocatalysts

Two powder components were prepared using sol–gel technique:
(i)
Cu2S was prepared by mixing 0.2 mol of copper nitrate (Cu(NO3)2, 99.9%, Scharlau, Barcelona, Spain) aqueous solution with 0.5 mol of sodium thiosulfate (Na2S2O3, 99.9%, Scharlau) aqueous solution. After 15 min of stirring the gel was formed and kept 3 h undisturbed to achieve the complete precipitation. The precipitate was centrifuged and thermally treated at 120 °C in a ceramic capsule in sulfured (Sulfur, 99%, Sigma Aldrich, Munich, Germany) atmosphere.
(ii)
WO3 was obtained by dissolving tungsten hexachloride (WCl6, 99.4%, Acros Organics, Geel, Belgium) in a mixture of ethanol (100%, Sigma Aldrich) and 2-propanol (100%, Sigma Aldrich). After 120 min of stirring a light yellow alcoholic solution was obtained. Then, 0.15 mol of natrium hydroxide (99.98%, Honeywell, Charlotte, NC, USA) was added drop by drop and the gel was formed. After precipitation and centrifugation the resulting powder was annealed for 8 h at 500 °C.
The TiO2 powder was purchased (99.99%, Scharlau, Barcelona, Spain) and used without any purification procedures.

2.1.2. Preparation of Multi-Component Photocatalysts

Three bi-component and one three-component heterostructures were prepared by sol–gel technique:
(i)
Cu2S_TiO2 sample was obtained using the same procedure as described for Cu2S, with the single modification that TiO2 powder was dispersed into copper nitrate solution, considering the Cu:Ti atomic ratio of 1:1. The final powder was thermally treated at 150 °C for 2 h.
(ii)
Cu2S_WO3 sample was obtained following the procedure described for Cu2S. The WO3 powder already prepared was added into copper nitrate solution, considering the Cu:W atomic ratio of 1:1. The final powder was thermally treated at 150 °C for 1.5 h.
(iii)
TiO2_WO3 sample was prepared as using a similar procedure as presented for WO3, and the TiO2 power was dispersed into tungsten hexachloride solution. The uniform TiO2 distribution was assured by adding polyethylene glycol (99%, Scharlau) and the Ti:W atomic ratio was 1:1. The annealing treatment was done at 500 °C for 5 h.
(iv)
Cu2S_TiO2_WO3 sample was synthesized by adding TiO2_WO3 powder, previously obtained, into copper nitrate solution and the mixture was stirred for 2 h. The Cu:Ti:W atomic ratio was 1:1:1. Then, 0.7 mol of sodium thiosulfate (Na2S2O3, 99.9%, Scharlau) was added drop by drop under continuous stirring. The precipitate was centrifuged and thermally treated at 150 °C for 2 h.

2.2. Photocatalytic Activity

The photocatalytic activity was tested using a closed reactor able to produce light irradiation from the top and lateral sides, which increase the radiation uniformity during the experiments. The reactor chamber use ventilation engines and thermocouples in order to preserve constant temperature (25 °C) and humidity (60%). Two kind of radiation sources were used (single or combined): 18W UVa black tubes (T8, 3Lx flux intensity, λUVa,max = 365 nm, range 310–390 nm, Philips) and 18W Vis cold tubes (TL-D Super 80/865, flux intensity 28Lx, λVis,max = 565 nm, range 400–700 nm, Philips). Two irradiation scenarios where verified and the corresponding total irradiance is presented in Table 1.
S-Metholachlor (S-MCh) herbicide was used as reference pollutant molecule, due to his reluctance to traditional wastewater treatments procedures as well as his toxicity and accumulation potential into the biota. High concentrations of S-MCh may induce cytotoxicity and genotoxicity effect localized in human lymphocytes. The photocatalytic experiments were done using 30 mg/L S-MCh aqueous solution and the photocatalyst dosage was 30 mg/50 mL. Quart recipients were considered due to high UV transmittance. The total time of experiment was 10 h: 2 h in dark to reach the absorption equilibrium and 8 h under irradiation. The pollutant concentration was measured hourly and compared with the UV–Vis is calibration curve (274 nm is the S-MCh absorption wavelength). Finally, the photocatalytic efficiency was calculated considering the initial (C0) and final (C) concentrations based on the following Equation (1):
η = [ ( C 0 C ) C 0 ] 100

2.3. Characterization Instruments

The presence of crystalline structure was evaluated using X-ray diffraction (XRD, Bruker, Model D8 Discover, Karlsruhe, Germany) with 0.003 degree scan step locked-couple technique and 0.015 s/step. The morphology characterizations were done with scanning electron microscopy (SEM, Hitachi model S–3400 N type 121 II, Tokyo, Japan) in high vacuum regime for mono-component samples and with field emission scanning electron microscopy (FESEM, SU8010, Fukuoka, Japan) operated at an accelerated voltage of 25 kV for multi-component samples. The light intensity and total irradiance values inside the photoreactor were measured using a class A high precision pyranometer (SR11, Hukselflux, Berlin, Germany). The optical properties as well as the photocatalytic activity were investigated using UV–Vis spectrometry (Perkin Elmer Lambda 950, Waltham, MA, USA).

3. Results and Discussions

3.1. Composition and Morphology

The evaluation of crystalline structure presented in Figure 1 indicates that TiO2 has maintained the anatase structure (ICCD 83-0951) and crystallites size (Table 2), as received from the manufacturer, regardless the chemical and thermal treatments followed during the heterostructures development. The crystallite sizes were calculated based on the Scherrer formula, Equation (2):
D = 0.9 λ β cos θ
where β is the observed angular width at half maximum intensity (FWHM) of the peak, λ is the X-ray wavelength (1.5406 Å for CuKα1), and θ is the Bragg’s angle.
The anatase TiO2 structure is considered as more photosensitive compared with rutile and brookite, and suitable for photocatalytic applications. Cu2S exhibit orthorhombic structure (djurleite, ICCD 20-0365) and the crystallite size was influenced by the insertion of metal oxides during the sol–gel procedure. The presence of TiO2 favors the formation of smaller Cu2S crystallites size, while WO3 induce an increase of the copper sulfide crystallite sizes [31]. This behavior was previously presented by other researchers [32,33] in relation with similar materials. The insertion of metal oxides particles during the Cu2S development will act as preferential nucleation sites. Using larger metal oxide crystallites size will favor the grow process, which is no longer restricted by space limitation [34]. Contrary, if the nucleation sites are based on smaller metal oxides crystallites these will also limits the Cu2S crystallite grow rate [35]. The similar crystallites size values obtained for copper sulfide in Cu2S_TiO2 and Cu2S_TiO2_WO3 samples is an indicator of the preferential Cu2S grow on TiO2 during the formation of three-component heterostructure. WO3 with monoclinic structure (ICCD 83-0951) show a similar behavior as Cu2S, and the crystallite size decrease in the presence of TiO2 during the sol–gel synthesis.
The morphological analyses presented in Figure 2 indicate that both mono and multi-components samples are composed on particles with irregular shape and sizes. The mono-component samples contain smaller particles, which have the tendency to form larger aggregates. The average value of mono-component particles was 0.5 µm for Cu2S, 30 nm for TiO2 and 0.2 µm for WO3. The bi-component heterostructures containing Cu2S exhibit pellets and particle-like structures closely combined in a relatively compact assembly. The TiO2_WO3 sample has a higher homogeneity due to use of polyethylene glycol (PG) additive which favor the uniform distribution of particles and decrease the aggregates size. A similar observation was made by Dudita M. [36], showing that the dispersive component of the PG can significantly decrease the aggregates formation. However, the PG removal procedure requires higher annealing temperature, in order to avoid carbon contamination, which is not suitable for samples containing Cu2S. Finally, the Cu2S_TiO2_WO3 sample morphology indicates a uniform distribution of Cu2S on the TiO2_WO3 assembly.
The EDS analysis was undertaken during the morphological investigations, in order to observe the elemental surface composition and the values are presented in Table 3. The results indicate that the atomic ratio between the metals was preserved as presented in the synthesis method. Small deviations of the atomic ratio for Cu2S_WO3 and Cu2S_TiO2_WO3 may be attributed to the non-uniform distribution at the sample surface. Additionally, the qualitative results were compared with theoretical sulfur and oxygen content, calculated based on the stoichiometric compounds identified during the X-ray diffraction investigations. The values indicate a similar tendency of sulfur deficit and oxygen excess. The results are consistent with other studies [37,38], showing that the heterostructures submitted to annealing in air may lose sulfur and develop higher oxygen content. During the annealing treatment part of the oxygen vacancy formed during the synthesis procedure will be passivated as presented in Equation (3).
V O + 1 2 O 2 O O x + 2 h

3.2. Photocatalytic Properties and Mechanism

The photocatalytic activity of the mono-component and multi-component samples was tested using two irradiation scenarios: UVa (Figure 3a) and combined UVa–Vis (Figure 3b). In all cases the S-MCh herbicide concentration was 30 mg/L and photocatalyst dosage was 30 mg/50 mL. As expected, under UV irradiation the samples containing TiO2 exhibit higher photocatalytic efficiency. The lowest photocatalytic efficiency (11.9%) was recorded for Cu2S sample which absorb mostly in Vis spectra. However, small photocatalytic efficiency differences where observed by comparing the Cu2S_TiO2_WO3 (36.6%) and TiO2_WO3 (30.8%) heterostructures, which means that the Cu2S contribution in UVa scenario is rather limited. This observation is valid for bi-component heterostructures containing Cu2S coupled with a TiO2 or WO3, where the photocatalytic efficiency increase with few percentage than that of mono-component samples. The combined UVa–Vis scenario brings the advantage of Cu2S component, which have a significant contribution in both bi and three-component heterostructures. The photocatalytic efficiency increase with additional 10% for Cu2S_TiO2 and Cu2S_WO3, comparing with UV scenario. The highest increase was observed for three-component heterostructure which reach 61.08% photocatalytic turnover rate, after 8 h of irradiation. These results are significant considering that S-MCh is considered to be a reluctant molecule towards photocatalytic decomposition due to the structural stability induced by the aromatic cycle and functional groups [39,40]. The formation of bi-products with pollutant potential cannot be excluded as partial oxidation may induce the development of carboxylic acids. Bare Cu2S, TiO2 and WO3 have lower photocatalytic efficiency due to the spectral absorption limitation and faster recombination charge. The experiments made without catalyst indicate a low S-MCh decomposition rate of 1.9% under UVa irradiation and 1.7% under UVa–Vis irradiations. The S-MCh ability of self-degradation in both irradiation scenarios is negligible.
The influence of photocatalysts composition and light irradiation scenario was further correlated with the kinetic investigation, based on the simplified Langmuir–Hinshelwood mathematical Equation (4) and the results are presented in Figure 4 and Table 4.
ln C C 0 = k t
The kinetic data help to better understand the influence of the heterostructure composition on the photocatalytic activity. Under UVa irradiation the photocatalytic activity of Cu2S_TiO2_WO3 heterostructure is 1.2x faster than TiO2_WO3, 2.1x faster than Cu2S_WO3, and 1.5x faster than Cu2S_TiO2 bi-component samples. The differences increase significantly under UVa–Vis irradiation where the three-component sample photocatalytic activity is 2.5x faster than TiO2_WO3, 2.7x faster than Cu2S_WO3, and 2.4x faster than Cu2S_TiO2 photocatalysts. It is worth noting that the differences between the constant rates of the bi-component heterostructures decrease in the combined irradiation scenario which confirms the contribution of Cu2S on the overall photocatalytic efficiency [41]. The kinetic evaluation made on bare Cu2S indicate that this component is 2x more active in UV–Vis scenario compared with UV, while TiO2 and WO3 exhibits small changes.
The L-H model (Equation (5)) proposed by Turchi and Ollis [42] was used to evaluate the influence of photon absorption based on both UV and UV–Vis irradiation scenarios.
r = d C d t = k r K S C 1 + K S C
where C represents the S-MCh concentration (mol·L−1), kr correspond to the apparent reaction rate constant (mol·L−1·min−1), the apparent adsorption constant is represented by Ks (L·mol−1) and the S-MCh removal rate is given by r (mol·L−1·min−1). It is useful to describe the apparent rate constant k (min−1) as kr·Ks which is a specific term for each heterostructure photocatalytic activity. The kr depends on the light radiation parameters and allows changing the Equation (5), accordingly:
1 r = 1 k r K S 1 C + 1 k r
The results obtained based on the linear plot 1/C vs. 1/r, considering the (1/kr) intercept and (/krKS) slope are presented in Table 5. The mathematical model indicators show that L-H fits well with the experimental photocatalytic results obtained under UV–Vis irradiation. Similar results were obtained in UV scenario for TiO2_WO3 and Cu2S_TiO2_WO3 samples. These conditions allows that the apparent reaction rates have the same order of magnitude with the apparent adsorption constant, while for the Cu2S_TiO2 and Cu2S_WO3 heterostructures, there is one order of magnitude difference. There are similar reports in the literature [43,44], indicating the correspondence between the light irradiation spectra and the energy bands of the heterostructure composition which contributes to the photogeneration of electrons and holes and the conversion to oxidative species.
In order to evaluate the heterostructures mechanism, the components band energies (Figure 5) are necessary to provide additional information regarding the potential compatibility between them. The band energy diagram was developed based on the experimental band gap values (Figure 5b,c), corresponding to each heterostructure component and the bands energy potential were evaluated in good agreement with the methodology presented in literature [45,46]. The shift of band gap values in heterostructure may be possible due to the internal energy field. The procedure use the following Equations (7)–(10) to determine the energy band position based on the semiconductor electronegativity (χsemiconductor), energy of free electrons vs. hydrogen (Ee), band gap energy (Eg), absolute cationic electronegativity (χcation), and specific cationic electronegativity χ c a t i o n ( P . u . ) where P.u. represent the Pauling units.
E V B = χ s e m i c o n d u c t o r E e + 0.5 E g
E C B = E V B E g
χ s e m i c o n d u c t o r ( e V ) = 0.45 χ c a t i o n ( e V ) + 3.36
χ c a t i o n ( e V ) = χ c a t i o n ( P . u ) + 0.206 0.336
The bands energy diagram represents an S-scheme charge transfer mechanism corresponding to the heterostructures semiconductor components. Under light excitation the photogenerated electrons from the Cu2S conduction band (−0.4 eV) will migrate to TiO2 conduction band (Cu2S_TiO2 and Cu2S_TiO2_WO3) and further on WO3 conduction band (+0.4 eV). However, the electrons from TiO2 and WO3 conduction bands (CB) cannot be involved in •O2 generation, and the holes from Cu2S valence band (+0.92 eV) cannot produce •OH, owing to their potential. Consequently, some of those charges which are not useful for photocatalytic reaction will recombine [47,48]. The useful electrons from the Cu2S valence band (−0.4 eV), as well as the holes from TiO2 (+2.9 eV) and WO3 (+3.1 eV) valence bands, have a stronger redox ability and are efficiently separated by the electric field in the charged space region. Under the combined drift and diffusion effect, the photogenerated charges will migrate through the heterostructure components, [49,50]. The higher photocatalytic efficiency exhibited by Cu2S_TiO2_WO3 is explained by the synergic activity of a dual S-scheme mechanism between the Cu2S/TiO2 and TiO2/WO3, respectively. The band gap values of each components shift, allowing the formation of a collective heterostructure band gap. The Tauc plot presented in Figure 5e indicate a band gap value of 1.45 eV for Cu2S_TiO2_WO3 heterostructure, confirming the absorbance in both UV and Vis light spectra. Therefore, the development of S-scheme heterostructure systems could efficiently promote the photoexcited charge carriers separation which determine the increase of (super)oxidative species concentration acting on S-MCh photocatalytic removal.

4. Conclusions

A dual Cu2S_TiO2_WO3 S-scheme heterostructure was prepared by sol-gel technique and the photocatalytic properties where tested and compared with other samples. The heterostructure is composed on orthorhombic Cu2S, anatase TiO2 and monoclinic WO3, with crystallite sizes varying from 65.2 Å for Cu2S to 97.1 Å for WO3. The heterostructure exhibit a dense morphology with pellets and particle-like morphology closely combined, in a relatively compact assembly. The surface elemental composition indicate that the heterostructure maintain a similar atomic ratio as established during the synthesis, with a slight sulfur deficit and oxygen excess due to the annealing treatments.
The photocatalytic activity toward S-MCh herbicide removal was evaluated using UVa and UVa–Vis irradiation scenarios. The results indicate a superior photocatalytic activity of Cu2S_TiO2_WO3, compared with Cu2S_TiO2, Cu2S_WO3, TiO2_WO3 or bare components. Under UVa–Vis irradiation the Cu2S_TiO2_WO3 heterostructure reach 61% photocatalytic efficiency with a constant rate of 0.1140 s−1 (2.5x faster than TiO2_WO3 and 2.7x faster than Cu2S_WO3). The higher photocatalytic activity of Cu2S_TiO2_WO3 sample was attributed to the formation of dual S-scheme mechanism, allowing the increase of photogenerated charge carrier concentration, efficient charge separation and migration.
Future studies will be dedicated to the improvement of the photocatalytic efficiency using two approaches: (1) the optimization of heterostructure components ratio in order to increase the concentration of oxidative species, and (2) the replacement of one components in order to maximize the photons conversion during the photocatalytic activity. Both ways will preserve the S-mechanisms which seems to be the most favorable for resilient organic pollutants molecules. Additionally, HPLC studies will be included to investigate the formation of potential pollutant bi-products during the photocatalytic activity.

Author Contributions

Conceptualization, A.E.; methodology, A.E.; validation, L.I.; formal analysis, L.I.; investigation, A.E.; resources, L.I.; data curation, L.I.; writing—original draft preparation, A.E.; writing—review and editing, A.E.; visualization, L.I.; supervision, L.I.; project administration, A.E.; funding acquisition, A.E. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by a grant of the Romanian Ministry of Education 309 and Research, CCCDI-UEFISCDI, project number PN-III-P2-2.1-PED-2019-2028, within PNCDI III.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data presented in this study are available by requesting from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Matheus, M.C.; Lourenço, G.R.; Solano, B.A.; Dezotti, M.W.C.; Bassin, J.P. Assessing the impact of hydraulic conditions and absence of pretreatment on the treatability of pesticide formulation plant wastewater in a moving bed biofilm reactor. J. Water Proc. Eng. 2020, 36, 101243. [Google Scholar] [CrossRef]
  2. Alistea, M.; Garridoa, I.; Floresa, P.; Hellína, P.; Velab, N.; Navarroc, S.; Fenoll, J. Reclamation of agro-wastewater polluted with thirteen pesticides by solar photocatalysis to reuse in irrigation of greenhouse lettice grown. J. Environ. Manag. 2020, 266, 110565. [Google Scholar] [CrossRef] [PubMed]
  3. Xiao, Z.; Wang, L.; Zhang, Y.; Wang, Y.; Saleha, A.S.M.; Zhu, M.; Gao, Y.; Xu, C.; Hassana, M.E.; Yang, Q.; et al. Synthesis and characterization of a novel rice bran protein-cerium complex for the removal of organophosphorus pesticide residues from wastewater. Food Chem. 2020, 320, 126604. [Google Scholar]
  4. Fang, H.; Zhang, H.; Han, L.; Mei, J.; Ge, Q.; Long, Z.; Yu, Y. Exploring bacterial communities and biodegradation genes in activated sludge from pesticide wastewater treatment plants via metagenomic analysis. Environ. Pollut. 2018, 243, 1206–1216. [Google Scholar] [CrossRef]
  5. Sharma, L.; Kakkar, R. Magnetically retrievable one-pot fabrication of mesoporous magnesium ferrite (MgFe2O4) for the remediation of chlorpyrifos and real pesticide wastewater. J. Environ. Chem. Eng. 2018, 6, 6891–6903. [Google Scholar] [CrossRef]
  6. Ghodsi, J.; Rafati, A.A. A voltammetric sensor for diazinon pesticide based on electrode modified with TiO2 nanoparticles covered multi walled carbon nanotube nanocomposite. J. Electroanal. Chem. 2017, 807, 1–9. [Google Scholar] [CrossRef]
  7. Fernández-Domene, R.M.; Sánchez-Tovar, R.; Lucas-Granados, B.; Muñoz-Portero, M.J.; García-Antón, J. Elimination of pesticide atrazine by photoelectrocatalysis using a photoanode based on WO3 nanosheets. Chem. Eng. J. 2018, 350, 1114–1124. [Google Scholar] [CrossRef]
  8. Huang, X.; Wang, L.; Sun, Y.; Meng, F.; Liu, J. Quantitative analysis of pesticide residue based on the dynamic response of a single SnO2 gas sensor. Sens. Acuat. B 2004, 99, 330–335. [Google Scholar] [CrossRef]
  9. Honarnezhad, R.; Fathinia, M.; Khataee, A. Mechanical production and sonocatalytic application of Cu2S nanoparticles for degradation of isopropylxanthic acid: Kinetic modeling via white and black box methods. J. Mol. Liq. 2019, 287, 110899. [Google Scholar] [CrossRef]
  10. Tunesi, M.M.; Kalwar, N.; Abbas, M.W.; Karakus, S.; Soomro, R.A.; Kilislioglu, A.; Abro, M.I.; Hallam, K.R. Functionalised CuO nanostructures for the detection of organophosphorus pesticides: A non-enzymatic inhibition approach coupled with nano-scale electrode engineering to improve electrode sensitivity. Sens. Acuat. B 2018, 260, 480–489. [Google Scholar] [CrossRef]
  11. Ramu, A.G.; Arun Kumari, M.L.; Elshikh, M.S.; Alkhamis, H.H.; Alrefaei, A.F.; Choi, D. A facile and green synthesis of CuO/NiO nanoparticles and their removal activity of toxic nitro compounds in aqueous medium. Chemosphere 2020, 271, 129475. [Google Scholar] [CrossRef]
  12. Zhu, D.; Zhou, Q. Action and mechanism of semiconductor photocatalysis on degradation of organic pollutants in water treatment: A review. Environ. Nanotech. Monit. Manag. 2019, 12, 100255. [Google Scholar] [CrossRef]
  13. Mouchaal, Y.; Enesca, A.; Mihoreanu, C.; Khelil, A.; Duta, A. Tuning the opto-electrical 350 properties of SnO2 thin films by Ag+1 and In+3 co-doping. Mater. Sci. Eng. B Adv. 2015, 199, 22–29. [Google Scholar] [CrossRef]
  14. Nicholls, T.P.; Bissember, A.C. Developments in visible-light-mediated copper photocatalysis. Tetrahedron Lett. 2019, 60, 150883. [Google Scholar] [CrossRef]
  15. Sheng, Y.; Wei, Z.; Miao, H.; Yao, W.; Li, H.; Zhu, Y. Enhanced organic pollutant photodegradation via adsorption/photocatalysis synergy using a 3D g-C3N4/TiO2 free-separation photocatalyst. Chem. Eng. J. 2019, 370, 287–294. [Google Scholar] [CrossRef]
  16. Meng, S.; Zhang, J.; Chen, S.; Zhang, S.; Huang, W. Perspective on construction of heterojunction photocatalysts and the complete utilization of photogenerated charge carriers. Appl. Surf. Sci. 2019, 476, 982–992. [Google Scholar] [CrossRef]
  17. Deng, H.; Fei, X.; Yang, Y.; Fan, J.; Yu, J.; Cheng, B.; Zhang, L. S-scheme heterojunction based on p-type ZnMn2O4 and n-type ZnO with improved photocatalytic CO2 reduction activity. Chem. Eng. J. 2021, 409, 127377. [Google Scholar] [CrossRef]
  18. Li, J.; Li, M.; Jin, Z. 0D CdxZn1-xS and amorphous Co9S8 formed S-scheme heterojunction boosting photocatalytic hydrogen evolution. Mol. Catal. 2021, 501, 111378. [Google Scholar] [CrossRef]
  19. Pham, T.T.; Shin, E.W. Thermal formation effect of g-C3N4 structure on the visible light driven photocatalysis of g-C3N4/NiTiO3 Z-scheme composite photocatalysts. Appl. Surf. Sci. 2018, 447, 757–766. [Google Scholar] [CrossRef]
  20. Chang, N.; Chen, Y.R.; Xie, F.; Liu, Y.P.; Wang, H.T. A promising Z-scheme heterojunction via loading Ag/AgCl into porous Co3O4 derived from ZIF-67 for visible light driven photocatalysis. Micropor. Mesopor. Mat. 2020, 307, 110530. [Google Scholar] [CrossRef]
  21. Dursun, S.; Koyuncu, S.N.; Kaya, I.C.; Kaya, G.G.; Kalem, V.; Akyildiz, H. Production of CuO–WO3 hybrids and their dye removal capacity/performance from wastewater by adsorption/photocatalysis. J. Water Proc. Eng. 2020, 36, 101390. [Google Scholar] [CrossRef]
  22. Ojha, D.P.; Karki, H.P.; Song, J.H.; Kim, H.J. Amine-assisted synthesis of FeWO4 nanorodg-C3N4 for enhanced visible light-driven Z-scheme photocatalysis. Compos. Part B Eng. 2019, 160, 277–284. [Google Scholar] [CrossRef]
  23. Li, Y.; He, R.; Han, P.; Hou, B.; Peng, S.; Ouyang, C. A new concept: Volume photocatalysis for efficient H2 generation using low polymeric carbon nitride as an example. Appl. Cat. B Environ. 2020, 279, 119379. [Google Scholar] [CrossRef]
  24. Chen, H.J.; Yang, Y.L.; Hong, M.; Chen, J.G.; Suo, G.Q.; Hou, X.J.; Feng, L.; Chen, Z.G. Separable and recyclable meso-carbon@TiO2/carbon fiber composites for visible-light photocatalysis and photoelectrocatalysis. Sustain. Mater. Technol. 2019, 21, e00105. [Google Scholar] [CrossRef]
  25. Li, X.; Hu, T.; Zhang, J.; Dai, K.; Liang, C. Novel 2D SnNb2O6/Ag3VO4 S-scheme heterojunction with enhanced visible-light photocatalytic activity. Ceram. Int. 2021, 47, 7169–7176. [Google Scholar] [CrossRef]
  26. Wang, R.; Shen, J.; Zhang, W.; Liu, Q.; Zhang, M.; Tang, H. Build-in electric field induced step-scheme TiO2/W18O49 heterojunction for enhanced photocatalytic activity under visible-light irradiation. Ceram. Int. 2020, 46, 23–30. [Google Scholar] [CrossRef]
  27. Hu, X.; Wang, G.; Wang, J.; Hu, Z.; Su, Y. Step-scheme NiO/BiOI heterojunction photocatalyst for rhodamine photodegradation. Appl. Surf. Sci. 2020, 511, 145499. [Google Scholar] [CrossRef]
  28. Xu, A.J.; Tu, W.; Shen, S.; Lin, Z.; Gao, N.; Zhong, W. BiVO4@MoS2 core-shell heterojunction with improved photocatalytic activity for discoloration of Rhodamine B. Appl. Surf. Sci. 2020, 528, 146949. [Google Scholar] [CrossRef]
  29. Pei, C.Y.; Chen, Y.G.; Wang, L.; Chen, W.; Huang, G.B. Step-scheme WO3/CdIn2S4 hybrid system with high visible light activity for tetracycline hydrochloride photodegradation. Appl. Surf. Sci. 2021, 535, 147682. [Google Scholar] [CrossRef]
  30. Wei, J.; Chen, Y.; Zhang, H.; Zhuang, Z.; Yu, Y. Hierarchically porous S-scheme CdS/UiO-66 photocatalyst for efficient 4-nitroaniline reduction. Chin. J. Catal. 2021, 42, 78–86. [Google Scholar] [CrossRef]
  31. Guan, R.; Zhai, H.; Li, J.; Qi, Y.; Li, M.; Song, M.; Zhao, Z.; Zhang, J.; Wang, D.; Tan, H. Reduced mesoporous TiO2 with Cu2S heterojunction and enhanced hydrogen production without noble metal cocatalyst. Appl. Surf. Sci. 2019, 507, 144772. [Google Scholar] [CrossRef]
  32. Singh, S.V.; Sharma, A.; Biring, S.; Pal, B.N. Solution processed Cu2S/TiO2 heterojunction for visible-near infrared photodetector. Thin Solid Films 2020, 710, 138275. [Google Scholar] [CrossRef]
  33. Baviskar, V.; Salunkhe, D.; Tarkas, H.; Wagh, R.; Baviskar, P.; Patil, R. Sensitization of TiO2 by chemically deposited Cu2S for solar cell: Effect of deposition time on photoelectrochemical performance. Optik 2020, 207, 163890. [Google Scholar] [CrossRef]
  34. Rao, V.N.; Reddy, N.L.; Kumari, M.M.; Ravi, P.; Sathish, M.; Kuruvilla, K.M.; Preethi, V.; Reddy, K.R.; Shettie, N.P.; Aminabhavi, T.M.; et al. Photocatalytic recovery of H2 from H2S containing wastewater: Surface and interface control of photo-excitons in Cu2S@TiO2 core-shell nanostructures. Appl. Catal. B Environ. 2019, 254, 174–185. [Google Scholar]
  35. Baneto, M.; Enesca, A.; Mihoreanu, C.; Lare, Y.; Jondo, K.; Napo, K.; Duta, A. Effects of the growth temperature on the properties of spray deposited CuInS2 thin films for photovoltaic applications. Ceram. Int. 2015, 41, 4742–4749. [Google Scholar] [CrossRef]
  36. Dudita, M.; Bogatu, C.; Enesca, A.; Duta, A. The influence of the additives composition and concentration on the properties of SnOx thin films used in photocatalysis. Mater. Lett. 2011, 65, 2185–2189. [Google Scholar] [CrossRef]
  37. Zhang, L.; Zhao, Y.; Zhong, L.; Wang, Y.; Chai, S.; Yang, T.; Han, X. Cu2S-Cu-TiO2 mesoporous carbon composites for the degradation of high concentration of methyl orange under visible light. Appl. Surf. Sci. 2017, 422, 1093–1101. [Google Scholar] [CrossRef]
  38. Tang, L.; Deng, Y.; Zeng, G.; Hu, W.; Wang, J.; Zhou, Y.; Wang, J.; Tang, J.; Fang, W. CdS/Cu2S co-sensitized TiO2 branched nanorod arrays of enhanced photoelectrochemical properties by forming nanoscale heterostructure. J. Alloy. Compound. 2016, 662, 516–527. [Google Scholar] [CrossRef]
  39. Marín-Benito, J.M.; Herrero-Hernández, E.; Ordaxa, J.M.; Sánchez-Martín, M.J.; Rodríguez-Cruz, M.S. The role of two organic amendments to modify the environmental fate of S-metolachlor in agricultural soils. Environ. Res. 2021, 195, 110871. [Google Scholar] [CrossRef] [PubMed]
  40. Yang, L.; Ivantsova, E.; Souders, C.L., II; Martyniuk, C.J. The agrochemical S-metolachlor disrupts molecular mediators and morphology of the swim bladder: Implications for locomotor activity in zebrafish (Danio rerio). Ecotox. Environ. Safe 2021, 208, 111641. [Google Scholar] [CrossRef]
  41. Sithole, R.K.; Machogo, L.F.E.; Moloto, M.J.; Gqoba, S.S.; Mubiayi, K.P.; Wyk, J.V.; Moloto, N. One-step synthesis of Cu3N, Cu2S and Cu9S5 and photocatalytic degradation of methyl orange and methylene blue. J. Photochem. Photobiol. A Chem. 2020, 397, 112577. [Google Scholar] [CrossRef]
  42. Turchi, C.S.; Ollis, D.F. Photocatalytic degradation of organic water contaminants: Mechanisms involving hydroxyl radical attack. J. Catal. 1990, 122, 178–192. [Google Scholar] [CrossRef]
  43. Xu, F.; Costa, M.F.; Sun, S.; Teixeira, V.; Wang, Y.; Bian, F.; Wang, H.; Cui, H. Structure Optimization on the Photoelectric and Photocatalytic Properties of Cu2S and ZnO Complex Films. Mater. Today Proc. 2015, 2, 253–260. [Google Scholar] [CrossRef] [Green Version]
  44. Herrmann, J.M.; Guillard, C.; Pichat, P. Heterogeneous photocatalysis: An emerging technology for water treatment. Catal. Today 1993, 17, 7–20. [Google Scholar] [CrossRef]
  45. Gao, C.; Li, J.; Shan, Z.; Huang, F.; Shen, H. Preparation and visible-light photocatalytic activity of In2S3/TiO2 composite. Mater. Chem. Phys. 2010, 122, 183–187. [Google Scholar] [CrossRef]
  46. Mise, T.; Nakada, T. Low temperature growth and properties of Cu–In–Te based 433 thin films for narrow bandgap solar cells. Thin Solid Films 2010, 518, 5604–5609. [Google Scholar] [CrossRef]
  47. Liu, Y.; Gong, Z.; Lv, H.; Ren, H.; Xing, X. Rational design of Au decorated Mn0.5Cd0.5S/WO3 step-scheme heterostructure with multichannel charge transfer and efficient H2 generation. Appl. Surf. Sci. 2020, 526, 146734. [Google Scholar] [CrossRef]
  48. Jia, X.; Han, Q.; Liu, H.; Li, S.; Bi, H. A dual strategy to construct flowerlike S-scheme BiOBr/BiOAc1−xBrx heterojunction with enhanced visible-light photocatalytic activity. Chem. Eng. J. 2020, 399, 125701. [Google Scholar] [CrossRef]
  49. He, R.; Liu, H.; Liu, H.; Xu, D.; Zhang, L. S-scheme photocatalyst Bi2O3/TiO2 nanofiber with improved photocatalytic performance. J. Mater. Sci. Technol. 2020, 52, 145–151. [Google Scholar]
  50. Enesca, A.; Isac, L.; Duta, A. Charge carriers injection in tandem semiconductors for dyes mineralization. Appl. Catal. B. 2015, 162, 352–363. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of the heterostructures.
Figure 1. XRD patterns of the heterostructures.
Materials 14 02231 g001
Figure 2. Scanning electron microscopy images of (a) Cu2S, (b) TiO2, (c) WO3, (d) Cu2S_TiO2, (e).Cu2S_WO3, (f) TiO2_WO3, and (g) Cu2S_TiO2_WO3.
Figure 2. Scanning electron microscopy images of (a) Cu2S, (b) TiO2, (c) WO3, (d) Cu2S_TiO2, (e).Cu2S_WO3, (f) TiO2_WO3, and (g) Cu2S_TiO2_WO3.
Materials 14 02231 g002aMaterials 14 02231 g002b
Figure 3. Photocatalytic removal efficiency under (a) UVa and (b) UVa–Vis irradiation.
Figure 3. Photocatalytic removal efficiency under (a) UVa and (b) UVa–Vis irradiation.
Materials 14 02231 g003
Figure 4. Kinetic evaluation of the photocatalytic activity under (a) UVa and (b) UVa–Vis irradiation.
Figure 4. Kinetic evaluation of the photocatalytic activity under (a) UVa and (b) UVa–Vis irradiation.
Materials 14 02231 g004
Figure 5. (a) Band energy diagram of the S-scheme heterostructure and the corresponding (b) Cu2S, (c) TiO2, (d) WO3, and (e) Cu2S_TiO2_WO3 band gaps.
Figure 5. (a) Band energy diagram of the S-scheme heterostructure and the corresponding (b) Cu2S, (c) TiO2, (d) WO3, and (e) Cu2S_TiO2_WO3 band gaps.
Materials 14 02231 g005
Table 1. Irradiation scenarios for the photocatalytic applications.
Table 1. Irradiation scenarios for the photocatalytic applications.
Irradiation SourcesUV Radiation Sources
(310–390 Nm)
Vis Radiation Sources
(400–800 Nm)
Total Irradiance (W/M2)
UVa light, 18 W8016.3
UVa–Vis light, 18 W4424.9
Table 2. Average crystallite size values calculated using the Scherrer formula.
Table 2. Average crystallite size values calculated using the Scherrer formula.
Samples CodeCrystallite Size (Å)
Cu2STiO2WO3
Cu2S_TiO26482-
Cu2S_WO379-103
TiO2_WO3-8196
Cu2S_TiO2_WO3658397
Table 3. Average atomic composition at the heterostructure surface (EDS).
Table 3. Average atomic composition at the heterostructure surface (EDS).
SchemeElemental Composition (% at)
CuTiWOOth 1SSth 1
Cu2S_TiO223.722.6-45.445.28.311.8
Cu2S_WO314.2-16.762.550.16.67.1
TiO2_WO3-13.812.174.163.9--
Cu2S_TiO2_WO314.39.89.659.948.46.47.1
1 Theoretic content calculated based on the stoichiometry.
Table 4. Kinetic data corresponding to S-MCh photocatalytic removal.
Table 4. Kinetic data corresponding to S-MCh photocatalytic removal.
Kinetic DataCu2STiO2WO3Cu2S_TiO2Cu2S_WO3TiO2_WO3Cu2S_TiO2_WO3
K [S−1]R2K [S−1]R2K [S−1]R2K [S−1]R2K [S−1]R2K [S−1]R2K [S−1]R2
UV0.01480.99580.03160.99640.02350.99920.03460.99710.02560.99840.04350.99660.05400.9977
UV–Vis0.02430.99620.03340.99610.02370.99720.04690.99490.04190.99670.04480.99640.11400.9899
Table 5. Heterostructures kinetic parameters for both radiation scenarios based on Equation (6).
Table 5. Heterostructures kinetic parameters for both radiation scenarios based on Equation (6).
Photocatalyst, Pollutantkr·108
(Mol·L−1·Min−1)
Ks·102
(Mol·L−1)
Cu2S_TiO2, UVa1.63624
Cu2S_TiO2, UVa–Vis2.941682
Cu2S_WO3, UVa1.42556
Cu2S_WO3, UVa–Vis2.981625
TiO2_WO3, UVa2.681044
TiO2_WO3,UVa–Vis2.731592
Cu2S_TiO2_WO3, UVa3.191788
Cu2S_TiO2_WO3, UVa–Vis5.262389
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Enesca, A.; Isac, L. Tuned S-Scheme Cu2S_TiO2_WO3 Heterostructure Photocatalyst toward S-Metolachlor (S-MCh) Herbicide Removal. Materials 2021, 14, 2231. https://doi.org/10.3390/ma14092231

AMA Style

Enesca A, Isac L. Tuned S-Scheme Cu2S_TiO2_WO3 Heterostructure Photocatalyst toward S-Metolachlor (S-MCh) Herbicide Removal. Materials. 2021; 14(9):2231. https://doi.org/10.3390/ma14092231

Chicago/Turabian Style

Enesca, Alexandru, and Luminita Isac. 2021. "Tuned S-Scheme Cu2S_TiO2_WO3 Heterostructure Photocatalyst toward S-Metolachlor (S-MCh) Herbicide Removal" Materials 14, no. 9: 2231. https://doi.org/10.3390/ma14092231

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