Next Article in Journal
Recent Progress in the Mechanism and Engineering of α/β Hydrolases for Chiral Chemical Production
Next Article in Special Issue
Enhanced Photocatalytic and Anticancer Activity of Zn-Doped BaTiO3 Nanoparticles Prepared through a Green Approach Using Banana Peel Extract
Previous Article in Journal
Investigation of Different Aqueous Electrolytes for Biomass-Derived Activated Carbon-Based Supercapacitors
Previous Article in Special Issue
Influence of Rapid Heat Treatment on the Photocatalytic Activity and Stability of Strontium Titanates against a Broad Range of Pollutants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hydrothermal Co-Crystallization of Novel Copper Tungstate-Strontium Titanate Crystal Composite for Enhanced Photocatalytic Activity and Increased Electron–Hole Recombination Time

Department of Physical Chemistry and Materials Sciences, University of Szeged, Rerrich B. Sqr.1, H-6720 Szeged, Hungary
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(2), 287; https://doi.org/10.3390/catal13020287
Submission received: 30 December 2022 / Revised: 20 January 2023 / Accepted: 24 January 2023 / Published: 27 January 2023

Abstract

:
The development of catalysts continues to have a significant influence on science today since we can utilize them to efficiently destroy some contaminants. A study in this field is justified because there is a dearth of comprehensive literature on the creation of SrTiO3-based photocatalysts. Related to this topic, here we report the facile preparation of a structure-modified SrTiO3 photocatalyst, by incorporating CuWO4. Within the case of the CuWO4-modified samples (0.5–3 wt% nominal CuWO4 content), the photo-oxidation of phenol, as a contaminant, was more than two times higher than the initial SrTiO3. However, the photocatalytic activity does not change linearly with increasing CuWO4 content, and the CWS2.5 (2.5 wt% nominal CuWO4 content and 4.25 wt% measured content) has the highest photo-activity under the applied conditions. The reason for the better activity was the increased recombination time of charge separation on the catalyst surface. Slower recombination can result in more water being oxidized to hydroxyl radicals, leading to the faster decomposition of the phenol.

1. Introduction

The study of photocatalysts is still a popular field of research today, because their good properties can be used in many fields, e.g., antimicrobial treatment, air cleaning and water purification [1,2,3,4,5]. Most photocatalysts are metal nanoparticles or semiconducting metal oxides, such as strontium titanate (SrTiO3), which is widely studied nowadays. SrTiO3 has many useful advantages such as low cost, inertness, stability and relatively high photocatalytic activity in the UV region. It is justified to deal with this material because the photocatalytic activity can be increased after various modifications. We have several options for increasing the photocatalytic activity, such as reducing the band gap (excitation with lower energy photons), increasing the number of charge separations created per unit time or increasing the recombination time (more reactive radicals are produced) [6,7]. The long-term goal is to learn more about the nature of modifications where we do not achieve higher photocatalytic activity by reducing the band gap but by increasing the recombination time so that we can later enhance this in the best way.
Heterogeneous photocatalysis is an advanced oxidation process (other AOPs: homogeneous photocatalysis, photolysis, Gamma-radiolysis) [8]. In the case of heterogeneous photocatalysis, the photocatalyst can be excited with photons, and this induces a hole–electron pair through the change in the energy state of the electron (the electron energy moves from the valence band to the conduction band) [9]. The residual hole is the oxidation side, where the hole oxidizes the adsorbed molecules (usually water molecules), and the result is the generation of reactive radicals. The excited electron is on the reduction side, where some molecules are able to accept electrons, which results in a negative oxidation potential (so it will have a reductive characteristic). Through the processes described, reactive oxygen species could be generated, e.g., hydroxyl radical, H2O2, peroxide radical, hydroxyl anion and superoxide anion [10]. One option is to reduce the band gap (excitation with visible light), while another option is to increase the recombination time (more reactive radicals are produced).
We have three useful ways for modification: decoration, which is when the surface of the photocatalyst is modified (e.g., metal deposition), doping, which is when the crystalline structure is modified, and morphological modifications [11,12,13]. During the surface modification, metals or metal oxides are usually deposited on the surface, and some transition metals or transition metal oxides (Ag, Au, CuO) have a plasmonic effect on the surface of the catalyst, which also increases the recombination time [13]. Doping photocatalysts with various elements (e.g., Ag, Ru, C, N, Cu) can reduce the band gap energy, and, therefore, the photocatalyst could be activated with a photon with lower energy, and the dopant can act as an electron trap, which slows down the recombination of the electron–hole pairs, which also favors photocatalytic activity [12]. In the case of morphological modification, generally, the shape and porosity are modified [11]. In the scientific literature, we can find many articles on modified SrTiO3. The comparison of photocatalytic activities measured by others in other labs is difficult, due to the different conditions, e.g., the light source, the concentration of the catalyst, the reactor, the construction of the system, the model pollutants and its concentration are different. For example, Syafiuddin et al. doped the SrTiO3 with sulfur and nitrogen (separated); in the former case, the catalyst decomposed ~12% of the phenol and, in the latter case, ~10% under 2 h [14]. Another study investigated the decomposition of Rhodamine B on a graphene oxide/strontium titanate composite and on pure SrTiO3; the former decomposed 94.5% of the pollutant and the latter 64.5% under the applied conditions [15]. According to another article, CuOx-modified SrTiO3 was used to investigate the degradation of phenol, achieving a degradation efficiency of ~19% [7].
In this work, we modified the SrTiO3 photocatalyst with different amounts of copper-tungstate (CuWO4) by a hydrothermal co-crystallization method, and the CuWO4 content was optimized. This is novel because CuWO4 has not yet been used to modify SrTiO3. Structure and optical characterization methods were applied to prove the success of the photocatalyst preparation and to determine the properties of the prepared photocatalysts. Next, we determined the photocatalytic activity for all the prepared photocatalysts under UV illumination following the degradation of the phenol test molecule and the ability to produce hydroxyl radicals through coumarin degradation. The results show that the CuWO4-modified SrTiO3 has higher photocatalytic activity than pure SrTiO3.

2. Results and Discussion

2.1. Characterization

The crystal phase of the obtained photocatalysts was determined by XRD measurements. Figure 1A shows the diffractogram of the pure SrTiO3 (CWS0) and the composite samples with increased CuWO4 content (CWSX). The patterns of the diffractograms are the same, and therefore the crystal structure is the same. However, the main reflection shifts slightly to the left (from 2Θ = 32.48° to 32.36°), like a trend, which suggests that the distance between the corresponding crystal faces has changed (Figure 1C). This is due to the presence of a crystalline foreign material [16], which in our case is probably W6+ instead of Ti4+ from CuWO4, because their ionic radius is similar (0.61 Å for Ti4+ and 0.60 Å for W6+), and by increasing the proportion of CuWO4, the crystal faces get closer and closer. Furthermore, the SEM pictures taken of some samples show that with the increasing CuWO4 content the surface of the catalyst becomes flaky (Figure S1A–C), while the EDX mapping indicates the increasing Cu content (Figure S1E,F). The trend of the shift is broken for the CWS3 sample, which indicates overdoping, so CuWO4 is present in a larger amount than can be integrated into the lattice; this also affects the crystallization, and probably amorphous CuWO4 remains on the sample. This phenomenon is also clearly visible when the results are presented later. As a check, we crystallized pure copper tungstate (Figure 1B), and its XRD pattern is different from the composite samples; hence, in the absence of SrTiO3, it crystallizes in a different crystal phase.
The results indicate that during co-crystallization, they crystallize in the same crystal structure, in the crystal structure of SrTiO3; that is, the crystallization of SrTiO3 affects the crystallization of CuWO4. There could be several reasons for this; for example, the crystallization energy of SrTiO3 [17], which is two orders of magnitude larger, affects the crystallization of CuWO4, or more crystal sources are formed from SrTiO3 and a composite crystal is formed, which would explain the reflection shift due to the different ion sizes [16]. The CWSX sample peaks belong to the SrTiO3 ((100), (110), (111), (200), (210), (211), (220), (300), (310) [18]), and these peaks were assigned with the help of the SrTiO3 JCPDS card (35-0734). The CuWO4 sample peaks were identified based on the CuWO4 JCPDS card (70-1732): (110), (111), (111*), (021), (012), (102), (220), (211), (022), (202), (003), (301) [19].
Furthermore, the CuWO4-modified samples were turquoise in color (the color intensifies with the CuWO4 content), which also indicates the successful modification of the initial white strontium titanate. The light absorption of the CWSX samples was measured by the diffuse reflectance spectroscopy (DRS) method (Figure 2). The band gap energy values were determined by the Kubelka–Munk method [20], and the result was around 3.20 eV for each sample. The band gap slightly changed, but not significantly, and therefore the increased photocatalytic activity (see later) has another reason, not the decreased band gap. Therefore, the incorporation of CuWO4 into the structure of SrTiO3 did not significantly affect (in the case of the samples produced in this work) the energy dependence of the excitability. It can also be observed that the spectra of the CWS3 sample with the highest functionalization degree are slightly different from the others, which is certainly due to overdoping, since more CuWO4 is present than can be incorporated into the lattice.
By measuring the photoluminescence, we can get information about the recombination after excitation [21], and we can compare the samples in the energy consumption received from the photon. On the basis of the measured emission spectra, it can be seen that the presence of CuWO4 reduced the degree of recombination, and it can also be seen that the more CuWO4, the more the recombination decreases (up to the level of copper tungstate concentrations used); the normalized results are in Figure 3. This means that the excited electron can return to the valence band in other ways, which is the reason for the increased photocatalytic activity (applied conditions). Due to the closeness of the potentials, the excited electron of strontium titanate (CB of SrTiO3 is also 3.4 eV [22]) can be transferred to the valence band of CuWO4 (VB of CuWO4 is also 2.3 eV [23]), which is possible because the outer electron of CuWO4 is also excited at the same time as the excitation of SrTiO3, so a positive hole remains in its valence band. The excited electron of CuWO4 is therefore more difficult to recombine, and it is much more likely to be taken up by oxygen, because it has more time for that, after which a hole remains at the end of the “chain” (SrTiO3 VB), which produces a hydroxyl radical through the oxidation of water [24]. In the case of pure SrTiO3, the dissolved oxygen (or other molecules from the medium) can only pick up the electron from the SrTiO3, so the electron has two options: it recombines or is picked up by a molecule of the medium. However, the CuWO4-SrTiO3 composite has a third option when the electron is switching to CuWO4 (which can happen, e.g., BiVO4-based photocatalysts [25]). Because of this, the possibility of recombination is statistically minimally decreased [26], which is clearly visible on the normal photoluminescence spectra (the measured intensity is reduced). Therefore, electron transfer and hydroxyl radical production also take place in the case of SrTiO3 (proven later in Section 2.3), but in the presence of CuWO4, there are more possibilities for this in a unit time with minimally suppressed recombination. This is the reason for the increased photocatalytic activity.
The shape of the emission spectra is very similar for CWS0 and the other CWSX samples, which also shows (in addition to XRD) the relationship between the two substances, so it is not a physical mixture, and they did not crystallize next to each other but crystallized “together” (co-crystallization). Another reason that the shape of the spectra is very similar is that the CuWO4 concentration used is very low, and the resulting tungsten concentration is therefore even lower. In the case of doping, the foreign element is usually not integrated in full amount, and in our case, the change is very small.
The specific surface area values are presented in Table 1. The samples have similar surface areas; each sample was measured three times, and the difference is comparable to the value of the standard deviations (Table 1). Thus, the specific surface area of the samples is approximately the same (around 40 m2/g), which also proves that it is the same structure and probably co-crystallization, namely CuWO4 incorporated into the structure of SrTiO3, and the result is composite crystallites. Due to the similarity of the values, the crystallite size probably did not change significantly with the modified structure.
Figure 4 shows the zeta-potential values of the CWSX samples; the standard deviation was calculated from five measurements. Compared to the initial SrTiO3 (CWS0) (ζ= −30.3 ± 1.6), the zeta potential of functionalized samples slightly decreased (~−33–38 mV); however, CWS3 with the highest CuWO4 content had a significantly lower value (−38.1 mV) because of the above-mentioned overdoping. Further increasing the ratio would probably further decrease the zeta potential, which would favor the adsorption of a positively charged pollutant, but in this work, we did not examine this observation, because we used phenol as a model pollutant and we examined how the photocatalytic activity changed for phenol; in any case, this information can also be useful for the literature and other researchers. The lower surface charge increases the stability of the suspension when the sample is used as a suspension. The pH was around 6.50 for each sample.
The measured amount of CuWO4 presented in a weight ratio is proven by EDX measurements; for demonstration, the CWS3 sample EDX spectrum is in Figure 5. The measured values do not always correspond to the nominal (theoretical amount of CuWO4) values but follow a well-managed trend; the results are in Table 1. It can also be read from the spectra that it does not contain significant amounts of other elements. The ratio of CuWO4 in the measured samples is probably higher than the theoretical value, because during the synthesis, the SrTiO3 precursors remained on the wall of the vessel during the pouring from one vessel to another and also during pouring from the vessel into the autoclave.

2.2. Photocatalytic Activity Measurements

The conditions during the photocatalytic experiments were the same for all samples. For reference, the CWS0 sample was used, in which the concentration of CuWO4 was 0 wt%. The results are summarized in Figure 6, and almost all of the synthesized photocatalyst samples had higher photocatalytic activity than the reference CWS0 (pure SrTiO3). CWS2.5 exhibited the highest photocatalytic activity (in Figure 6A are the decay curves, and in Figure 6B is the degraded amount of phenol). The reference CWS0 degraded only a small amount of phenol, and the conversion rate was ~10% (1.063 mgpheonl/gcat). In the case of CWSX samples, significantly better results were obtained, with conversion rates of ~10–24%. The highest value, which belongs to the CWS2.5, was 24% (2.340 mgpheonl/gcat). Because neither the band gap nor the surface charge exhibits an “extreme” value in this sample, the superior photocatalytic activity has to be explained in another way. Because hydroxyl radicals are involved in the conversion of phenol [27], their quantity in our system can be crucial. According to the XRD results, the trend of the main reflection shift indicates doping (see Section 2.1), which means a small structural change. This trend continues linearly to the CWS2.5 sample and then breaks in the last sample (CWS3), due to overdoping, which has a negative effect on the photocatalytic activity. Due to the mentioned reasons, fewer pollutants can be oxidized on the hole side after the charge separation (direct effect), and fewer hydroxyl radicals can be formed by the oxidation of water. The next section discusses this occurrence through the hydroxyl radical production. The photocatalytic activity does not follow the usual trend, the effect of the modification is not immediately visible in the first member of the modified sample series. The increase in photocatalytic activity starts from the third member (CWS1.5, 1.5 wt.% nominal and 2.16 wt.% measured CuWO4 content) of the modified sample series and then “breaks off” at the last member. This can be explained that at the beginning when the doping is so small, it cannot yet have a significant effect on the processes after charge separation; however, the later samples already have this “threshold concentration”. For the last sample, the explanation may be the phenomenon of overdoping described earlier. The results show that, under certain conditions, the CWS2.5 sample has the optimal CuWO4 content (measured to be 4.28% by weight) to achieve maximum phenol degradation activity.

2.3. Detection of Hydroxyl Radical

Without a catalyst, the 7-hydroxyoumarin molecule was not detectable, and only the hydroxyl radicals created during the photocatalytic reaction can provide 7-hydroxycoumarin [28]. The measurements were performed on all samples, and in Figure 7, the 7-hydroxycoumarin formation kinetic can be seen. The formation kinetics of 7-hydroxycoumarin is the highest in the case of CWS2.5; the corresponding rate constant (k7-hydroxycoumarin) was 1.52 × 10−5 µmol·L−1·s−1 which is an order of magnitude greater than the reference (CWS0) sample (5.29 × 10−6 µmol·L−1·s−1). The mentioned values were calculated based on the measurements presented in Figure 7. The trend closely follows the trend obtained for photocatalytic efficiency, for the same reason as described in the previous section (Section 2.2).

3. Materials and Methods

3.1. Catalyst Preparation

Analytical grade chemicals were used for synthesis without further purification: Sr(NO3)2 (Molar Chemicals, Halásztelek, Hungary; >99%), titanium isopropoxid (C12H28O4Ti) (Sigma Aldrich, St. Louis, MO, USA; >99%), Cu(NO3)2 × 3 H2O (Molar Chemicals, Halásztelek, Hungary; >99%), Na2WO4 × 2 H2O (Chinoin, Budapest; >99%), KOH (Molar Chemicals, Halásztelek, Hungary; >99%) and ultrapure Millipore Milli-Q water (Budapest, Hungary). CuWO4-modified strontium titanate (CWSX) samples were prepared using a hydrothermal co-crystallization procedure, which is a relatively cheap and resizable method. The synthesis was performed by combining two recipes [16,29]. First modified sample, CWS0.5, was prepared as follows: First, we prepared solution A: 2.842 g C12H28O4Ti dissolved in 50 mL 2-propanol. Solution B: 2.111 g Sr(NO3)2 was dissolved in 50 mL H2O. Dispersion C: 0.00922 g of CuWO4 dispersion was precipitated by adding the necessary copper and tungsten precursors (Cu(NO3)2 × 3 H2O and Na2WO4 × 2 H2O) in stoichiometric amounts to 10 mL of water. Solution D: 50 mL 0.4 mol/L KOH solution. During the synthesis, dispersion C was added to solution B, and then the obtained dispersion E was added to solution A dropwise under vigorous stirring (dispersion F). Next, solution D was added to dispersion F. The mixture was kept at 90 °C for 1 h, and then it was loaded into a Teflon-lined autoclave which was kept at 200 °C for 3 h. After washing and drying the photocatalyst, the sample was calcined at 500 °C for 3 h in the presence of air. The product was 1.844 g of SrTiO3 with 0.5 wt% CuWO4. In the case of the other samples, the preparation procedure was the same, with the only difference being that we systematically prepared more CuWO4 precipitates; the values are in Table 1. CWS0 was the reference; no CuWO4 precipitate was added during the synthesis. According to this procedure, we tried to carry out the synthesis with the minimum number of steps, and the energy requirement was mainly due to the double heat treatment.

3.2. Characterization

The CuWO4-SrTiO3 sample crystallinity was determined with X-ray diffraction measurements, which were carried out with a Philips X-ray diffractometer (PW 1830 generator, PW 1820 goniometer, CuKα: λ = 0.1542 nm, 40–50 kV, 30–40 mA) (Philips, Germany), from 20 to 80° (2Θ), at room temperature.
For optical characterization, the diffuse reflectance spectra were recorded with UV–VIS USB4000 type (Ocean Optics Inc., Dunedin, FL, USA), diode array spectrophotometer. The spectra were recorded from 200 to 850 nm wavelength. The band gap values were determined using the Kubelka–Munk method.
The room-temperature photoluminescence (PL) emission spectra of the samples were recorded at 300 nm excitation wavelength using a Horiba Jobin Yvon Fluoromax-4 type spectrofluorometer (Horiba, Kyoto, Japan).
Energy-dispersive X-ray spectra were measured with Hitachi S-4700 scanning electron microscope (Tokyo, Japan) with Röntec EDX detector (30 keV). We also performed the mapping measurements with this instrument.
The specific surface area values of the samples were measured by the BET method from N2 adsorption isotherms (77 ± 0.5 K) with Micromeritics Gemini 2375 Surface Area Analyzer (Micromeritics, Norcross, GA, USA).
The electrophoretic mobility of the samples was also determined using a HORIBA SZ-100 Nanoparticle Analyzer (Horiba, Kyoto, Japan); a disposable carbon-coated electrode cell was used. The zeta-potential values were determined using the Smoluchowski model, from the measured data. The measured dispersion concentration was 0.001 w/v%. In addition to the CWS samples, we also measured the pure CuWO4. The zeta-potential values provide useful information about the aqueous-phase hydrodynamic stability as well [30].

3.3. Photocatalytic Activity Measurements

The photocatalytic activity of the CWSX samples was measured in 50 mL aqueous suspension. Phenol was the model pollutant, and the experiments were performed in an opened glass reactor, at room temperature. The used UV light source (local λmax = 366 nm (Figure S2) was fixed at 5 cm distance from the surface of continuously stirred (3000 rpm) suspensions. The photocatalyst concentration in the suspension was 0.1 wt%, while the initial phenol concentration was 10 mg/L. The suspensions were irradiated for 120 min during the tests. The concentration of phenol was followed with an Agilent 1290 Infinity II liquid chromatograph with UV detector (detection wavelength was 254 nm), the column was an InfinityLab Poroshell 120 ECsingle bondC18, while the mobile phase was 50/50 acetonitrile/water with 1 mL/min flow rate. Before the irradiation, the suspensions were kept in the dark for 30 min to ensure the adsorption equilibrium. Before chromatographic measurement, the collected samples were centrifuged two times (14,500 rpm, 5 min) and filtered through 0.1 μm Millex-VV PVDF filter.

3.4. Detection of Hydroxyl Radical

The reaction between coumarin and hydroxyl radical produces 7-hydroxycoumarin (Figure 8) [28], which can be monitored spectrofluorometrically (λem, max = 453 nm), and the amount of produced 7-hydroxycoumarin depends only on the hydroxyl radical concentration. The capacity of the hydroxyl radical production was measured on reference CWS0 and the best CWSX (the best photocatalyst in this work) sample, which was CWS2.5. The experiments were preceded by the calibration of spectrofluorometer with 7-hydroxycoumarin; the concentration dependence was linear over the range used (1 × 10−8 mol/L–5 × 10−6 mol/L), R2 = 0.9999.
For measurements, the initial coumarin concentration was 1 × 10−4 mol/L, and the photocatalyst concentration was 0.1 wt%. The UV light source (Figure S2) was placed at 5 cm distance from the suspension surface. The 7-hydroxycoumarin concentration change was determined with a Horiba Jobin Yvon Fluoromax-4 type spectrofluorometer. The excitation wavelength of 7-hydroxycoumarin was 345 nm, while the detection wavelength was 453 nm. Before the measurements, the samples were centrifuged two times (14,500 rpm, 5 min) to completely remove the dispersed photocatalyst particles.

4. Conclusions

SrTiO3 photocatalysts were successfully modified by incorporating them into a CuWO4 crystal lattice through a simple co-crystallization hydrothermal method. Due to this, SrTiO3 was also doped by exchanging an element (described in the XRD section), probably the exchange of Ti4+ with W6+. The characterization results show that compared to initial SrTiO3, the structure of the composite crystallites did not change. Instead, the addition of the new material only slightly affected the spacing between some crystal faces, proving that the new material was successfully incorporated (XRD). The presence of CuWO4 could be confirmed by EDX measurements. The other characterization methods indicated that the change is minimal and does not significantly affect the extent of some properties (e.g., specific surface area (~40 m2/g), band gap (~3.2 eV)); however, the photoluminescence measurements showed that the probability of recombination decreased in the case of the modified samples.
The photocatalytic activity of the modified samples is higher than that of pure strontium titanate, as shown by the photocatalytic tests, which clearly show the weaker hole–electron recombination. Sample CWS2.5 has the highest activity, its breakdown efficiency during phenol decomposition is 2.340 mgphenol/gcat, while the reference CWS0 (pure SrTiO3) has 1.063 mgphenol/gcat, so the photocatalytic activity was more than doubled. We discovered that the origin of 7-hydroxycoumarin, which is connected to the quantity of hydroxyl radical generated, can be used to explain the cause of the rise in photocatalytic activity. The formation rate constant of 7-hydroxycoumarins in the case of CWS2.5 was 1.52 × 10−5 µmol·L−1·s−1, and in the case of CWS0, it was 5.29 × 10−6 µmol·L−1·s−1. The reason for this is the reduced probability of charge recombination (which is due to the foreign element in the lattice of SrTiO3), due to which more hydroxyl radicals are formed on the oxidation side. It can be concluded that the CWS2.5 sample has an optimal CuWO4 content (the measured content was 4.28 wt.%) in terms of photocatalytic activity, under the conditions used.
In summary, we have successfully synthesized a strontium titanate-based photocatalyst, SrTiO3-CuWO4 composite crystallites with low CuWO4 content, which degrades polluting molecules more efficiently than the initial, pure SrTiO3 due to its better hydroxyl radical production capacity. We optimized the amount of CuWO4 used during the synthesis and proved the success of CuWO4 modification. The obtained results may contribute to increasing the photocatalytic efficiency of SrTiO3 in the future, without reducing the band gap.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal13020287/s1, Figure S1: SEM images of the selected samples. (A): CWS0, (B): CWS1, (C): CWS2.5, (D): CWS3, EDX-mapping: (E): CWS1, (F): CWS2.5; Figure S2: Emission spectrum of the light source used during the photocatalytic experiments.

Author Contributions

Á.Á.: Conceptualization, methodology, investigation, original draft. L.J.: Funding acquisition, writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

The authors are very thankful for the financial support from the National Research, Development and Innovation Office (GINOP-1.1.2-PIACI-KFI-2021–00193) and the Hungarian Scientific Research Fund (OTKA) FK 142437. This paper was also supported by the UNKP-22-5 New National Excellence Program of the Ministry for Innovation and Technology from the National Research, Development and Innovation Fund and by the János Bolyai Research Scholarship of the Hungarian Academy of Sciences. Áron Ágoston thanks for the support of the Hungarian NTP-NFTÖ-22-B-0025 scholarship.

Data Availability Statement

Not applicable.

Conflicts of Interest

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

References

  1. Rodriguez, C.; Di Cara, A.; Renaud, F.N.R.; Freney, J.; Horvais, N.; Borel, R.; Puzenat, E.; Guillard, C. Antibacterial Effects of Photocatalytic Textiles for Footwear Application. Catal. Today 2014, 230, 41–46. [Google Scholar] [CrossRef]
  2. Mills, A.; Davies, R.H.; Worsley, D. Water Purification by Semiconductor Photocatalysis. Chem. Soc. Rev. 1993, 22, 417–425. [Google Scholar] [CrossRef]
  3. Ali, H.; Zaman, S.; Majeed, I.; Kanodarwala, F.K.; Nadeem, M.A.; Stride, J.A.; Nadeem, M.A. Porous Carbon/RGO Composite: An Ideal Support Material of Highly Efficient Palladium Electrocatalysts for the Formic Acid Oxidation Reaction. ChemElectroChem 2017, 4, 3126–3133. [Google Scholar] [CrossRef]
  4. Zhou, Y.; Wang, Z.; Huang, L.; Zaman, S.; Lei, K.; Yue, T.; Li, Z.A.; You, B.; Xia, B.Y. Engineering 2D Photocatalysts toward Carbon Dioxide Reduction. Adv. Energy Mater. 2021, 11, 2003159. [Google Scholar] [CrossRef]
  5. Noureen, L.; Xie, Z.; Gao, Y.; Li, M.; Hussain, M.; Wang, K.; Zhang, L.; Zhu, J. Multifunctional Ag3PO4-RGO-Coated Textiles for Clean Water Production by Solar-Driven Evaporation, Photocatalysis, and Disinfection. ACS Appl. Mater. Interfaces 2020, 12, 6343–6350. [Google Scholar] [CrossRef]
  6. Sasikala, R.; Kandasamy, M.; Suresh, S.; Ragavendran, V.; Sasirekha, V.; Pearce, J.M.; Murugesan, S.; Mayandi, J. Enhanced Dye-Sensitized Solar Cell Performance Using Strontium Titanate Perovskite Integrated Photoanodes Modified with Plasmonic Silver Nanoparticles. J. Alloys Compd. 2021, 889, 161693. [Google Scholar] [CrossRef]
  7. Ágoston, Á.; Balog, Á.; Narbutas, Š.; Dékány, I.; Janovák, L. Optimization of Plasmonic Copper Content at Copper-Modified Strontium Titanate (Cu-SrTiO3): Synthesis, Characterization, Photocatalytic Activity. Catalysts 2022, 12, 1041. [Google Scholar] [CrossRef]
  8. Farré, M.J.; Franch, M.I.; Malato, S.; Ayllón, J.A.; Peral, J.; Doménech, X. Degradation of Some Biorecalcitrant Pesticides by Homogeneous and Heterogeneous Photocatalytic Ozonation. Chemosphere 2005, 58, 1127–1133. [Google Scholar] [CrossRef] [PubMed]
  9. Reydellet, L.-H.; Roche, P.; Glattli, D.C.; Etienne, B.; Jin, Y. Quantum Partition Noise of Photon-Created Electron-Hole Pairs. Phys. Rev. Lett. 2003, 90, 176803. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Vieira, Y.; Leichtweis, J.; Foletto, E.L.; Silvestri, S. Reactive Oxygen Species-Induced Heterogeneous Photocatalytic Degradation of Organic Pollutant Rhodamine B by Copper and Zinc Aluminate Spinels. J. Chem. Technol. Biotechnol. 2020, 95, 791–797. [Google Scholar] [CrossRef]
  11. He, R.A.; Cao, S.W.; Yu, J.G. Recent Advances in Morphology Control and Surface Modification of Bi-Based Photocatalysts. Wuli Huaxue Xuebao/Acta Phys.-Chim. Sin. 2016, 32, 2841–2870. [Google Scholar] [CrossRef]
  12. Cravanzola, S.; Cesano, F.; Gaziano, F.; Scarano, D. Sulfur-Doped TiO2: Structure and Surface Properties. Catalysts 2017, 7, 214. [Google Scholar] [CrossRef] [Green Version]
  13. Sakthivel, S.; Shankar, M.V.; Palanichamy, M.; Arabindoo, B.; Bahnemann, D.W.; Murugesan, V. Enhancement of Photocatalytic Activity by Metal Deposition: Characterisation and Photonic Efficiency of Pt, Au and Pd Deposited on TiO2 Catalyst. Water Res. 2004, 38, 3001–3008. [Google Scholar] [CrossRef] [PubMed]
  14. Syafiuddin, A.; Hadibarata, T.; Zon, N.F. Salmiati Characterization of Titanium Dioxide Doped with Nitrogen and Sulfur and Its Photocatalytic Appraisal for Degradation of Phenol and Methylene Blue. J. Chin. Chem. Soc. 2017, 64, 1333–1339. [Google Scholar] [CrossRef]
  15. Rosy, A.; Kalpana, G. Reduced Graphene Oxide/Strontium Titanate Heterostructured Nanocomposite as Sunlight Driven Photocatalyst for Degradation of Organic Dye Pollutants. Curr. Appl. Phys. 2018, 18, 1026–1033. [Google Scholar] [CrossRef]
  16. Xu, J.; Wei, Y.; Huang, Y.; Wang, J.; Zheng, X.; Sun, Z.; Fan, L.; Wu, J. Solvothermal Synthesis Nitrogen Doped SrTiO3 with High Visible Light Photocatalytic Activity. Ceram. Int. 2014, 40, 10583–10591. [Google Scholar] [CrossRef]
  17. Choi, Y.; Jung, M.; Lee, Y.K. Effect of Heating Rate on the Activation Energy for Crystallization of Amorphous Ge2Sb2Te5 Thin Film. Electrochem. Solid-State Lett. 2009, 12, F17. [Google Scholar] [CrossRef]
  18. Da Silva, L.F.; Avansi, W.; Moreira, M.L.; Mesquita, A.; Maia, L.J.Q.; Andrés, J.; Longo, E.; Mastelaro, V.R. Relationship between Crystal Shape, Photoluminescence, and Local Structure in SrTiO3 Synthesized by Microwave-Assisted Hydrothermal Method. J. Nanomater. 2012, 2012, 890397. [Google Scholar] [CrossRef] [Green Version]
  19. WEITZEL, H. Kristallstrukturverfeinerung von Wolframiten Und Columbiten. Zeitschrift für Krist.-Cryst. Mater. 1976, 144, 238–258. [Google Scholar] [CrossRef]
  20. Gillespie, J.B.; Lindberg, J.D.; Laude, L.S. Kubelka-Munk Optical Coefficients for a Barium Sulfate White Reflectance Standard. Appl. Opt. 1975, 14, 807–809. [Google Scholar] [CrossRef]
  21. Lévy, F.; Mercier, A.; Voitchovsky, J.P. Band-Edge Photoluminescence of PbI2. Solid State Commun. 1974, 15, 819–822. [Google Scholar] [CrossRef]
  22. Siebenhofer, M.; Viernstein, A.; Morgenbesser, M.; Fleig, J.; Kubicek, M. Photoinduced Electronic and Ionic Effects in Strontium Titanate. Mater. Adv. 2021, 2, 7583–7619. [Google Scholar] [CrossRef] [PubMed]
  23. Wu, Z.; Zhao, Z.; Cheung, G.; Doughty, R.M.; Ballestas-Barrientos, A.R.; Hirmez, B.; Han, R.; Maschmeyer, T.; Osterloh, F.E. Role of Surface States in Photocatalytic Oxygen Evolution with CuWO 4 Particles. J. Electrochem. Soc. 2019, 166, H3014–H3019. [Google Scholar] [CrossRef]
  24. Balassa, L.; Ágoston, Á.; Kása, Z.; Hornok, V.; Janovák, L. Surface Sulfate Modified TiO2 Visible Light Active Photocatalyst for Complex Wastewater Purification: Preparation, Characterization and Photocatalytic Activity. J. Mol. Struct. 2022, 1260, 132860. [Google Scholar] [CrossRef]
  25. Shukla, S.; Pandey, H.; Singh, P.; Tiwari, A.K.; Baranwal, V.; Pandey, A.C. Synergistic Impact of Photocatalyst and Dopants on Pharmaceutical-Polluted Waste Water Treatment: A Review. Environ. Pollut. Bioavailab. 2021, 33, 347–364. [Google Scholar] [CrossRef]
  26. Zhang, M.; Yu, H.; Lyu, M.; Wang, Q.; Yun, J.-H.; Wang, L. Composition-Dependent Photoluminescence Intensity and Prolonged Recombination Lifetime of Perovskite CH3NH3PbBr3−xClx Films. Chem. Commun 2014, 50, 11727. [Google Scholar] [CrossRef] [Green Version]
  27. Jayathilaka, P.B.; Pathiraja, G.C.; Bandara, A.; Subasinghe, N.D.; Nanayakkara, N. Theoretical Study of Phenol and Hydroxyl Radical Reaction Mechanism in Aqueous Medium by the DFT/B3LYP/6-31+G(d,p)/CPCM Model. Can. J. Chem. 2014, 92, 809–813. [Google Scholar] [CrossRef]
  28. Maier, A.C.; Iglebaek, E.H.; Jonsson, M. Confirming the Formation of Hydroxyl Radicals in the Catalytic Decomposition of H2O2 on Metal Oxides Using Coumarin as a Probe. ChemCatChem 2019, 11, 5435–5438. [Google Scholar] [CrossRef]
  29. Pourmortazavi, S.M.; Rahimi-Nasrabadi, M.; Khalilian-Shalamzari, M.; Ghaeni, H.R.; Hajimirsadeghi, S.S. Facile Chemical Synthesis and Characterization of Copper Tungstate Nanoparticles. J. Inorg. Organomet. Polym. Mater. 2014, 24, 333–339. [Google Scholar] [CrossRef]
  30. Gallardo, V.; Morales, M.E.; Ruiz, M.A.; Delgado, A.V. An Experimental Investigation of the Stability of Ethylcellulose Latex: Correlation between Zeta Potential and Sedimentation. Eur. J. Pharm. Sci. 2005, 26, 170–175. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of SrTiO3 (CWS0) and CuWO4-SrTiO3 composite samples (CWSX) (A), CuWO4 (B) and the main reflection shift (C).
Figure 1. XRD patterns of SrTiO3 (CWS0) and CuWO4-SrTiO3 composite samples (CWSX) (A), CuWO4 (B) and the main reflection shift (C).
Catalysts 13 00287 g001
Figure 2. Kubelka–Munk representation of SrTiO3 (CWS0) and SRTX samples for band gap energy determination.
Figure 2. Kubelka–Munk representation of SrTiO3 (CWS0) and SRTX samples for band gap energy determination.
Catalysts 13 00287 g002
Figure 3. The PL spectra of the CWSX samples.
Figure 3. The PL spectra of the CWSX samples.
Catalysts 13 00287 g003
Figure 4. The zeta-potential values of 0.001 wt% aqueous dispersion of the initial SrTiO3 (CWS0) and CWSX samples (pH = ~6.50).
Figure 4. The zeta-potential values of 0.001 wt% aqueous dispersion of the initial SrTiO3 (CWS0) and CWSX samples (pH = ~6.50).
Catalysts 13 00287 g004
Figure 5. Representative EDX spectrum of the CWS3 sample.
Figure 5. Representative EDX spectrum of the CWS3 sample.
Catalysts 13 00287 g005
Figure 6. Run of decay curves, photocatalytic activity of the CWSX photocatalysts and reference SrTiO3 (CWS0) (A) and the exact amount of phenol that has been destroyed (mgphenol/gcat) (B).
Figure 6. Run of decay curves, photocatalytic activity of the CWSX photocatalysts and reference SrTiO3 (CWS0) (A) and the exact amount of phenol that has been destroyed (mgphenol/gcat) (B).
Catalysts 13 00287 g006
Figure 7. Production kinetics of 7-hydroxycoumarin from coumarin in the case of CWSX samples.
Figure 7. Production kinetics of 7-hydroxycoumarin from coumarin in the case of CWSX samples.
Catalysts 13 00287 g007
Figure 8. Coumarin and hydroxyl radical reaction to produce 7-hydroxycoumarin.
Figure 8. Coumarin and hydroxyl radical reaction to produce 7-hydroxycoumarin.
Catalysts 13 00287 g008
Table 1. Nominal and measured CuWO4 concentrations and specific surface area of the samples.
Table 1. Nominal and measured CuWO4 concentrations and specific surface area of the samples.
Catalyst NameNominal Cu Content (wt%)Measured CuWO4 Content (wt%) *Specific Surface Area (m2/g)
CWS00040.30 ± 1.60
CWS0.50.500.2138.38 ± 0.40
CWS11.001.3441.04 ± 0.35
CWS1.51.502.1639.30 ± 0.55
CWS22.002.8640.81 ± 1.95
CWS2.52.504.2840.92 ± 5.43
CWS33.005.2240.13 ± 1.18
* determined by EDX measurements.
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

Ágoston, Á.; Janovák, L. Hydrothermal Co-Crystallization of Novel Copper Tungstate-Strontium Titanate Crystal Composite for Enhanced Photocatalytic Activity and Increased Electron–Hole Recombination Time. Catalysts 2023, 13, 287. https://doi.org/10.3390/catal13020287

AMA Style

Ágoston Á, Janovák L. Hydrothermal Co-Crystallization of Novel Copper Tungstate-Strontium Titanate Crystal Composite for Enhanced Photocatalytic Activity and Increased Electron–Hole Recombination Time. Catalysts. 2023; 13(2):287. https://doi.org/10.3390/catal13020287

Chicago/Turabian Style

Ágoston, Áron, and László Janovák. 2023. "Hydrothermal Co-Crystallization of Novel Copper Tungstate-Strontium Titanate Crystal Composite for Enhanced Photocatalytic Activity and Increased Electron–Hole Recombination Time" Catalysts 13, no. 2: 287. https://doi.org/10.3390/catal13020287

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