Next Article in Journal
Sustainable Nanomedicine: Enhancement of Asplatin’s Cytotoxicity In Vitro and In Vivo Using Green-Synthesized Zinc Oxide Nanoparticles Formed via Microwave-Assisted and Gambogic Acid-Mediated Processes
Previous Article in Journal
CO2 Hydrogenation to Methanol over In2O3 Decorated by Metals of the Iron Triad
Previous Article in Special Issue
Hydrothermally Synthesized ZnCr- and NiCr-Layered Double Hydroxides as Hydrogen Evolution Photocatalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Visible Light-Driven Photocatalysis of Al-Doped SrTiO3: Experimental and DFT Study

1
Department of Materials Science, Nanotechnology and Engineering Physics, Satbayev University, Almaty 050032, Kazakhstan
2
Institute of Nuclear Physics, Almaty 050032, Kazakhstan
3
Joint Institute for Nuclear Research, Dubna 141980, Russia
4
Bes Saiman Group, Almaty 050057, Kazakhstan
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(22), 5326; https://doi.org/10.3390/molecules29225326
Submission received: 17 October 2024 / Revised: 7 November 2024 / Accepted: 8 November 2024 / Published: 12 November 2024

Abstract

:
Environmental problems associated with water pollution caused by organic dyes have raised serious concerns. In this context, photocatalytic processes have proven to be promising and environmentally friendly methods for water purification utilising abundant solar energy. In this study, a SrTiO3-based photocatalyst was modified by doping with Al ions and the deposition of dual co-catalysts (Rh/Cr2O3 and CoOOH) to enhance the photocatalytic decomposition efficiency of methylene blue (MB). Pure perovskite SrTiO3 was synthesised by chemical precipitation followed by calcination at 1100 °C. Al-doped SrTiO3 with deposited co-catalysts showed 3.2 times higher photocatalytic activity compared to unalloyed SrTiO3 with co-catalysts in MB decomposition under visible radiation. This study highlights the effectiveness of using dual co-catalysts and low-valence metal doping to enhance the efficiency of the photocatalytic decomposition of organic pollutants. The density functional theory analysis results show that the Al doping of SrTiO3 improves charge separation and increases the lifetime of photogenerated electrons and holes while maintaining the size of the forbidden band, which confirms its effectiveness for enhancing photocatalytic activity.

1. Introduction

In recent years, the pollution of aquatic ecosystems by organic substances including dyes, antibiotics, and pesticides has significantly worsened the environmental situation [1]. These pollutants not only disrupt aquatic ecosystems but also pose a threat to human health because they can accumulate in living organisms and cause serious diseases, including cancer [2]. Dyes used in textiles, cosmetics, plastics, and other industries are particularly prevalent, resulting in wastewater containing these substances [3,4]. Owing to their high stability, dyes such as methylene blue (MB) persist in aqueous environments and are difficult to degrade. MB is toxic and carcinogenic, thus posing a threat to human health and the environment. For example, Gahlot et al. [5] report cellular stress and apoptosis at elevated MB levels. Therefore, effective methods for its removal from water bodies must be developed [6]. Various methods are used to treat wastewater from dyes, such as electrocatalysis [7], coagulation–flocculation [8], ionic exchange [9], oxidation [10], biodegradation [11], and adsorption [12]. These technologies vary considerably in terms of their efficiency and environmental impact, and each has its own drawbacks. For example, adsorption, despite its efficiency, has several disadvantages, such as high investment, high operating costs, low efficiency, and the problems of recycling and reusability for actual application in wastewater treatment [13]. Similarly, ion exchange, though very effective in removing certain contaminants such as heavy metals or hardness ions, may possess a much lower effectiveness for other types of contaminants [14].
Over the past few decades, photocatalysis [15] has been extensively studied as an effective method for removing toxic substances and pathogens from water and air, producing hydrogen by splitting water, synthesising organic compounds for pharmaceuticals, treating cancer cells, and creating self-cleaning coatings. [16,17]. Photocatalysis is widely recognised as one of the most effective technologies for the efficient removal of toxic pollutants. The reaction, accelerated by a catalyst upon the absorption of light, is called photocatalysis. The photogeneration of e/h+ pairs leads to the formation of hydroxyl and superoxide radicals, which interact with dye molecules and reduce pollutants [18]. To date, various options are known for composite structures used for the photocatalysis process, including CdZnS/TiO2 [19], GaN–ZnO [20], SrTiO3 [21], Y2Ti2O5S2 [22], PAN/SrTiO3 [23], BaTaO2N [24], TNT@SrTiO3 [25], and SrTaO2N [26]. However, most modern photocatalysts are limited by their poor efficiency, owing to excessive defects appearing in the crystal structure. Specifically, during the thermal treatment of SrTiO3, oxygen vacancies occur, provoking undesirable defects in the form of Ti3+, which act as recombination centres in the band structure. To solve this problem, previous studies on defect engineering [27] have shown that doping with low-valence metal cations at the B node is effective. As aforementioned, the replacement of Ti3+ by Ga3+ or Na+ to substitute Sr2+ can introduce oxygen vacancies into the perovskite structure, resulting in a decrease in the Ti3+ concentration associated with the deactivation of photocatalysis. In addition, the molten flux method, conducted at relatively high temperatures, allows for the efficient alloying of perovskites with low-valence metal cations while avoiding the formation of side impurities [28]. Most studies on the defect engineering of perovskite photocatalysts based on doped SrTiO3 have focused on the photocatalytic decomposition of water [29] and CO2 reduction [30]. However, studies on the photocatalytic decomposition of organic pollutants are limited, emphasising the need for further research in this area.
On the other hand, metal-doped SrTiO3, in its pure form, has shown limited efficiency in photocatalytic applications due to sluggish surface redox reactions and suboptimal charge separation. For instance, pure Al-doped SrTiO3 achieved only a 54.91% degradation of methylene blue over 5 h under UV light [31]. In another study, Al-doped SrTiO3 synthesised via the flux method achieved a 58% degradation of congo red under visible light within 90 min—a relatively extended interval considering the goals of photocatalytic applications [32]. These findings underscore the challenges associated with doped SrTiO3 in its pure form and highlight the need for further modifications to enhance activity [33]. To overcome these limitations, one effective approach is the deposition of cocatalysts and the construction of heterostructures. These strategies facilitate the efficient separation of photogenerated charges and introduce active reaction centres, thereby accelerating redox reactions on the photocatalyst surface. Similar studies have reported the improved photocatalytic degradation of organic dyes through such modifications, underscoring the potential of cocatalyst-decorated, doped SrTiO3 composites [32,34,35]. This work builds on these insights by examining the combined impact of Al doping and dual cocatalyst deposition on the photocatalytic degradation of methylene blue, aiming to bridge gaps in the current literature and demonstrate a more efficient degradation pathway.
In this study, Al-doped SrTiO3 photocatalyst SrTiO3 was successfully synthesised by the molten flux method followed by the photodeposition of dual co-catalysts (Rh/Cr2O3 and CoOOH). This photocatalyst exhibited high efficiency in the decomposition of MB under visible light. Highly crystalline SrTiO3 was prepared by a simple chemical precipitation method followed by calcination at 1100 °C. Various electron microscopy and structural analysis studies were carried out to evaluate the changes in the physicochemical properties of the photocatalysts before and after doping. The elementary supercell of SrTiO3@Al is shown in Figure 1. The analytical results indicate the promising application of dual co-catalysts and doping with low-valence metals to improve the efficiency of the photocatalytic decomposition of organic pollutants.

2. Results and Discussion

Figure 2 presents the morphological images of SrTiO3 samples and doped SrTiO3@Al particles synthesised by the molten salt method at 1100 °C for 10 h, obtained using scanning electron microscopy (SEM) and transmission electron microscopy (TEM). The SrTiO3 samples shown in Figure 2a,b have well-defined block-like structures, with sizes ranging from 150 to 300 nm. The doping of SrTiO3@Al (Figure 2c,d) results in the particles acquiring more rounded faces and a homogeneous distribution, as seen in the TEM images. These particles exhibit a smooth, cubic morphology with sizes approximately between 100 and 150 nm. The cubic structure of SrTiO3@Al with truncated faces is formed because of the selective adsorption of Cl on the {111} faces when SrCl2 is used as flux. This adsorption lowers the surface energy of the facets, which in turn contributes to the expansion and reduction in the facet size {100} [36]. Further, EDX mapping confirmed the presence of Ti, O, Sr, and Al without any other impurities in the composition (Supplementary Figures S1 and S2).
As photocatalytic reactions occur on the catalyst surface, the presence of anisotropic crystalline faces with different surface energies significantly affects the photocatalytic activity. The anisotropic structure of cubic SrTiO3@Al improves the separation and transfer of photogenerated electrons and holes on the catalyst surface, which in turn promotes the formation of superoxide (•O2) and hydroxyl (•OH) radicals in the reaction medium [37].
Figure 3a shows the X-ray diffraction (XRD) spectra of the SrTiO3, SrTiO3@Al, and Rh/Cr2O3/SrTiO3@Al/CoOOH samples. The spectrum of SrTiO3@Al shows the standard perovskite structure of the cubic SrTiO3 phase (JCPDS #35-0734) [38,39,40], which confirms the successful synthesis of the material. The comparison of the spectra of doped and undoped SrTiO3 shows a significant increase in the peak intensity of SrTiO3@Al, indicating improved crystallinity and increased crystal size because of Al doping performed using the molten salt method. In addition, the presence of aluminium and Rh/Cr2O3/CoOOH does not affect the crystal structure of SrTiO3, as their characteristic peaks are not observed in the spectrum. This may be due to their low content, high dispersibility, and low crystallinity [30,37]. X-ray photoelectron spectroscopy (XPS) is used to study the chemical composition and structure of the flux-doped SrTiO3@Al surface photocatalyst. Figure 3b–d present spectra showing the changes in chemical composition and elemental states in the Al-doped sample compared to pure SrTiO3. The Al 2p spectrum presented in Figure 3b shows a peak in the binding energy range of 72.1–76.4 eV, indicating the presence of Al3+. This confirms the incorporation of aluminium into the SrTiO3 lattice in the oxidised state through bonding with oxygen. Similar studies [41] indicate that high-temperature treatment causes oxygen vacancies in the SrTiO3 structure, leading to the formation of Ti3+ ions, as also shown in Figure 3c. In the undoped SrTiO3 sample synthesised at 1100 °C, Ti 2p peaks are observed corresponding to Ti3+ ions, which can serve as recombination centres [42]. In contrast, in the doped SrTiO3@Al sample, Ti3+ ions are completely absent, and the spectrum shows only Ti4+ peaks with binding energies of 457.9 and 463.6 eV. This indicates the stabilisation of titanium in the oxidised Ti4+ state. Figure 3d shows the O 1s level spectra, where two Gaussian peaks are revealed: one at 528.4 eV, which is associated with oxygen in the crystal lattice, and the other at 530.0 eV, corresponding to oxygen defects. Doping with Al significantly alters the ratio of the lattice and defect oxygen, increasing the defect oxygen content from 15.9% to 54.2% [21]. Thus, the results of XPS and XRD analysis confirm the successful synthesis of SrTiO3@Al. Al doping effectively suppresses the formation of Ti3+ ions and promotes the oxygen vacancies in an optimal amount. These findings agree with previous studies [43,44], which highlight the improved photocatalytic performance of SrTiO3 upon doping with aluminium.

2.1. Photocatalytic Performance

To evaluate the photocatalytic activity of the synthesised catalysts, experiments were performed on the photocatalytic degradation of the cationic dye MB in aqueous solution. In all the experiments, the samples were irradiated under visible light (λ ≥ 400 nm) for 60 min. The results illustrating the photocatalytic activity of the samples as a function of the irradiation time are presented in Figure 4a. The MB dye removal rates (initial concentration of 10 mg/L, 50 mL) were 13.5%, 18.2%, 27.9%, and 88.7% for the SrTiO3, SrTiO3@Al, Rh/Cr2O3/SrTiO3/CoOOH, and Rh/Cr2O3/SrTiO3@Al/CoOOH photocatalysts, respectively. The differences in photodegradation efficiency between SrTiO3 and SrTiO3@Al indicate the influence of the crystallite size effect related to Al doping, which was also discussed in the context of the sample morphology. An increase in the particle size can promote the recombination of photogenerated charges, which negatively affects the photocatalytic efficiency [45]. The highest photocatalytic activity was recorded for the Rh/Cr2O3/SrTiO3@Al/CoOOH sample, the efficiency of which exceeds that of SrTiO3, SrTiO3@Al, and Rh/Cr2O3/SrTiO3/CoOOH by a factor of 6.6, 4.9, and 3.2, respectively. This was attributed to the efficient charge separation and charge transfer provided by the dual separately photo-deposited co-catalysts in the catalyst/photocatalyst system.
According to a previous study [42], Al doping suppresses Ti3+ recombination centres, which promotes the spatial separation of photogenerated electrons and holes and their transfer to different active centres to participate in photocatalytic pollutant decomposition reactions [42].
The study of the kinetic characteristics plays an important role in understanding the mechanism of the photocatalytic decomposition of dyes. The pseudo-first-order model described in Equation (1) was used to estimate the rate of photocatalytic MB removal.
ln ( C t C 0 ) = k p t ,
where C0 and Ct (mg/L) denote the initial and current dye concentrations at time t, and kp is the reaction rate constant. Figure 4b shows linear plots of the dependence of ( C t C 0 ) on reaction time. The regression coefficient (R2) shows good correlation with the pseudo-first-order kinetics for different photocatalysts, as follows: R2 = 0.995 for SrTiO3, R2 = 0.972 for SrTiO3@Al, R2 = 0.980 for Rh/Cr2O3/SrTiO3/CoOOH, and R2 = 0.982 for Rh/Cr2O3/SrTiO3@Al/CoOOH. The calculations showed that the values of the constant kp for SrTiO3, SrTiO3@Al, Rh/Cr2O3/SrTiO3/CoOOH, and Rh/Cr2O3/SrTiO3@Al/CoOOH are 0.0009, 0.0017, 0.0026, and 0.0312 min−1, respectively. The obtained data confirm that the Rh/Cr2O3/SrTiO3@Al/CoOOH sample is the most efficient photocatalyst for MB decomposition, which is consistent with the results of photodegradation studies.

2.2. Degradation Mechanism

Similar studies using radical scavengers such as isopropanol, benzoquinone, and EDTA to investigate the photocatalytic mechanism of SrTiO3 with cocatalysts have shown that superoxide (•O2) and hydroxyl (•OH) radicals play a key role in the degradation of methylene blue [46,47].
The mechanism of action for the photocatalyst Rh/Cr2O3/SrTiO3@Al/CoOOH can be described as follows: upon the absorption of solar photons by SrTiO3@Al particles, electrons (e) and holes (h+) are generated. The electrons are directed towards the reduction cocatalyst Rh/Cr2O3, while the holes migrate to the oxidation cocatalyst CoOOH. The strategic placement of these cocatalysts on different facets of SrTiO3@Al facilitates the effective separation of electron–hole pairs, thereby preventing their recombination and enabling redox reactions [48]. Superoxide radicals (•O2) are formed as a result of the reaction between electrons and adsorbed oxygen, while hydroxyl radicals (•OH) are generated through the interaction of holes with water or hydroxide ions. Additionally, methylene blue absorbs photons, initiating the formation of reactive species. The interaction of these radicals with dye molecules leads to the production of CO2 and H2O, delineating the key stages of the photocatalytic reaction [21].
R h / C r 2 O 3 / S r T i O 3 @ A l / C o O O H + h ν e + h +
e + O 2 O 2
h + + H 2 O O H
O H + M B i n t e r m e d i a t e   p r o d u c t s C O 2 + H 2 O
O 2 + M B i n t e r m e d i a t e   p r o d u c t s C O 2 + H 2 O

2.3. Density Functional Theory (DFT) Calculation

To analyse the electronic structure of SrTiO3 and Al-doped SrTiO3, particularly in terms of band structure and density of states, DFT simulations were conducted. These results align with those of previous studies [49,50], indicating a calculated band gap of 1.99 eV using generalised gradient approximation (GGA), which is lower than the experimental value because of the well-known underestimation of on-site Coulomb interactions for d- and f-electrons. By incorporating the Hubbard U parameter (U = 5, J = 4), this limitation was mitigated, yielding a bandgap of 3.2 eV. Figure 5a illustrates the calculated band structure of pure SrTiO3, showing an indirect band gap of 3.19 eV, consistent with our previous experimental findings [21,51]. The anisotropy of the Ti 3d bands in the conduction zone was evident, with a nearly dispersionless character in the GX direction, whereas the quasiparticle energy increased rapidly in other directions. Thus, the excited electrons in the conduction band were likely to occupy states in the GX direction, where the kinetic energy was minimised. This corresponded to the distribution of conduction electron momenta perpendicular to the crystal’s cubic faces, as observed experimentally.
Al doping in SrTiO3 primarily replaces Ti atoms owing to the similar ionic radii of Al (54 pm) and Ti (61 pm) within the SrTiO3 lattice. The substitution of Ti4+ with Al3+ induces p-type conductivity by introducing impurity levels associated with Al near the valence band edge [52,53,54,55]. Our DFT calculations show that while Al doping does not significantly alter the band gap size, it does extend the lifetimes of photogenerated electrons and holes by reducing Ti3+ recombination centres. This finding is corroborated by XPS analysis (Figure 3), which indicates a decrease in recombination sites, thereby enhancing charge separation efficiency. Experimentally, this enhanced charge separation is evident in the photocatalytic performance of SrTiO3, as the prolonged lifetime of charge carriers reduces recombination and increases the availability of reactive species.
Consequently, the DFT results provide direct theoretical confirmation of the observed experimental behaviour, illustrating how Al doping modifies the electronic structure to extend charge carrier lifetimes and enhance the separation of photogenerated electron–hole pairs. Specifically, in the photocatalytic degradation of methylene blue, the SrTiO3@Al composite with deposited co-catalysts demonstrated a 3.2-fold improvement in efficiency compared to undoped SrTiO3 with an equivalent co-catalyst ratio. This enhanced performance is attributed to the increased charge separation and reduced recombination facilitated by Al doping, which is critical to achieving the observed degradation efficiency. These findings substantiate the superior photocatalytic activity of Al-doped SrTiO3, linking the DFT predictions to the improved experimental outcomes.

3. Methodology

3.1. Materials

Titanium (IV) oxide, anatase (TiO2, particle sizes < 25 nm, 99.7%), strontium nitrate (Sr(NO3)2, ≥ 98%), nitric acid (HNO3, < 90%), strontium titanate (SrTiO3, particle sizes < 100 nm, 99%), aluminium oxide (Al2O3, particle sizes < 50 nm, 99.8%), rhodium (III) chloride hydride (RhCl3·6H2O, Rh 38–40%), cobalt (II) nitrate hexahydrate (Co(NO3)3·6H2O, ≥98%), and methylene blue (C16H18ClN3S·xH2O, dye content, ≥82%) were purchased from Sigma–Aldrich (St. Louis, MO, USA). Oxalic acid ((COOH)2·2H2O, >98%), strontium chloride hexahydrate (SrCl2·6H2O, 99.7%), and potassium chromate (K2CrO4, 99.5%) were purchased from Laborpharma (Almaty, Kazakhstan). All chemicals were used without pretreatment.

3.2. Synthesis of SrTiO3 Calcined at 1100 °C

SrTiO3 powder was synthesised using a simple chemical precipitation method followed by calcination at 1100 °C as described in Ref. [56]. Briefly, a 0.12 M solution of Sr(NO3)2 was mixed with TiO2 in distilled water at a Sr/Ti molar ratio of 1:1 and subjected to ultrasonic treatment for 30 min. A 0.4 M solution of clear oxalic acid (COOH)2-2H2O was used as a reducing agent. The two solutions were combined in a chemical cylinder with constant stirring using a magnetic stirrer. Then, a 10% aqueous ammonia solution was added to the solution to achieve a pH of 6–7. The resulting precipitate was washed and dried at 60 °C for 16 h. In the final step, the product was calcined in a muffle furnace at 1100 °C for one hour, resulting in a white powder of SrTiO3.

3.3. Synthesis of SrTiO3@Al

The alloying of the SrTiO3 powder with aluminium was performed using the fluxing method. To synthesise SrTiO3@Al, SrTiO3, Al2O3, and SrCl2 powders were mixed in a molar ratio of 1:0.02:10 in an agate mortar for half an hour until a uniform mass was obtained. The mixture was then placed in an aluminium oxide crucible and heat-treated at 1100 °C in a muffle furnace for 10 h. After cooling, hot distilled water was added to the crucible and the mixture was ultrasonicated to separate the samples from the crucible. The precipitate was then washed with hot distilled water for five centrifugation cycles to remove residual SrCl2. The final step was drying the powder at 60 °C for 16 h to ensure the high purity of the final product.

3.4. Photodeposition of Double Co-Catalysts

SrTiO3 and SrTiO3@Al samples were further modified with dual co-catalysts, Rh/Cr2O3 and CoOOH, using the photodeposition method [21]. Then, 0.1 g of SrTiO3@Al powder was added to distilled water (50 mL) and subjected to ultrasonication for 30 min. The resulting suspension was placed in a reactor for photochemical reaction and 50 μL of RhCl3·6H2O was added, followed by irradiation (10 min) under constant magnetic stirring. Then, 25 μL of K2CrO4 (2 mg (Cr)/mL) and 25 μL of Co(NO3)3 (2 mg(Co)/mL) were added and irradiated for another 10 min. The resulting samples were washed several times and dried for 16 h at 60 °C. The amounts of added aqueous solutions of the co-catalysts were calculated for mass concentrations of Rh, Cr, and Co of 0.1%, 0.05%, and 0.05%, respectively. This allowed us to obtain samples with the photodeposition of the dual co-catalysts Rh/Cr2O3/SrTiO3@Al/CoOOH and Rh/Cr2O3/SrTiO3/CoOOH.

3.5. Characteristics

Data on the phase compositions of the obtained samples and modified composites were obtained using XRD using a Drone-8 with angles of 5–70° and steps of 0.01°. The morphology and elemental composition were studied using SEM (Zeiss Crossbeam 540 isCarl Zeiss Microscopy GmbH, Oberkochen, Germany) at 5–20 kV and energy-dispersive X-ray spectroscopy (EDX; INCA X-Sight). TEM (JEM-2100 LaB6 HRTEM) at 80 kV was used to study the morphologies of the co-catalysts. XPS was performed on a Microtech Multilab 3000 VG instrument with Mg and Al as X-ray sources for valence state analysis, calibrated by the C1s peak at 284.8 eV. Ultraviolet reflectance spectra (UV–Vis DRS) were recorded on a Perkin Elmer Lambda 35 spectrophotometer in the range of 200–800 nm.

3.6. Photocatalytic Test

The photocatalytic decomposition of MB in aqueous solution was carried out at room temperature (25 °C) using a photochemical reactor (Shanghai Leewen Scientific Instrument Co., Ltd., Shanghai, China) with a high-pressure mercury lamp (10 W; λmax = 664 nm) and a cutoff filter (λ ≈ 400 nm). The distance between the quartz flask (50 mL) and lamp was set to 10 cm. The duration of photocatalytic irradiation was 60 min for each sample, and fractions (1 mL each) were collected every 15 min and filtered through a PVDF syringe filter. The resulting fractions were examined using the UV–Vis method (I5 Hanon Advanced Technology Group Co., Ltd., Jinan, China). All photocatalytic diagnostics were performed twice, and the mean values were considered to account for deviations.

3.7. Computational Details

The open-source Quantum ESPRESSO package was utilised for DFT calculations to investigate the electronic structure of SrTiO3 [57]. Al-doped SrTiO3 was modelled using a supercell consisting of 3 × 3 × 3 elementary cubic cells with a lattice parameter of 3.94 Å, as shown in Figure 1. In an ideal crystal, this supercell contains 135 atoms. Structural optimisation was performed using the Perdew–Burke–Ernzerhof functional within the GGA, with a plane-wave cutoff energy of 40 Ry and a 3 × 3 × 3 uniform k-point grid in the first Brillouin zone. The relaxation continued until the atomic forces were reduced below 10−4 Ry/Å. The standard DFT method encounters challenges in modelling d- and f-electron materials owing to self-interaction errors in the localised states. Therefore, the GGA + U method (Dudarev formulation) [58] was applied for electronic structure calculations on SrTiO3 and Al-doped SrTiO3. Given the hybridisation of Ti 3d and O 2p states [49,59,60], on-site Coulomb corrections, with U values of 5.9 eV for Ti and 4.2 eV for O, were applied to ensure accuracy. The cutoff energy was set to 60 Ry and a 5 × 5 × 5 k-point grid was used for the calculations.

4. Conclusions

In this study, the photocatalytic activity of synthesised SrTiO3-based catalysts was evaluated for MB degradation under visible light. The Rh/Cr2O3/SrTiO3@Al/CoOOH sample showed the best results, with the dye removal efficiency reaching 88.7% within 60 min, which exceeds the activity of Rh/Cr2O3O3/SrTiO3/CoOOH by 3.2 times. These results indicated that Al doping reduces Ti3+ recombination centres and promotes the efficient separation and transfer of photogenerated charges. Kinetic studies confirmed the photocatalytic decomposition mechanism as pseudo-first-order, with a decomposition rate constant of 0.0312 min−1 for the Rh/Cr2O3/SrTiO3@Al/CoOOH sample, highlighting its outstanding efficiency. DFT analysis revealed that the Al doping of SrTiO3 improves charge separation and extends the lifetime of photogenerated electrons and holes while maintaining the bandgap size. However, to achieve substantial enhancement in photocatalytic activity, additional measures such as the use of dual co-catalysts are required. These co-catalysts create active centres and facilitate effective charge transfer, significantly boosting pollutant degradation efficiency.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29225326/s1, Figure S1: EDX mapping of SrTiO3@Al; Figure S2: TEM images.

Author Contributions

Conceptualization, methodology, writing—original draft preparation, writing—review and editing, U.A., M.B. and A.S.; formal analysis, resources, N.M. and A.I.; project administration, funding acquisition I.K. and T.A.; formal analysis, investigation, K.M. and E.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research is funded by the Science Committee of the Ministry of Science and Higher Education and of the Republic of Kazakhstan (Grant No. BR18574073).

Institutional Review Board Statement

The study did not require ethical approval.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

Author Aigerim Serik was employed by the company Bes Saiman Group. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Daulbayev, C.; Nursharip, A.; Tauanov, Z.; Busquets, R.; Baimenov, A. Mechanisms of Mercury Removal from Water with Highly Efficient MXene and Silver-Modified Polyethyleneimine Cryogel Composite Filters. Adv. Compos. Hybrid Mater. 2024, 7, 139. [Google Scholar] [CrossRef]
  2. Molla, M.A.I.; Ahmed, A.Z.; Kaneco, S. Chapter 3-Reaction Mechanism for Photocatalytic Degradation of Organic Pollutants. In Nanostructured Photocatalysts; Nguyen, V.-H., Vo, D.-V.N., Nanda, S., Eds.; Elsevier: Amsterdam, The Netherlands, 2021; pp. 63–84. ISBN 978-0-12-823007-7. [Google Scholar]
  3. Du, F.; Yang, D.; Kang, T.; Ren, Y.; Hu, P.; Song, J.; Teng, F.; Fan, H. SiO2/Ga2O3 Nanocomposite for Highly Efficient Selective Removal of Cationic Organic Pollutant via Synergistic Electrostatic Adsorption and Photocatalysis. Sep. Purif. Technol. 2022, 295, 121221. [Google Scholar] [CrossRef]
  4. Serik, A.; Idrissov, N.; Baratov, A.; Dikov, A.; Kislitsin, S.; Daulbayev, C.; Kuspanov, Z. Recent Progress in Photocatalytic Applications of Electrospun Nanofibers: A Review. Molecules 2024, 29, 4824. [Google Scholar] [CrossRef]
  5. Gehlot, S.; Gupta, A.; Gupta, R. A CNN-Based Unified Framework Utilizing Projection Loss in Unison with Label Noise Handling for Multiple Myeloma Cancer Diagnosis. Med. Image Anal. 2021, 72, 102099. [Google Scholar] [CrossRef]
  6. Ye, H.; Luo, Y.; Yang, T.; Xue, M.; Yin, Z.; Gao, B. Effects of Ball Milling on Hydrochar for Integrated Adsorption and Photocatalysis Performance. Sep. Purif. Technol. 2025, 354, 128687. [Google Scholar] [CrossRef]
  7. Qian, W.; Xu, S.; Zhang, X.; Li, C.; Yang, W.; Bowen, C.R.; Yang, Y. Differences and Similarities of Photocatalysis and Electrocatalysis in Two-Dimensional Nanomaterials: Strategies, Traps, Applications and Challenges. Nano-Micro Lett. 2021, 13, 156. [Google Scholar] [CrossRef]
  8. Katrivesis, F.K.; Karela, A.D.; Papadakis, V.G.; Paraskeva, C.A. Revisiting of Coagulation-Flocculation Processes in the Production of Potable Water. J. Water Process Eng. 2019, 27, 193–204. [Google Scholar] [CrossRef]
  9. Sivaranjanee, R.; Kumar, P.S.; Mahalaxmi, S. A Review on Agro-Based Materials on the Separation of Environmental Pollutants from Water System. Chem. Eng. Res. Des. 2022, 181, 423–457. [Google Scholar] [CrossRef]
  10. Bhat, A.P.; Gogate, P.R. Degradation of Nitrogen-Containing Hazardous Compounds Using Advanced Oxidation Processes: A Review on Aliphatic and Aromatic Amines, Dyes, and Pesticides. J. Hazard. Mater. 2021, 403, 123657. [Google Scholar] [CrossRef]
  11. Vu, D.H.; Åkesson, D.; Taherzadeh, M.J.; Ferreira, J.A. Recycling Strategies for Polyhydroxyalkanoate-Based Waste Materials: An Overview. Bioresour. Technol. 2020, 298, 122393. [Google Scholar] [CrossRef]
  12. Mabalane, K.; Shooto, N.D.; Thabede, P.M. A Novel Permanganate and Peroxide Carbon-Based Avocado Seed Waste for the Adsorption of Manganese and Chromium Ions from Water. Case Stud. Chem. Environ. Eng. 2024, 10, 100782. [Google Scholar] [CrossRef]
  13. Regel-Rosocka, M.; Kruszelnicka, I.; Góra, W.; Baraniak, M.; Lota, G.; Ginter-Kramarczyk, D.; Staszak, K. Removal of Nickel(II) from Industrial Wastewater Using Selected Methods: A Review. Chem. Process Eng. 2022, 43, 437–448. [Google Scholar]
  14. Jafarinejad, S. 6-Treatment of Oily Wastewater. In Petroleum Waste Treatment and Pollution Control; Jafarinejad, S., Ed.; Butterworth-Heinemann: Oxford, UK, 2017; pp. 185–267. ISBN 978-0-12-809243-9. [Google Scholar]
  15. Nunes, M.J.; Lopes, A.; Pacheco, M.J.; Ciríaco, L. Visible-Light-Driven AO7 Photocatalytic Degradation and Toxicity Removal at Bi-Doped SrTiO3. Materials 2022, 15, 2465. [Google Scholar] [CrossRef]
  16. Boyjoo, Y.; Sun, H.; Liu, J.; Pareek, V.K.; Wang, S. A Review on Photocatalysis for Air Treatment: From Catalyst Development to Reactor Design. Chem. Eng. J. 2017, 310, 537–559. [Google Scholar] [CrossRef]
  17. Kuspanov, Z.; Bakbolat, B.; Baimenov, A.; Issadykov, A.; Yeleuov, M.; Daulbayev, C. Photocatalysts for a Sustainable Future: Innovations in Large-Scale Environmental and Energy Applications. Sci. Total Environ. 2023, 885, 163914. [Google Scholar] [CrossRef] [PubMed]
  18. Keerthana, S.P.; Yuvakkumar, R.; Ravi, G.; Kungumadevi, L.; Ravi Sankar, V. Hydrothermal Synthesis of CeVO4/g-C3N4 Z-Scheme Photocatalyst for Removal of Dyes under Photocatalysis. Inorg. Chem. Commun. 2024, 168, 112928. [Google Scholar] [CrossRef]
  19. Tian, J.; Qian, F.; Zhang, Y.; Li, W.; Li, J.; Chen, S.; Wang, L. Z-Scheme Membrane CdZnS/TiO2 Heterojunction Photocatalyst for Efficient Photocatalytic Removal of Microcystis aeruginosa under Simulated Sunlight: Adjustable Suspended Depth and Flexible Assembly. J. Mater. Sci. Technol. 2025, 217, 70–79. [Google Scholar] [CrossRef]
  20. Nguyen Van, K.; Nguyen Thi, V.N.; Tran Thi, T.P.; Truong, T.T.; Lieu Le Thi, T.; Tran Huu, H.; Nguyen, V.T.; Vo, V. A Novel Preparation of GaN-ZnO/g-C3N4 Photocatalyst for Methylene Blue Degradation. Chem. Phys. Lett. 2021, 763, 138191. [Google Scholar] [CrossRef]
  21. Kuspanov, Z.; Serik, A.; Tattibay, A.; Baratov, A.; Abdikarimova, U.; Bissenova, M.; Yeleyov, M.; Sakhiyev, S.; Daulbayev, C. Investigating and Correlating the Photocatalytic Activity of Synthesised Strontium Titanate Nanopowder with Calcination Temperature. Environ. Technol. Innov. 2024, 36, 103852. [Google Scholar] [CrossRef]
  22. Wang, Q.; Nakabayashi, M.; Hisatomi, T.; Sun, S.; Akiyama, S.; Wang, Z.; Pan, Z.; Xiao, X.; Watanabe, T.; Yamada, T.; et al. Oxysulfide Photocatalyst for Visible-Light-Driven Overall Water Splitting. Nat. Mater. 2019, 18, 827–832. [Google Scholar] [CrossRef]
  23. Serik, A.; Kuspanov, Z.; Bissenova, M.; Idrissov, N.; Yeleuov, M.; Umirzakov, A.; Daulbayev, C. Effective Photocatalytic Degradation of Sulfamethoxazole Using Pan/SrTiO3 Nanofibers 2024. J. Water Process Eng. 2024, 66, 106052. [Google Scholar] [CrossRef]
  24. Li, H.; Xiao, J.; Vequizo, J.J.M.; Hisatomi, T.; Nakabayashi, M.; Pan, Z.; Shibata, N.; Yamakata, A.; Takata, T.; Domen, K. One-Step Excitation Overall Water Splitting over a Modified Mg-Doped BaTaO2N Photocatalyst. ACS Catal. 2022, 12, 10179–10185. [Google Scholar] [CrossRef]
  25. Bissenova, M.; Umirzakov, A.; Mit, K.; Mereke, A.; Yerubayev, Y.; Serik, A.; Kuspanov, Z. Synthesis and Study of SrTiO3/TiO2 Hybrid Perovskite Nanotubes by Electrochemical Anodization. Molecules 2024, 29, 1101. [Google Scholar] [CrossRef] [PubMed]
  26. Chen, K.; Xiao, J.; Vequizo, J.J.M.; Hisatomi, T.; Ma, Y.; Nakabayashi, M.; Takata, T.; Yamakata, A.; Shibata, N.; Domen, K. Overall Water Splitting by a SrTaO2N-Based Photocatalyst Decorated with an Ir-Promoted Ru-Based Cocatalyst. J. Am. Chem. Soc. 2023, 145, 3839–3843. [Google Scholar] [CrossRef] [PubMed]
  27. Chen, X.; Shen, S.; Guo, L.; Mao, S.S. Semiconductor-Based Photocatalytic Hydrogen Generation. Chem. Rev. 2010, 110, 6503–6570. [Google Scholar] [CrossRef]
  28. Qin, Y.; Fang, F.; Xie, Z.; Lin, H.; Zhang, K.; Yu, X.; Chang, K. La,Al-Codoped SrTiO3 as a Photocatalyst in Overall Water Splitting: Significant Surface Engineering Effects on Defect Engineering. ACS Catal. 2021, 11, 11429–11439. [Google Scholar] [CrossRef]
  29. Fang, F.; Zhang, J.; Su, Z.; Xu, F.; Li, J.; Chang, K. Boost of Solar Water Splitting on SrTiO3 by Designing V-Ions Center for Localizing Defect Charge to Suppress Deep Trap. J. Catal. 2023, 425, 422–431. [Google Scholar] [CrossRef]
  30. Abdi, M.; Mahdikhah, V.; Sheibani, S. Visible Light Photocatalytic Performance of La-Fe Co-Doped SrTiO3 Perovskite Powder. Opt. Mater. 2020, 102, 109803. [Google Scholar] [CrossRef]
  31. Iriani, Y.; Sandi, D.K.; Hikmah, D.N.; Afriani, R.; Nurosyid, F.; Handoko, E.; Faquelle, D. Comparison Study of Aluminum (Al)-Doped Strontium Titanate (SrAlxTi1-xO3; x = 3% and 5%) Photocatalyst for Methylene Blue Degradation. Mater. Today Proc. 2024. [Google Scholar] [CrossRef]
  32. Abd Elkodous, M.; El-Khawaga, A.M.; Abouelela, M.M.; Abdel Maksoud, M.I.A. Cocatalyst Loaded Al-SrTiO3 Cubes for Congo Red Dye Photo-Degradation under Wide Range of Light. Sci. Rep. 2023, 13, 6331. [Google Scholar] [CrossRef]
  33. Kuspanov, Z.; Serik, A.; Matsko, N.; Bissenova, M.; Issadykov, A.; Yeleuov, M.; Daulbayev, C. Efficient Photocatalytic Degradation of Methylene Blue via Synergistic Dual Co-Catalyst on SrTiO3@Al under Visible Light: Experimental and DFT Study. J. Taiwan Inst. Chem. Eng. 2024, 165, 105806. [Google Scholar] [CrossRef]
  34. Bae, H.S.; Manikandan, V.; Hwang, J.H.; Seo, Y.-S.; Chung, H.-S.; Ryu, H.I.; Chae, W.-S.; Cho, M.; Ekambe, P.S.; Jang, J.S. Photocatalytic Degradation of Organic Pollutants and Inactivation of Pathogens under Visible Light via CoOx Surface-Modified Rh/Sb-Doped SrTiO3 Nanocube. J. Mater. Sci. 2021, 56, 17235–17253. [Google Scholar] [CrossRef]
  35. Anitha, B.G.; Devi, L.G. Study of Reaction Dynamics of Photocatalytic Degradation of 4-Chlorophenol Using SrTiO3, Sulfur Doped SrTiO3, Silver Metallized SrTiO3 and Silver Metallized Sulfur Doped SrTiO3 Catalysts: Detailed Analysis of Kinetic Results. Surf. Interfaces 2019, 16, 50–58. [Google Scholar] [CrossRef]
  36. Shen, Q.; Kang, W.; Ma, L.; Sun, Z.; Jin, B.; Li, H.; Miao, Y.; Jia, H.; Xue, J. Tuning the Anisotropic Facet of SrTiO3 to Promote Spatial Charge Separation for Enhancing Photocatalytic CO2 Reduction Properties. Chem. Eng. J. 2023, 478, 147338. [Google Scholar] [CrossRef]
  37. Wang, S.; Teramura, K.; Hisatomi, T.; Domen, K.; Asakura, H.; Hosokawa, S.; Tanaka, T. Effective Driving of Ag-Loaded and Al-Doped SrTiO3 under Irradiation at λ > 300 Nm for the Photocatalytic Conversion of CO2 by H2O. ACS Appl. Energy Mater. 2020, 3, 1468–1475. [Google Scholar] [CrossRef]
  38. Faisal, M.; Harraz, F.A.; Ismail, A.A.; El-Toni, A.M.; Al-Sayari, S.A.; Al-Hajry, A.; Al-Assiri, M.S. Polythiophene/Mesoporous SrTiO3 Nanocomposites with Enhanced Photocatalytic Activity under Visible Light. Sep. Purif. Technol. 2018, 190, 33–44. [Google Scholar] [CrossRef]
  39. Konstas, P.-S.; Konstantinou, I.; Petrakis, D.; Albanis, T. Development of SrTiO3 Photocatalysts with Visible Light Response Using Amino Acids as Dopant Sources for the Degradation of Organic Pollutants in Aqueous Systems. Catalysts 2018, 8, 528. [Google Scholar] [CrossRef]
  40. Kiran, K.S.; Shashanka, R.; Lokesh, S.V. Enhanced Photocatalytic Activity of Hydrothermally Synthesized Perovskite Strontium Titanate Nanocubes. Top. Catal. 2022. [Google Scholar] [CrossRef]
  41. Zhao, Z.; Goncalves, R.V.; Barman, S.K.; Willard, E.J.; Byle, E.; Perry, R.; Wu, Z.; Huda, M.N.; Moulé, A.J.; Osterloh, F.E. Electronic Structure Basis for Enhanced Overall Water Splitting Photocatalysis with Aluminum Doped SrTiO3 in Natural Sunlight. Energy Environ. Sci. 2019, 12, 1385–1395. [Google Scholar] [CrossRef]
  42. Li, R.; Takata, T.; Zhang, B.; Feng, C.; Wu, Q.; Cui, C.; Zhang, Z.; Domen, K.; Li, Y. Criteria for Efficient Photocatalytic Water Splitting Revealed by Studying Carrier Dynamics in a Model Al-Doped SrTiO3 Photocatalyst. Angew. Chem. Int. Ed. 2023, 62, e202313537. [Google Scholar] [CrossRef]
  43. Suwannaruang, T.; Kidkhunthod, P.; Butburee, T.; Shivaraju, H.P.; Shahmoradi, B.; Wantala, K. Facile Synthesis of Cooperative Mesoporous-Assembled CexSr1-xFexTi1-xO3 Perovskite Catalysts for Enhancement Beta-Lactam Antibiotic Photodegradation under Visible Light Irradiation. Surf. Interfaces 2021, 23, 101013. [Google Scholar] [CrossRef]
  44. Baek, J.-Y.; Duy, L.T.; Lee, S.Y.; Seo, H. Aluminum Doping for Optimization of Ultrathin and High-k Dielectric Layer Based on SrTiO3. J. Mater. Sci. Technol. 2020, 42, 28–37. [Google Scholar] [CrossRef]
  45. Bakbolat, B.; Daulbayev, C.; Sultanov, F.; Beissenov, R.; Umirzakov, A.; Mereke, A.; Bekbaev, A.; Chuprakov, I. Recent Developments of TiO2-Based Photocatalysis in the Hydrogen Evolution and Photodegradation: A Review. Nanomaterials 2020, 10, 1790. [Google Scholar] [CrossRef]
  46. Asgari, S.; Mohammadi Ziarani, G.; Badiei, A.; Vasseghian, Y. A Ternary Composite Nanofibrous Photocatalyst: FcLR-gC3N4/Polyisothianaphthene/Polyacrylonitrile for Degradation of Organic Dyes. J. Taiwan Inst. Chem. Eng. 2024, 163, 105672. [Google Scholar] [CrossRef]
  47. Sohrabian, M.; Mahdikhah, V.; Alimohammadi, E.; Sheibani, S. Improved Photocatalytic Performance of SrTiO3 through a Z-Scheme Polymeric-Perovskite Heterojunction with g-C3N4 and Plasmonic Resonance of Ag Mediator. Appl. Surf. Sci. 2023, 618, 156682. [Google Scholar] [CrossRef]
  48. Wang, B.; Li, P.; Du, C.; Wang, Y.; Gao, D.; Li, S.; Zhang, L.; Wen, F. Synergetic Effect of Dual Co-Catalysts on the Activity of BiVO4 for Photocatalytic Carbamazepine Degradation. RSC Adv. 2019, 9, 41977–41983. [Google Scholar] [CrossRef]
  49. Wang, Y.R.; Tao, H.L.; Cui, Y.; Liu, S.M.; He, M.; Song, B.; Jian, J.K.; Zhang, Z.H. Investigations on Tuning the Band Gaps of Al Doped SrTiO3. Chem. Phys. Lett. 2020, 757, 137879. [Google Scholar] [CrossRef]
  50. Fang, F.; Xu, F.; Su, Z.; Li, X.; Han, W.; Qin, Y.; Ye, J.; Chang, K. Understanding Targeted Modulation Mechanism in SrTiO3 Using K+ for Solar Water Splitting. Appl. Catal. B Environ. 2022, 316, 121613. [Google Scholar] [CrossRef]
  51. Kuspanov, Z.; Serik, A.; Baratov, A.; Abdikarimova, U.; Idrissov, N.; Bissenova, M.; Daulbayev, C. Efficient Photocatalytic Hydrogen Evolution via Cocatalyst Loaded Al-Doped SrTiO3. Eurasian Chem. Technol. J. 2024, 26, 133–140. [Google Scholar] [CrossRef]
  52. Amor, L.B.; Belgacem, B.; Filhol, J.-S.; Doublet, M.-L.; Yahia, M.B.; Hassen, R.B. New P-Type Al-Substituted SrSnO3 Perovskites for TCO Applications? Chem. Commun. 2020, 56, 2566–2569. [Google Scholar] [CrossRef]
  53. Toyoda, T.; Yabe, M. The Temperature Dependence of the Refractive Indices of Fused Silica and Crystal Quartz. J. Phys. D Appl. Phys. 1983, 16, L97. [Google Scholar] [CrossRef]
  54. Higuchi, T.; Tsukamoto, T.; Yamaguchi, S.; Kobayashi, K.; Sata, N.; Ishigame, M.; Shin, S. Observation of Acceptor Level of P-Type SrTiO3 by High-Resolution Soft-X-Ray Absorption Spectroscopy. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2003, 199, 255–259. [Google Scholar] [CrossRef]
  55. Guo, H.; Liu, L.; Fei, Y.; Xiang, W.; Lü, H.; Dai, S.; Zhou, Y.; Chen, Z. Optical Properties of P-Type In-Doped SrTiO3 Thin Films. J. Appl. Phys. 2003, 94, 4558–4562. [Google Scholar] [CrossRef]
  56. Kudaibergen, A.D.; Kuspanov, Z.B.; Issadykov, A.N.; Beisenov, R.E.; Mansurov, Z.A.; Yeleuov, M.A.; Daulbayev, C.B. Synthesis, Structure, and Energetic Characteristics of Perovskite Photocatalyst SrTiO3: An Experimental and DFT Study. Eurasian Chem. Technol. J. 2023, 25, 139–146. [Google Scholar] [CrossRef]
  57. Giannozzi, P.; Baroni, S.; Bonini, N.; Calandra, M.; Car, R.; Cavazzoni, C.; Ceresoli, D.; Chiarotti, G.L.; Cococcioni, M.; Dabo, I.; et al. Quantum Espresso: A Modular and Open-Source Software Project for Quantum Simulations of Materials. J. Phys. Condens. Matter 2009, 21, 395502. [Google Scholar] [CrossRef]
  58. Dudarev, S.L.; Botton, G.A.; Savrasov, S.Y.; Humphreys, C.J.; Sutton, A.P. Electron-Energy-Loss Spectra and the Structural Stability of Nickel Oxide: An LSDA+U Study. Phys. Rev. B 1998, 57, 1505–1509. [Google Scholar] [CrossRef]
  59. Ghosez, P.; Gonze, X.; Lambin, P.; Michenaud, J.-P. Born Effective Charges of Barium Titanate: Band-by-Band Decomposition and Sensitivity to Structural Features. Phys. Rev. B 1995, 51, 6765–6768. [Google Scholar] [CrossRef]
  60. Shah, S.H.; Bristowe, P.D.; Kolpak, A.M.; Rappe, A.M. First Principles Study of Three-Component SrTiO3/BaTiO3/PbTiO3 Ferroelectric Superlattices. J. Mater. Sci. 2008, 43, 3750–3760. [Google Scholar] [CrossRef]
Figure 1. Elementary supercell of SrTiO3@Al.
Figure 1. Elementary supercell of SrTiO3@Al.
Molecules 29 05326 g001
Figure 2. Morphology of samples: (a) SrTiO3 (1100 °C) obtained by SEM; (b) SrTiO3@Al obtained by TEM; (c) SrTiO3@Al obtained by SEM; and (d) SrTiO3@Al obtained by TEM.
Figure 2. Morphology of samples: (a) SrTiO3 (1100 °C) obtained by SEM; (b) SrTiO3@Al obtained by TEM; (c) SrTiO3@Al obtained by SEM; and (d) SrTiO3@Al obtained by TEM.
Molecules 29 05326 g002
Figure 3. X-ray radiographs, synthesised photocatalysts of (a) SrTiO3, SrTiO3@Al, and Rh/Cr2O3/SrTiO3@Al/CoOOH; and XPS spectra of (b) Al 2p, (c) Ti 2 p, and (d) O 1s of SrTiO3@Al sample.
Figure 3. X-ray radiographs, synthesised photocatalysts of (a) SrTiO3, SrTiO3@Al, and Rh/Cr2O3/SrTiO3@Al/CoOOH; and XPS spectra of (b) Al 2p, (c) Ti 2 p, and (d) O 1s of SrTiO3@Al sample.
Molecules 29 05326 g003
Figure 4. (a) Photocatalytic decomposition activity and (b) plots of ln ( C t C 0 ) versus time.
Figure 4. (a) Photocatalytic decomposition activity and (b) plots of ln ( C t C 0 ) versus time.
Molecules 29 05326 g004
Figure 5. (a) DFT + U electronic band structure of bare SrTiO3; and (b) density of states of Al-doped SrTiO3.
Figure 5. (a) DFT + U electronic band structure of bare SrTiO3; and (b) density of states of Al-doped SrTiO3.
Molecules 29 05326 g005
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

Abdikarimova, U.; Bissenova, M.; Matsko, N.; Issadykov, A.; Khromushin, I.; Aksenova, T.; Munasbayeva, K.; Slyamzhanov, E.; Serik, A. Visible Light-Driven Photocatalysis of Al-Doped SrTiO3: Experimental and DFT Study. Molecules 2024, 29, 5326. https://doi.org/10.3390/molecules29225326

AMA Style

Abdikarimova U, Bissenova M, Matsko N, Issadykov A, Khromushin I, Aksenova T, Munasbayeva K, Slyamzhanov E, Serik A. Visible Light-Driven Photocatalysis of Al-Doped SrTiO3: Experimental and DFT Study. Molecules. 2024; 29(22):5326. https://doi.org/10.3390/molecules29225326

Chicago/Turabian Style

Abdikarimova, Ulzhan, Madina Bissenova, Nikita Matsko, Aidos Issadykov, Igor Khromushin, Tatyana Aksenova, Karlygash Munasbayeva, Erasyl Slyamzhanov, and Aigerim Serik. 2024. "Visible Light-Driven Photocatalysis of Al-Doped SrTiO3: Experimental and DFT Study" Molecules 29, no. 22: 5326. https://doi.org/10.3390/molecules29225326

APA Style

Abdikarimova, U., Bissenova, M., Matsko, N., Issadykov, A., Khromushin, I., Aksenova, T., Munasbayeva, K., Slyamzhanov, E., & Serik, A. (2024). Visible Light-Driven Photocatalysis of Al-Doped SrTiO3: Experimental and DFT Study. Molecules, 29(22), 5326. https://doi.org/10.3390/molecules29225326

Article Metrics

Back to TopTop