Next Article in Journal
Origins of a Low-Sulfur Superalloy Al2O3 Scale Adhesion Map
Previous Article in Journal
Single-Longitudinal-Mode Laser at 1123 nm Based on a Twisted-Mode Cavity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Efficient Photocatalytic Degradation of RhB by Constructing Sn3O4 Nanoflakes on Sulfur-Doped NaTaO3 Nanocubes

1
School of Materials Science and Engineering, Shandong Jianzhu University, Fengming Road, Jinan 250101, China
2
State Key Laboratory of Crystal Materials, Shandong University, Jinan 250100, China
*
Author to whom correspondence should be addressed.
Submission received: 5 November 2020 / Revised: 6 January 2021 / Accepted: 6 January 2021 / Published: 13 January 2021
(This article belongs to the Section Crystal Engineering)

Abstract

:
Band structure engineering and heterojunction photocatalyst construction are efficient approaches to improve the separation of photo-induced electrons and holes, along with enhancing light response ability. By sulfur doping, sodium tantalite (NaTaO3) showed an improved photocatalytic property for the degradation of Rhodamine B (RhB). Sn3O4 nanoflakes were constructed on the surface of NaTaO3 nanocubes, forming a surface heterostructure via a simple hydrothermal process, initially. This heterostructure endows the photocatalyst with an enhanced charge separation rate, resulting in an improved photocatalytic degradation of RhB. Moreover, a possible mechanism over Sn3O4/NaTaO3 and the photodegradation pathway of RhB were proposed as the combined effect of photo-induced electrons and holes. This facile process for band structure engineering and heterostructure construction provides the possibility for the practical application of high-efficiency photocatalysts.

1. Introduction

In recent years, the extensive treatment of wastewater containing organic pollutants has led to a negative impact on the environment and serious consequences for human health. Although considerable efforts and strategies have been taken to alleviate the above-mentioned problems, the photocatalytic oxidation of water as an advanced sewage disposal mechanism is of confocal interest and is considered an effective solution [1,2,3,4,5,6,7]. For this reason, there has been an upsurge in research on a variety of photocatalysts. Among these, ABO3 perovskite-type oxides have been widely studied, due to their high performance in photocatalytic reactions, such as with overall water splitting and the photodegradation of organic pollutants, which provide a potential approach to solve our urgent energy and environmental problems [8,9]. Sodium tantalite (NaTaO3) is a prominent ABO3 perovskite-like oxide with a high performance in photocatalytic reactions, including abundance, photochemical stability, photodegradation of organic pollutants, water splitting, and low environmental impact, which attracts a lot of attention for its favorable layered structure and distinctive separation effectiveness in charge separation [10,11,12,13,14,15,16,17,18,19,20]. Hence, NaTaO3-based photocatalysts can intrinsically decrease the recombination rate of photo-induced carriers and promote the efficiency of photocatalytic systems. It has been well-documented that tantalates are superior to other metal oxides. Moreover, NaTaO3 is a new type of wide bandgap semiconductor material, which has a bandgap of about 4.0 eV at room temperature [10]. To enhance the photocatalytic performance, metal or non-metal ions doped with NaTaO3 (dopants: Sr2+ [15,17], La3+ [20,21], Ba2+ [22], Ca2+ [23], N or S [12,23]) or semiconductors loaded on NaTaO3, such In2S3 [10], WO3 [11], NiO [24], AgCl/Ag2O [25], g-C3N4 [18,26], RuO2 [27], and CdS [28] have been typically studied. These works were developed to improve the photocatalytic activity of NaTaO3-based materials. The assessment of composite photocatalysts included photocatalytic water splitting and photodegradation of organic pollutants. These techniques and other strategies make NaTaO3 a better photocatalyst. Concerning the aforementioned, the potentiality of NaTaO3-based photocatalysts is apparent.
To date, heterovalent tin oxide Sn3O4 formed of mixed valences of Sn2+ and Sn4+ has garnered noticeable attention. One third of Sn atoms are located in Sn (II) tetrahedral coordination sites and two-thirds are located in Sn (IV) octahedral sites [29,30,31]. Additionally, the Sn3O4 tin structure has been reported to influence the physical properties of tin oxide. As proposed by theoretical calculations, Sn3O4 has been experimentally confirmed to possess absorption bands in a limited visible-light region, which may have promising applications as a potential photocatalyst in environmental remediation and energy conversion. With a direct bandgap energy of around 2.56 eV, Sn3O4 has also been widely studied for its relatively negative conduction band edge and moderate charge transport features. The proper energy band makes it a better material for constructing heterojunction catalysts with other semiconductor materials [32,33,34,35,36,37]. Hence, it is very vital to broaden the photocatalytic active spectral range and improve the separation of carriers of Sn3O4.
To the best of our knowledge, there is no research on the photocatalytic performance of the Sn3O4/NaTaO3 heterostructure. Therefore, in the current work, sulfur-doped NaTaO3 cubes were synthesized to modify the band structure of NaTaO3. Based on the appropriate valance band (EVB ~3.01 eV) and conduction band edge (ECB ~0.99 eV) of NaTaO3 [12,28], which can match well with that of Sn3O4 (EVB ~1.07 eV and ECB ~1.55 eV vs. NHE, pH = 7) [30], the heterojunction of Sn3O4/NaTaO3 was synthesized. An enhanced photo-induced charge separation was highly expected, which results in the enhanced photocatalytic degradation of RhB. This work provides an ordinary and low-cost method for the large-scale production of NaTaO3-based materials in various applications. It is expected to offer an upfront approach for developing highly stable and effective heterostructures for organic pollutant degradation.

2. Materials and Methods

2.1. Materials

Tantalum oxide (Ta2O5), hydrochloric acid (HCl), sulfuric acid (H2SO4), sodium hydroxide (NaOH), tin(II) chloride dehydrate (SnCl2·2H2O), ethanol (C2H5OH), sodium citrate dihydrate (Na3C6H5O7·2H2O), and Rhodamine B (RhB) were purchased from Sinopharm Chemical Reagent Co., Ltd. and used without any further purification. Deionized water was used throughout the study.

2.2. Synthesis

NaTaO3 and sulfur anion doped in NaTaO3 were prepared via the reported effortless hydrothermal approach. In a typical synthesis process, 0.442 g of Ta2O5, 1.2 g of NaOH, and a certain amount of Na2S2O3·5H2O were added into a Teflon-lined stainless steel autoclave (50 mL capacity) that was filled with deionized water to 75% of the total volume. The autoclave was sealed and put into a preheated oven to perform hydrothermal treatment at 180 °C for 12 h. After cooling to room temperature, the acquired precipitates were collected by centrifugation and thoroughly washed with deionized water several times, and then dried at 80 °C for 12 h before further characterization and photocatalytic reaction. The obtained products were NaTaO3 powders and S-doped NaTaO3 powders, respectively; the latter had a calculated S:Ta molar ratio of 5%, where the amount of Na2S2O3•5H2O was 0.0248 g.
The Sn3O4/NaTaO3 and Sn3O4/NaTaO3-S were prepared via a facile hydrothermal coprecipitation method. First, 5.0 mmol of SnCl2•2H2O and 12.5 mmol of Na3C6H5O7•2H2O were dissolved in 12.5 mL of deionized water and stirred for 5 min to get a transparent solution, where a certain amount of NaTaO3 or NaTaO3-S (0.175 g) was added. Then, 12.5 mL of 0.2 M NaOH aqueous solution was added to the above-mentioned solution, with vigorous stirring followed by ultrasonic treatment and the solution was transferred to a 50 mL Teflon-lined stainless steel autoclave and maintained at 180 °C for 12 h. The acquired powder was washed with deionized water and ethanol several times and then finally dried at 60 °C for 12 h. Sn3O4/NaTaO3 and Sn3O4/NaTaO3-S composites were prepared with the molar ratio of Sn:Ta 1:1. Pure Sn3O4 was synthesized without the addition of cubic NaTaO3 in a similar process.
The samples acquired above are denoted as NaTaO3, NTO-S, Sn/NTO, Sn/NTO-S, and Sn3O4, respectively.

2.3. Characterizations

The crystalline structures of the products were characterized by X-ray diffraction (XRD) on a Bruker D8 Advance powder X-ray diffractometer with Cu Kα radiation (λ = 0.15406 nm). The morphologies of the as-obtained catalysts were recorded by field emission scanning electron microscopy (FE-SEM, HITACHIS-4800). The microstructures of the samples were recorded using high-resolution transmission electron microscopy (HRTEM, JOEL JEM 2100). UV–vis diffuse reflectance spectra (DRS) of the catalysts were acquired on a UV–vis spectrophotometer (UV-2550, Shimadzu) with an integrating sphere attachment within the wavelength range from 200 nm to 800 nm along with BaSO4 as the reflectance standard. The photoluminescence (PL) emission spectra were carried out with an FLS920 fluorescence spectrometer at room temperature under the excitation wavelength of 300 nm.

2.4. Photocatalytic Degradation of RhB

The photocatalytic activities of as-obtained products determined by the degradation of RhB were conducted in a photoreaction apparatus, an XPA-II photochemical reactor (XPA-II, Nanjing Xujiang Machine-electronic Plant, China). In typical processing, a 30 mg catalyst was added to 30 mL of RhB solution (20 mg/L). The mixer was stirred for 30 min in the dark to achieve the adsorption–desorption equilibrium between the dye and the surface of the catalyst, then the reaction system was performed via the irradiation of a 300 W mercury lamp and by continuously stirring under favorable ambient conditions. At certain intervals of time, 4 mL of the mixed solution was collected and taken out for centrifugation to remove the catalyst and analyzed using a UV–vis spectrophotometer (UV-6100, Metash). The residual samples were collected for repeated photocatalytic reactions.

3. Results and Discussion

To know the phase composition of hetero photocatalyst, crystallinity, and purity, XRD was performed and is presented in Figure 1. The XRD pattern of NaTaO3 shows that the peaks at 22.9°, 32.5°, 40.1°, 46.6°, 52.5°, 57.8°, 57.9°, and 67.9° are attributed to the (100), (101), (111), (200), ( 1 ¯ 02), ( 1 ¯ 12), ( 1 ¯ 21), and (022) diffraction planes, respectively. All the diffraction peaks can be readily indexed as a pure perovskite NaTaO3 with a monoclinic structure (JCPDS No. 74-2478) [10,12], which can facilitate the separation of photo-induced electron–hole pairs. As has been reported, NaTaO3 may exist in three polymorphs, which are the orthorhombic, cubic, and monoclinic phases [10,12]. All the phases share TaO6 octahedra frameworks at the corners, along with sodium ions residing in the dodecahedral interspaces, but the local structure distortion is highly different among various polymorphs [12]. The TaO6 octahedra of the monoclinic phase were close to the ideal perovskite, favoring the separation of photo-induced electrons and holes and representing better photocatalytic activity [14]. After doping the S anion, it can also be seen that the diffraction peaks of NTO-S series samples still matched the monoclinic structure. As for bare Sn3O4, its XRD pattern showed poorer crystallinity and lower diffraction peak intensity. The principal diffraction peaks may correspond to the (101), ( 1 ¯ 20), (111), ( 2 ¯ 10), ( 1 ¯ 21), and (210) crystal planes of standard triclinic-phase Sn3O4 (JCPDS No.16-0737) [29,30,31,34]. Additionally, the XRD patterns of Sn/NTO and Sn/NTO-S have no detected impurities, corresponding to the aforementioned NaTaO3 and Sn3O4.
The morphologies and microstructural details of the as-prepared NaTaO3, NTO-S, Sn/NTO, and Sn/NTO-S were discussed based on the SEM and TEM images. Figure 2 contains representative SEM images of the as-obtained samples. As shown (inset of Figure 2a,b), NaTaO3 and NTO-S presented an approximate cubic structure with an average size between 300 nm and 500 nm, while the surfaces of some of them were a little coarse. Figure 2c,d shows the synthesized Sn3O4 clusters with low and high magnification, respectively, which are composed of thin flakes with a size range between 100 nm and 400 nm and a thickness of approximately 5–10 nm. Then the SEM images of Sn/NTO (Figure 2e) and Sn/NTO-S (Figure 2f) are displayed. Their surfaces were rough, which consisted of Sn3O4 nanoflakes and might provide still more active faces and a larger specific area beneficial to heterogeneous nucleation and assembly growth of Sn3O4 nanoflakes.
As shown in Figure 3a,b, NaTaO3 always shows the cubic morphology before and after constructing with Sn3O4. Additionally, the average size of those composites changes little, approximately 300 nm. Evidently, the cubic NaTaO3 was essentially surrounded by sheet-like Sn3O4 and a portion of small slices that were well dispersed on the outside surface of the cubic NaTaO3 to compose a representative Sn/NTO heterojunction. Hence, Figure 3c illustrates a schematic diagram of the growth process of Sn/NTO in two steps. The first step is heterogeneous nucleation of Sn3O4 seeds on the outside surface of the cubic NaTaO3, driven by the adsorption connecting -OH groups on the surface of the cubic NaTaO3 with Sn2+ in the precursor [31]. The second step is Sn3O4 nanoflakes successfully growing an assembly on the outside surface of cubic NaTaO3. The heterojunction between Sn3O4 and NaTaO3 would facilitate photoresponse and interfacial electron transfer and therefore enhance the photocatalytic activity.
The light properties of a material are very sensitive to and seriously influenced by their intrinsic microstructures and, therefore, any changes in electronic structure. Figure 4a determines the UV–vis diffuse reflectance spectra (DRS) of NaTaO3, NTO-S, Sn/NTO, and Sn/NTO-S (molar ratio Sn/Ta = 1/1) heterostructures, along with pure Sn3O4. The bare NaTaO3 presented steep absorption only in the UV region located at 308 nm. Sulfur ion doping in NaTaO3 showed an intense absorption with an absorption red-shift of 323 nm. The pure Sn3O4 could absorb the light irradiation as its optical absorption was determined to be around 473 nm. After coupling NaTaO3 or NTO-S, it was found that Sn/NTO and Sn/NTO-S heterostructures exhibited a broader absorption peak in contrast to pure NaTaO3. The Sn/NTO composites presented a significant shift toward 484.7 nm. However, the Sn/NTO-S heterostructure showed a continuous red-shift to 523.5 nm. According to hv = 1240/λ and the Kubelka–Munk (KM) method, with the equation:   α h v = A   ( h v E g ) 2 , in which α , hv, Eg, and A are the absorption coefficient, the photon energy, indirect bandgap, and a constant, respectively [10,11,31], the bandgap energies of NaTaO3, NTO-S, Sn3O4, Sn/NTO, and Sn/NTO-S shown in Figure 4b were calculated to be 4.03 eV, 3.84 eV, 2.62 eV, 2.56 eV, and 2.37 eV, respectively. Hence, the bandgaps of the composites were greatly reduced. Interestingly, the bandgap values of Sn/NTO and Sn/NTO-S composites were not between pure NaTaO3 and Sn3O4; the optical transitions of the composites extend much wider. Consequently, it can be inferred that the bandgap value of the composite narrowed via sulfur doping; then, the light absorption was further enhanced by coupling Sn3O4. The enhanced light response could be attributed to the formation of a heterojunction between them [10,11,18,26,27,28,29,31].
To investigate the photocatalytic degradation capabilities of the as-obtained catalysts, we carried out a degradation properties assessment using different samples on RhB in a photocatalytic device that had been specially designed. The degradation of RhB and the corresponding kinetic plots are displayed in Figure 5a,b, respectively. Figure 5a shows the different relative concentration variations of the RhB over 140 min. C0 is the initial concentration of the RhB, and Ct is the concentration of the RhB after a reaction for t. The concentration of the RhB was almost constant, which implies that the RhB had attained the adsorption equilibrium in a short time and prior reaction. Therefore, it was confirmed that the decrease of RhB was a result of degradation as time went by. Cubic NaTaO3 presented a steady degradation rate under UV light irradiation. After 120 min, the concentration of the RhB decreased by 63%. The photodegradation of the RhB in the presence of pure Sn3O4 performed relatively slowest, as only 18% of the RhB was degraded in the 120 min illumination period. The reduced degradation rate possibly contributed to the inhibited electron transportation efficiency and narrow spectral response in bare Sn3O4, which led to the fast recombination of photo-induced electron–hole pairs and poor activity. It is well known that NaTaO3 exhibits low photocatalytic activity, even under UV illumination, due to its large bandgap. However, after having been doped in sulfur ion or coupled with Sn3O4, the photocatalytic degradation ability of NTO-S or Sn/NTO was very much enhanced in contrast to that of the pure NaTaO3 or the Sn3O4 catalyst. Interestingly, both displayed similar photocatalytic degradation rates of RhB under UV light irradiation. The Sn/NTO-S showed the best performance. The RhB was degraded almost entirely in 120 min, which also demonstrated that heterojunction structure plays a decisive role in the transportation and separation of photo-induced carriers.
Furthermore, the degradation rate for RhB can be acquired from the slope (min−1) of –Ln(Ct/C0) ~ t plots (Figure 5b); the calculated results of the decolorization rate constant κ (min−1) for NaTaO3, NTO-S, Sn/NTO, Sn/NTO-S, and Sn3O4 were 0.3366, 0.5790, 0.6060, 1.3350, and 0.0555 min−1, respectively. Evidently, the Sn/NTO-S catalyst exhibited the highest κ values (1.3550 min−1), which were about 4.0 times those of NaTaO3.
Based on the experimental results mentioned above, a schematic model of a proposed reaction pathway of degraded products is illustrated in Scheme 1. The band edge positionings of NaTaO3 and Sn3O4 were carefully made based on experimental results and earlier studies. Via DRS, the bandgap energy of NaTaO3 (4.03 eV) was wider than that of Sn3O4 (2.62 eV) [38]. The CB of NaTaO3 was estimated to be −0.99 eV, as documented [10,39]. The narrowing of bandgap energy as a result of doping S anion is a very key factor. The photodegradation efficiency of NTO-S was found to be greater than that of pure NaTaO3. The reason may be that the substitution of resident O atoms within the NaTaO3 lattice by sulfur atoms induced the local states just above the valence level as the new hybridized orbitals were formed [12,40]. Aside from the bandgap width, the photocatalytic degradation activity of a sample highly depended on the recombination of electron (e)–hole (h+) pairs [12,41,42,43,44]. Under photo-irradiation, the photo-induced e−/h+ pairs on the CB/VB of each semiconductor were created. As reported previously, the principal active species for the photocatalytic degradation of RhB included hydroxyl radicals (•OH) and superoxide radicals (•O2), which were composed of the creation of h+ and e [34], and some researchers have explored intermediates and key reaction species [45,46,47]. S-doped catalysts displayed a less intense emission than pure NaTaO3 samples, indicating a reduced e/h+ recombination and an enhanced RhB photodegradation performance [12].
This plausible mechanism comprised excitation of the CB of the Sn3O4; then, the e on the CB of the NaTaO3 reacted with the available O2 to produce the superoxide radicals •O2. Then, the •O2 degraded RhB, or the RhB intermediate, into degraded products [45]. The h+ of the NTO-S transmitted to the VB of the Sn3O4 [31], and holes contributed more to the high photocatalytic performance, while •OH showed less importance in its ability to degrade RhB [34]. In this case, the RhB reacted with h+ and formed the intermediate RhB that was finally converted to degraded products. Comprehensively, the photocatalytic degradation mechanisms could be depicted as the interfacial charge transfer of the composite-enhanced RhB photosensitization process under irradiation, which could effectively separate and transfer photogenerated carriers. Therefore, the resultant •O2 and h+ favored the photodegradation procedure, and the coupled semiconducting materials formed more dynamic catalytic centers that improved the efficiency of RhB dye removal. This suggested that high-quality interfaces in the heterostructure composites played a vital role in enhancing the catalytic performance.
To prove the rationality of the proposed reaction pathway and investigate the capabilities of photo-induced e−/h+ pairs in the semiconductors, PL emission spectra of NaTaO3, NTO-S, Sn/NTO, Sn/NTO-S, and Sn3O4 were carried out, respectively. PL emission arises from the recombination of free carriers, and a higher PL intensity reveals a higher recombination rate of photogenerated e−/h+ pairs. By contrast, a weaker PL intensity indicates a lower recombination rate of photo-excited e−/h+ pairs, and accordingly, much more photo-induced carriers can participate in the photocatalytic reaction [10,34]. As shown in Figure 6, the spectrum of NaTaO3 nanocubes showed three peaks at around 395 nm, 451 nm, and 469 nm. The emission at 451 nm and 469 nm was due to band-to-band mixing of Ta4+-O states in the octahedral TaO6 motifs of NaTaO3 [48]. After doping the S anion in NaTaO3, the PL intensity of NTO-S was weaker than that of pure NaTaO3. It indicated that the NTO-S could suppress the recombination of e−/h+ pairs. Noticeably, the intensities of the PL spectra of Sn/NTO and the Sn/NTO-S heterojunction photocatalysts strongly decreased with the Sn/NTO heterojunctions formed as compared to pure cubic NaTaO3; this can be seen at the maximum PL intensity at around 469 nm. The lower PL intensity implied the efficient inhibition of photogenerated pairs and therefore higher photocatalytic activity in the UV light region, which could effectively explain the variation of the photocatalytic activities of the different samples.

4. Conclusions

On the whole, we have successfully synthesized a novel Sn3O4/NaTaO3 heterojunction with doping sulfur anion via a simple hydrothermal method. The bandgap value of NaTaO3 and the composite narrowed via sulfur doping, which induced an increase in light absorption. With coupling Sn3O4, it was found that the as-obtained Sn/NTO and Sn/NTO-S composites exhibited cubic NaTaO3, or that NTO-S assembled by countless interlaced fine Sn3O4 nanoflakes were successfully synthesized for the first time. The fabrication of enough interfaces with high qualities played a critical role in light absorption and enhancing photocatalytic performance. Assessed by the degradation of the RhB solution, the Sn/NTO-S composite presents a superior photoreactivity toward RhB degradation under UV light illumination in contrast to individual Sn3O4 or NaTaO3. The calculated decolorization rate constant is much greater (1.3550 min−1) than single Sn3O4 (0.0555 min−1) or NaTaO3 (0.3366 min−1). This could be due to the enhanced light absorbance by the larger specific surface area of the Sn3O4 nanoflakes and the formation of the heterojunction between NTO-S and Sn3O4, which can separate photo-induced carriers efficiently. It is also proposed that the formed •O2 and h+ favored the photocatalytic degradation of RhB. This work highlights the importance of heterostructured Sn3O4/NaTaO3, which may provide a lead to develop high potential and NaTaO3-based photocatalysts for the degradation of organic dyes.

Author Contributions

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

Funding

This research was financially supported by the 2018 Doctoral Research Funds of Shandong Jianzhu University (X18064Z), Joint Fund Project for Natural Science Foundation of Shandong Province (ZR2019LLZ002).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

This work received technical support from the State Key Laboratory of Crystal Materials in Shandong University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Akbal, F. Photocatalytic degradation of organic dyes in the presence of titanium dioxide under UV and solar light: Effect of operational parameters. Env. Prog. 2005, 24, 317–322. [Google Scholar] [CrossRef]
  2. Schneider, J.; Matsuoka, M.; Takeuchi, M.; Zhang, J.; Horiuchi, Y.; Anpo, M.; Bahnemann, D.W. Understanding TiO2 photocatalysis: Mechanisms and materials. Chem. Rev. 2014, 114, 9919–9986. [Google Scholar] [CrossRef] [PubMed]
  3. Yang, J.; Wang, D.; Han, H.; Li, C. Roles of Cocatalysts in Photocatalysis and Photoelectrocatalysis. Acc. Chem. Res. 2013, 46, 1900–1909. [Google Scholar] [CrossRef] [PubMed]
  4. Matatov-Meytal, Y.I.; Sheintuch, M. Catalytic Abatement of Water Pollutants. Ind. Eng. Chem. Res. 1998, 37, 309–326. [Google Scholar] [CrossRef]
  5. Kabra, K.; Chaudhary, R.; Sawhney, R.L. Treatment of Hazardous Organic and Inorganic Compounds through Aqueous-Phase Photocatalysis:  A Review. Ind. Eng. Chem. Res. 2004, 43, 7683–7696. [Google Scholar] [CrossRef]
  6. Chang, S.; Yang, X.; Sang, Y.; Liu, H. Highly Efficient Photocatalysts and Continuous-Flow Photocatalytic Reactors for Degradation of Organic Pollutants in Wastewater. Chem. Asian J. 2016, 11, 2352–2371. [Google Scholar] [CrossRef] [PubMed]
  7. Chang, S.; Wang, Q.; Liu, B.; Sang, Y.; Liu, H. Hierarchical TiO2 nanonetwork–porous Ti 3D hybrid photocatalysts for continuous-flow photoelectrodegradation of organic pollutants. Catal. Sci. Technol. 2017, 7, 524–532. [Google Scholar] [CrossRef]
  8. Someshwar Pola and Ramesh Gade. Significant Role of Perovskite Materials for Degradation of Organic Pollutants. Available online: https://www.intechopen.com/online-first/significant-role-of-perovskite-materials-for-degradation-of-organic-pollutants (accessed on 3 April 2020).
  9. Orak, C.; Atalay, S.; Ersöz, G. Photocatalytic and photo-Fenton-like degradation of methylparaben on monolith-supported perovskite-type catalysts. Sep. Sci. Technol. 2017, 52, 1310–1320. [Google Scholar] [CrossRef]
  10. Xu, J.; Luo, B.; Gu, W.; Jian, Y.; Wu, F.; Tang, Y.; Shen, H. Fabrication of In2S3/NaTaO3 composites for enhancing the photocatalytic activity toward the degradation of tetracycline. New J. Chem. 2018, 42, 5052–5058. [Google Scholar] [CrossRef]
  11. Qu, L.; Lang, J.; Wang, S.; Chai, Z.; Su, Y.; Wang, X. Nanospherical composite of WO3 wrapped NaTaO3: Improved photodegradation of tetracycline under visible light irradiation. Appl. Surf. Sci. 2016, 388, 412–419. [Google Scholar] [CrossRef]
  12. Li, H.; Shi, X.; Liu, X.; Li, X. Synthesis of novel, visible-light driven S,N-doped NaTaO3 catalysts with high photocatalytic activity. Appl. Surf. Sci. 2020, 508, 145306. [Google Scholar] [CrossRef]
  13. Kato, H.; Kudo, A. Highly efficient decomposition of pure water into H2 and O2 over NaTaO3 photocatalysts. Catal. Lett. 1999, 58, 153–155. [Google Scholar] [CrossRef]
  14. Hu, C.; Tsai, C.; Teng, H. Structure Characterization and Tuning of Perovskite-Like NaTaO3 for Applications in Photoluminescence and Photocatalysis. J. Am. Ceram. Soc. 2009, 92, 460–466. [Google Scholar] [CrossRef]
  15. An, L.; Sasaki, T.; Weidler, P.G.; Wöll, C.; Ichikuni, N.; Onishi, H. Local Environment of Strontium Cations Activating NaTaO3 Photocatalysts. ACS Catal. 2018, 8, 880–885. [Google Scholar] [CrossRef] [Green Version]
  16. He, Y.; Zhu, Y.; Wu, N. Synthesis of nanosized NaTaO3 in low temperature and its photocatalytic performance. J. Solid State Chem. 2004, 177, 3868–3872. [Google Scholar] [CrossRef]
  17. An, L.; Onishi, H. Electron–Hole Recombination Controlled by Metal Doping Sites in NaTaO3 Photocatalysts. ACS Catal. 2015, 5, 3196–3206. [Google Scholar] [CrossRef]
  18. Yang, F.; Yan, L.; Zhang, B.; He, X.; Li, Y.; Tang, Y.; Ma, C.; Li, Y. Fabrication of ternary NaTaO3/g-C3N4/G heterojunction photocatalyst with enhanced activity for Rhodamine B degradation. J. Alloys Compd. 2019, 805, 802–810. [Google Scholar] [CrossRef]
  19. Zhou, X.; Chen, Y.; Mei, H.; Hu, Z.; Fan, Y. A facile route for the preparation of morphology-controlled NaTaO3 films. Appl. Surf. Sci. 2008, 255, 2803–2807. [Google Scholar] [CrossRef]
  20. Torresmartinez, L.M.; Cruzlopez, A.; Juarezramirez, I.; Meza-de la Rosa, M.E. Methylene blue degradation by NaTaO3 sol-gel doped with Sm and La. J. Hazard. Mater. 2009, 165, 774–779. [Google Scholar] [CrossRef]
  21. Kato, H.; Asakura, A.K.; Kudo, A. Highly Efficient Water Splitting into H2 and O2 over Lanthanum-Doped NaTaO3 Photocatalysts with High Crystallinity and Surface Nanostructure. J. Am. Chem. Soc. 2003, 125, 3082–3089. [Google Scholar] [CrossRef]
  22. Iwase, A.; Kato, H.; Kudo, A. The Effect of Alkaline Earth Metal Ion Dopants on Photocatalytic Water Splitting by NaTaO3 Powder. ChemSusChem 2009, 2, 873–877. [Google Scholar] [CrossRef] [PubMed]
  23. Liu, D.; Jiang, Y.; Gao, G. Photocatalytic degradation of an azo dye using N-doped NaTaO3 synthesized by one-step hydrothermal process. Chemosphere 2011, 83, 1546–1552. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, M.; Ma, Y.; Fo, Y.; Lyu, Y.; Zhou, X. Theoretical insights into the origin of highly efficient photocatalyst NiO/NaTaO3 for overall water splitting. Int. J. Hydrogen Energy 2020, 45, 19357–19369. [Google Scholar] [CrossRef]
  25. Xu, D.; Shi, W.; Yang, S.; Chen, B.; Bai, H.; Xiao, L. Fabrication of ternary p-n heterostructures AgCl/Ag2O/NaTaO3 photocatalysts: Enhanced charge separation and photocatalytic properties under visible light irradiation. Catal. Commun. 2016, 84, 163–166. [Google Scholar] [CrossRef]
  26. Tang, L.; Feng, C.; Deng, Y.; Zeng., G.; Wang, J.; Liu, Y.; Feng, H.; Wang, J. Enhanced photocatalytic activity of ternary Ag/g-C3N4/NaTaO3 photocatalysts under wide spectrum light radiation: The high potential band protection mechanism. Appl. Catal. B Environ. 2018, 230, 102–114. [Google Scholar] [CrossRef]
  27. Gomezsolis, C.; Ballesteros, J.C.; Torresmartinez, L.M.; Juárez-Ramírez, I. RuO2–NaTaO3 heterostructure for its application in photoelectrochemical water splitting under simulated sunlight illumination. Fuel 2016, 166, 36–41. [Google Scholar] [CrossRef]
  28. Singh, A.P.; Kumar, S.; Thirumal, M. Efficient Charge Transfer in Heterostructures of CdS/NaTaO3 with Improved Visible-Light-Driven Photocatalytic Activity. ACS Omega 2019, 4, 12175–12185. [Google Scholar] [CrossRef] [Green Version]
  29. Xia, W.; Wang, H.; Zeng, X.; Han, J.; Zhu, J.; Zhou, M.; Wu, S. High-efficiency photocatalytic activity of type II SnO/Sn3O4 heterostructures via interfacial charge transfer. CrystEngComm 2014, 16, 6841–6847. [Google Scholar] [CrossRef]
  30. He, Y.; Li, D.; Chen, J.; Shao, Y.; Xian, J.; Zheng, X.; Wang, P. Sn3O4: A novel heterovalent-tin photocatalyst with hierarchical 3D nanostructures under visible light. Rsc Adv. 2014, 4, 1266–1269. [Google Scholar] [CrossRef]
  31. Chen, G.; Ji, S.; Sang, Y.; Chang, S.; Wang, Y.; Hao, P.; Claverie, J.; Liu, H.; Yu, G. Synthesis of scaly Sn3O4/TiO2 nanobelt heterostructures for enhanced UV-visible light photocatalytic activity. Nanoscale 2015, 7, 3117–3125. [Google Scholar] [CrossRef]
  32. Park, S.H.; Son, Y.C.; Willis, W.S.; Suib, S.L.; Creasy, K.E. Tin oxide films made by physical vapor deposition thermal oxidation and spray pyrolysis. Chem. Mater. 1998, 10, 2389–2398. [Google Scholar] [CrossRef]
  33. Yu, X.; Zhao, Z.; Ren, N.; Liu, J.; Sun, D.; Ding, L.; Liu, H. Top or Bottom, Assembling Modules Determine the Photocatalytic Property of the Sheetlike Nanostructured Hybrid Photocatalyst Composed with Sn3O4 and rGO (GQD). ACS Sustain. Chem. Eng. 2018, 6, 11775–11782. [Google Scholar] [CrossRef]
  34. Hu, J.; Li, X.; Wang, X.; Li, Q.; Wang, F. Novel hierarchical Sn3O4/BiOX (X = Cl, Br, I) p–n heterostructures with enhanced photocatalytic activity under simulated solar light irradiation. Dalton Trans. 2019, 48, 8937–8947. [Google Scholar] [CrossRef] [PubMed]
  35. Yu, H.; Li, J.; Luo, W.; Li, Z.; Tian, Y.; Yang, Z.; Gao, Z.; Liu, H. Hetero-structure La2O3-modified SnO2-Sn3O4 from tin anode slime for highly sensitive and ppb-Level formaldehyde detection. Appl. Surf. Sci. 2020, 513, 145825. [Google Scholar] [CrossRef]
  36. Manikandan, M.; Tanabe, T.; Li, P.; Ueda, S.; Ramesh, G.V.; Kodiyath, R. Photocatalytic Water Splitting under Visible Light by Mixed-Valence Sn3O4. ACS Appl. Mater. Interfaces 2014, 6, 3790–3793. [Google Scholar] [CrossRef] [PubMed]
  37. Hu, J.; Tu, J.; Li, X.; Wang, Z.; Li, Y.; Li, Q.; Wang, F. Enhanced UV-Visible Light Photocatalytic Activity by Constructing Appropriate Heterostructures between Mesopore TiO2 Nanospheres and Sn3O4 Nanoparticles. Nanomaterials 2017, 7, 336. [Google Scholar] [CrossRef] [Green Version]
  38. Portugal, G.R.; Santos, S.F.; Arantes, J.T. NaTaO3 cubic and orthorhombic surfaces: An intrinsic improvement of photocatalytic properties. Appl. Surf. Sci. 2020, 502, 144206. [Google Scholar] [CrossRef]
  39. Modak, B.; Srinivasu, K.; Ghosh, S.K. Band gap engineering of NaTaO3 using density functional theory: A charge compensated codoping strategy. Phys. Chem. Chem. Phys. 2014, 16, 17116–17124. [Google Scholar] [CrossRef]
  40. Li, F.; Liu, D.; Gao, G.; Xue, B.; Jiang, Y. Improved visible-light photocatalytic activity of NaTaO3 with perovskite-like structure via sulfur anion doping. Appl. Catal. B Environ. 2015, 166, 104–111. [Google Scholar] [CrossRef]
  41. Li, Z.; Ma, B.; Zhang, X.; Sang, Y.; Liu, H. One-pot synthesis of BiOCl nanosheets with dual functional carbon for ultra-highly efficient photocatalytic degradation of RhB. Environ. Res. 2020, 182, 109077. [Google Scholar] [CrossRef]
  42. Alido, J.P.M.; Sari, F.N.I.; Ting, J.-M. Synthesis of Ag/hybridized 1T-2H MoS2/TiO2 heterostructure for enhanced visible-light photocatalytic activity. Ceram. Int. 2019, 45, 23651–23657. [Google Scholar] [CrossRef]
  43. Sari, F.N.I.; Lu, S.-H.; Ting, J.-M. Wide-bandgap HfO2-V2O5 nanowires heterostructure for visible light driven photocatalytic degradation. J. Am. Ceram. Soc. 2020, 103, 2252–2261. [Google Scholar] [CrossRef]
  44. Sari, F.N.I.; Yen, D.T.K.; Ting, J.-M. Enhanced photocatalytic performance of TiO2 through a novel direct dual Z-scheme design. Appl. Surf. Sci. 2020, 533, 147506. [Google Scholar] [CrossRef]
  45. Munusamy, T.D.; Yee, C.S.; Khan, M.M.R. Construction of hybrid g-C3N4/CdO nanocomposite with improved photodegradation activity of RhB dye under visible light irradiation. Adv. Powder Technol. 2020, 31, 2921–2931. [Google Scholar] [CrossRef]
  46. Zhang, Z.; Feng, Y.; Liu, N.; Zhao, Y.; Wang, X.; Yang, S.; Long, Y.; Qiu, L. Preparation of Sn/Mn loaded steel slag zeolite particle electrode and its removal effect on rhodamine B(RhB). J. Water Process Eng. 2020, 37, 101417. [Google Scholar] [CrossRef]
  47. Ding, X.; Gutierrez, L.; Croue, J.-P.; Li, M.; Wang, L.; Wang, Y. Hydroxyl and sulfate radical-based oxidation of RhB dye in UV/H2O2 and UV/persulfate systems: Kinetics, mechanisms, and comparison. Chemosphere 2020, 253, 126655. [Google Scholar] [CrossRef]
  48. Lee, Y.C.; Teng, H.; Hu, C.C.; Hu, S.Y. Temperature-dependent photoluminescence in NaTaO3 with different crystalline structures. Electrochem. Solid State Lett. 2008, 11, P1–P4. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of as-obtained samples: NaTaO3, NTO-S, Sn/NTO, Sn/NTO-S composites, and Sn3O4.
Figure 1. XRD patterns of as-obtained samples: NaTaO3, NTO-S, Sn/NTO, Sn/NTO-S composites, and Sn3O4.
Crystals 11 00059 g001
Figure 2. SEM images of NaTaO3 (a), NTO-S (b), Sn3O4 (c,d), Sn/NTO composites (e), and Sn/NTO-S (f), respectively.
Figure 2. SEM images of NaTaO3 (a), NTO-S (b), Sn3O4 (c,d), Sn/NTO composites (e), and Sn/NTO-S (f), respectively.
Crystals 11 00059 g002
Figure 3. TEM images of NaTaO3 (a), Sn/NTO (b), and a schematic diagram of the growth process of Sn/NTO (c).
Figure 3. TEM images of NaTaO3 (a), Sn/NTO (b), and a schematic diagram of the growth process of Sn/NTO (c).
Crystals 11 00059 g003
Figure 4. UV–vis diffuse reflectance spectra (DRS) (a) and the calculated band gaps (b) of different samples: NaTaO3, NTO-S, Sn/NTO, Sn/NTO-S (molar ratio Sn:Ta = 1:1) heterostructure, and Sn3O4.
Figure 4. UV–vis diffuse reflectance spectra (DRS) (a) and the calculated band gaps (b) of different samples: NaTaO3, NTO-S, Sn/NTO, Sn/NTO-S (molar ratio Sn:Ta = 1:1) heterostructure, and Sn3O4.
Crystals 11 00059 g004
Figure 5. Photocatalytic degradation activities (a) and plots (b) of –Ln (Ct/C0) vs. irradiation time for RhB degradation under ultraviolet light irradiation, measured at regular time intervals in the presence of NaTaO3, NTO-S, Sn/NTO, Sn/NTO-S, and Sn3O4, respectively.
Figure 5. Photocatalytic degradation activities (a) and plots (b) of –Ln (Ct/C0) vs. irradiation time for RhB degradation under ultraviolet light irradiation, measured at regular time intervals in the presence of NaTaO3, NTO-S, Sn/NTO, Sn/NTO-S, and Sn3O4, respectively.
Crystals 11 00059 g005
Scheme 1. Schematic model of a proposed reaction pathway of degraded products.
Scheme 1. Schematic model of a proposed reaction pathway of degraded products.
Crystals 11 00059 sch001
Figure 6. Room temperature photoluminescence (PL) spectra of NaTaO3, NTO-S, Sn/NTO, Sn/NTO-S, and Sn3O4, respectively.
Figure 6. Room temperature photoluminescence (PL) spectra of NaTaO3, NTO-S, Sn/NTO, Sn/NTO-S, and Sn3O4, respectively.
Crystals 11 00059 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chang, S.; Sang, Y.; Liu, H. Efficient Photocatalytic Degradation of RhB by Constructing Sn3O4 Nanoflakes on Sulfur-Doped NaTaO3 Nanocubes. Crystals 2021, 11, 59. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst11010059

AMA Style

Chang S, Sang Y, Liu H. Efficient Photocatalytic Degradation of RhB by Constructing Sn3O4 Nanoflakes on Sulfur-Doped NaTaO3 Nanocubes. Crystals. 2021; 11(1):59. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst11010059

Chicago/Turabian Style

Chang, Sujie, Yuanhua Sang, and Hong Liu. 2021. "Efficient Photocatalytic Degradation of RhB by Constructing Sn3O4 Nanoflakes on Sulfur-Doped NaTaO3 Nanocubes" Crystals 11, no. 1: 59. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst11010059

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