Next Article in Journal
Evaluation of Structural and Functional Properties of La0.6Sr0.4MnO3 Perovskite Prepared by the Fast Solution Combustion Approach
Next Article in Special Issue
Photocatalytically Active Semiconductor Cu3P Unites with Flocculent TiN for Efficient Removal of Sulfamethoxazole
Previous Article in Journal
Regulating the Assembly of Precursors of Carbon Nitrides to Improve Photocatalytic Hydrogen Production
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of Calcination Temperature on Photocatalyst Performances of Floral Bi2O3/TiO2 Composite

1
Academy of Art & Design, Nanchang Institute of Technology, Nanchang 330044, China
2
College of Environment and Chemical Engineering, Nanchang Hangkong University, Nanchang 330063, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work and should be considered as co-first authors.
Catalysts 2022, 12(12), 1635; https://doi.org/10.3390/catal12121635
Submission received: 3 November 2022 / Revised: 7 December 2022 / Accepted: 8 December 2022 / Published: 13 December 2022

Abstract

:
Heterojunction photocatalytic materials show excellent performance in degrading toxic pollutants. This study investigates the influence of calcination temperature on the performances of floral Bi2O3/TiO2 composite photocatalyst crystal, which was prepared with glycerol, bismuth nitrate, and titanium tetrachloride as the major raw materials via the solvothermal method. XRD, SEM/TEM, BET, Uv-vis, and XPS were employed to analyze the crystal structure, morphology, specific surface area, band gap, and surface chemical structure of the calcined temperature catalysts. The calcination temperature influence on the catalytic performance of composite photocatalysis was tested with rhodamine B (RhB) as the degradation object. The results revealed the high catalytic activity and higher photocatalytic performance of the Bi2O3/TiO2 catalyst. The degradation efficiency of the Bi2O3/TiO2 catalyst to RhB was 97%, 100%, and 91% at 400 °C, 450 °C, and 500 °C calcination temperatures, respectively, in which the peak degradation activity appeared at 450 °C. The characterization results show that the appropriate calcination temperature promoted the crystallization of the Bi2O3/TiO2 catalyst, increased its specific surface area and the active sites of catalytic reaction, and improved the separation efficiency of electrons and holes.

Graphical Abstract

1. Introduction

Due to rapid industrial development, pollution has become an ever-growing issue [1]. In particular, water pollution as a major basic environmental factor affects and restricts human survival, safety, and social development worldwide [2]. In recent years, a tremendous number of industrial pollutants, phenolic pesticides, fertilizers, dyes, and organic pollutants have been discharged into waters [3]. Even the low concentration of the pollutants remaining in the wastewater can still affect plants, aquatic animals, and humans due to their non-degradability and toxicity. Therefore, organic wastewater purification is necessary for sustainable development and environmental protection. At present, a variety of techniques have been used for the degradation of organic pollutants in water, such as photocatalysis [4,5,6], adsorption [7], the Fenton method [8], hydrolysis [9], ionizing radiation [10], ultrasonic radiation [11], and oxygen plasma treatment [12]. Among them, photocatalysis can completely eliminate organic pollutants with its outstanding reaction rate, low processing cost, moderate operating conditions, and environment-friendly nature, making it one of the most effective methods in the removal of organic pollution [13,14,15]. Photocatalytic degradation is to mineralize and purify water for organic pollutants with solar energy and is especially suitable for wastewater that contains a small amount of refractory organic matter; thus, it has a promising application prospect.
Due to the nature of the stable physical and chemical structure, high oxidizing capacity, and being low-cost and non-toxic, titanium dioxide (TiO2) has a broad application prospect in photocatalytic volatile organic compounds [16]. Nevertheless, defects such as large width of the forbidden band (3.2 eV), narrow absorption light wavelength range, and a tendency to merge with photoelectrons and holes often lead to the reduction of photocatalytic efficiency [17]. The methods to improve the photocatalytic efficiency of TiO2 mainly include precious metal doping [18], structure regulation [19], and heterojunction construction [20,21]. Among them, heterojunction is widely used to achieve the spatial separation of redox reaction sites and to retain the high redox capability of charge carriers, while extending the spectral response to wider wavelengths. For example, it has been reported that the photocatalytic performance of TiO2/gC3N4 [22], CuO/TiO2 [23], TiO2/CeO2 [24], and TiO2/Bi2WO6 [25] heterojunctions was greatly improved under light.
Bismuth oxide (Bi2O3) is an important metal oxide p-type semiconductor. It has four main crystal types, namely, α, β, γ, and δ, in which the α phase is stable at low temperature, δ is stable at high temperature, and the others are metastable phases [26]. In general, β-Bi2O3 has better photocatalytic activity than α-Bi2O3, as it has lower band gap energy and higher visible light absorption efficiency [27]. Bismuth oxide is easy to synthesize, with the advantages of low quantum efficiency, appropriate band gap (2.58~2.8 eV), and that it can be activated by visible light [28]. However, the rapid recombination of photocarriers may occur in pure Bi2O3 material; therefore, it is often used to construct heterojunctions with other semiconductors in order to effectively improve the photocatalytic performance [29]. Attempts to improve the photocatalytic activity of TiO2 and Bi2O3 have been made in preparing Bi2O3/TiO2 heterojunctions, which not only enhanced the photogenerated charge separation, but also increased the light absorption range [30,31]. Since the phase transitions of titanium dioxide and bismuth oxide are accompanied by temperature changes, the calcination temperature is an important factor in determining the photocatalytic activity in the preparation of the Bi2O3/TiO2 catalyst [32,33,34]. However, few studies have reported the effect of the calcination temperature on the photocatalytic activity of Bi2O3/TiO2 composite photocatalysts.
Therefore, this study investigates the effects of different calcination temperatures on the morphology, crystal shape, and performance of Bi2O3/TiO2 composite photocatalysts. The RhB was adopted as the model wastewater.

2. Experiment

2.1. Reagent

Titanium tetrachloride was purchased from Shanghai Nuotai Chemical Co., Ltd., Shanghai, China. Nitric acid, bismuth nitrate pentahydrate, and sodium bicarbonate were from Sinopharm Chemical Reagent Co., Ltd., Shanghai, China. Anhydrous ethanol was from Aladdin Bio-Chem Technology Co., Ltd., Shanghai, China. Sodium hydroxide and propylene glycol were from Xilong Chemical Co., Ltd., Guangzhou, China. RhB was purchased from Shanghai Yuanye Biotechnology Co., Ltd., Shanghai, China. All reagents were analytically pure without any further processing.

2.2. Floral Bi2O3/TiO2 Preparation

To start with, 30 mL of glycerol, 50 ml bismuth nitrate (Bi/Ti mol of 2.1%, the concentration of bismuth nitrate was 4.16 g/L), and 20 mL of absolute ethanol were poured into a 100 mL beaker, stirred for 30 min, and ultrasonic processed for 1 min. Subsequently, the mixture was poured into a Teflon-sealed reactor. Then, titanium tetrachloride was added in dropwise, and the mixture was reacted into an autoclave at 110 °C for 48 h. Afterwards, the suspension was cooled down to room temperature, centrifuged, and washed with absolute ethanol, then vacuum-dried at 80 °C for 8 h. The product was divided into five portions, calcined at 375 °C, 400 °C, 425 °C, 450 °C, and 500 °C for 4 h, respectively, and then cooled to room temperature. Thus, the Bi2O3/TiO2 samples after heat treatment at different temperatures were obtained.
For comparison, pure TiO2 was also prepared. Firstly, 30 mL glycerol and 20 mL ethanol were added into a 100 mL beaker with continuous stirring for 30 min and ultrasounding for 1 min at room temperature. Then, the mixed solution was added into the Teflon reactor. Next, 1 mL titanium tetrachloride solution was added into the above solution drop by drop in the fume hood. Then, the solution was transferred into the oven and heated at 110 °C for 48 h. After cooling down to room temperature, the materials were washed and centrifuged repeatedly with ethanol and then dried at 80 °C for 8 h. Finally, the samples were obtained after calcinating at 500 °C for 4 h.

2.3. Characterization

X-ray diffraction (XRD; RiGdku, RINT2000, CuK α emission line of λ = 0.15418 nm) was used to detect the crystal structure of the sample, and the sample morphology was identified using a field-emission scanning electron microscope (FESEM) in secondary electron scattering mode of JSM-6700F at 15 kV. High-resolution TEM mapping of the lattice structure and lattice stripes of the prepared samples was obtained via JEM-2010F transmission electron microscope at an accelerating voltage of 200 kV. The optical properties of the samples were tested by UV-visible spectrophotometer (DRS; Spec-3700 DUV Shimadzu). The specific surface area and porosity of the samples were measured using the Brunauer–Emmett–Teller (BET) method (Micromeritics, ASAP2020) at 77 K and N2 adsorption and desorption. The UV-visible diffuse reflection spectrum (UV-vis/DRS) of the samples was measured with a Shimadzu UV-2450 spectrometer. The steady-state photoluminescence spectrogram (PL) was tested by a single Hitachi F4500 fluorescence spectrometer at an excitation wavelength of 320 nm. Finally, the surface composition of the samples and the chemical state of the containing elements were measured through X-ray photoelectron spectroscopy (XPS; ESCALAB MK II).

2.4. Photocatalytic Performance Test

The photocatalytic properties of the Bi2O3/TiO2 composites were tested through the degradation of RhB, an organic dye, under a xenon lamp. In this process, the filter was employed to filter out light with a wavelength of less than 420 nm. The power of the xenon lamp’s (PLS-SXE300, Beijing Park Fei Lay Technology Co., LTD, Beijing, China. light intensity of 100 mW/cm2) was set for 300 W. The distance between RhB and the xenon lamp was 5 cm. The experiment was divided into five groups (five temperature-treated samples); steps of each group were as follows: Firstly, 50 mg prepared catalyst was mixed with 60 mL 20 mg/L RhB and stirred for 30 min, during which any light source was completely cut off in order to create the adsorption equilibrium. Secondly, 3 mL mixture was extracted through a syringe and placed in darkness as the initial sample, then another 3 mL sample was extracted from the mixture after every 20 min, and the residual catalyst separated with a filter. The entire catalytic process was performed under a xenon lamp equipped with an optical filter, with an on-going stirrer throughout the illumination period. Finally, the absorbance intensity of all collected samples was measured at the corresponding wavelength via UV spectrophotometer (the maximum absorption wavelength of RhB is 554 nm). In recycling experiments, the Bi2O3/TiO2 composite was isolated from the suspension by repeated centrifugation and washing. The photodegradation efficiency was calculated by Formula (2), as follows:
degradation   efficiency   % =   C 0 C 1 C 0 100 %
where C0 is the initial concentration of the solution before the reaction, respectively; C1 is the concentration of the solution after a period of reaction.

3. Results and Discussion

3.1. Crystal Structure Analysis

To explore the possible structural changes caused by calcination, the Bi2O3/TiO2 prepared at gradient calcination temperatures and the pure TiO2 calcinated at 500 °C were analyzed by XRD, as is shown in Figure 1. As the calcination temperature increases, the diffraction peak of the Bi2O3/TiO2 catalyst narrows down, and the height of the peak, namely the intensity, increases accordingly. This is caused by the increase in the crystal size as the temperature rises. The XRD diagram of the Bi2O3/TiO2 composite catalyst was 2 θ of 25.3°, 37.8°, 48.2°, and 54.7°, with the corresponding TiO2 lattice planes of (101), (004), (200), and (211), typically characteristic of anatase phase titanium dioxide [35]. Only the diffraction peaks of anatase phase TiO2 were observed in the XRD map of the Bi2O3/TiO2 composite catalyst, while corresponding ones were not detected, which indicates that Bi2O3 did not form a separate crystallization phase. The reason can be either the even dispersion of bismuth ions in the anatase phase TiO2 crystal, or the almost undetectable low Bi2O3 content (2.1 mol%) [36]. As for the pure TiO2 treated at 500 °C, in addition to the diffraction peak containing anatase, a distinct characteristic peak at 2 θ = 27.4° for rutile phase titanium dioxide (JCPDS 75-1749) was captured, indicating that TiO2 had a phase transition into an anatase and rutile mixed phase at 500 °C [37]. However, the characteristic peak of the rutile phase was not observed for the Bi2O3/TiO2 treated at 500 °C, which is possibly due to the Bi2O3 inhibition of the transition from the anatase phase to the anatase/rutile mixed phase. According to the previous studies, it was found that the crystallinity of titania samples was of crucial importance to the photocatalytic efficiency [38,39]. The Scherrer’s equation was employed to calculate the crystallite size (L in nm) [40]. The crystallite size was 17.72 nm.
L = kλ/βcosθ
where k is a constant and it is equivalent to 1.
λ-X-ray wavelength: its value is 0.15418 nm.
θ-half angle diffraction, which is 25.294.
β-full width at maximum half intensity (FWHM) and the value is 0.0089 (Rad).
Figure 1. XRD pattern of Bi2O3/TiO2 calcined at different temperatures and TiO2 at 500 °C.
Figure 1. XRD pattern of Bi2O3/TiO2 calcined at different temperatures and TiO2 at 500 °C.
Catalysts 12 01635 g001

3.2. Micromorphology and Analysis

Figure 2 shows the SEM images of the Bi2O3/TiO2 composite, catalyzing at different calcination temperatures. It can be clearly seen that the morphology of Bi2O3/TiO2 is a petal-like spheroid. The petals are composed of many small particles, and the surface is relatively rough. The temperature increase did not change the texture of the composite. The figures indicate that the diameter of the petal at 425 °C was 54.2 nm, and 88.3 nm at 500 °C. The crystallinity of titanium dioxide increased due to the temperature rise, and thus the size of small petal particles increased. This is consistent with the conclusion of XRD. The sample treated at 425 °C had relatively intact shapes, whereas some petals at 500 °C were fragmented.
Figure 3a,b show the TEM images of the 450 °C-treated Bi2O3/TiO2 composite catalyst taken at two different sites of the same sample, in which anatase phase titanium dioxide revealed a lattice stripe diameter of 0.36 nm, β phase bismuth oxide of 0.327 nm, and α phase bismuth oxide of 0.371 nm. Interestingly, the β phase bismuth oxide might have transformed to the α phase due to the temperature rise; therefore, after calcination for 4 h, both the α phase and β phase bismuth oxide might exist in the product. Figure 3c exhibits the element scanning area, and Figure 3d–f are the element distribution image of Bi, O, and Ti, respectively. All three elements can be observed clearly in the image, which revealed relatively scattered distribution.

3.3. Analysis of Light Absorption Region and Band Gap

The optical absorption capacity and response range of the Bi2O3/TiO2 catalyst with different calcination temperatures were studied through UV-visible diffuse reflection spectroscopy (UV-Vis DRS). The results are shown in Figure 4a. From the figure, all samples exhibited high absorption capacity within the visible light range. Relatively obvious absorption initiation changes could be seen near the wavelength of 410 nm, and the light absorption performance of the 450 °C-treated sample was slightly better than that of the others. The band gap of catalytic material can be calculated by the formula: α h ν = A (h ν-Eg)n/2, where the α, h, ν, and Eg represent the absorption coefficient, the Planck constant, the photon frequency, and the band gap of the semiconductor, respectively [41]. Figure 4b shows the band gap energy of the Bi2O3/TiO2 catalyst. The band gap energy of the 450 °C- and 500 °C-treated samples were 2.7 eV and 2.86 eV, respectively. It is evident that the 450 °C-treated sample has a lower band gap and more enhanced light absorption capacity compared to the 500 °C-treated one.

3.4. Analysis of Surface Elements

The element composition and element valence states of the Bi2O3/TiO2 samples synthesized at 450 °C and 500 °C were analyzed in comparison via high-resolution XPS mapping, the results of which are shown in Figure 5. Figure 5a shows the XPS full-survey spectrum of the Bi2O3/TiO2 photocatalyst. The characteristic peaks of Bi, Ti, and O are evidently presented, which demonstrates the successful synthesis of TiO2/Bi2O3 photocatalyst. The Bi4f spectral map (Figure 5b) was divided into four peaks, in which the 159.09 eV and 164.39 eV were attributed to the Bi4f 7/2 and Bi4 f 5/2 characteristic peaks, indicating the existence of Bi3+; the characteristic peaks of binding energy at 157.41 eV and 162.73 eV belonged to the metal Bi, indicating the simultaneous presence of Bi3+ and elementary Bi [42]. In Figure 5c, O1 s in 529.67 eV corresponded to the Bi-O bonds; the characteristic peaks at 530.40 eV and 531.59 eV were attributed to Ti-O bonds and hydroxyl oxygen produced in the composite possibly due to water in the chemisorption, respectively [43]. In the spectra of Ti2p (Figure 5d), the binding energy of Ti 2p 1/2 and Ti 2p 3/2 spin-orbit splitting photoelectrons were 464.38 eV and 458.53 eV, respectively, reflecting the presence of Ti in the Bi2O3/TiO2 composite as Ti4+, which tetrahedral coordinated with oxygen [44]. Notably, when comparing the element maps of Ti, O, and Bi in Bi2O3/TiO2 at 500 °C and 450 °C, it was found that the element characteristic peaks at 500 °C shifted right to those of Ti2p and Bi4f at 450 °C, while the O1s peak shifted towards lower binding energy. This may be caused by an increased crystallinity at a higher calcination temperature.

3.5. Analysis of the Specific Surface Area

The surface area of the photocatalyst is a crucial factor in terms of creating active sites that associate with the enhancement of photocatalytic degradation efficiency. According to previous studies, photocatalysts with a large surface area provide more surface reaction sites. The porosity and structural properties of catalytic materials were systematically investigated using the N2 adsorption–desorption isotherm. Figure 6 shows the adsorption and desorption curve of the Bi2O3/TiO2 composite catalyst at 400 °C, 450 °C, and 500 °C. Typical type IV isotherms of type H3 hysteresis loop can be observed for all samples at different temperatures, indicating the mesoporous structure of the prepared Bi2O3/TiO2 composite catalysts. The specific surface areas of the Bi2O3/TiO2 composite catalysts of 400 °C, 450 °C, and 500 °C were 35.2 m2/g, 40.1 m2/g, and 25.7 m2/g, as shown in Table 1, respectively, in which is the maximum value was at 450 °C, consistent with the results covered in the document [45]. This may be attributed to the crystalline grain growth of bismuth oxide, alongside phase transition, as calcination temperature increases [26,35]. The TEM image indicates that when it exceeded 450 °C, bismuth oxide underwent a phase transition and size expansion; at 500 °C, anatase phase titanium dioxide also transformed into rutile phase titanium dioxide. As can be seen from the XRD pattern, the grain size of titanium dioxide increased and its effective specific surface area decreased.

3.6. Photoluminescence Spectroscopy Analysis

A photoluminescence (PL) spectroscopy was used to study the carrier separation of the catalyst. Figure 7 is the PL profiles of Bi2O3/TiO2 when calcined at 375 °C, 400 °C, 425 °C, 450 °C, and 500 °C, respectively. All samples had a strong emission peak at 470 nm, resulting from the energy released from the electrons rapidly combining the holes in the valence band after the transition from valence band to conduction band. As we can see, the PL peak intensity of Bi2O3/TiO2 calcined at 450 °C was significantly lower than that of the other samples. To our knowledge, the lower the PL peak intensity, the higher the separation efficiency of the catalyst photogenerated electron–hole pairs to a certain extent, and the more efficient the corresponding photocatalysis [46]. PL peak intensity gradually decreased as the calcination temperature increased up to 450 °C, indicating that temperature rise favors the separation of the electron–hole pairs. When it further increased to 500 °C, peak intensity reached the maximum, indicating that excess temperature accelerates the composite rate of the electron–hole pairs, which is possibly due to the phase transition of TiO2 in the composite induced by high temperature. The results above are consistent with the XRD and the photocatalytic efficiency. In summary, among all prepared samples, the 450 °C calcined one showed the lowest PL peak intensity and had higher electron–hole separation efficiency, with the best photocatalytic performance. Furthermore, the fluctuation of separation efficiency also demonstrates the presence of heterojunctions.

3.7. Photocatalytic Performance Analysis

The photocatalytic activity of the Bi2O3/TiO2 composite was tested through RhB degradation under visible light, which was conducted in two steps: (1) dark reaction for the first 30 min to reach concentration equilibrium where the samples adsorbed the dye; (2) using Xenon lamp to simulate sunlight. Figure 8a is a curve graph of the RhB degradation trends of Bi2O3/TiO2 samples treated at a different temperature. It can be seen that at 60 min, the degradation rate of the 450 °C-treated sample was almost 100%, whereas the 400 °C- and 500 °C-treated samples showed 97% and 91% degradation rates, respectively. Based on the TEM results, in the samples calcined above 450 °C, some bismuth oxide converted into the α phase, the photocatalytic activity of which was very weak under visible light, as previously mentioned. This is likely to be one of the reasons for the reduced photocatalytic efficiency of the 500 °C-treated sample. In addition, according to the BET analysis, sample grains treated at 500 °C had a smaller specific surface area, while treatment below 450 °C led to insufficient crystallinity of the Bi2O3/TiO2 composite, and consequently weakened photocatalytic efficiency.
To present the catalytic performance of each sample more intuitively, it is assumed that all photocatalytic processes follow a quasi-first-order kinetic model. As is shown in Figure 8b, the reaction rate constant of Bi2O3/TiO2 was k(375°C) > k(400°C) > k(425 °C) > k(450°C) > k(500°C). Among them, k(450°C) (0.11316) is more than twice larger than k(500°C)(0.05482), which is possibly due to the transition from anatase phase titanium dioxide to rutile induced by temperature increase, and the consequent catalytic efficiency reduction. The temperature rises also led to the transition from the β phase to α phase bismuth oxide, and thus photocatalysis was reduced.
Figure 8c shows the cycle stability experiment of Bi2O3/TiO2. The sample used in this cycle experiment was Bi2O3/TiO2 (450 °C), and the RhB was used for photodegradation. The experimental procedure was the same as the photocatalytic experiment, with a sampling time of every 20 min, thus, 60 min in total. As is shown in Figure 8c, in all five cycles, the degradation efficiency of RhB was almost unchanged, which remained at around 98%. This shows that the flower-shaped Bi2O3/TiO2 composite samples prepared at different temperatures in this experiment had proper photocatalytic cycle stability and great potential in processing pollutants contained in sewage.
When Bi2O3 was involved, proven by various methods of characterization as well as experimental phenomena, the separation efficiency of photogenic electron and holes of TiO2 increased, indicating the formation of heterojunctions in composite. Furthermore, the possible mechanism is displayed in Figure 8d. First, RhB molecules were adsorbed on the surface and cavity of Bi2O3/TiO2 composites. Then, Bi2O3/TiO2 composites were irradiated by visible light. Subsequently, the photogenerated electrons in the VB of Bi2O3 transferred to the CB of Bi2O3. The holes on the VB of Bi2O3 were generated due to the electrons moved. Meanwhile, the holes on the VB of Bi2O3 were moved to the VB of TiO2, and the photogenerated holes were effectively gathered in the VB of TiO2. Finally, the holes of TiO2 were active species, and oxidized RhBs were oxidized to other products [47,48].

4. Conclusions

Briefly, a solvothermal route was employed to synthesize the Bi2O3/TiO2 photocatalysts, and the influence of calcination temperature was studied systematically. The results of this study implied that the calcination temperature had an effect on the performances of the Bi2O3/TiO2 photocatalysts. The XRD results implied that 450 °C was a suitable temperature. The photocatalytic performances of catalysts were assessed through degradation RhB. When the calcination temperature was 450 °C, the Bi2O3/TiO2 photocatalyst had the best degradation efficiency, which was about 100% degradation of RhB at 60 min.

Author Contributions

Experiments, M.W. and C.L.; Analysed, B.L. and W.Q.; Writing, M.W., C.L. and B.L.; Revised, W.Q.; Funding, Y.X. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (No. 22066017 and 21667019), the Science and Technology Department of Jiangxi Province (20181ACG70025), and the Key Project of Science and Technology Research of Jiangxi Provincial Department of Education (No. GJJ191044). This work was also supported by the Shanghai Summit Discipline in Design (SJGFXK-2019-001), Jiangxi Province “double thousand plan” project (2020009410085, jxsq2019201007) and Jiangxi Province’s Scientific Planning Project (No. YG2021057).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tang, Q.-Y.; Yang, M.-J.; Yang, S.-Y.; Xu, Y.-H. Enhanced photocatalytic degradation of glyphosate over 2D CoS/BiOBr heterojunctions under visible light irradiation. J. Hazard. Mater. 2020, 407, 124798. [Google Scholar] [CrossRef] [PubMed]
  2. Luo, J.; Dai, Z.; Feng, M.; Gu, M.; Xie, Y. Graphitic carbon nitride/ferroferric oxide/reduced graphene oxide nanocomposite as highly active visible light photocatalyst. Nano Res. 2022, 1–6. [Google Scholar] [CrossRef]
  3. Perumal, K.; Shanava, S.; Karthigeyan, A.; Ahamad, T.; Alshehri, S.M.; Murugakoothan, P. Hydrothermal assisted precipitation synthesis of highly stable g-C3N4/BiOBr/CdS photocatalyst with enhanced visible light photocatalytic degradation of tetracycline. Diam. Relat. Mater. 2020, 110, 108091. [Google Scholar] [CrossRef]
  4. Tong, H.; Ji, Y.; He, T.; He, R.; Chen, M.; Zeng, J.; Wu, D. Preparation and photocatalytic performance of UIO-66/La-MOF composite. Water Sci. Technol. 2022, 86, 95–109. [Google Scholar] [CrossRef] [PubMed]
  5. Sujinnapram, S.; Wongrerkdee, S. Synergistic effects of structural, crystalline, and chemical defects on the photocatalytic performance of Y-doped ZnO for carbaryl degradation. J. Environ. Sci. 2022, 124, 667–677. [Google Scholar] [CrossRef] [PubMed]
  6. Song, X.; Su, R.; Wang, Y.; Zhang, Y.; Gao, B.; Wang, Y.; Ma, D.; Li, Q. Visible light-driven chlorite activation process for enhanced sulfamethoxazole antibiotics degradation, antimicrobial resistance reduction and biotoxicity elimination. Chem. Eng. J. 2022. [Google Scholar] [CrossRef]
  7. Dong, J.; Yang, F.; Wang, Q.; Jiang, C.; Wang, H.; Cao, X.; Li, Z.; Wang, Z. Preparation of porous carbon@TiO2 composites for the adsorption/sonocatalytic degradation of organic dyes. J. Mol. Liq. 2022, 367, 120469. [Google Scholar] [CrossRef]
  8. Shi, W.; Sun, W.; Liu, Y.; Zhang, K.; Sun, H.; Lin, X.; Hong, Y.; Guo, F. A self-sufficient photo-Fenton system with coupling in-situ production H2O2 of ultrathin porous g-C3N4 nanosheets and amorphous FeOOH quantum dots. J. Hazard. Mater. 2022, 436, 129141. [Google Scholar] [CrossRef]
  9. Huang, M.; Wang, X.; Zhu, C.; Zhu, F.; Liu, P.; Wang, D.; Fang, G.; Chen, N.; Gao, S.; Zhou, D. Efficient chlorinated alkanes degradation in soil by combining alkali hydrolysis with thermally activated persulfate. J. Hazard. Mater. 2022, 438, 129571. [Google Scholar] [CrossRef]
  10. Fang, L.; Chu, L.; Wang, J.; Yang, Q. Treatment of polyacrylamide-containing wastewater by ionizing radiation: Efficient reduction of viscosity and degradation of polyacrylamide. Radiat. Phys. Chem. 2023, 202, 110547. [Google Scholar] [CrossRef]
  11. Wei, B.; Qi, H.; Zou, J.; Li, H.; Wang, J.; Xu, B.; Ma, H. Degradation mechanism of amylopectin under ultrasonic irradiation. Food Hydrocoll. 2020, 111, 106371. [Google Scholar] [CrossRef]
  12. Li, S.; Wang, X.; Li, L.; Liu, J.; Ding, Y.; Zhao, T.; Zhang, Y. Atomic-scale simulations of the deoxynivalenol degradation induced by reactive oxygen plasma species. Food Res. Int. 2022, 162, 111939. [Google Scholar] [CrossRef] [PubMed]
  13. Fu, W.; Zhao, Y.; Wang, H.; Chen, X.; Liu, K.; Zhang, K.; Wei, Q.; Wang, B. Study on preparation, photocatalytic performance and degradation mechanism of polymeric carbon nitride/Pt/nano-spherical MoS2 composite. J. Phys. Chem. Solids 2022, 166, 110700. [Google Scholar] [CrossRef]
  14. Kumar, G.; Dutta, R.K. Sunlight-induced enhanced photocatalytic reduction of chromium (VI) and photocatalytic degradation of methylene blue dye and ciprofloxacin antibiotic by Sn3O4/SnS2 nanocomposite. Environ. Sci. Pollut. Res. 2022, 29, 57758–57772. [Google Scholar] [CrossRef]
  15. Xiang, W.; Yuan, J.; Wu, Y.; Luo, H.; Xiao, C.; Zhong, N.; Zhao, M.; Zhong, D.; He, Y. Working principle and application of photocatalytic optical fibers for the degradation and conversion of gaseous pollutants. Chin. Chem. Lett. 2022, 33, 3632–3640. [Google Scholar] [CrossRef]
  16. Gu, S.; Liu, X.; Wang, H.; Liu, Z.; Xing, H.; Yu, L. Preparation and characterization of TiO2 photocatalytic composites supported by blast furnace slag fibres for wastewater degradation. Ceram. Int. 2022. [Google Scholar] [CrossRef]
  17. Hu, L.; Chen, J.; Wei, Y.; Wang, M.; Xu, Y.; Wang, C.; Gao, P.; Liu, Y.; Liu, C.; Song, Y.; et al. Photocatalytic degradation effect and mechanism of Karenia mikimotoi by non-noble metal modified TiO2 loading onto copper metal organic framework (SNP-TiO2@Cu-MOF) under visible light. J. Hazard. Mater. 2023, 442, 130059. [Google Scholar] [CrossRef]
  18. Khairy, M.; Kamar, E.M.; Mousa, M.A. Photocatalytic activity of nano-sized Ag and Au metal-doped TiO2 embedded in rGO under visible light irradiation. Mater. Sci. Eng. B 2022, 286, 116023. [Google Scholar] [CrossRef]
  19. Liu, B.; Zhang, B.; Ji, J.; Li, K.; Cao, J.; Feng, Q.; Huang, H. Effective regulation of surface bridging hydroxyls on TiO2 for superior photocatalytic activity via ozone treatment. Appl. Catal. B Environ. 2022, 304, 120952. [Google Scholar] [CrossRef]
  20. Gordanshekan, A.; Arabian, S.; Nazar, A.R.S.; Farhadian, M.; Tangestaninejad, S. A comprehensive comparison of green Bi2WO6/g-C3N4 and Bi2WO6/TiO2 S-scheme heterojunctions for photocatalytic adsorption/degradation of Cefixime: Artificial neural network, degradation pathway, and toxicity estimation. Chem. Eng. J. 2023, 451, 139067. [Google Scholar] [CrossRef]
  21. Feizpoor, S.; Habibi-Yangjeh, A.; Luque, R. Design of TiO2/Ag3BiO3 n-n heterojunction for enhanced degradation of tetracycline hydrochloride under visible-light irradiation. Environ. Res. 2022, 215, 114315. [Google Scholar] [CrossRef] [PubMed]
  22. Kane, A.; Chafiq, L.; Dalhatou, S.; Bonnet, P.; Nasr, M.; Gaillard, N.; Dikdim, J.M.D.; Monier, G.; Assadi, A.A.; Zeghioud, H. g-C3N4/TiO2 S-scheme heterojunction photocatalyst with enhanced photocatalytic Carbamazepine degradation and mineralization. J. Photochem. Photobiol. A Chem. 2022, 430, 113971. [Google Scholar] [CrossRef]
  23. Liang, C.; Li, C.; Zhu, Y.; Du, X.; Zeng, Y.; Zhou, Y.; Zhao, J.; Li, S.; Liu, X.; Yu, Q.; et al. Light-driven photothermal catalysis for degradation of toluene on CuO/TiO2 Composite: Dominating photocatalysis and auxiliary thermalcatalysis. Appl. Surf. Sci. 2022, 601, 154144. [Google Scholar] [CrossRef]
  24. Hou, J.; Yang, H.; He, B.; Ma, J.; Lu, Y.; Wang, Q. High photocatalytic performance of hydrogen evolution and dye degradation enabled by CeO2 modified TiO2 nanotube arrays. Fuel 2022, 310, 122364. [Google Scholar] [CrossRef]
  25. Li, L.; Yang, J.; Yang, L.; Fu, F.; Xu, H.; Fan, X. Photocatalytic performance of TiO2/Bi2WO6 photocatalysts with trace Fe3+ dopant for gaseous toluene decomposition. J. Environ. Chem. Eng. 2022, 10, 107708. [Google Scholar] [CrossRef]
  26. Xiao, X.; Hu, R.; Liu, C.; Xing, C.; Qian, C.; Zuo, X.; Nan, J.; Wang, L. Facile large-scale synthesis of  -Bi2O3 nanospheres as a highly efficient photocatalyst for the degradation of acetaminophen under visible light irradiation. Appl. Catal. B Environ. 2013, 140–141, 433–443. [Google Scholar] [CrossRef]
  27. Cheng, H.; Huang, B.; Lu, J.; Wang, Z.; Xu, B.; Qin, X.; Zhang, X.; Dai, Y. Synergistic effect of crystal and electronic structures on the visible-light-driven photocatalytic performances of Bi2O3 polymorphs. Phys. Chem. Chem. Phys. 2010, 12, 15468–15475. [Google Scholar] [CrossRef]
  28. Gao, J.; Rao, S.; Yu, X.; Wang, L.; Xu, J.; Yang, J.; Liu, Q. Dimensional-matched two dimensional/two dimensional TiO2/Bi2O3 step-scheme heterojunction for boosted photocatalytic performance of sterilization and water splitting. J. Colloid Interface Sci. 2022, 628, 166–178. [Google Scholar] [CrossRef]
  29. Ke, T.; Shen, S.; Yang, K.; Lin, D. In situ fabrication of Bi2O3/C3N4/TiO2@C photocatalysts for visible-light photodegradation of sulfamethoxazole in water. Appl. Surf. Sci. 2022, 580, 152302. [Google Scholar] [CrossRef]
  30. Chen, J.; Tang, T.; Feng, W.; Liu, X.; Yin, Z.; Zhang, X.; Chen, J.; Cao, S. Large-Scale Synthesis of p−n Heterojunction Bi2O3/TiO2 Nanostructures as Photocatalysts for Removal of Antibiotics under Visible Light. ACS Appl. Nano Mater. 2022, 5, 1296–1307. [Google Scholar] [CrossRef]
  31. Ren, C.; Qiu, W.; Zhang, H.; He, Z.; Chen, Y. Degradation of benzene on TiO2/SiO2/Bi2O3 photocatalysts under UV and visible light. J. Mol. Catal. A Chem. 2015, 398, 215–222. [Google Scholar] [CrossRef]
  32. Lal, M.; Sharma, P.; Ram, C. Calcination temperature effect on titanium oxide (TiO2) nanoparticles synthesis. Opt. Int. J. Light Electron Opt. 2021, 241, 166934. [Google Scholar] [CrossRef]
  33. Li, L.; Tao, R.; Liu, Y.; Zhou, K.; Fan, X.; Han, Y.; Tang, L. Co3O4 nanoparticles/Bi2O3 nanosheets: One step synthesis, high-efficiency thermal catalytic performance, and catalytic mechanism research. Mol. Catal. 2022, 528, 112483. [Google Scholar] [CrossRef]
  34. Bao, Y.; Guo, R.; Gao, M.; Kang, Q.; Ma, J. Morphology control of 3D hierarchical urchin-like hollow SiO2@TiO2 spheres for photocatalytic degradation: Influence of calcination temperature. J. Alloy. Compd. 2021, 853, 157202. [Google Scholar] [CrossRef]
  35. Huang, Q.; Wang, Q.; Tao, T.; Zhao, Y.; Wang, P.; Ding, Z.; Chen, M. Controlled synthesis of Bi2O3/TiO2 catalysts with mixed alcohols for the photocatalytic oxidation of HCHO. Environ. Technol. 2019, 40, 1937–1947. [Google Scholar] [CrossRef] [PubMed]
  36. Shamaila, S.; Sajjad, A.K.L.; Chen, F.; Zhang, J. Study on highly visible light active Bi2O3 loaded ordered mesoporous titania. Appl. Catal. B Environ. 2010, 94, 272–280. [Google Scholar] [CrossRef]
  37. Shen, J.; Wang, H.; Song, Y.; Zhou, Y.; Ye, N.; Fang, L.; Wang, L. Amorphous carbon coated TiO2 nanocrystals embedded in a carbonaceous matrix derived from polyvinylpyrrolidone decomposition for improved Li-storage performance. Chem. Eng. J. 2014, 240, 379–386. [Google Scholar] [CrossRef]
  38. Kőrösi, L.; Bognar, B.; Horvath, M.; Schneider, G.; Kovacs, J.; Scarpellini, A.; Castelli, A.; Colombo, M.; Prato, M. Hydrothermal evolution of PF-co-doped TiO2 nanoparticles and their antibacterial activity against carbapenem-resistant Klebsiella pneumoniae. Appl. Catal. B Environ. 2018, 231, 115–122. [Google Scholar] [CrossRef]
  39. Korösi, L.; Pertics, B.; Schneider, G.; Bognár, B.; Kovács, J.; Meynen, V.; Scarpellini, A.; Pasquale, L.; Prato, M. Photocatalytic inactivation of plant pathogenic bacteria using TiO2 nanoparticles prepared hydrothermally. Nanomaterials 2020, 10, 1730. [Google Scholar] [CrossRef]
  40. Pelaez, M.; Nolan, N.T.; Pillai, S.C.; Seery, M.K.; Falaras, P.; Kontos, A.G.; Dunlop, P.S.M.; Hamilton, J.W.J.; Byrne, J.A.; O'Shea, K.; et al. A review on the visible light active titanium dioxide photocatalysts for environmental applications. Appl. Catal. B Environ. 2012, 125, 331–349. [Google Scholar] [CrossRef] [Green Version]
  41. Hu, R.; Xiao, X.; Tu, S.; Zuo, X.; Nan, J. Synthesis of flower-like heterostructured  -Bi2O3/Bi2O2CO3microspheres using Bi2O2CO3self-sacrifice precursor and itsvisible-light-induced photocatalytic degradation of o-phenylphenol. Appl. Catal. B Environ. 2015, 163, 510–519. [Google Scholar] [CrossRef]
  42. Xu, D.; Hai, Y.; Zhang, X.; Zhang, S.; He, R. Bi2O3 cocatalyst improving photocatalytic hydrogen evolution performance of TiO2. Appl. Surf. Sci. 2017, 400, 530–536. [Google Scholar] [CrossRef]
  43. Alhaddad, M.; Ismail, A.A.; Alghamdi, Y.G.; Al-Khathami, N.D.; Mohamed, R.M. Co3O4 Nanoparticles Accommodated Mesoporous TiO2 framework as an Excellent Photocatalyst with Enhanced Photocatalytic Properties. Opt. Mater. 2022, 131, 112643. [Google Scholar] [CrossRef]
  44. Hu, J.; Zhao, R.; Li, H.; Xu, Z.; Dai, H.; Gao, H.; Yu, H.; Wang, Z.; Wang, Y.; Liu, Y.; et al. Boosting visible light photocatalysis in an Au@TiO2 yolk-in-shell nanohybrid. Appl. Catal. B Environ. 2022, 303, 120869. [Google Scholar] [CrossRef]
  45. Duan, X.; Yang, J.; Hu, G.; Yang, C.; Chen, Y.; Liu, Q.; Ren, S.; Li, J. Optimization of TiO2/ ZSM-5 photocatalysts: Energy band engineering by solid state diffusion method with calcination. J. Environ. Chem. Eng. 2021, 9, 105563. [Google Scholar] [CrossRef]
  46. Majhi, D.; Das, K.; Mishra, A.; Dhiman, R.; Mishra, B.G. One pot synthesis of CdS/BiOBr/Bi2O2CO3: A novel ternary double Zscheme heterostructure photocatalyst for efficient degradation of atrazine. Appl. Catal. B Environ. 2020, 260, 118222. [Google Scholar] [CrossRef]
  47. Kőrösi, L.; Bognár, B.; Bouderias, S.; Castelli, A.; Scarpellini, A.; Pasquale, L.; Prato, M. Highly-efficient photocatalytic generation of superoxide radicals by phase-pure rutile TiO2 nanoparticles for azo dye removal. Appl. Surf. Sci. 2019, 493, 719–728. [Google Scholar] [CrossRef]
  48. Bina, B.; Fatehizadeh, A.; Taheri, E.; Heydari, M.; Darvishmotevalli, M.; Bazmeh, A. Atenolol removal from aqueous solutions using Bi2O3/TiO2 under UV-C and visible light irradiations. Int. J. Environ. Anal. Chem. 2022, 1–22. [Google Scholar] [CrossRef]
Figure 2. SEM images of a Bi2O3/TiO2 composite catalyst treated at different calcination temperatures: (a) 375 °C, (b) 400 °C, (c) 425 °C, (d) 500 °C.
Figure 2. SEM images of a Bi2O3/TiO2 composite catalyst treated at different calcination temperatures: (a) 375 °C, (b) 400 °C, (c) 425 °C, (d) 500 °C.
Catalysts 12 01635 g002
Figure 3. (a,b) are 450 °C Bi2O3/TiO2HRTEM images, (c) scan area, and distribution of (d) Bi, (e) O, and (f) Ti.
Figure 3. (a,b) are 450 °C Bi2O3/TiO2HRTEM images, (c) scan area, and distribution of (d) Bi, (e) O, and (f) Ti.
Catalysts 12 01635 g003
Figure 4. The optical diagram of (a) the UV-vis spectrum and the (b) the band gap energy spectrum of the Bi2O3/TiO2 samples with different calcination temperatures.
Figure 4. The optical diagram of (a) the UV-vis spectrum and the (b) the band gap energy spectrum of the Bi2O3/TiO2 samples with different calcination temperatures.
Catalysts 12 01635 g004
Figure 5. The 500 °C- versus 450 °C-treated Bi2O3/TiO2 (a) XPS full-spectrum map, high-resolution spectrum map of (b) Bi4f, (c) O1s, (d) Ti2p.
Figure 5. The 500 °C- versus 450 °C-treated Bi2O3/TiO2 (a) XPS full-spectrum map, high-resolution spectrum map of (b) Bi4f, (c) O1s, (d) Ti2p.
Catalysts 12 01635 g005
Figure 6. N2 adsorption–desorption isotherm of Bi2O3/TiO2 calcined at 400 °C, 450 °C, and 500 °C.
Figure 6. N2 adsorption–desorption isotherm of Bi2O3/TiO2 calcined at 400 °C, 450 °C, and 500 °C.
Catalysts 12 01635 g006
Figure 7. PL result of Bi2O3/TiO2 samples at different calcination temperatures.
Figure 7. PL result of Bi2O3/TiO2 samples at different calcination temperatures.
Catalysts 12 01635 g007
Figure 8. (a) RhB degradation trends of Bi2O3/TiO2 composite catalyst at different calcination temperatures, (b) reaction rate constant of Bi2O3/TiO2 composite catalyst for RhB at different calcination temperatures, (c) cycle stability experiment of Bi2O3/TiO2 composite catalyst treated at 450 °C, (d) mechanism diagram of Bi2O3/TiO2 heterojunction photocatalysis.
Figure 8. (a) RhB degradation trends of Bi2O3/TiO2 composite catalyst at different calcination temperatures, (b) reaction rate constant of Bi2O3/TiO2 composite catalyst for RhB at different calcination temperatures, (c) cycle stability experiment of Bi2O3/TiO2 composite catalyst treated at 450 °C, (d) mechanism diagram of Bi2O3/TiO2 heterojunction photocatalysis.
Catalysts 12 01635 g008
Table 1. Specific surface area, mean pore size, and pore volume of TiO2, Bi2O3, and Bi2O3/TiO2 composites.
Table 1. Specific surface area, mean pore size, and pore volume of TiO2, Bi2O3, and Bi2O3/TiO2 composites.
SamplesSpecific Surface Area (m²/g)Mean Pore Size (nm)Pore Volume (cm3 g−1)
TiO29.110.50.1
Bi2O33.26.20.01
Bi2O3/TiO2 400 °C35.211.90.15
Bi2O3/TiO2 450°C 40.112.40.21
Bi2O3/TiO2 500 °C25.711.10.13
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, M.; Li, C.; Liu, B.; Qin, W.; Xie, Y. Influence of Calcination Temperature on Photocatalyst Performances of Floral Bi2O3/TiO2 Composite. Catalysts 2022, 12, 1635. https://doi.org/10.3390/catal12121635

AMA Style

Wang M, Li C, Liu B, Qin W, Xie Y. Influence of Calcination Temperature on Photocatalyst Performances of Floral Bi2O3/TiO2 Composite. Catalysts. 2022; 12(12):1635. https://doi.org/10.3390/catal12121635

Chicago/Turabian Style

Wang, Mingjun, Che Li, Bingfang Liu, Wenzhen Qin, and Yu Xie. 2022. "Influence of Calcination Temperature on Photocatalyst Performances of Floral Bi2O3/TiO2 Composite" Catalysts 12, no. 12: 1635. https://doi.org/10.3390/catal12121635

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