Next Article in Journal
Enhanced Electrochemical Performance of Metallic CoS-Based Supercapacitor by Cathodic Exfoliation
Previous Article in Journal
DFT Study of Adsorption Behavior of Nitro Species on Carbon-Doped Boron Nitride Nanoribbons for Toxic Gas Sensing
Previous Article in Special Issue
Natural Cotton Cellulose-Supported TiO2 Quantum Dots for the Highly Efficient Photocatalytic Degradation of Dyes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Photocatalytic Activity of CaTiO3 Derived from the Microwave-Melting Heating Process of Blast Furnace Slag

1
School of Resource and Environmental Engineering, Wuhan University of Science and Technology, Wuhan 430081, China
2
Hubei Key Laboratory for Efficient Utilization and Agglomeration of Metallurgic Mineral Resources, Wuhan 430081, China
3
School of Materials and Metallurgy, Wuhan University of Science and Technology, Wuhan 430081, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(8), 1412; https://doi.org/10.3390/nano13081412
Submission received: 26 March 2023 / Revised: 14 April 2023 / Accepted: 17 April 2023 / Published: 19 April 2023
(This article belongs to the Special Issue Photocatalytic Nanomaterials in Water Decontamination)

Abstract

:
The extraction of titanium-bearing components in the form of CaTiO3 is an efficient utilization of blast furnace slag. The photocatalytic performance of this obtained CaTiO3 (MM-CaTiO3) as a catalyst for methylene blue (MB) degradation was evaluated in this study. The analyses indicated that the MM-CaTiO3 had a completed structure with a special length–diameter ratio. Furthermore, the oxygen vacancy was easier to generate on a MM-CaTiO3(110) plane during the photocatalytic process, contributing to improving photocatalytic activity. Compared with traditional catalysts, MM-CaTiO3 has a narrower optical band gap and visible-light responsive performance. The degradation experiments further confirmed that the photocatalytic degradation efficiency of pollutants by using MM-CaTiO3 was 3.2 times that of pristine CaTiO3 in optimized conditions. Combined with molecular simulation, the degradation mechanism clarified that acridine of MB molecular was stepwise destroyed by using MM-CaTiO3 in short times, which is different from demethylation and methylenedioxy ring degradation by using TiO2. This study provided a promising routine for using solid waste to obtain catalysts with excellent photocatalytic activity and was found to be in keeping with sustainable environmental development.

1. Introduction

Titanium-bearing blast furnace slag is a kind of alkaline solid waste formed in the pyrometallurgy process from titanomagnetite and has almost 100 million tons of total quality in the world [1]. The comprehensive utilization of blast furnace slag has become one key sustainability issue confronting the titanium industry in recent years [2]. Generally, the utilization of blast furnace slag is always based on the extraction titanium, which can be divided into acid leaching [3,4], molten salt decomposition [5,6] and carbonization chlorination [7,8,9]. It is further found that the Ca-bearing components are associated with Ti-bearing minerals in blast furnace slag, leading to the difficulty in separation extraction [10]. As perovskite is easily formed in the pyrometallurgical process, it is worth considering the extraction of titanium-bearing components in the form of CaTiO3 from the slag.
It has been reported that CaTiO3 is a good industrial catalyst alternative for TiO2 in photocatalytic degradation pollutants in the environmental remediation field [11,12]. As is well-known, photocatalytic degradation is an effective method to eliminate organic pollutants [13] and exhibits considerable advantages over other physiochemical techniques such as chemical precipitation, adsorption and coagulation. In detail, it can directly degrade pollutants into small molecules represented by water and carbon dioxide to prevent secondary pollution. In addition, photocatalytic degradation relies on solar energy to avoid the environmental problems caused by fossil fuels. Actually, the photocatalytic efficiency depends on the catalytic activity of the photocatalyst [14]. Various chemical methods have been developed to synthesize CaTiO3 photocatalysts, including the solid-state reaction [15,16], sol–gel process [17,18], hydrothermal process [19,20,21], etc. Furthermore, precious metal doping [22,23] and organic framework assembly [24,25] are also applied to optimize its catalytic activity. The solid-state synthesis method is based on the solid diffusion reaction at high temperatures. Due to the fast growth rate of an intermediate product, the synthesis of CaTiO3 contains impurities with a low specific surface. Solution-phase synthesis methods commonly suffer due to their complex process and are difficult to control, which limits the extensive development. Therefore, the extraction of titanium in the form of CaTiO3 from secondary resources to obtain a photocatalyst is an effective solution.
The CaTiO3-enriched products have been successfully collected from titanium-bearing blast furnace slag. The obtained catalyst was shown to have a narrower optical band gap than pristine TiO2 [26], Ag-doped TiO2 NPs [27], and CuO/TiO2 [28], which is conducive to higher catalytic activity. MB is a kind of typical pollutant to explore photocatalysis performance due to its recognized carcinogenicity, wide application, and stability under ultraviolet light. In terms of the optimized degradation conditions, the catalytic experiments proved that its performance of degrading pollutants is 2.2 times that of the pristine TiO2. Furthermore, the correlations between the crystal structure and catalytic ability are also investigated. In detail, the CaTiO3 particles gather and grow to form a complete crystal structure with a special length–diameter ratio induced by the microwave heating effect. In this study, the photocatalytic activity of the obtained CaTiO3 from the microwave-melting heating process of a blast furnace was systematically explored through the degradation of MB. It is found that the special crystal structure of the CaTiO3 product has a lower banding gap and electron–hole pair recombination rate. Furthermore, the obtained CaTiO3 products show the catalytic response under the visible light band based on characterization ability and experimental performance. In addition, combined with molecular simulation and theoretical calculation, the photocatalytic activity of the obtained CaTiO3 was compared with that of pristine CaTiO3, and the corresponding mechanism for their difference was further discussed and explicated. The obtained catalyst was also found to achieve faster degradation of MB than the pristine TiO2, so the degradation mechanism was compared by in situ IR combined with calculated Gibbs free energy. In conclusion, the catalyst with excellent photocatalytic activity was derived from titanium-bearing blast furnace slag in this study, which provided a new idea for solid-waste resource recycling and environmental governance.

2. Materials and Methods

2.1. Material Preparation

The blast furnace slag of vanadium titanium magnetite used in the experiment came from Panzhihua City in Sichuan Province, China. And contents of TiO2 and SiO2 were 20.75 wt.% and 25.36 wt.%, respectively. Based on our study, microwave-melting heating followed by gravity separation has been carried out to extract CaTiO3 product from blast furnace slag (melting temperature at 1300 °C for pretreatment and microwave heating temperature at 1200 °C for 20 min). The CaTiO3 component was enriched and precipitated at the bottom of the heating process. The obtained CaTiO3 products were collected and dried in an oven at 105 °C for 4 h, which was named MM-CaTiO3.

2.2. Experimental Design of Photocatalysis

The photocatalytic reaction was carried out in a circulating reactor equipped with quartz immersion wells, and the photocatalytic reaction device is shown in Figure 1. Experiments were performed using an adjustable monochrome light source system (PL-KT300D, UV region (<390 nm) output power: 2.6 W, Beijing Precise Technology, Beijing, China) to ensure that the photocatalytic process was performed at a known fixed wavelength. The quartz immersion well was experimented with in a camera obscura (Beijing Precise Technology, Beijing, China) to rule out visible light interference. The peristaltic pump (100 rpm, Longer Precision Pump Co. Ltd., Baoding, China) was used to ensure that the photocatalyst powder was in suspension during the experiment. After the photocatalytic degradation experiment, the aqueous mixture was collected and centrifuged to remove the solid catalyst. The MB degradation efficiency (ε) and degradation quantity (Q) were calculated by Equations (1) and (2) [29]. Specifically, the volume of solution used in the experiment is 10 mL; The illumination wavelength was 200 nm to 550 nm, the initial concentration of MB was 2 mg/L to 10 mg/L, and the MM-CaTiO3 dosage was 1 g/L to 10 g/L. MB calibration curve formulas such as Equation (3) show that its correlation coefficient (R2) is 0.9995 and ranges from 2 mg/L to 10 mg/L,
ε(MB, %) = (C0 − Ct)/C0 × 100%
Q(MB, μg) = (Q0 − Qt)
y = 0.0902x − 0.0094
where C0 and Ct are initial and residue MB concentrations, respectively; t is time; Q0 and Qt are initial and residue MB quantity; y and x represent the absorbance and concentration of MB, respectively.

2.3. Instrumental Characterization

X-ray diffraction (XRD, Shimadzu, XD5A, Kyoto, Japan) was used to determine the chemical composition of the catalysts. The morphology and microstructure of catalysts were characterized by scanning electron microscopy (SEM, Fei Company, Nova 400, Hillsboro, OR, USA) and combined with energy dispersive spectroscopy (EDS, Oxford corporation, Penta FET X-3, Abington, UK) analysis to determine the chemical composition. The transmission electron microscope (TEM, JEOL, JEM-2100 UHR STEM/EDS, Kyoto, Japan) was used to determine the orientation of the crystal planes and interplanar spacing of catalysts. The light absorption characteristics and the forbidden band width of catalysts were studied by ultraviolet-visible diffuse reflectance spectroscopy (UV-vis DRS, Shimadzu, UV-2600, Kyoto, Japan). Photoluminescence (PL, Edinburgh Instruments, FLS1000, Edinburgh, UK) spectroscopy with an excitation wavelength of 335 nm was used to study the photoluminescence properties of different catalysts. Photoluminescence quantum yield (PLQY, Edinburgh Instruments, FLS980, Edinburgh, UK) was used to study the utilization efficiency of catalysts for photons. The MB concentration was monitored by measuring the absorbance at 664 nm using the ultraviolet-visible spectrometer (UV-Vis, Shanghai Yidian, Shanghai, China, N4). In situ infrared spectroscopy (In situ IR, Thermo Scientific, Nicolet iS50, Waltham, MA, USA) uses a light wavelength of 300 nm to observe dynamic functional groups in MB solutions. An ion chromatograph (IC, Thermofisher, ICS-6000, Waltham, MA, USA) was used to measure the final product of photocatalytic degradation of methylene blue. GaussView 6.0 was used to calculate the Gibbs free energy of elementary reaction on the degradation process.

2.4. First-Principles Theoretical Analysis

The photocatalytic process depends on the reaction in which the catalyst absorbs photons to generate high-energy electrons and holes, then reduction and oxidation reactions are initiated [30]. To reveal the effect of the crystal structure of CaTiO3 on the photocatalytic activity, the formation of oxygen vacancy (Eov) on different lattice planes was determined using DFT calculation as implemented in the Vienna ab initio Simulation Package (VASP). The Eov values were calculated by Equation (4) [31,32],
Eov = Eov-1 − Epure + 1/2 EO2
where Eov-1 and Epure represent the gross energy with and without the loss of one oxygen atom, respectively; EO2 represents the energy of a single oxygen molecule in the gas phase.

3. Results and Discussion

3.1. Characterization

Explicating the main phase and morphology structure of MM-CaTiO3 is helpful in explaining the difference in their distinguished photocatalytic performance with synthesized CaTiO3. The phase composition of blast furnace slag and MM-CaTiO3 are shown in Figure 2. The results show that the main Ti-bearing components in the slag are directional transfer to CaTiO3 on a microwave-melting heating process. Compared with the predominant orientation of CaTiO3 (121) in slag, the MM-CaTiO3 has a different plane (112) and a (110) crystal plane. The detailed lattice parameters of CaTiO3 are shown in Table 1. The relevant literature reported that the CaTiO3 (121) in the space group of Pnma could be transformed into CaTiO3 (112) in the space group of Pbnm by rotating 90°, which has the same crystal parameters [33,34]. Hence, CaTiO3 (110) is the newly-generated dominant crystal plane in MM-CaTiO3. The SEM-EDS analysis of MM-CaTiO3 showed that the main component of MM-CaTiO3 was CaTiO3, as shown in Figure 3a. The micromorphology further indicated that the CaTiO3 particle gathers and grows with dense crystal branches and a special length–diameter ratio. In addition, a clear boundary was formed between CaTiO3 particles and diopside particles. The results of SEM-EDS analysis with larger magnification (Figure 3b) show that the size of CaTiO3 grains is about 5 μm.
According to the XRD analyses of MM-CaTiO3, the interplanar spacing(s) of various CaTiO3 crystal orientations were calculated based on the Braggs equation [35,36], as shown in Table 1. Meanwhile, the TEM analyses were used to further determine the newly-generated crystal orientation of MM-CaTiO3, as shown in Figure 4. The regional spectrum analyses showed that the main component of the catalyst was CaTiO3. The typical oblique downward (110) crystal plane and the almost vertical (112) crystal plane of CaTiO3 were founded in the HRTEM image, as shown in Figure 4b,c. The interplanar spacing of (110) and (112) crystal planes were 0.266 nm and 0.272 nm, respectively, which agreed well with the XRD analyses.
The photocatalytic property depends on the ability to generate photogenerated electrons and the recombination rate of electron–hole pair, which can be characterized by UV-vis DRS spectrum and PL spectroscopy, as shown in Figure 5 and Figure 6. The UV-vis DRS spectrum analyses show that the threshold wavelength of MM-CaTiO3 reaches 551 nm (Figure 5a), indicating that MM-CaTiO3 has a wider response frequency band than that of pristine CaTiO3 (354 nm) [37] and TiO2 (400 nm) [38]. In addition, the threshold wavelength of 551 nm is within the wavelength range of visible light (400 to 760 nm), which indicates that MM-CaTiO3 can be used as a visible light-responsive photocatalyst. The band gap (Eg) is further calculated by the Tauc plot method [39], as shown in Figure 5b. The results show that the Eg of MM-CaTiO3 is 2.25 eV, which is significantly lower than that of pristine CaTiO3 (3.5 eV) [20] and TiO2 (3.1 eV) [26]. The relationship between the lattice structure and the band gap is further investigated based on molecular simulation and theoretical calculation. According to the preponderant orientation of MM-CaTiO3 and pristine CaTiO3, the oxygen vacancy formation energy (Eov) on different lattice planes was determined using DFT calculation, as shown in Table 2. The theoretical calculation analyses indicate that the oxygen vacancy is easier to generate on the CaTiO3 (110) plane.
Besides the band gap, the recombination of the photogenerated electron–hole pair is also an essential issue impacting the photocatalytic ability of the catalyst. The analyses of photogenerated electron–hole pairs recombination by PL spectroscopy are shown in Figure 6. The MM-CaTiO3 and pristine CaTiO3 have characteristic peaks at the same wavelength (614 nm), while the characteristic peak wavelength of pristine TiO2 at 830 nm. As confirmed in Figure 6a, the MM-CaTiO3 has a lower recombination rate of photogenerated electron–hole pairs than both pristine CaTiO3 and TiO2, which would contribute to its higher photocatalytic activity. Here, the possible mechanism for the recombination rate of the MM-CaTiO3 is discussed. As MM-CaTiO3 still contains small amounts of diopside, the photogenerated electrons are transferred at the interface of different components that suppress the recombination rate of photogenerated electrons and holes [40].
On the other hand, the SiO2 tetrahedron structure of the diopside may promote the separation of photogenerated electron–hole pairs [41,42]. The utilization efficiency of photons is further characterized by quantum yield, as shown in Figure 6b. The results show that the quantum yield of MM-CaTiO3 is 0.11%, which is much higher than that of pristine CaTiO3, indicating a higher photon utilization efficiency, although quantum yields of TiO2 reach 0.2%, which is higher than both of MM-CaTiO3 and pristine CaTiO3. MM-CaTiO3 still shows a good catalytic performance considering its lower band gap and recombination rate of photogenerated electron–hole pairs.

3.2. Photocatalytic Performance of MM-CaTiO3

The experiments on the photocatalytic degradation of MB by using MM-CaTiO3 are systemically investigated, as shown in Figure 7, Figure 8 and Figure 9. The effect of illumination wavelength on the photocatalysis of MM-CaTiO3 was investigated on the conditions of initial MB concentration of 6 mg/L and MM-CaTiO3 dosage of 6 g/L for illumination time of 1000 s in the light range of 200 nm to 550 nm. Figure 7 shows that the degradation efficiency of MB reached the peak value at 300 nm, which is related to the fixed energy required for photogenerated electrons in MM-CaTiO3 to achieve the energy level transition. In addition, it also generated an inflection point at 500 nm in the visible light range [43]. Combined with the previous analysis of photocatalytic property, the newly generated CaTiO3 (110) lattice plane was helpful for the response in the visible light range to achieve partial MB degradation, agreed well with the threshold wavelength analysis of MM-CaTiO3 (λ = 551 nm) in Figure 5a.
The effect of MB initial concentration was further investigated with a MM-CaTiO3 dosage of 6 g/L for an illumination time of 1000 s illumination wavelength of 300 nm. Within the MB initial concentration range of 2 mg/L to 10 mg/L, the degradation efficiency of MB decreased with the increase in the MB initial concentration. The initial excess concentration of MB reduced the transparency of the solution and hindered the UV light absorption of MM-CaTiO3, whereas the degradation amount of MB obviously increased as the increasing of MB initial concentration. Based on this, the initial MB concentration of 4 mg/L was appropriate.
The effect of the MM-CaTiO3 dosage on photocatalysis was investigated at the illumination wavelength of 300 nm with an initial MB concentration of 4 mg/L for an illumination time of 1000 s. The degradation efficiency of MB gradually increased and then decreased as the MM-CaTiO3 dosage increased. The degradation efficiency reached a peak value of 67% with the dosage of 8 g/L. As further increasing the photocatalyst dosage, the degradation efficiency obviously decreased. In addition, the degradation amount of MB showed a similar trend with degradation efficiency. As an excessive dosage of MM-CaTiO3 would affect the UV light absorption, the MM-CaTiO3 dosage of 8 g/L was suitable for MB photocatalytic degradation.
The results of the kinetic study on the photocatalytic degradation of the MB solution are shown in Figure 10 and Table 3. As shown in Figure 10a, the MB degradation efficiency was about 20% when the photocatalytic time was 0. This phenomenon indicated that seldom MB molecular was absorbed on the surface of MM-CaTiO3. When the initial MB concentration was 4 mg/L and MM-CaTiO3 dosage was 8 g/L, the photocatalytic degradation efficiency of MB increased significantly as the extension of the illumination time. When the photocatalytic time exceeded 3 min, the photocatalytic degradation efficiency of MB increased slowly and then became almost stable at 10 min. In order to clarify the degradation behavior of MB, the typical kinetic models were applied to test the fitness of experimental data. The corresponding equations and parameters are shown in Figure 10. The first-order reaction kinetics of ln(Ct/C0) vs. t resulted in a straight line with higher R2 values, which indicated the better applicability of other models. According to the first-order reaction ln(Ct/C0) = −0.2358 t − 0.2364, the rate constant of the photocatalytic degradation reaction was −0.2358 in this study.
According to the above analyses of the photocatalytic process, the MB degradation efficiency was further compared under the optimized condition by using different photocatalysts, as shown in Figure 11. The photocatalytic degradation efficiency of MB in the presence of MM-CaTiO3 was much higher, which was 3.2 times than that of pristine CaTiO3 at an illumination wavelength of 300 nm with an initial MB concentration of 4 mg/L and catalyst dosage of 8 g/L for illumination for 1000 s. The degradation efficiency of MB by pristine TiO2 is less than 30%. Based on the crystal structure and photocatalytic property analyses, the MM-CaTiO3 with different preponderant lattice orientation contributed to the improved photocatalytic degradation efficiency. In addition, the lower recombination rate of photogenerated electron–hole pairs delays the annihilation of photogenerated electrons and further promotes the degradation of pollutants.
Due to the inconsistency of photocatalytic experimental conditions, especially wavelength and light source power, researchers cannot measure the photocatalytic ability of catalysts by comparing the degradation efficiencies of pollutants. At present, the commonly used measurement method is to compare the band gap of different catalysts. The synthesis process, including raw materials and method of MM-CaTiO3, are compared with the previous literature, as shown in Table 4. The results show that the photocatalyst obtained in this study was derived from secondary resources with a narrow band gap by using a simple synthesis process.

3.3. Degradation Mechanism of MB Using MM-CaTiO3

Furthermore, the mechanism of the photocatalytic degradation of MB using MM-CaTiO3 was explored. The in situ IR was used to observe the dynamic behaviors of the functional group on the degradation process, as shown in Figure 12. The MB solution exhibited bands for the C-H asymmetric stretching of CH3 at 2921 cm−1, the C-H symmetric stretching of CH3 at 2850 cm−1, the C=N central ring stretching at 1645 cm−1, the C=C side ring stretching at 1537 cm−1, the acridines stretching at 1461 cm−1, the N-C stretching at 1041, 868 and 746 cm−1 and the single ring stretching 539 and 466 cm−1. The acridines stretching at 1461 cm−1 disappeared immediately in the presence of MM-CaTiO3 under UV light illumination. As the extension of illumination time, the intensity of C-H asymmetric stretching of CH3 at 2921 cm−1, the C-H symmetric stretching of CH3 at 2850 cm−1 and the C=C side ring stretching at 1537 cm−1 gradually weakened and then disappeared. However, the N-C stretching at 1041, 868, and 746 cm−1 and the single ring stretching at 539 and 466 cm−1 were significantly enhanced [46]. This phenomenon indicated that the acridine was directly destroyed by using MM-CaTiO3 in a short amount of time. Then, the destroyed acridines further broke the C=C band and demethylated, which then generated small molecules.
The intermediates product and their Gibbs free energies of MB elementary degradation are presented in Table S1 and Figure 13. Obviously, the two hydroxyl radicals preferentially combine with carbon atoms located symmetrically, which has a lower Gibbs free energy reaction than that of a single hydroxyl radical (−75.8 eV). In addition, more hydroxyl radicals react in the symmetry of the benzene ring to promote the further destruction of acridine structure and chromophores cleavage. Furthermore, the enhancement of single-ring stretching foreshadowed the opening ring reactions. Combined with the in situ IR analysis, the demethylation and opening ring reaction almost react at the same time. Finally, the functional group further decomposed to form inorganic ions, H2O and CO2 [47].
The relevant literature reported that the first step of the demethylation of the MB molecular used TiO2. Furthermore, the breaking of the central aromatic ring relied on the transfer of photogenerated electrons and H+, thus slowing down the conversion of MB to CO2, H2O, NH4+, and SO42− [48,49]. In contrast, high-activity species were produced in the presence of MM-CaTiO3 under UV light illumination, which directly reacted with acridines to achieve stepwise destruction and improve degradation efficiency.

4. Conclusions

The titanium-bearing components in the form of CaTiO3 derived from blast furnace slag by microwave-melting had excellent photocatalytic properties. This study showed that the MM-CaTiO3 obtained a complete crystal structure with a special length–diameter ratio and various predominant crystal orientations. Based on the DFT theoretical calculation, the new-generated CaTiO3 (110) plane was easier to form the oxygen vacancy and contributed to the lower band gap than that of pristine CaTiO3. The UV-vis DRS spectrum showed that the MM-CaTiO3 had a photocatalytic response under visible light, which was further confirmed by the light wavelength experiment of MB degradation. The photocatalytic experiments clarified that the degradation of MB reached 67% at the wavelength of 300 nm with an initial MB concentration of 4 mg/L and dosage of 8 g/L for an illumination time of 1000 s. On the optimized condition, the photocatalytic degradation efficiency by using MM-CaTiO3 was 3.2 times that of pristine CaTiO3 and 2.2 times that of pristine TiO2. The degradation mechanism also indicated that the hydroxyl radical was directly reacted with acridines to achieve stepwise destruction by using MM-CaTiO3 in short time, which was different from the gradual decomposition process of demethylation and the MB ring structure breaking by using TiO2. Generally, the degradation efficiency of MB using MM-CaTiO3 could be further increased by using light with more power. This study provided a promising routine for using solid waste to obtain catalysts with excellent photocatalytic activity and was thereby favorable to sustainable environmental development.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/nano13081412/s1, Table S1: The Gibbs free energy of MB molecular and intermediate products.

Author Contributions

J.X.: conceptualization, methodology, investigation, writing—original draft. Q.Y.: methodology, formal analysis, writing—review and editing. G.Q.: conceptualization, writing—review & editing, supervision, funding acquisition. J.Z.: validation and formal analysis. Y.L.: validation and formal analysis. 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 [grant number 52204282] and “The 14th Five Year Plan,” Hubei Provincial Advantaged Characteristic Disciplines (Groups) Project of Wuhan University of Science and Technology [grant number 2023A0401].

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

We are grateful to the High-Performance Computing Center of Central South University for assistance with the computations.

Conflicts of Interest

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

References

  1. Fan, G.; Wang, M.; Dang, J.; Zhang, R.; Lv, Z.; He, W.; Lv, X. A novel recycling approach for efficient extraction of titanium from high-titanium-bearing blast furnace slag. Waste Manag. 2021, 120, 626–634. [Google Scholar] [CrossRef] [PubMed]
  2. Zheng, F.; Guo, Y.; Qiu, G.; Chen, F.; Wang, S.; Sui, Y.; Jiang, T.; Yang, L. A novel process for preparation of titanium dioxide from Ti-bearing electric furnace slag: NH4HF2-HF leaching and hydrolyzing process. J. Hazard. Mater. 2018, 344, 490–498. [Google Scholar] [CrossRef] [PubMed]
  3. Bian, Z.; Feng, Y.; Li, H.; Wu, H. Efficient separation of vanadium, titanium, and iron from vanadium-bearing titanomagnetite by pressurized pyrolysis of ammonium chloride-acid leaching-solvent extraction process. Sep. Purif. Technol. 2021, 255, 117169. [Google Scholar] [CrossRef]
  4. Zhu, X.; Zheng, S.; Zhang, Y.; Fang, Z.Z.; Zhang, M.; Sun, P.; Li, Q.; Zhang, Y.; Li, P.; Jin, W. Potentially more ecofriendly chemical pathway for production of high-purity TiO2 from titanium slag. ACS Sustain. Chem. Eng. 2019, 7, 4821–4830. [Google Scholar] [CrossRef]
  5. Li, Z.; Lei, Y.; Ma, W.; Zhang, Y.; Wang, S.; Ren, Y.; Lv, G. An approach to prepare high-purity TiSi2 for clean utilization of Ti-bearing blast furnace slag. Green Chem. 2022, 24, 3344–3357. [Google Scholar] [CrossRef]
  6. Tian, M.; Liu, Y.; Wang, L.; Chen, D.; Zhao, H.; Meng, F.; Zhen, Y.; Qi, T. Role of aluminum salt on thermal hydrolysis of titanyl sulfuric–chloric mixture acid solution. J. Mater. Res. Technol. 2021, 14, 2486–2496. [Google Scholar] [CrossRef]
  7. Li, L.; Zhu, F.; Deng, P.; Zhang, D.; Jia, Y.; Li, K.; Kong, L.; Liu, D. Behavior of magnesium impurity during carbochlorination of magnesium-bearing titanium slag in chloride media. J. Mater. Res. Technol. 2021, 13, 204–215. [Google Scholar] [CrossRef]
  8. Deng, P.; Liu, D.; Chen, X.; Jiang, W.; Kong, L.; Li, L. Carbochlorination mechanism of low-grade titanium slag: Ab initio molecular dynamic simulation. J. Mater. Res. Technol. 2022, 17, 459–466. [Google Scholar] [CrossRef]
  9. Zhang, Y.; Fang, Z.Z.; Xia, Y.; Huang, Z.; Lefler, H.; Zhang, T.; Sun, P.; Free, M.L.; Guo, J. A novel chemical pathway for energy efficient production of Ti metal from upgraded titanium slag. Chem. Eng. J. 2016, 286, 517–527. [Google Scholar] [CrossRef]
  10. Zhang, L.; Zhang, L.N.; Wang, M.Y.; Li, G.Q.; Sui, Z.T. Recovery of titanium compounds from molten Ti-bearing blast furnace slag under the dynamic oxidation condition. Miner. Eng. 2007, 20, 684–693. [Google Scholar] [CrossRef]
  11. Passi, M.; Pal, B. A review on CaTiO3 photocatalyst: Activity enhancement methods and photocatalytic applications. Powder Technol. 2021, 388, 274–304. [Google Scholar] [CrossRef]
  12. Wei, K.; Faraj, Y.; Yao, G.; Xie, R.; Lai, B. Strategies for improving perovskite photocatalysts reactivity for organic pollutants degradation: A review on recent progress. Chem. Eng. J. 2021, 414, 128783. [Google Scholar] [CrossRef]
  13. Yadav, A.A.; Hunge, Y.M.; Kang, S.-W.; Fujishima, A.; Terashima, C. Enhanced Photocatalytic Degradation Activity Using the V2O5/RGO Composite. Nanomaterials 2023, 13, 338. [Google Scholar] [CrossRef]
  14. Wang, W.; Wang, D.; Song, H.; Hao, D.; Xu, B.; Ren, J.; Wang, M.; Dai, C.; Wang, Y.; Liu, W. Size effect of gold nanoparticles in bimetallic ZIF catalysts for enhanced photo-redox reactions. Chem. Eng. J. 2023, 455, 140909. [Google Scholar] [CrossRef]
  15. Singh, P.; Yadav, R.S.; Rai, S.B. Enhanced photoluminescence in a Eu3+ doped CaTiO3 perovskite phosphor via incorporation of alkali ions for white LEDs. J. Phys. Chem. Solids 2021, 151, 109916. [Google Scholar] [CrossRef]
  16. Miranda-Hernández, J.G.; Vargas-Hernández, J.; Casarrubias-Vargas, H.; González-Morán, C.O.; Refugio-García, E.; Flores Cuautle, J.d.J.A. Synthesis and effect of CaTiO3 formation in CaO·Al2O3 by solid-state reaction from CaCO3·Al2O3 and Ti. Mater. Chem. Phys. 2019, 232, 57–64. [Google Scholar] [CrossRef]
  17. Shawky, A.; Alhaddad, M.; Al-Namshah, K.S.; Mohamed, R.M.; Awwad, N.S. Synthesis of Pt-decorated CaTiO3 nanocrystals for efficient photoconversion of nitrobenzene to aniline under visible light. J. Mol. Liq. 2020, 304, 112704. [Google Scholar] [CrossRef]
  18. Cesconeto, F.R.; Borlaf, M.; Nieto, M.I.; de Oliveira, A.P.N.; Moreno, R. Synthesis of CaTiO3 and CaTiO3/TiO2 nanoparticulate compounds through Ca2+/TiO2 colloidal sols: Structural and photocatalytic characterization. Ceram. Int. 2018, 44, 301–309. [Google Scholar] [CrossRef]
  19. Han, J.; Liu, Y.; Dai, F.; Zhao, R.; Wang, L. Fabrication of CdSe/CaTiO3 nanocomposties in aqueous solution for improved photocatalytic hydrogen production. Appl. Surf. Sci. 2018, 459, 520–526. [Google Scholar] [CrossRef]
  20. Jiang, E.; Yang, L.; Song, N.; Zhang, X.; Liu, C.; Dong, H. Multi-shelled hollow cube CaTiO3 decorated with Bi12O17Cl2 towards enhancing photocatalytic performance under the visible light. J. Colloid Interface Sci. 2020, 576, 21–33. [Google Scholar] [CrossRef]
  21. Feng, L.; Liu, J.; Huynen, I.; Mahmoud, M.Z.; Akhtar, M.N. Electromagnetic performance, optical and physiochemical features of CaTiO3/NiO and SrFe12O19/NiO nanocomposites based bilayer absorber. J. Colloid Interface Sci. 2022, 610, 879–892. [Google Scholar] [CrossRef] [PubMed]
  22. Ji, Y.; Ding, B.; Ni, W.; Li, X.; He, X.; Chen, Z.; Ran, S.; Lü, H. Tailoring the crystal structure of CaTiO3 by multielement doping for photo-assisted activation of NO. Chem. Eng. J. 2022, 450, 138255. [Google Scholar] [CrossRef]
  23. Song, P.; Wang, J.; Wang, X.; Li, N.; Zhang, Z.; Liu, Y. The active pairs of Co-Co2C adjusted by La-doped CaTiO3 with perovskite phase for higher alcohol synthesis from syngas. Chem. Eng. J. 2022, 439, 135635. [Google Scholar] [CrossRef]
  24. Chamanehpour, E.; Hossein Sayadi, M.; Hajiani, M. Metal-organic framework coordinated with g-C3N4 and metal ions for boosting photocatalytic H2 production under sunlight. J. Photochem. Photobiol. A Chem. 2023, 434, 114221. [Google Scholar] [CrossRef]
  25. Zhang, Z.; Yang, X.; Li, P.; Wang, Y.; Zhao, X.; Safaei, J.; Tian, H.; Zhou, D.; Li, B.; Kang, F.; et al. Biomimetic Dendrite-Free Multivalent Metal Batteries. Adv. Mater. 2022, 34, 2206970. [Google Scholar] [CrossRef]
  26. Sun, R.; Wang, Z.; Saito, M.; Shibata, N.; Ikuhara, Y. Atomistic mechanisms of nonstoichiometry-induced twin boundary structural transformation in titanium dioxide. Nat. Commun. 2015, 6, 7120. [Google Scholar] [CrossRef]
  27. Abbad, S.; Guergouri, K.; Gazaout, S.; Djebabra, S.; Zertal, A.; Barille, R.; Zaabat, M. Effect of silver doping on the photocatalytic activity of TiO2 nanopowders synthesized by the sol-gel route. J. Environ. Chem. Eng. 2020, 8, 103718. [Google Scholar] [CrossRef]
  28. Nekooie, R.; Shamspur, T.; Mostafavi, A. Novel CuO/TiO2/PANI nanocomposite: Preparation and photocatalytic investigation for chlorpyrifos degradation in water under visible light irradiation. J. Photochem. Photobiol. A Chem. 2021, 407, 113038. [Google Scholar] [CrossRef]
  29. Vanthana Sree, G.; Nagaraaj, P.; Kalanidhi, K.; Aswathy, C.A.; Rajasekaran, P. Calcium oxide a sustainable photocatalyst derived from eggshell for efficient photo-degradation of organic pollutants. J. Clean. Prod. 2020, 270, 122294. [Google Scholar] [CrossRef]
  30. Gao, C.; Wang, J.; Xu, H.; Xiong, Y. Coordination chemistry in the design of heterogeneous photocatalysts. Chem. Soc. Rev. 2017, 46, 2799–2823. [Google Scholar] [CrossRef]
  31. Xie, C.; Zhu, B.; Sun, Y.; Li, F.; Song, W. Understanding the roles of copper dopant and oxygen vacancy in promoting nitrogen oxides removal over iron-based catalyst surface: A collaborative experimental and first-principles study. J. Colloid Interface Sci. 2022, 612, 584–597. [Google Scholar] [CrossRef]
  32. Wang, S.; He, T.; Chen, P.; Du, A.; Ostrikov, K.; Huang, W.; Wang, L. In Situ Formation of Oxygen Vacancies Achieving Near-Complete Charge Separation in Planar BiVO4 Photoanodes. Adv. Mater. 2020, 32, 2001385. [Google Scholar] [CrossRef]
  33. Chen, L.; Xu, B.; Yang, Y.; Bellaiche, L. Macroscopic and Microscopic Structures of Cesium Lead Iodide Perovskite from Atomistic Simulations. Adv. Funct. Mater. 2020, 30, 1909496. [Google Scholar] [CrossRef]
  34. Kahmann, S.; Nazarenko, O.; Shao, S.; Hordiichuk, O.; Kepenekian, M.; Even, J.; Kovalenko, M.V.; Blake, G.R.; Loi, M.A. Negative Thermal Quenching in FASnI3 Perovskite Single Crystals and Thin Films. ACS Energy Lett. 2020, 5, 2512–2519. [Google Scholar] [CrossRef]
  35. Lu, M.; Zhang, Y.; Chen, J.; Han, W.; Zhang, W.; Li, H.; Zhang, X.; Zhang, B. K+ alkalization promoted Ca2+ intercalation in V2CTx MXene for enhanced Li storage. J. Energy Chem. 2020, 49, 358–364. [Google Scholar] [CrossRef]
  36. Zhang, Z.; Chen, Z.; Wang, Y.; Chi, J.; Wang, Y.; Zhao, Y. Bioinspired Bilayer Structural Color Hydrogel Actuator with Multienvironment Responsiveness and Survivability. Small Methods 2019, 3, 1900519. [Google Scholar] [CrossRef]
  37. Wang, R.; Ni, S.; Liu, G.; Xu, X. Hollow CaTiO3 cubes modified by La/Cr co-doping for efficient photocatalytic hydrogen production. Appl. Catal. B Environ. 2018, 225, 139–147. [Google Scholar] [CrossRef]
  38. Hu, S.; Shaner, M.R.; Beardslee, J.A.; Lichterman, M.; Brunschwig, B.S.; Lewis, N.S. Amorphous TiO2 coatings stabilize Si, GaAs, and GaP photoanodes for efficient water oxidation. Science 2014, 344, 1005–1009. [Google Scholar] [CrossRef]
  39. Yang, G.; Ren, Z.; Liu, K.; Qin, M.; Deng, W.; Zhang, H.; Wang, H.; Liang, J.; Ye, F.; Liang, Q.; et al. Stable and low-photovoltage-loss perovskite solar cells by multifunctional passivation. Nat. Photonics 2021, 15, 681–689. [Google Scholar] [CrossRef]
  40. Ahmad, N.; Anae, J.; Khan, M.Z.; Sabir, S.; Yang, X.J.; Thakur, V.K.; Campo, P.; Coulon, F. Visible light-conducting polymer nanocomposites as efficient photocatalysts for the treatment of organic pollutants in wastewater. J. Environ. Manag. 2021, 295, 113362. [Google Scholar] [CrossRef]
  41. Wu, Y.; Lu, J.; Li, M.; Yuan, J.; Wu, P.; Chang, X.; Liu, C.; Wang, X. Bismuth silicate photocatalysts with enhanced light harvesting efficiency by photonic crystal. J. Alloy. Compd. 2019, 810, 151839. [Google Scholar] [CrossRef]
  42. Guan, X.; Zhang, X.; Zhang, C.; Li, R.; Wang, Y.; Fan, C. One-step synthesis of novel Bi/Bi2SiO5 flower-like composites with highly-efficient CO2 photoreduction performance. Compos. Commun. 2020, 20, 100366. [Google Scholar] [CrossRef]
  43. Wang, H.; Zhao, R.; Hu, H.; Fan, X.; Zhang, D.; Wang, D. 0D/2D Heterojunctions of Ti3C2 MXene QDs/SiC as an Efficient and Robust Photocatalyst for Boosting the Visible Photocatalytic NO Pollutant Removal Ability. ACS Appl. Mater. Interfaces 2020, 12, 40176–40185. [Google Scholar] [CrossRef] [PubMed]
  44. Yang, H.; Han, C.; Xue, X. Photocatalytic activity of Fe-doped CaTiO3 under UV–visible light. J. Environ. Sci. 2014, 26, 1489–1495. [Google Scholar] [CrossRef] [PubMed]
  45. Fang, F.; Liu, Y.; Sun, X.; Fu, C.; Prakash Bhoi, Y.; Xiong, W.; Huang, W. TiO2 Facet-dependent reconstruction and photocatalysis of CuOx/TiO2 photocatalysts in CO2 photoreduction. Appl. Surf. Sci. 2021, 564, 150407. [Google Scholar] [CrossRef]
  46. Spectroscopic Tools. Available online: https://www.science-and-fun.de/tools/ (accessed on 14 April 2023).
  47. Yu, Z.; Chuang, S.S.C. Probing Methylene Blue Photocatalytic Degradation by Adsorbed Ethanol with In Situ IR. J. Phys. Chem. C 2007, 111, 13813–13820. [Google Scholar] [CrossRef]
  48. He, J.; Du, Y.-E.; Bai, Y.; An, J.; Cai, X.; Chen, Y.; Wang, P.; Yang, X.; Feng, Q. Facile Formation of Anatase/Rutile TiO2 Nanocomposites with Enhanced Photocatalytic Activity. Molecules 2019, 24, 2996. [Google Scholar] [CrossRef]
  49. Xiang, H.; Ren, G.; Zhong, Y.; Xu, D.; Zhang, Z.; Wang, X.; Yang, X. Fe3O4@C Nanoparticles Synthesized by In Situ Solid-Phase Method for Removal of Methylene Blue. Nanomaterials 2021, 11, 330. [Google Scholar] [CrossRef]
Figure 1. Flow chart of the photocatalytic reaction device.
Figure 1. Flow chart of the photocatalytic reaction device.
Nanomaterials 13 01412 g001
Figure 2. The main phase composition of blast furnace slag and MM-CaTiO3.
Figure 2. The main phase composition of blast furnace slag and MM-CaTiO3.
Nanomaterials 13 01412 g002
Figure 3. SEM-EDS analysis of MM-CaTiO3. (a) Mag. 300×; (b) mag. 2400×.
Figure 3. SEM-EDS analysis of MM-CaTiO3. (a) Mag. 300×; (b) mag. 2400×.
Nanomaterials 13 01412 g003
Figure 4. (a) TEM image of MM-CaTiO3; (b,c) HRTEM image of CaTiO3.
Figure 4. (a) TEM image of MM-CaTiO3; (b,c) HRTEM image of CaTiO3.
Nanomaterials 13 01412 g004
Figure 5. The UV-vis DRS spectrum of MM-CaTiO3. (a) The threshold wavelength analysis of MM-CaTiO3, and (b) the forbidden band gap analysis of MM-CaTiO3.
Figure 5. The UV-vis DRS spectrum of MM-CaTiO3. (a) The threshold wavelength analysis of MM-CaTiO3, and (b) the forbidden band gap analysis of MM-CaTiO3.
Nanomaterials 13 01412 g005
Figure 6. (a) The photoluminescence (PL) spectroscopy and (b) photoluminescence quantum yield (PLQY) analyses of catalysts.
Figure 6. (a) The photoluminescence (PL) spectroscopy and (b) photoluminescence quantum yield (PLQY) analyses of catalysts.
Nanomaterials 13 01412 g006
Figure 7. The effect of illumination wavelength on the photocatalysis of MM-CaTiO3.
Figure 7. The effect of illumination wavelength on the photocatalysis of MM-CaTiO3.
Nanomaterials 13 01412 g007
Figure 8. The effect of MB initial concentration on the photocatalysis degradation of MM-CaTiO3.
Figure 8. The effect of MB initial concentration on the photocatalysis degradation of MM-CaTiO3.
Nanomaterials 13 01412 g008
Figure 9. The effect of photocatalyst dosage on the photocatalysis degradation.
Figure 9. The effect of photocatalyst dosage on the photocatalysis degradation.
Nanomaterials 13 01412 g009
Figure 10. (a) The effect of photocatalytic time on the photocatalysis of MM-CaTiO3. (b) Zero-order reaction kinetics, (c) first-order reaction kinetics, and (d) second-order reaction kinetics.
Figure 10. (a) The effect of photocatalytic time on the photocatalysis of MM-CaTiO3. (b) Zero-order reaction kinetics, (c) first-order reaction kinetics, and (d) second-order reaction kinetics.
Nanomaterials 13 01412 g010
Figure 11. The photocatalytic degradation performance of catalysts.
Figure 11. The photocatalytic degradation performance of catalysts.
Nanomaterials 13 01412 g011
Figure 12. (a) In situ IR absorption spectra of MB solution at different time periods; (b) flow diagram of photocatalytic degradation of MB by MM-CaTiO3.
Figure 12. (a) In situ IR absorption spectra of MB solution at different time periods; (b) flow diagram of photocatalytic degradation of MB by MM-CaTiO3.
Nanomaterials 13 01412 g012
Figure 13. Theoretical calculation and Gibbs free energy of MB-degradation by MM-CaTiO3.
Figure 13. Theoretical calculation and Gibbs free energy of MB-degradation by MM-CaTiO3.
Nanomaterials 13 01412 g013
Table 1. The lattice parameters of CaTiO3 in various materials.
Table 1. The lattice parameters of CaTiO3 in various materials.
MaterialsTitanium-Bearing Blast Furnace SlagMM-CaTiO3Pristine CaTiO3
Space group|numberPnma|62Pbnm|62Pm-3m|221Pbnm|62
a(Å)5.44245.37963.79505.3780
b(Å)7.64175.44233.79505.4440
c(Å)5.38077.64013.79507.6370
α/β/γ(°)90.000090.000090.000090.0000
Z4.004.001.004.00
(h k l)(121)(112)(110)(112)
d(Å)2.70252.70302.68352.7030
Table 2. The oxygen vacancy formation energy of CaTiO3 (112) and (110) lattice plane.
Table 2. The oxygen vacancy formation energy of CaTiO3 (112) and (110) lattice plane.
Lattice Plane(112)(110)
Molecular modelNanomaterials 13 01412 i001Nanomaterials 13 01412 i002
Epure−609.95 eV−595.48 eV
Eov-1−601.76 eV−587.59 eV
Oxygen vacancy formation energy8.152 eV7.852 eV
Table 3. Kinetic reaction model and correlation of photocatalytic degradation.
Table 3. Kinetic reaction model and correlation of photocatalytic degradation.
ModelEquationR2
Zero-order reaction kineticsCt = −0.5220 t + 3.07700.98422
First-order reaction kineticsln(Ct/C0) = −0.2358 t − 0.23640.99902
Second-order reaction kinetics1/Ct = 0.0998 t + 0.31010.98969
Table 4. Comparison of the band gap values of photocatalysts in different synthesis processes.
Table 4. Comparison of the band gap values of photocatalysts in different synthesis processes.
Raw MaterialSynthesis MethodBand GapRef.
Calcium carbonate and titanium dioxide.Calcium carbonate and titanium dioxide are firstly mixed and then calcined at 1400 °C for 2 h.3.50 eV[44]
Titanium isoprenoid, calcium acetate and nitric acid.The Ca2+/TiO2 sol is first synthesized using calcium acetate, nitric acid and titanium isoprenoid, and then aging and calcined after freeze-drying.3.44 eV[18]
Ti(C4H9O)4, AgNO3, and CH3COOH.Ag-doped TiO2 NPs were synthesized by sol–gel technology.2.67 eV[27]
C12H28O4Ti, Cu(NO3)2, CH3COOH, C3H7OH, and NaBH4.Cu(NO3)2 is added to the ultrasonically dispersed TiO2 NPs aqueous suspension and magnetically stirred. Subsequently, add NaBH4 solution and continue stirring for 24 h. Finally, centrifuge, wash, and dry.2.40 eV[28]
TiO2 (001), TiO2 (100), TiO2 (101), and Cu(NO3)2 (0.1 mol/L).CuOx/TiO2 catalysts were synthesized by the initial impregnation method using TiO2 (001), TiO2 (100) and TiO2 (101) as the vector and Cu(NO3)2 (0.1 mol/L) solution as the load.2.25 eV[45]
Titanium-bearing blast furnace slag.Melting pretreatment at 1300 °C for 1 h and then roasted in a microwave field at 1200 °C with a heating time of 20 min.2.25 eVThis study
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

Xie, J.; Ye, Q.; Zhou, J.; Liao, Y.; Qian, G. The Photocatalytic Activity of CaTiO3 Derived from the Microwave-Melting Heating Process of Blast Furnace Slag. Nanomaterials 2023, 13, 1412. https://doi.org/10.3390/nano13081412

AMA Style

Xie J, Ye Q, Zhou J, Liao Y, Qian G. The Photocatalytic Activity of CaTiO3 Derived from the Microwave-Melting Heating Process of Blast Furnace Slag. Nanomaterials. 2023; 13(8):1412. https://doi.org/10.3390/nano13081412

Chicago/Turabian Style

Xie, Jun, Qing Ye, Jianghao Zhou, Yue Liao, and Gongming Qian. 2023. "The Photocatalytic Activity of CaTiO3 Derived from the Microwave-Melting Heating Process of Blast Furnace Slag" Nanomaterials 13, no. 8: 1412. https://doi.org/10.3390/nano13081412

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