Next Article in Journal
Uncontrolled and Controlled Destruction of Acetylene Pressure Cylinders
Next Article in Special Issue
Structural Performance Optimization Design of Continuously Accelerating Electromagnetic Weft Insertion System
Previous Article in Journal
Experimental Study on the Influence of Slurry Concentration and Standing Time on the Friction Characteristics of a Steel Pipe-Soil Interface
Previous Article in Special Issue
Research on Key Factors of Sealing Performance of Combined Sealing Ring
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Study on Photocatalytic Degradation of Acid Red 73 by Fe3O4@TiO2 Exposed (001) Facets

1
Hubei Key Laboratory of Digital Textile Equipment, Wuhan Textile University, Wuhan 430073, China
2
Key Laboratory of Materials Physics, Institute of Solid State Physics, Chinese Academy of Sciences, Hefei 230031, China
*
Author to whom correspondence should be addressed.
Appl. Sci. 2022, 12(7), 3574; https://doi.org/10.3390/app12073574
Submission received: 5 March 2022 / Revised: 25 March 2022 / Accepted: 29 March 2022 / Published: 31 March 2022
(This article belongs to the Special Issue Selected Papers from MMSE 2021)

Abstract

:
Water pollution can be treated through the photocatalytic reaction of TiO2 or TiO2 compounds. A solvothermal method was used to prepare Fe3O4 and Fe3O4@TiO2 composite photocatalyst with (001) high-energy facets exposed in the anatase phase. TiO2 and Fe3O4@TiO2 were characterized by field emission scanning electron microscopy, ultraviolet–visible diffuse reflectance spectroscopy, X-ray diffraction spectroscopy and Raman spectroscopy. It was found that the composite Fe3O4@TiO2 can reduce the band gap and maintain a certain proportion of (001) high-energy facet exposure. The band gaps of Fe3O4@TiO2 and TiO2 are 2.5 eV and 2.9 eV, respectively. The exposure percentages of (001) facets of Fe3O4@TiO2 and TiO2 are about 25.2% and 12.1%, respectively. Fe3O4@TiO2 was used for photocatalytic degradation of Acid Red 73, and it was found that Fe3O4@TiO2 could improve the efficiency of photocatalytic degradation of Acid Red 73. The photocatalytic degradation rates of Fe3O4@TiO2 and TiO2 at 24 min were 93.56% and 74.47%, respectively. The cycle experiment of photocatalytic degradation of Acid Red 73 by Fe3O4@TiO2 showed that at the fifth cycle, the rate of dye degradation decreased to 77.05%, but the rate of dye degradation can reach more than 90% after self-cleaning treatment. The photocatalytic degradation mechanism is explained by the energy band theory and the first-order kinetic equation model.

1. Introduction

Due to the rapid development of technology and industry, it is found that many organic pollutants (dyes, drugs, pesticides, etc.) in natural water have gradually increased and will affect the ecosystem for a long time [1]. Acid Red 73 (AR73) is an azo dye with a good dyeing effect and low price and is widely used in the dyeing of leather, plastics and wool fabrics. However, Acid Red 73 is toxic, and the waste liquid discharged in the dyeing process will lead to environmental pollution and harm to human beings. Therefore, the degradation of Acid Red 73 in the discharged wastewater is particularly important. For example, solar energy has the characteristics of environmental protection. Utilizing solar energy efficiently is one of the hotspots of scientific research. Because solar energy is difficult to use, the research can only focus on solar conversion. In 1972, A. Fujishima and K. Honda first proposed the photocatalytic phenomenon. A crystalline TiO2 electrode and a Pt electrode were placed in water. It was found that the whole circuit would produce photocurrent under light, and the water was successfully cracked into H2 and O2 [2]. In 1976, Craey et al. reported the photocatalytic degradation of polychlorinated biphenyls and haloalkanes by TiO2 as a photocatalyst [3]. In 1983, Purden et al. published research on the degradation of trichloroethylene by TiO2. A series of research results showed that the TiO2 semiconductor catalyst had a strong degradation ability for a variety of organic pollutants, which opened a new direction for the application of semiconductor catalytic materials in environmental protection research [4]. An effective pollution control method is the semiconductor photocatalytic degradation of pollutants in water. Common photocatalysts include C3N4 [5], TiO2 [1] and SiC [6]. TiO2 has the characteristics of good chemical stability, strong oxidation and high photocatalytic activity, but the particles of TiO2 are small and cannot precipitate rapidly in solution, which will cause problems such as difficult recovery and secondary pollution. As a typical magnetic material, Fe3O4 has become the most popular magnetic carrier material due to its excellent magnetism and simple preparation process. The photocatalyst can be separated and recovered by a magnetic field by loading magnetic materials on TiO2. Recyclable Fe3O4@TiO2 composites can be prepared by the sol–gel method [7]. Fe3O4@TiO2 core–shell nanocomposites with rapid magnetic separation were prepared by the solvothermal method, and the photocatalytic performance was significantly enhanced [8]. An Fe3O4@TiO2 magnetic photocatalyst was prepared for photocatalytic degradation of bisphenol A under visible and long-wavelength UV light irradiation [9]. Fe3O4/TiO2 nanocomposites were prepared by precipitation method and sol–gel mixed method. The prepared nanocomposites have good optical properties and can be used for photocatalytic degradation of dyes [10]. Compared with TiO2, the prepared composites have higher photocatalytic performance and better reusability. The methods to improve the photocatalytic effect of TiO2 mainly include increasing the exposure ratio of (001) high-energy facets [11], composites [12,13] and doping [14]. TiO2 has three different crystal forms: rutile, plate titanium and anatase. For anatase-type TiO2, the (001) high-energy facets have better photocatalytic performance than the (101) low-energy facets. It was found that there were physical and chemical adsorption and decomposition reactions between water molecules and (001) facets, but only physical adsorption on (101) facets. Since separated water can facilitate the transfer of photogenerated carriers and the formation of active free radicals, it may make (001) facets have a stronger redox effect than (101) facets, causing the (001) facets to exhibit higher photocatalytic activity [15]. Many of the Fe3O4 and TiO2 composites studied now are core–shell structures. When the outer surface of Fe3O4 powder is coated with a layer of TiO2, it inhibits the transport of photoexcited electrons to a certain extent.
In this paper, the photocatalytic performance of TiO2 was improved by increasing the exposure ratio of (001) high-energy facets and compounding with Fe3O4 to change its band gap. HF was used as an inhibitor, which on the one hand destroyed the spherical structure of the previous Fe3O4 and on the other hand inhibited the growth of (101) surfaces. TiO2 and Fe3O4@TiO2 composite materials were prepared by the solvothermal method as photocatalysts. TiO2 and Fe3O4@TiO2 were characterized by X-ray diffraction spectroscopy, field emission scanning electron microscopy, ultraviolet–visible diffuse reflectance spectroscopy and Raman spectroscopy. Acid Red 73 (AR73) solution was used to simulate dye wastewater. The degradation ability of TiO2 and Fe3O4@TiO2 can be observed through the experiment of photocatalytic degradation of Acid Red 73. Fe3O4@TiO2 was recycled by magnetic recovery to degrade AR73. The photocatalytic degradation mechanism was studied.

2. Methods and Experiment Details

Tetrabutyl titanate AR 98%, polyethylene glycol-4000 AR, ferric chloride hexahydrate (FeCl3·6H2O) AR 99% and L-ascorbic acid 99.99% were purchased from Shanghai Macklin Biochemical Technology Co., Ltd. (Shanghai, China). Sodium hydroxide (NaOH) AR, hydrofluoric acid (HF) GR 40%, sodium acetate AR, ethylene glycol AR and absolute ethanol AR were obtained from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). The test water was ultrapure water. The power of the high-pressure mercury lamp (GGZ300) was 300 W. The concentration of Acid Red 73 solution before and after the photocatalytic experiment was determined with a TU-1950 Beijing Puxi UV-VIS spectrophotometer at the maximum absorption wavelength of 508 nm. The morphology of all samples was characterized by a field emission scanning electron microscope (FESEM: JSM-7800, Tokyo, Japan) operated at 3 kV. The elemental analysis of samples was performed with an energy dispersive spectrometer (EDS). The Raman spectra were obtained using a Raman spectrometer, and the excitation source was a 532 nm Ar laser. The crystalline structure of all samples was characterized by X-ray diffraction (XRD: Empyrean, Panaco Netherlands, Cu-Kα radiation). The 2θ scan range was 20 to 80°, the step length was 0.06 and the scan speed was 4°/min. The diffuse reflectance spectra (DRS) of all samples were measured using a ultraviolet–visible–near-infrared spectrophotometer (UV-VIS-NIR: Shimadzu Solid Spece-3700) equipped with an integrating sphere; the test wavelength range was 200–800 nm at room temperature, and barium sulfate was used as a reference. The DRS were converted to absorption coefficient F(R) by Kubelka–Munk function. The energy band gap (Ebg) of the samples was determined according to the Tauc method from the plot of (F(R)h𝜈)1/2 versus h𝜈 [1].
The preparation method of TiO2 was the solvothermal method, with tetrabutyl titanate as the titanium source, HF as (101) surface inhibitor of the anatase phase and absolute ethanol as dispersant [16]. Firstly, 80 mL ethanol and 3 mL HF were stirred magnetically for 10 min. After stirring evenly, 10 mL tetrabutyl titanate was slowly put into the solution and stirred for 30 min. Then the mixture was moved to a Teflon-lined autoclave. The reaction temperature was 180 °C, and the reaction time was 10 h. After the reactor was cooled, TiO2 was washed alternately with absolute ethanol and 1% NaOH solution and finally washed with ultrapure water and filtered. Then, the white TiO2 powder was put into a vacuum drying box; the drying temperature was 40 °C, and the drying time was 24 h.
The preparation method of Fe3O4 particles was as follows: Firstly, 80 mL ethylene glycol and 2.7 g FeCl3·6H2O were magnetically stirred for 30 min, and then 2 g polyethylene glycol-4000 was added into the above solution and stirred until the solution color was orange red. Finally, 7.2 g sodium acetate was added and stirred until the color of the solution was orange red, and then the solution was moved to the Teflon-lined autoclave. The reaction temperature was 200 °C, and the reaction time was 10 h. After the reaction, the black Fe3O4 powder was selected by magnetic separation, washed three times with absolute ethanol, and put into the vacuum drying box and dried at 40 °C.
The preparation process of Fe3O4@TiO2 particles was as follows: Firstly, 0.1 g of L-ascorbic acid and 0.25 g of Fe3O4 prepared nanoparticles were dispersed into 80 mL of ethanol; then, 10 mL tetrabutyl titanate and 3 mL HF were added successively and stirred until they become evenly dispersed, and the above solution was transferred to the Teflon-lined autoclave. The reaction temperature was 180 °C, and the reaction time was 10 h. After cooling, Fe3O4@TiO2 nanoparticles were washed alternately with absolute ethanol and 1% NaOH solution and magnetically separated.
In order to study the photocatalytic performance of Fe3O4@TiO2 and TiO2, dye wastewater was simulated by AR73 solution. The photocatalytic degradation effect of TiO2 and Fe3O4@TiO2 on AR73 was studied. Photocatalytic experiments were carried out in a photocatalytic workstation. The power of the high-pressure mercury lamp used in the experiment was 300 W and the main radiation wavelength was 365 nm. The lamp was 50 mm above the sample. Quantities of 50 mg Fe3O4@TiO2 and 50 mg TiO2 were weighed and placed at the bottom of two quartz tubes, respectively. Then, 25 mL AR73 solution with a concentration of 20 mg L−1 was added to the two quartz tubes mentioned above. The mixtures in the two quartz tubes were first placed in a dark chamber for 25 min and then in a photocatalytic reaction chamber.
Recyclability of photocatalysts is also important in practical applications. After each photocatalytic degradation experiment, the Fe3O4@TiO2 particles were collected by a permanent magnet, washed alternately with ultrapure water and absolute ethanol three times, and dried at 40 °C in a vacuum drying oven for 24 h.
The rate of dye degradation (η) in wastewater was measured according to the following equation:
η = C 0 C t C 0 × 100 %
where C0 (mg L−1) and Ct (mg L−1) are the initial concentration of dye and the concentration of dye at time t (min), respectively.

3. Results and Discussion

XRD patterns of TiO2, Fe3O4 and Fe3O4@TiO2 are shown in Figure 1. The prepared TiO2 powder displayed a good crystallinity; the diffraction peaks of TiO2 exhibited at 25°, 38°, 48°, 54°, 55.14° and 62.5° corresponded to the (101), (004), (200), (105), (211) and (204) lattice planes of TiO2 (JCPDS No. 65-3369), respectively. The results show that no crystals are formed except the anatase crystals. The diffraction peaks of Fe3O4 exhibited at 30.15°, 35.59°, 43.12°, 53.45°, 56.99° and 62.53° corresponded to the (220), (311), (400), (422), (511) and (440) lattice planes of the magnetite phase (JCPDS card 02-0387) [17], respectively. As shown in the XRD pattern of Fe3O4@TiO2, there are (101), (004), (200), (105), (211) and (204) diffraction peaks of TiO2 and (220), (422) and (400) diffraction peaks of Fe3O4.
The crystallite sizes of TiO2 and Fe3O4@TiO2 were calculated by applying the Scherrer formula to the diffraction peaks of the anatase crystallite plane. The calculation of crystallite sizes (D) was [18]:
D = K λ β cos θ
where D is the grain size (nm), λ is the X-ray wavelength (nm), β is the full width at half maximum (rad), K is the Scherrer constant and θ is the diffraction angle (°). According to the above formula, the grain sizes of TiO2 and Fe3O4@TiO2 are about 13.64 nm and 23.97 nm, respectively.
The results of Raman spectroscopy are shown in Figure 2. The peaks of basic vibration modes of anatase TiO2 were 144, 394, 514 and 636 cm−1 [19]. The vibration at 144 cm−1 of the Eg peak was due to the symmetric stretching vibration of the O-Ti-O bond in TiO2, and its intensity decreases with the increase in exposed (001) facets. The A1g peak at 514 cm−1 was generated by the antisymmetric bending vibration of O-Ti-O in TiO2 and its intensity increased with the increase in exposed (001) facets. Therefore, the percentage of exposed (001) facets exposed for TiO2 and Fe3O4@TiO2 can be calculated by measuring the peak intensity ratio of the Eg and A1g modes [20]. The percentages of exposed (001) facets of TiO2 and Fe3O4@TiO2 were estimated to be 12.1% and 25.2%, respectively.
The morphologies of Fe3O4, TiO2 and Fe3O4@TiO2 are shown in Figure 3. The particle diameters of Fe3O4, TiO2 and Fe3O4@TiO2 are about 390 nm, 100 nm and 120 nm, respectively. The size of Fe3O4@TiO2 and TiO2 particles is uniform, but there is particle agglomeration. The possible reason is that the spherical structure is corroded by acid during the compounding process. Combined with the results of FESEM and XRD, it is found that the particle size inferred from FESEM is larger than that obtained by XRD. It is suggested that the shape of larger particles is determined by the agglomeration of adjacent smaller crystallites [21]. EDS patterns and FESEM elemental mapping of Fe3O4@TiO2 are shown in Figure 4. The results indicate that Ti and Fe elements are evenly distributed, which firmly confirms the incorporation of TiO2 and Fe3O4 in the composite.
The UV-VIS diffuse reflectance spectra of Fe3O4, TiO2 and Fe3O4@TiO2 particles are shown in Figure 5a,b. The Fe3O4@TiO2 spectrum shows a ‘red shift’ compared with the TiO2 spectrum, indicating that Fe3O4@TiO2 can absorb light in a wider wavelength range compared with TiO2, and its photocatalytic activity is significantly improved. According to the Kubelka–Munk function, the absorption threshold is estimated. The energy band gap (Ebg) of Fe3O4@TiO2 and TiO2 is determined by the relationship between (F(R)h𝜈)1/2 and h𝜈 (Tauc plot). The value of Ebg is measured in the high absorption region. The zero F(R) (i.e., Ebg) is measured by extrapolating the linear region to the horizontal axis in the high absorption region [22]. Figure 5c shows band gap spectra of TiO2 and Fe3O4@TiO2 obtained by Kubelka–Munk method. The band gaps of TiO2 and Fe3O4@TiO2 are 2.9 eV and 2.5 eV, respectively. The possible reasons for the decrease in the band gap of composite materials are that after HF ionization, H+ ions react with Fe3O4 to promote ionization and increase the concentration of F¯, causing it to have a stronger inhibitory effect on the (101) surface. The photogenerated electron–hole rate of Fe3O4@TiO2 is higher than that of TiO2, and Fe3O4@TiO2 has good light absorption in the wider wavelength range.
The dark reaction showed that both Fe3O4@TiO2 and TiO2 had good adsorption of AR73, as shown in Figure 6. In the initial stage of photocatalytic reaction, the concentration of AR73 was high, and the rate of dye degradation was not high, but the rate of dye degradation increased rapidly. The intermediate products generated by the degraded AR73 occupied the activity of the photocatalyst and increased as time went on, so the photocatalytic degradation rate was high, but the rate of dye degradation increased slowly, as shown in Figure 6c. After the combination of Fe3O4 and TiO2, the band gap decreased and the exposure ratio of (001) high-energy facets increased, which improved the photocatalytic activity. This result is consistent with Figure 2 and Figure 5c. With the extension of the photocatalytic time, there were too many intermediates generated by AR73, which led to a significant reduction in the contact dimensions between the AR73 and Fe3O4@TiO2, inhibiting the photocatalytic degradation and causing the degradation rate to increase slowly or even not increase. The photocatalytic degradation rates of TiO2 and Fe3O4@TiO2 at 24 min were 74.47% and 93.56%, respectively, which indicated that under the same experimental conditions, the rate of dye degradation by Fe3O4@TiO2 was better than that by TiO2. Fe3O4@TiO2 was recovered by magnetism, as shown in Figure 6d.
The recovered photocatalyst degraded AR73 under the same experimental conditions. The photocatalytic effect of the recycled photocatalyst is shown in Figure 7. The abscissae 0–24 min, 24–42 min, 42–60 min, 60–78 min, 78–96 min and 96–114 min correspond to the first, second, third, fourth, fifth and sixth photocatalytic reactions, respectively. The result indicates that the rate of photocatalytic degradation of AR73 by the photocatalyst recycled for the fifth time is lower than those of the previous four times, but it can still reach 77.05%, as shown in Figure 7. This phenomenon may be caused by the adsorption of some insoluble or insoluble intermediates on the catalyst surface, which occupies the activity of the photocatalyst, affects the absorbance of the catalyst and reduces the yield of photogenerated electron–hole pairs. Fe3O4@TiO2 in water was irradiated with a high-pressure mercury lamp for 40 min to achieve self-cleaning, and the conditions for self-cleaning were the same as those for photocatalytic degradation of AR73. In the sixth photocatalytic degradation experiment of AR73, the concentration of AR73 was 0.91 mg L−1 and the degradation rate was 95.45% after irradiation for 18 min. This indicated that through self-cleaning, the precipitation on the surface was further degraded by the photocatalyst to soluble small molecules, restoring the photocatalytic activity of Fe3O4@TiO2.
Figure 8 shows the linear fitting of first-order kinetics of photocatalytic degradation of AR73 by Fe3O4@TiO2 and TiO2. The degradation reaction rate constant K can be acquired with the following equation [23]:
ln C t C 0 = Kt
where C0 (mg L−1) and Ct (mg L−1) have the same meaning as in formula (1). The fitting results are shown in Figure 8. It shows that the photocatalytic process has a concentration-controlled reaction rate. The degradation reaction rate constant of Fe3O4@TiO2 was better than that of TiO2. This result is consistent with Figure 6c.
When the ultraviolet light irradiates on the surface of the sample, the electrons in the conduction band come from electrons of Fe3O4@TiO2 in the valence band absorbing the energy of photons. Photogenerated electrons are formed in the conduction band, photogenerated holes are formed in the valence band, and highly active hole–electron pairs are formed, forming a redox system. Under the action of an electric field, photogenerated electrons and holes are separated and transferred to Fe3O4@TiO2 (001) and (101) crystal planes, respectively [24]. Photogenerated holes (h+) can oxidize OH and H2O molecules absorbed on the surface of TiO2 to generate ·OH and ·O2. Photogenerated electrons (e) can also react with dissolved oxygen (O2) in water to form superoxide radicals (·O2), and ·O2 further reacts with H2O to form OH [25]. ·OH, h+ and·O2 can decompose AR73 dye under UV irradiation, and a schematic illustration is shown in Figure 9. The reaction process is as follows:
T iO 2 + h 𝜈 h + + e
h + + O H O H
e + O 2 O 2
· O 2 + H 2 O O H + HO 2 ·
H O 2 + H 2 O H 2 O 2 + O H
H 2 O 2 + e O H + O H
H 2 O 2 + O 2 O H + H +
O H / O 2 / h + + A R 73 C O 2 + H 2 O + D e g r a t i o n p r o d u c t s

4. Conclusions

TiO2 and Fe3O4@TiO2 were prepared by the solvothermal method. The structures and photocatalytic properties of TiO2 and Fe3O4@TiO2 were characterized. The photocatalytic rates of AR73 degradation by TiO2 and Fe3O4@TiO2 were compared under high-pressure mercury lamp irradiation. The complex of Fe3O4@TiO2 can reduce the band gap to extend the light absorption range, maintain a certain proportion of (001) high-energy facet exposure, and make the absorption band of the photocatalyst move to the visible band. The band gaps of Fe3O4@TiO2 and TiO2 are 2.5 eV and 2.9 eV, respectively. As revealed by XRD analysis of TiO2 and Fe3O4@TiO2, Fe3O4@TiO2 still has a certain proportion of (001) high-energy surface exposure. The percentages of (001) exposed surface of TiO2 and Fe3O4@TiO2 were estimated to be 12.1% and 25.2% by Raman spectroscopy, respectively. The photocatalytic degradation rates of Fe3O4@TiO2 and TiO2 at 24 min were 93.56% and 74.47%, respectively. The cycle experiment of photocatalytic degradation of Acid Red 73 by Fe3O4@TiO2 showed that at the fifth cycle, the rate of dye degradation can reach 77.05%, but the rate of dye degradation can reach more than 90% after self-cleaning treatment. In the photocatalytic cycle experiment, the reason for the decrease in degradation rate is that some intermediates insoluble in water were adsorbed on the surface of the photocatalyst. However, the photocatalyst can realize self-cleaning under long-time irradiation from a high-pressure mercury lamp, and the intermediate products on its surface are degraded into small molecules and dissolved in water, improving the photocatalytic activity. According to the analysis of the results of the first-order kinetic equation model and photocatalytic degradation of AR73 by Fe3O4@TiO2 and TiO2, the photocatalytic effect of Fe3O4@TiO2 is obviously better than that of TiO2. The Fe3O4@TiO2 photocatalyst can be recycled by magnetism and will not pollute the environment, so it has a certain development prospect in wastewater treatment.

Author Contributions

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

Funding

This research was funded by the National Natural Science Foundation of China, grant number 11904268, and the Hubei Province Education Department of China, grant number Q20181702.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Makropoulou, T.; Panagiotopoulou, P.; Venieri, D. N-doped TiO2 photocatalysts for bacterial inactivation in water. J. Chem. Technol. Biotechnol. 2018, 93, 2518–2526. [Google Scholar] [CrossRef]
  2. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  3. Carey, J.H.; Lawrence, J.; Tosine, H.M. Advances in photolysis of persistent organic pollutants in water. Bull. Environ. Contam. Toxicol. 1976, 16, 697–701. [Google Scholar] [CrossRef] [PubMed]
  4. Pruden, A.L.; Ollis, D.F. Photoassisted heterogeneous catalysis: The degradation of trichloroethylene in water. J. Catal. 1983, 82, 404–417. [Google Scholar] [CrossRef]
  5. Alharbi, O.M.; Khattab, R.A.; Ali, I. Health and environmental effects of persistent organic pollutants. J. Mol. Liq. 2018, 263, 442–453. [Google Scholar] [CrossRef]
  6. Li, P.; Chen, Q.; Chen, B.; Liu, Z. Preparation of phosphotungstic acid/SiC and their photocatalytic activity for rhodamine B. Micro Nano Lett. 2020, 15, 779–783. [Google Scholar] [CrossRef]
  7. Zhang, Q.; Yu, L.; Xu, C.; Zhang, W.; Chen, M.; Xu, Q.; Diao, G. A novel method for facile preparation of recoverable Fe3O4@TiO2 core-shell nanospheres and their advanced photocatalytic application. Chem. Phys. Lett. 2020, 761, 138073. [Google Scholar] [CrossRef]
  8. Xin, T.; Ma, M.; Zhang, H.; Gu, J.; Wang, S.; Liu, M.; Zhang, Q. A facile approach for the synthesis of magnetic separable Fe3O4@TiO2, core-shell nanocomposites as highly recyclable photocatalysts. Appl. Surf. Sci. 2014, 288, 51–59. [Google Scholar] [CrossRef]
  9. Chu, A.C.; Sahu, R.S.; Chou, T.H.; Shih, Y.H. Magnetic Fe3O4@TiO2 nanocomposites to degrade bisphenol A, one emerging contaminant, under visible and long wavelength UV light irradiation. J. Environ. Chem. Eng. 2021, 9, 105539. [Google Scholar] [CrossRef]
  10. Gnanasekaran, L.; Hemamalini, R.; Rajendran, S.; Qin, J.; Yola, M.L.; Atar, N.; Gracia, F. Nanosized Fe3O4 incorporated on a TiO2 surface for the enhanced photocatalytic degradation of organic pollutants. J. Mol. Liq. 2019, 287, 110967. [Google Scholar] [CrossRef]
  11. Liu, G.; Yang, H.G.; Wang, X.; Cheng, L.; Lu, H.; Wang, L.; Cheng, H.M. Enhanced photoactivity of oxygen-deficient ana- tase TiO2 sheets with dominant (001) facets. J. Phys. Chem. B 2009, 113, 21784–21788. [Google Scholar]
  12. Mourao, H.A.; Ribeiro, C. TiO2 and SnO2 magnetic nanocomposites: Influence of semiconductors and synthetic methods on photoactivity. J. Nanosci. Nanotechnol. 2011, 11, 7876–7883. [Google Scholar] [CrossRef] [PubMed]
  13. Liu, G.; Wang, L.; Yang, H.G.; Cheng, H.M.; Lu, G.Q.M. Titania-based photocatalysts-crystal growth, doping and heterostructuring. J. Nanosci. Nanotechnol. 2010, 20, 831–843. [Google Scholar] [CrossRef]
  14. Su, X.; He, Q.; Yang, Y.E.; Cheng, G.; Dang, D.; Yu, L. Free-standing nitrogen-doped TiO2 nanorod arrays with enhancedcapacitive capability for supercapacitors. Diam. Relat. Mater. 2021, 114, 108–168. [Google Scholar] [CrossRef]
  15. Vittadini, A.; Selloni, A.; Rotzinger, F.P.; Grätzel, M. Structure and energetics of water adsorbed at TiO2 anatase (101) and (001) surfaces. Phys. Rev. Lett. 1998, 81, 2954. [Google Scholar] [CrossRef]
  16. Yang, H.G.; Sun, C.H.; Qiao, S.Z.; Zou, J.; Liu, G.; Smith, S.C.; Lu, G.Q. Anatase TiO2 single crystals with a large percentage of reactive facets. Nature 2018, 453, 638–641. [Google Scholar] [CrossRef] [Green Version]
  17. Xu, L.; Wang, J. Fenton-like degradation of 2,4-dichlorophenol using Fe3O4 magnetic nanoparticles. Appl. Catal. B-Environ. 2012, 123–124, 117–126. [Google Scholar] [CrossRef]
  18. Kisch, H.; Sakthivel, S.; Janczarek, M.; Mitoraj, D. A low-band gap, nitrogen-modified titania visible-light photocatalyst. J. Phys. Chem. B 2007, 111, 11445–11449. [Google Scholar] [CrossRef]
  19. Wu, Y.; Zhang, J.; Xiao, L.; Chen, F. Preparation and characterization of TiO2 photocatalysts by Fe3+ doping together with Au deposition for the degradation of organic pollutants. Appl. Catal. B-Environ. 2009, 88, 525–532. [Google Scholar] [CrossRef]
  20. Tian, F.; Zhang, Y.; Zhang, J.; Pan, C. Raman spectroscopy: A new approach to measure the percentage of anatase TiO2 exposed (001) facets. Phys. Chem. C 2012, 116, 7515–7519. [Google Scholar] [CrossRef]
  21. Shah, H.N.; Jayaganthan, R.; Kaur, D. Influence of reactive gas and temperature on structural properties of magnetron sputtered CrSiN coatings. Appl. Surf. Sci. 2011, 257, 5535–5543. [Google Scholar] [CrossRef]
  22. Kadam, A.N.; Dhabbe, R.S.; Kokate, M.R.; Gaikwad, Y.B.; Garadkar, K.M. Preparation of N doped TiO2 via microwave-assisted method and its photocatalytic activity for degradation of Malathion. Spectrochim. Acta A 2014, 133, 669–676. [Google Scholar] [CrossRef] [PubMed]
  23. Butterworth, P.J.; Warren, F.J.; Grassby, T.; Patel, H.; Ellis, P.R. Analysis of starch amylolysis using plots for first-order kinetics. Carbohyd. Polym. 2012, 87, 2189–2197. [Google Scholar] [CrossRef]
  24. Khaki, M.R.D.; Shafeeyan, M.S.; Raman, A.A.A.; Daud, W.M.A.W. Application of doped photocatalysts for organic pollutant degradation—A review. J. Environ. Manag. 2017, 198, 78–94. [Google Scholar] [CrossRef]
  25. Rafieezadeh, M.; Kianfar, A.H. Fabrication of heterojunction ternary Fe3O4/TiO2/CoMoO4 as a magnetic photocatalyst for organic dyes degradation under sunlight irradiation. J. Photochem. Photobiol. A 2022, 423, 113596. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of TiO2, Fe3O4 and Fe3O4@TiO2.
Figure 1. XRD patterns of TiO2, Fe3O4 and Fe3O4@TiO2.
Applsci 12 03574 g001
Figure 2. Raman spectra in the range of 100–800 cm−1 for TiO2 and Fe3O4@TiO2.
Figure 2. Raman spectra in the range of 100–800 cm−1 for TiO2 and Fe3O4@TiO2.
Applsci 12 03574 g002
Figure 3. FESEM images. (a) Fe3O4; (b) TiO2; (c) Fe3O4@TiO2.
Figure 3. FESEM images. (a) Fe3O4; (b) TiO2; (c) Fe3O4@TiO2.
Applsci 12 03574 g003
Figure 4. FESEM elemental mapping and EDS patterns of Fe3O4@TiO2. (a) oxygen elements; (b) Fe elements; (c) Ti elemnts; (d) EDS patterns of Fe3O4@TiO2.
Figure 4. FESEM elemental mapping and EDS patterns of Fe3O4@TiO2. (a) oxygen elements; (b) Fe elements; (c) Ti elemnts; (d) EDS patterns of Fe3O4@TiO2.
Applsci 12 03574 g004
Figure 5. UV-VIS DRS of (a) Fe3O4, (b) TiO2 and Fe3O4@TiO2; (c) Plot of transferred Kubelka−Munk values versus the energy of the light absorbed by TiO2 and Fe3O4@TiO2.
Figure 5. UV-VIS DRS of (a) Fe3O4, (b) TiO2 and Fe3O4@TiO2; (c) Plot of transferred Kubelka−Munk values versus the energy of the light absorbed by TiO2 and Fe3O4@TiO2.
Applsci 12 03574 g005
Figure 6. Photocatalytic degradation of AR73 by (a) Fe3O4@TiO2 and (b) TiO2; (c) the rate of degradation of AR73 by Fe3O4@TiO2 and TiO2 in a dark reaction time of 25 min and irradiation time of 24 min; (d) magnetically adsorbed Fe3O4@TiO2.
Figure 6. Photocatalytic degradation of AR73 by (a) Fe3O4@TiO2 and (b) TiO2; (c) the rate of degradation of AR73 by Fe3O4@TiO2 and TiO2 in a dark reaction time of 25 min and irradiation time of 24 min; (d) magnetically adsorbed Fe3O4@TiO2.
Applsci 12 03574 g006
Figure 7. The rate of AR73 degradation by Fe3O4@TiO2 for six cycles.
Figure 7. The rate of AR73 degradation by Fe3O4@TiO2 for six cycles.
Applsci 12 03574 g007
Figure 8. Linear fitting of first-order kinetics of photocatalytic degradation of AR73.
Figure 8. Linear fitting of first-order kinetics of photocatalytic degradation of AR73.
Applsci 12 03574 g008
Figure 9. Schematic illustration of photocatalytic degradation mechanism of Fe3O4@TiO2.
Figure 9. Schematic illustration of photocatalytic degradation mechanism of Fe3O4@TiO2.
Applsci 12 03574 g009
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sun, L.; Zhou, Q.; Mao, J.; Ouyang, X.; Yuan, Z.; Song, X.; Gong, W.; Mei, S.; Xu, W. Study on Photocatalytic Degradation of Acid Red 73 by Fe3O4@TiO2 Exposed (001) Facets. Appl. Sci. 2022, 12, 3574. https://doi.org/10.3390/app12073574

AMA Style

Sun L, Zhou Q, Mao J, Ouyang X, Yuan Z, Song X, Gong W, Mei S, Xu W. Study on Photocatalytic Degradation of Acid Red 73 by Fe3O4@TiO2 Exposed (001) Facets. Applied Sciences. 2022; 12(7):3574. https://doi.org/10.3390/app12073574

Chicago/Turabian Style

Sun, Li, Quan Zhou, Jiaheng Mao, Xingyu Ouyang, Zhigang Yuan, Xiaoxiang Song, Wenbang Gong, Shunqi Mei, and Wei Xu. 2022. "Study on Photocatalytic Degradation of Acid Red 73 by Fe3O4@TiO2 Exposed (001) Facets" Applied Sciences 12, no. 7: 3574. https://doi.org/10.3390/app12073574

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