Next Article in Journal
Investigating the Performance of a Novel Modified Binary Black Hole Optimization Algorithm for Enhancing Feature Selection
Previous Article in Journal
FSN: Feature Shift Network for Load-Domain (LD) Domain Generalization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation of Fe3O4@SiO2@N-TiO2 and Its Application for Photocatalytic Degradation of Methyl Orange in Na2SO4 Solution

Hubei Key Laboratory of Digital Textile Equipment, Wuhan Textile University, Wuhan 430073, China
*
Author to whom correspondence should be addressed.
Appl. Sci. 2024, 14(12), 5205; https://doi.org/10.3390/app14125205
Submission received: 19 April 2024 / Revised: 29 May 2024 / Accepted: 13 June 2024 / Published: 14 June 2024

Abstract

:
In this paper, Fe3O4@SiO2@TiO2 and N-doped Fe3O4@SiO2@N-TiO2 photocatalysts with magnetic core-shell structures were prepared using a multi-step synthesis method. The materials were analyzed using various techniques, such as X-ray diffraction (XRD), ultraviolet-visible diffuse reflectance spectroscopy (DRS), transmission electron microscopy (TEM), field-emission scanning electron microscopy (FESEM), selected-area electron diffraction patterns (SAED), and X-ray photoelectron spectroscopy (XPS). The results indicated that the prepared samples had an anatase structure, and N was successfully doped. Fe3O4@SiO2@TiO2 and Fe3O4@SiO2@N-TiO2 with different amounts of nitrogen doping were used for the study of photocatalytic degradation of methyl orange (MO) in pure MO solution, and in MO and Na2SO4 (MO-Na2SO4) mixed solution, respectively. The average photocatalytic degradation rate of MO in pure MO solution with three different batches each of Fe3O4@SiO2@TiO2 and Fe3O4@SiO2@N-TiO2 (3 mL of NH4OH used for doping) under high-pressure mercury lamp irradiation reached 85.25% ± 2.23% and 95.53% ± 0.53%, respectively. The average photocatalytic degradation rate of MO in the MO-Na2SO4 mixed solution with three different batches each of Fe3O4@SiO2@TiO2 and Fe3O4@SiO2@N-TiO2 (3 mL of NH4OH used for doping) under the same irradiation condition reached 90.46% ± 3.33% and 97.79% ± 2.09%, respectively. The results showed that Na2SO4 can promote photocatalytic degradation of MO. The experiment of recycling photocatalysts showed that there was still a good degradation effect after five cycles. Finally, the first-order kinetic model and the photocatalytic degradation mechanism were investigated.

1. Introduction

With the rapid development of industrial technology, wastewater and emissions from factories are constantly released into the environment, energy is increasingly scarce today, and organic pollutants are found in natural water bodies [1,2]. Dye wastewater is complex, highly colored, toxic, and difficult to degrade, which can make wastewater treatment more difficult and has a fatal impact on the environment [3]. Usually, physical or chemical methods are used to treat dye wastewater, but the cost is relatively high. Compared with the above methods, semiconductor photocatalytic technology has the advantages of non-pollution, good effect, and simple treatment methods, and has been verified by other researchers [4,5]. Photocatalytic technology is currently an efficient method for treating environmental pollution, and further research is a very meaningful task [6,7]. TiO2 has stable chemical properties, low cost, high activity, and nontoxicity, making it the most promising photocatalyst in semiconductor photocatalysts [8,9]. However, TiO2 also has shortcomings, which are mainly manifested in three aspects: Firstly, the recombination rate of photogenerated electron–hole pairs is very fast, resulting in low photocatalytic activity of TiO2 [10]. Secondly, the bandgap is wide, and the absorption range of the light is very narrow, and it can only absorb the ultraviolet range [10]. Thirdly, it is difficult to quickly separate TiO2 from solution after photocatalytic degradation, which makes it difficult to recover the photocatalyst. If the photocatalyst cannot be recovered, it will also cause secondary pollution [11]. Therefore, how to improve the bandgap of TiO2 to broaden the photosensitive range of light and increase the utilization rate of photocatalysts have become the focus of research in the field of photocatalysis. Currently, the main methods include metal ion doping [12,13], non-metallic doping [14,15,16], and compounding with other materials [17]. Therefore, in this paper, N-doped TiO2 multilayer core-shell structure photocatalysts were designed, and the preparation process of the magnetic photocatalysts Fe3O4@SiO2@TiO2 and Fe3O4@SiO2@N-TiO2 was explored experimentally. The samples were characterized by transmission electron microscopy (TEM), X-ray diffraction spectroscopy (XRD), UV-VIS diffuse reflectance spectroscopy (DRS), and X-ray photoelectron spectroscopy (XPS). The photocatalytic degradation experiments of MO were also carried out in a pure MO solution and a MO-Na2SO4 mixed solution, respectively. The mechanism of photocatalytic degradation of MO and MO-Na2SO4 was meticulously examined.

2. Materials and Methods

2.1. Materials and Reagents

Absolute ethanol (AR 99%), FeCl3·6H2O (AR 99%), sodium acetate (CH3COONa) (AR), ammonia (NH4OH; AR, 25–28%), tetraethyl orthosilicate (AR 98%), sodium sulphate absolute (Na2SO4; AR), and methyl orange (MO, AR) were purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). Tetrabutyl titanate (AR 98%) and polyethylene glycol-4000 (AR) were sourced from Shanghai Macklin Biochemical Technology Co., Ltd., located in Shanghai, China. Water used during the experiment was ultrapure water.

2.2. Preparation of Nanocomposite

Fe3O4 was prepared using the solvothermal method. Here, 2.7 g of FeCl3·6H2O was dissolved in 80 mL of ethylene glycol and stirred magnetically until fully dissolved in a beaker. Then, we poured 2 g of polyethylene glycol-4000 into the above solution and stirred magnetically until it was completely dissolved. Finally, 7.2 g of sodium acetate was poured into a beaker and stirred magnetically until it was dissolved. After complete dissolution, the solution was poured into a reaction kettle and heated at 180 °C for 10 h, then washed 3 times with ethanol, placed in a vacuum drying oven, and heated at 40 °C until completely dried.
Fe3O4@SiO2 was prepared using the sol-gel method. Here, 0.2 g of Fe3O4 powder was added into 80 mL of absolute ethanol and ultrasonically dispersed for 30 min until the Fe3O4 powder was fully mixed with the absolute ethanol. Next, 20 mL of ultrapure water was added into the above suspension, then 5 mL of ammonia was dropped slowly into the mixture, and the suspension was ultrasonically dispersed and mechanically stirred for 30 min and 30 min, respectively. After that, 4 mL of tetraethyl orthosilicate was dropped slowly into the above suspension, and the mixture was continually mechanically stirred for 6 h. Finally, the solution was magnetically separated by permanent magnets, the upper layer was poured out, and the black magnetic fluid was washed three times with absolute ethanol and placed into the vacuum drying box and dried at 40 °C.
Preparation of Fe3O4@SiO2@TiO2 (FS-NT (0 mL), where ‘0 mL’ represents the amount of N source in NH4OH, which was 0 mL) nanomaterials was carried out using the sol-gel method. Here, 7 mL of tetrabutyl titanate was added dropwise into 120 mL of absolute ethanol and ultrasonically dispersed for 10 min. Then, 0.2 g of Fe3O4@SiO2 was added into the above suspension and continued to be ultrasonicated for 20 min. The beaker containing the suspension was then placed in a water bath heated to a temperature of 40 °C, and the suspension was mechanically stirred for 8 h. After the reaction, the solid powder and solution in the beaker were magnetically separated with a magnet. The separated solid powder was washed three times with absolute ethanol, and then the washed powder was dried in a vacuum drying oven at 40 °C for 10 h. After drying, the FS-NT (0 mL) was placed in an atmosphere furnace for annealing. Argon was selected as the protective gas and the annealing temperature was set at 450 °C and the heating time was 3 h.
A batch of Fe3O4@SiO2@N-TiO2 (FS-NT (2–5 mL), where ‘2–5 mL’ represents the amount of N source in NH4OH, which was 2, 3, 4, and 5 mL, respectively) doped with different nitrogen contents was prepared using the sol-gel method. FS-NT (2–5 mL) with different nitrogen contents only required changing the amount of NH4OH added, and the sample preparation process was the same. The preparation of FS-NT (2–5 mL) used the process of preparing FS-NT (3 mL) as an example. Firstly, mixture A (7 mL of tetrabutyl titanate was added into 120 mL of absolute ethanol and dispersed by ultrasound for 10 min, then 0.27 g of Fe3O4@SiO2 was added into the above solution and dispersed by ultrasound for 15 min) and solution B (3 mL of NH4OH and 2 mL of deionized water were added into 20 mL of absolute ethanol) were prepared. Then, solution B was slowly added dropwise to mixture A. The mixture of A and B was stirred for 7 h. The magnetic separation was carried out with a magnet. The obtained grey–black powder was washed three times with absolute ethanol and then the washed powder was placed in a vacuum drying oven at a drying temperature of 60 °C for 10 h. The dried samples were placed in an atmosphere furnace for annealing. Argon was chosen as the protective gas, with a heating temperature of 450 °C and a heating time of 3 h.

2.3. Characterization

The photocatalytic experimental device was a high-pressure mercury lamp (GGZ300, Shanghai Jiguang, Shanghai, China) with a light intensity of 23 mW/cm2. The concentrations of MO dye were quantitatively analyzed using a TU-1950 Beijing Puxi UV-VIS spectrophotometer from Beijing, China, specifically at a peak absorption wavelength of 508 nm, both before and after conducting the photocatalytic tests. Morphological examinations of the samples were conducted using a JSM-7800 field-emission scanning electron microscope from Tokyo, Japan, operating at 3 kV. Additionally, the crystal structures of the samples were detailed through X-ray diffraction analysis using an Empyrean device from Panaco, the Netherlands, employing Cu-Kα radiation, with the 2θ scanning ranging from 10 to 80 degrees, a step length of 0.06, and a scanning speed of 4 degrees per minute. Diffuse reflectance spectra for the samples were recorded on a Shimadzu Solid Spec-3700 UV-VIS-NIR spectrophotometer from Kyoto, Japan, which was coupled with an integrating sphere and operated in a wavelength range of 200–800 nm at ambient temperature. BaSO4 served as the reference standard. XPS measurements were calibrated against the binding energy standard of C 1s at 284.8 eV using a Thermo Scientific ESCALAB 250Xi from Waltham, MA, USA. The total organic carbon (TOC) was measured in a Shimadzu TOC-L from Kyoto, Japan analyzer coupled with an autosampler.

2.4. Photodynamic Activity Test

Here, 20 mg of photocatalyst was added into 10 mL of pure MO solution at 10 mg/L and MO-Na2SO4 solution in a petri dish, respectively. After that, the petri dish was placed in a dark environment to reach adsorption–desorption equilibrium. The dish was then irradiated with a high-pressure mercury lamp as a light source. The calculation of the degradation rate of the dye was:
η % = C 0 C t C 0 × 100 %
where C0 (mg L−1) is the initial concentration and Ct (mg L−1) is the concentration at time t (min).

3. Results and Discussion

XRD patterns of TiO2, Fe3O4, Fe3O4@SiO2, and FS-NT (0, 2, 3, 4, and 5 mL) are shown in Figure 1. The prepared TiO2 displayed good crystallinity: the diffraction peaks of TiO2 were exhibited at 25.2°, 38.7°, 48°, 54°, 55.14°, and 62.5°, corresponding to the (101), (004), (200), (105), (211), and (204) lattice planes of TiO2 (JCPDS No. 65-3369) [18]. The results suggest that TiO2 has anatase crystals. The diffraction peaks of Fe3O4 were exhibited at 30.1°, 35.5°, 43.1°, 53.4°, 56.9°, and 62.5°, corresponding to the (220), (311), (400), (422), (511), and (440) lattice planes of the magnetite phase (JCPDS No. 02-0387) [19]. Fe3O4@SiO2 maintained the characteristic diffraction peaks of Fe3O4 in Figure 1a. XRD patterns of FS-NT (0 mL) are shown in Figure 1a. It can be seen that compared with Fe3O4@SiO2, the peak intensity of Fe3O4 in FS-NT decreased due to the thickness of the TiO2 coating, and the characteristic diffraction peaks of TiO2 appeared at 2θ = 25.2°, 38.7°, 48.1°, 54.1°, and 62.5°, corresponding to (101), (004), (200), (211), and (204) crystal planes, which indicates that TiO2 on the surface of FS-NT (0 mL) has anatase crystal TiO2. The XRD patterns of FS-NT (2–5 mL) are shown in Figure 1b. The intensity of the Fe3O4 characteristic diffraction peaks was still high, which suggests that the prepared FS-NT (2–5 mL) can provide a strong magnetic response for the subsequent magnet recovery samples. The characteristic diffraction peaks of TiO2 appeared at 2θ = 25.2° and 48.1°, corresponding to the (101) and (200) planes of TiO2, respectively. Compared with FS-NT (0 mL), the diffraction peaks of FS-NT (2 and 3 mL) shifted slightly toward the small-angle direction. The reason for this is that N ions are larger than O ions, and both substitutional and interstitial N ions can result in an increase in the lattice parameters of TiO2 [20]. The possible reason for the insignificant shift of FS-NT (4 and 5 mL) is that the addition of 4 or 5 mL of NH4OH can increase the pH of the solution and accelerate the hydrolysis of tetraethyl titanate [21], so that N ions do not or rarely enter into the lattice structure.
The TEM micrographs of FS-NT (0 mL) and FS-NT (3 mL) are shown in Figure 2a,b, where it can be seen that the nanocomposite samples clearly have a core-shell structure. The HRTEM image of FS-NT (0 mL) is shown in Figure 2c. The observed lattice spacings of 0.355 nm could be indexed to the (101) plane of TiO2 [22], whereas those of 0.176 nm corresponded to the (422) plane of Fe3O4 [23]. The HRTEM image of FS-NT (3 mL) is shown in Figure 2d. The observed lattice spacings of 0.355 nm could be indexed to the (101) plane of TiO2 [22], whereas those of 0.263 nm corresponded to the (311) plane of Fe3O4 [23]. Moreover, the corresponding selected-area electron diffraction (SAED) pattern (Figure 2e,f) displayed diffuse diffraction rings that were well indexed to the characteristic planes of TiO2 and Fe3O4. The lattice spacings of (101) and (200) planes of FS-NT (3 mL) in Figure 2f were larger than those of FS-NT (0 mL) in Figure 2e, which may be due to the fact that N ions were larger than O ions in the lattice structure. The results are consistent with those of XRD.
Figure 3a,b show the UV-VIS DRS spectra of FS-NT (0 mL) and FS-NT (2~5 mL), respectively. The maximum of titania absorption located at 320–367 nm could be attributed to isolated Ti4+ sites in octahedral coordination [24] related to full connectivity of Ti–O–Ti linkages in Figure 3a. In Figure 3b, the maximum of titania absorption located at 336–378 nm was observed, and the UV-VIS DRS spectra of FS-NT (2~5 mL) showed a red-shifted onset of the low-energy edge, possibly due to the influence of a small amount of N. The prepared samples had strong absorption in the region at 400–800 nm due to the presence of Fe3O4 [18]. The DRS were converted to absorption coefficients, F(R), via the Kubelka–Munk function. The energy bandgap (Eg) of the samples was determined according to the Tauc method from the plot of (F(R)hv)2 versus hv by Equation (2):
α h v = C ( h ν Ε g ) 2
The bandgaps are shown in Figure 3c,d. The values of bandgaps of FS-NT (0 mL), FS-NT (2 mL), FS-NT (3 mL), FS-NT (4 mL), and FS-NT (5 mL) were 2.92 eV, 2.76 eV, 2.67 eV, 2.6 eV, and 2.52 eV, respectively. Compared with FS-NT (0 mL), FS-NT (2–5 mL) had a smaller bandgap, and although the light absorption range was in the ultraviolet range, it was still expanded. The reason for this is that N atoms in FS-NT (2–5 mL) replaced some O atoms in the lattice of TiO2 and formed a O-Ti-N bond, which created an additional energy level and further reduced the bandgap width. This result is consistent with the literature [25,26].
The chemical state of FS-NT (0 mL) and FS-NT (3 mL) was investigated using XPS analysis, as shown in Figure 4. Figure 4a,b present the N 1s spectra of FS-NT (0 mL) and FS-NT (3 mL), respectively. In Figure 4a,b, binding energy of 401.55 eV was attributed to the surface molecularly chemisorbed dinitrogen [27], while binding energy of 400 eV may be attributed to the presence of NH4OH (N-H) [28], possibly due to the presence of NH4OH in the Fe3O4@SiO2 residual during the preparation process. In Figure 4b, compared with FS-NT (0 mL), two additional peaks of binding energy were observed at 396.8eV and 405.4eV, respectively. N 1s peaks at 396.8 eV were attributed to substitutional nitrogen (O-Ti-N) [27,29,30], while a binding energy of 405.4eV may be attributed to interstitial nitrogen [29,31,32]. The results are consistent with those of XRD and SAED. Figure 4c shows the Ti 2p XPS spectra of FS-NT (0 mL). The binding energies at 458.8 and 464.5 eV were associated with Ti4+ 2p3/2 and Ti4+ 2p1/2, respectively, as confirmed by the spin orbit splitting of 5.7 eV [33]. Therefore, it could be inferred that titanium ion was in the form of Ti4+, and Ti3+ did not exist in the FS-NT. Figure 4d shows the binding energies of 458.6 and 464.3 eV corresponding to Ti4+ 2p3/2 and Ti4+ 2p1/2 for FS-NT (3 mL), respectively. It can be clearly seen that Ti 2p peak values of FS-NT (3 mL) were smaller than those of FS-NT (0 mL), which may be due to N substituting for O, causing a decrease in the binding energy. The derived electronic interaction between Ti and N anions resulted in partial electron transfer from N to Ti. At the same time, the electron density of Ti was also increasing due to the fact that the electronegative energy of N was less than that of O. This revealed that N was incorporated into TiO2 and replaced O [34]. In the O 1s XPS spectra, the presence of Ti-O and surface-adsorbed O2 or OH can be confirmed by the deconvoluted peaks at binding energies of 530.2 and 532.7 eV in Figure 4e,f [34,35], respectively.
The photocatalytic activity of the samples was studied by detecting the UV-VIS spectra of MO dye after the photocatalytic degradation reaction under the same conditions. Figure 5a shows the UV-VIS absorption spectra of MO before (Co) and after photocatalytic degradation by FS-NT (0 and 2–5 mL) in pure MO solution. The results indicate that the photocatalytic effect was best when 3 mL of NH4OH was used as the nitrogen-doping source. Figure 5b shows UV-VIS absorption spectra of MO with an initial concentration of 10 mg L−1 under different illumination times during the photocatalytic degradation of MO using a random batch of FS-NT (3 mL). Figure 5c shows the average degradation rate of MO degraded by three batches of FS-NT (0 and 2–5 mL), and the experimental results of the three different batches were averaged and standard errors were added. In a dark environment, dye in water was removed by adsorption, and the removal rates of MO by FS-NT (0 and 2–5 mL) were 10.05%, 9.34%, 10.91%, 8.33%, and 7.84%, respectively. The results showed that the average degradation rate of MO degraded by three different batches of FS-NT (3 mL) reached 95.53% ± 0.53% within 32 min under irradiation conditions, and this degradation effect was better than that of other samples. The results suggest that proper non-metallic doping can improve the photocatalytic ability of TiO2. The photocatalytic activity of some related materials for MO is presented in Table 1. Although there are differences in the materials [36,37,38], the photocatalytic degradation rate of FS-NT (3 mL) was commendable. The degradation effect of FS-NT (3 mL) in 32 min was favorable. In particular, the degradation rate was still 92% in recycling.
The recovered FS-NT (0 and 3 mL) was used to degrade MO under the same experimental conditions. Figure 6 shows the photocatalytic activity of the recycled FS-NT (0 and 3 mL). The first to fifth cycle experiments corresponded to horizontal coordinates of 0–32 min, 32–64 min, 64–96 min, 96–128 min, and 128–160 min, respectively. The degradation rates of MO degraded by a batch of FS-NT (0 mL) recovered for the five cycles were 89.25%, 80.68%, 75.48%, 84.47%, and 86.30%, as shown in Figure 6a, respectively, and the average degradation rate was 83.24%. The degradation rates of MO degraded by a batch of FS-NT (3 mL) recovered for the five cycles were 94.44%, 91.92%, 91.91%, 90.09%, and 92.20%, as shown in Figure 6b, respectively, and the average degradation rate was 92.11% for the five cycles. The results indicate that the FS-NT (3 mL) photocatalyst exhibited better degradation performance through magnetic recovery. In the second, third, and fourth cycle experiments, the possible reason for the decrease in the degradation rate was the production of insoluble intermediates on the surface of the photocatalyst during the cyclic degradation process [18]. The insoluble intermediates hindered the contact between the photocatalyst and the dye and reduced the absorption of the light by the photocatalyst surface, which resulted in a decrease in photocatalytic activity. In order to improve the degradation effect of cycle test, at the end of the third cycle of the experiment, the photocatalyst was self-cleaning. The self-cleaning process was as follows: the recovered photocatalyst was placed in deionized water and illuminated with a high-pressure mercury lamp for about 1 h, then washed by deionized water and dried.
The UV-VIS absorption spectra of MO degraded by FS-NT (0 and 2–5 mL) in MO-Na2SO4 mixed solution are shown in Figure 7a. The results suggest that the photocatalytic degradation effect of FS-NT (3 mL) was the best. Figure 7b shows UV-VIS absorption spectra of MO degraded by a random batch of FS-NT (3 mL) in MO-Na2SO4 mixed solution at different degradation times. Figure 7c shows the average degradation rate of MO degraded by three different batches of FS-NT (0 and 2–5 mL) in MO-Na2SO4 mixed solution, and the experimental results of the three different batches were averaged and standard errors were added. In a dark environment, the removal rates of MO-Na2SO4 by FS-NT (0 and 2–5 mL) were 10.51%, 10.74%, 11.42%, 8.83%, and 8.53%, respectively. The results showed that the average degradation rate of MO using three different batches of FS-NT (3 mL) in MO-Na2SO4 mixed solution reached 97.79% ± 2.09% within 32 min under irradiation conditions. The results indicate that the FS-NT (3 mL) photocatalyst exhibited better degradation performance in the MO-Na2SO4 mixed solution. It was found that when FS-NT has an appropriate amount of N, Na2SO4 can promote photocatalytic degradation. The possible reason is that Na2SO4 can improve the separation efficiency of photogenerated electron–hole pairs and promote the formation of free radicals with high activity [39]. The pH value of the MO solution was 4.0, and the pH value of the MO-Na2SO4 mixed solution was 6.22. MO is an anionic molecule that exhibits higher adsorption efficiency on the photocatalyst at lower pH values, which is very beneficial to photocatalytic degradation. After the addition of Na2SO4, the increase in pH was not conducive to adsorption, but SO42– reacted with h+ to form SO4·, which enhanced the photocatalytic activity of the photocatalyst and improved the photocatalytic degradation effect [39,40].
In order to study the effect of photocatalyst recycling in the MO-Na2SO4 mixed solution, the photocatalytic degradation of MO degraded by recycled FS-NT (0 and 3 mL) was studied. The degradation rate of MO using a batch of recycled FS-NT (0 mL) in the MO-Na2SO4 mixed solution is shown in Figure 8a, and the average degradation rate of MO for the five-cycle experiment was 91.38%. Figure 8b shows the degradation rate of MO by a batch of FS-NT (3 mL) recovered in the MO-Na2SO4 mixed solution, and the average degradation rate was 98.78%. It was found that in the MO-Na2SO4 mixed solution, the recovered FS-NT (3 mL) showed better photocatalytic degradation than recovered FS-NT (0 mL). The photocatalytic degradation effect of the recycled FS-NT (3 mL) and the recycled FS-NT (0 mL) in the MO-Na2SO4 mixed solution was better than that in the pure MO solution, respectively. The results indicate that the photocatalytic activity of FS-NT (3 mL) was very stable.
The first-order kinetic model is given by the following:
ln ( C 0 C t ) = k t
where t (minutes) is the reaction time, k is the rate constant, and C0 (mg L−1) and Ct (mg L−1) are the same as those in Equation (1).
As shown in Figure 9, the kinetic curves were highly linear, and the correlation coefficients (R2) was close to 1. This means that the photocatalytic reaction followed the first-order reaction kinetic model. The initial rate constant k was obtained using experimental data up to 24 min of simulated irradiation because subsequent data seemed to deviate from the linear plot of ln(C0/Ct) vs. t, owing to the decreasing MO concentration in the bulk solution. The rate constant k of FS-NT (0 mL) and FS-NT (3 mL) in pure MO solution were calculated to be 0.0575 and 0.08114 min−1, respectively. By comparing the k values, it was evident that N doping could improve photocatalytic activity. The corresponding degradation rate constant k values of FS-NT (0 mL) and FS-NT (3 mL) in the MO-Na2SO4 mixed solution were calculated to be 0.07334 and 0.08209 min−1, respectively. The results indicate that FS-NT (3 mL) showed better performance in the MO-Na2SO4 mixed solution. Comparing the k values of the MO and MO-Na2SO4 tests, it was seen that Na2SO4 could accelerate the reaction rate.
Figure 10 shows the schematic diagram of the photocatalytic degradation mechanism of MO in pure MO solution. Photogenerated holes (h+) can oxidize OH and H2O molecules to produce ·OH and ·O2. Photogenerated electrons (e) can indeed react with dissolved oxygen in water, leading to the formation of superoxide radicals ·O2, which can react with H2O to form ·OH [41]. As shown in Figure 10, doping with appropriate N content may introduce lattice defects and oxygen vacancies in the photocatalyst. These lattice defects and oxygen vacancies can enhance the visible light absorption ability of photocatalysts, which can improve the photocatalytic degradation effect [42]. The TOC of MO was 5.395 mg/L in pure MO solution, 3.425 mg/L after degradation by FS-NT (0 mL) in pure MO solution, and 5.089 mg/L after degradation by FS-NT (3 mL) in pure MO solution. Combined with the results of Figure 5, the results suggest that FS-NT (3 mL) has good decolorization ability, but its ability to completely degrade MO is weak.
Figure 11 shows the schematic diagram of the photocatalytic degradation mechanism of MO in the MO-Na2SO4 mixed solution. Under high-pressure mercury lamp irradiation, on the one hand, ·OH and ·O2 were produced, and on the other hand, SO42– was adsorbed on the surface of the photocatalyst, and SO42– reacted with h+ to produce SO4·, as shown in Figure 11. Thus, SO4· and ·OH jointly participated in the redox reaction. Dye will be oxidized by free radicals in solution or on the surface of the photocatalyst and SO4·, which will increase the reaction rate [39]. The synergistic effect of N doping and SO4· further enhanced the photocatalytic degradation effect in the MO-Na2SO4 mixed solution, as shown in Figure 11.

4. Conclusions

Fe3O4@SiO2@TiO2 (FS-NT (0 mL)) and N-doped Fe3O4@SiO2@N-TiO2 (FS-NT (2–5 mL)) photocatalysts with magnetic core-shell structures were prepared using a multi-step synthesis method. The prepared materials were characterized by DRS, XRD, FESEM, XPS, TEM, and SAED. The results of the characterizations showed that the prepared samples had an anatase structure, and N was successfully doped. FS-NT (0 mL) and FS-NT (2–5 mL) with different amounts of nitrogen doping were used for the study of photocatalytic degradation of methyl orange (MO) in a pure MO solution and a MO-Na2SO4 mixed solution, respectively. Under high-pressure mercury lamp irradiation, the average photocatalytic degradation rate of MO in the pure MO solution using three different batches each of FS-NT (0 mL) and FS-NT (3 mL) for the first time reached 85.25% ± 2.23% and 95.53% ± 0.53%, respectively. The average photocatalytic degradation rate of MO in the MO-Na2SO4 mixed solution using three different batches each of FS-NT (0 mL) and FS-NT (3 mL) for the first time under the same irradiation conditions reached 90.46% ± 3.33% and 97.79% ± 2.09%, respectively. The results showed that Na2SO4 can promote photocatalytic degradation of MO by FS-NT (0 mL) and FS-NT (3 mL). The reason for this may be that SO4· was produced by the interaction with h+, and SO4· can degrade MO. The experiment of recycling photocatalysts showed that there was still a good degradation effect after five cycles. The results of the recycle experiment indicated that the FS-NT (3 mL) photocatalyst exhibited better degradation performance in the MO solution and MO-Na2SO4 mixed solution, respectively. Finally, the first-order kinetic model and the photocatalytic degradation mechanism showed that the synergistic effects of N doping and SO4· can enhance the photocatalytic degradation effect of MO in the MO-Na2SO4 mixed solution. The FS-NT photocatalyst can be magnetically recycled and reused without pollution to the environment; thus, it has broad development prospects in wastewater treatment.

Author Contributions

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

Funding

This work has been subsidized by the National Natural Science Foundation of China (Grant No. 11904268) and the Science and Technology Program of Hubei Province (CN; Grant No. 2019CFB613).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Wang, J.; Kim, S.-D.; Lee, J.-Y.; Kim, J.-S.; Jang, N.; Kim, H.; Kim, D.-Y.; Nam, Y.; Han, M.; Kong, S.-H. Photocatalytic Oxidization Based on TiO2/Au Nanocomposite Film for the Pretreatment of Total Phosphorus (TP). Appl. Sci. 2024, 14, 1774. [Google Scholar] [CrossRef]
  2. Huayna, G.; Laura, A.; Churata, R.; Lazo, L.; Guzmán, R.; Ramos, P.G.; Rodriguez, J.M. Synthesis and Characterization of a Photocatalytic Material from TiO2 Nanoparticles Supported on Zeolite Obtained from Ignimbrite Residue Used in Decolorization of Methyl Orange. Appl. Sci. 2024, 14, 3146. [Google Scholar] [CrossRef]
  3. Liu, Q. Pollution and Treatment of Dye Waste-Water. IOP Conf. Ser. Earth Environ. Sci. 2020, 514, 052001. [Google Scholar] [CrossRef]
  4. Yang, L.L.; Zhou, H.; Fan, T.X.; Zhang, D. Semiconductor photocatalysts for water oxidation: Current status and challenges. Phys. Chem. Chem. Phys. 2014, 16, 6810–6826. [Google Scholar] [CrossRef]
  5. Zhang, F.; Wang, X.; Liu, H.; Liu, C.; Wan, Y.; Long, Y.; Cai, Z. Recent Advances and Applications of Semiconductor Photocatalytic Technology. Appl. Sci. 2019, 9, 2489. [Google Scholar] [CrossRef]
  6. El Habra, N.; Visentin, F.; Russo, F.; Galenda, A.; Famengo, A.; Rancan, M.; Losurdo, M.; Armelao, L. Supported MOCVD TiO2 Thin Films Grown on Modified Stainless Steel Mesh for Sensing Applications. Nanomaterials 2023, 13, 2678. [Google Scholar] [CrossRef]
  7. He, C.; He, J.; Cui, S.; Fan, X.; Li, S.; Yang, Y.; Tan, X.; Zhang, X.; Mao, J.; Zhang, L.; et al. Novel Effective Photocatalytic Self-Cleaning Coatings: TiO2-Polyfluoroalkoxy Coatings Prepared by Suspension Plasma Spraying. Nanomaterials 2023, 13, 3123. [Google Scholar] [CrossRef]
  8. Wang, F.; Chen, Z.; Yao, K.; Cai, Z.; Zhang, Q. Study on the photocatalytic mechanism and detoxicity of gemfibrozil by a sunlight-driven TiO2/carbon dots photocatalyst: The significant roles of reactive oxygen species. Appl. Catal. B Environ. Int. J. Devoted Catal. Sci. Its Appl. 2017, 204, 250–259. [Google Scholar]
  9. Fernandes, A.; Mako, P.; Wang, Z.; Boczkaj, G. Synergistic Effect of TiO2 Photocatalytic Advanced Oxidation Processes in the Treatment of Refinery Effluents. Chem. Eng. J. 2019, 391, 123488. [Google Scholar]
  10. Kiwaan, H.A.; Atwee, T.M.; Azab, E.A.; El-Bindary, A.A. Photocatalytic degradation of organic dyes in the presence of nanostructured titanium dioxide. J. Mol. Struct. 2020, 1200, 127115. [Google Scholar] [CrossRef]
  11. Zaleska, A. Doped-TiO2: A Review. Recent Pat. Eng. 2008, 2, 157–164. [Google Scholar] [CrossRef]
  12. Fuerte, A.; Hernandez-Alonso, M.D.; Maira, A.J.; Martinez-Arias, A. Visible Light-Activated Nanosized Doped-TiO2 Photocatalysts. Cheminform 2002, 24, 2718–2719. [Google Scholar] [CrossRef]
  13. Apostolova, I.; Apostolov, A.; Wesselinowa, J. Band Gap Tuning in Transition Metal and Rare-Earth-Ion-Doped TiO2, CeO2, and SnO2 Nanoparticles. Nanomaterials 2023, 13, 145. [Google Scholar] [CrossRef]
  14. Rengifo Herrera, J.A. Preparation, Characterization and Photocatalytic Activity of Commercial TiO2 Powders Co-Doped by N and S; EPFL: Lausanne, Switzerland, 2009. [Google Scholar]
  15. Chen, M.; Chu, W. Degradation of antibiotic norfloxacin in aqueous solution by visible-light-mediated C-TiO2 photocatalysis. J. Hazard. Mater. 2012, 219–220, 183–189. [Google Scholar] [CrossRef]
  16. Ohno, T.; Murakami, N.; Tsubota, T.; Nishimura, H. Development of metal cation compound-loaded S-doped TiO2 photocatalysts having a rutile phase under visible light. Appl. Catal. A. Gen. 2008, 349, 70–75. [Google Scholar] [CrossRef]
  17. Wu, L.; Yu, Y.; Song, L.; Zhi, J. MTiO2 (M=Au, Ag) transparent aqueous sols and its application on polymeric surface antibacterial post-treatment. J. Colloid Interface Sci. 2015, 446, 213–217. [Google Scholar] [CrossRef]
  18. 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. [Google Scholar] [CrossRef]
  19. Xu, L.; Wang, J. Fenton-like degradation of 2, 4-dichlorophenol using Fe3O4 magnetic nanoparticles. Appl. Catal. B Environ. 2012, 123, 117–126. [Google Scholar] [CrossRef]
  20. Yuan, J.; Wang, E.; Chen, Y.; Yang, W.; Yao, J.; Cao, Y. Doping mode, band structure and photocatalytic mechanism of B–N-codoped TiO2. Appl. Surf. Sci. 2011, 257, 7335–7342. [Google Scholar] [CrossRef]
  21. Serpone, N.; Lawless, D.; Khairutdinov, R. Size Effects on the Photophysical Properties of Colloidal Anatase TiO2 Particles: Size Quantization versus Direct Transitions in This Indirect Semiconductor? J. Phys. Chem. 1995, 99, 16646–16654. [Google Scholar] [CrossRef]
  22. Liu, W.; Wei, C.; Wang, G.; Cao, X.; Tan, Y.; Hu, S. In situ synthesis of plasmonic Ag@AgI/TiO2 nanocomposites with enhanced visible photocatalytic performance. Ceram. Int. 2019, 45, 17884–17889. [Google Scholar] [CrossRef]
  23. Ma, J.Q.; Guo, S.B.; Guo, X.H.; Ge, H.G. Liquid-phase deposition of TiO2 nanoparticles on core–shell Fe3O4@SiO2 spheres: Preparation, characterization, and photocatalytic activity. J. Nanoparticle Res. 2015, 17, 307. [Google Scholar] [CrossRef]
  24. Barrera, M.; Viniegra, M.; Escobar, J.; Vrinat, M.; De Los Reyes, J.; Murrieta, F.; García, J. Highly active MoS2 on wide-pore ZrO2–TiO2 mixed oxides. Catal. Today 2004, 98, 131–139. [Google Scholar] [CrossRef]
  25. Asahi, R.; Morikawa, T.; Ohwaki, T.; Aoki, K.; Taga, Y. Visible-light photocatalysis in nitrogen-doped titanium oxides. Science 2001, 293, 269–271. [Google Scholar] [CrossRef] [PubMed]
  26. Griffith, M.J.; Sunahara, K.; Wagner, P.; Wagner, K.; Wallace, G.G.; Officer, D.L.; Furube, A.; Katoh, R.; Mori, S.; Mozer, A.J. Porphyrins for dye-sensitised solar cells: New insights into efficiency-determining electron transfer steps. Chem. Commun. 2012, 48, 4145–4162. [Google Scholar] [CrossRef] [PubMed]
  27. Khore, S.K.; Tellabati, N.V.; Apte, S.K.; Naik, S.D.; Ojha, P.; Kale, B.B.; Sonawane, R.S. Green sol–gel route for selective growth of 1D rutile N–TiO2: A highly active photocatalyst for H2 generation and environmental remediation under natural sunlight. RSC Adv. 2017, 7, 33029–33042. [Google Scholar] [CrossRef]
  28. Ertl, G.; Thiele, N. XPS studies with ammonia synthesis catalysts. Appl. Surf. Sci. 1979, 3, 99–112. [Google Scholar] [CrossRef]
  29. Ananpattarachai, J.; Kajitvichyanukul, P.; Seraphin, S. Visible light absorption ability and photocatalytic oxidation activity of various interstitial N-doped TiO2 prepared from different nitrogen dopants. J. Hazard. Mater. 2009, 168, 253–261. [Google Scholar] [CrossRef] [PubMed]
  30. Di Valentin, C.; Finazzi, E.; Pacchioni, G.; Selloni, A.; Livraghi, S.; Paganini, M.C.; Giamello, E. N-doped TiO2: Theory and experiment. Chem. Phys. 2007, 339, 44–56. [Google Scholar] [CrossRef]
  31. Panepinto, A.; Dervaux, J.; Cormier, P.A.; Boujtita, M.; Odobel, F.; Snyders, R. Synthesis of p-type N-doped TiO2 thin films by co-reactive magnetron sputtering. Plasma Process. Polym. 2020, 17, 1900203. [Google Scholar] [CrossRef]
  32. Gomes, J.; Lincho, J.; Domingues, E.; Quinta-Ferreira, R.M.; Martins, R.C. N–TiO2 photocatalysts: A review of their characteristics and capacity for emerging contaminants removal. Water 2019, 11, 373. [Google Scholar] [CrossRef]
  33. Chi, Y.; Yuan, Q.; Li, Y.; Zhao, L.; Li, N.; Li, X.; Yan, W. Magnetically separable Fe3O4@ SiO2@ TiO2-Ag microspheres with well-designed nanostructure and enhanced photocatalytic activity. J. Hazard. Mater. 2013, 262, 404–411. [Google Scholar] [CrossRef]
  34. Chang, T.; Hou, Y.; Wang, F.; Wang, Y.; Bi, Y. Novel N-TiO2/SiO2/Fe3O4 nanocomposite as a photosensitizer with magnetism and visible light responsiveness for photodynamic inactivation of HeLa cells in vitro. Ceram. Int. 2019, 45, 13393–13400. [Google Scholar] [CrossRef]
  35. Tryba, B.; Wozniak, M.; Zolnierkiewicz, G.; Guskos, N.; Morawski, A.; Colbeau-Justin, C.; Wrobel, R.; Nitta, A.; Ohtani, B. Influence of an electronic structure of N-TiO2 on its photocatalytic activity towards decomposition of acetaldehyde under UV and fluorescent lamps irradiation. Catalysts 2018, 8, 85. [Google Scholar] [CrossRef]
  36. Ge, M.; Guo, C.; Zhu, X.; Ma, L.; Han, Z.; Hu, W.; Wang, Y. Photocatalytic degradation of methyl orange using ZnO/TiO2 composites. Front. Environ. Sci. Eng. China 2009, 3, 271–280. [Google Scholar] [CrossRef]
  37. Huang, M.; Xu, C.; Wu, Z.; Huang, Y.; Lin, J.; Wu, J. Photocatalytic discolorization of methyl orange solution by Pt modified TiO2 loaded on natural zeolite. Dye. Pigment. 2008, 77, 327–334. [Google Scholar] [CrossRef]
  38. Zhang, L.; Lv, F.; Zhang, W.; Li, R.; Zhong, H.; Zhao, Y.; Zhang, Y.; Wang, X. Photo degradation of methyl orange by attapulgite–SnO2–TiO2 nanocomposites. J. Hazard. Mater. 2009, 171, 294–300. [Google Scholar] [CrossRef] [PubMed]
  39. Al-Rasheed, R.; Cardin, D.J. Photocatalytic degradation of humic acid in saline waters. Part 1. Artificial seawater: Influence of TiO2, temperature, pH, and air-flow. Chemosphere 2003, 51, 925–933. [Google Scholar] [CrossRef]
  40. Geng, Q.; Cui, W. Adsorption and photocatalytic degradation of reactive brilliant red K-2BP by TiO2/AC in bubbling fluidized bed photocatalytic reactor. Ind. Eng. Chem. Res. 2010, 49, 11321–11330. [Google Scholar] [CrossRef]
  41. 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 Chem. 2022, 423, 113596. [Google Scholar] [CrossRef]
  42. Zhang, X.-R.; Lin, Y.-H.; Zhang, J.-F.; He, D.-Q.; Wang, D.-J. Photoinduced charge carrier properties and photocatalytic activity of N-doped TiO2 nanocatalysts. Acta Phys.-Chim. Sin. 2010, 26, 2733–2738. [Google Scholar]
Figure 1. XRD patterns of (a) TiO2, Fe3O4, Fe3O4@SiO2, and FS-NT (0 mL), and (b) FS-NT (0 and 2–5 mL).
Figure 1. XRD patterns of (a) TiO2, Fe3O4, Fe3O4@SiO2, and FS-NT (0 mL), and (b) FS-NT (0 and 2–5 mL).
Applsci 14 05205 g001
Figure 2. TEM micrographs of (a) FS-NT (0 mL) and (b) FS-NT (3 mL). HRTEM images of (c) FS-NT (0 mL) and (d) FS-NT (3 mL). SAED of (e) FS-NT (0 mL) and (f) FS-NT (3 mL).
Figure 2. TEM micrographs of (a) FS-NT (0 mL) and (b) FS-NT (3 mL). HRTEM images of (c) FS-NT (0 mL) and (d) FS-NT (3 mL). SAED of (e) FS-NT (0 mL) and (f) FS-NT (3 mL).
Applsci 14 05205 g002
Figure 3. UV-VIS diffuse reflectance spectra of (a) FS-NT (0 mL) and (b) FS-NT (2–5 mL). Tauc plots obtained for (c) FS-NT (0 mL) and (d) FS-NT (2–5 mL).
Figure 3. UV-VIS diffuse reflectance spectra of (a) FS-NT (0 mL) and (b) FS-NT (2–5 mL). Tauc plots obtained for (c) FS-NT (0 mL) and (d) FS-NT (2–5 mL).
Applsci 14 05205 g003
Figure 4. XPS spectra: N 1s of (a) FS-NT (0 mL) and (b) FS-NT (3 mL). Ti 2p of (c) FS-NT (0 mL) and (d) FS-NT (3 mL). O 1s of (e) FS-NT (0 mL) and (f) FS-NT (3 mL).
Figure 4. XPS spectra: N 1s of (a) FS-NT (0 mL) and (b) FS-NT (3 mL). Ti 2p of (c) FS-NT (0 mL) and (d) FS-NT (3 mL). O 1s of (e) FS-NT (0 mL) and (f) FS-NT (3 mL).
Applsci 14 05205 g004
Figure 5. UV-VIS spectra of MO solution after the photocatalytic degradation reaction by (a) FS-NT (0 and 2–5 mL). (b) Time-dependent UV-VIS spectra of MO degraded by a random batch of FS-NT (3 mL) in MO during the photocatalytic degradation process. (c) The average degradation rate of MO degraded by three different batches of FS-NT (0 and 2–5 mL).
Figure 5. UV-VIS spectra of MO solution after the photocatalytic degradation reaction by (a) FS-NT (0 and 2–5 mL). (b) Time-dependent UV-VIS spectra of MO degraded by a random batch of FS-NT (3 mL) in MO during the photocatalytic degradation process. (c) The average degradation rate of MO degraded by three different batches of FS-NT (0 and 2–5 mL).
Applsci 14 05205 g005
Figure 6. The degradation rate of MO by (a) FS-NT (0 mL) and (b) FS-NT (3 mL) for five cycles.
Figure 6. The degradation rate of MO by (a) FS-NT (0 mL) and (b) FS-NT (3 mL) for five cycles.
Applsci 14 05205 g006
Figure 7. UV-VIS spectra of MO in MO-Na2SO4 solution after the photocatalytic degradation reaction by (a) FS-NT (0 and 2–5 mL). (b) Time-dependent UV-VIS spectra of MO degraded by a random batch of FS-NT (3 mL) in MO-Na2SO4 during the photocatalytic degradation process. (c) The average degradation rate of MO degraded by three different batches of FS-NT (0 and 2–5 mL) in MO-Na2SO4 solution.
Figure 7. UV-VIS spectra of MO in MO-Na2SO4 solution after the photocatalytic degradation reaction by (a) FS-NT (0 and 2–5 mL). (b) Time-dependent UV-VIS spectra of MO degraded by a random batch of FS-NT (3 mL) in MO-Na2SO4 during the photocatalytic degradation process. (c) The average degradation rate of MO degraded by three different batches of FS-NT (0 and 2–5 mL) in MO-Na2SO4 solution.
Applsci 14 05205 g007
Figure 8. The degradation rate of MO in MO-Na2SO4 by (a) FS-NT (0 mL) and (b) FS-NT (3 mL) for five cycles.
Figure 8. The degradation rate of MO in MO-Na2SO4 by (a) FS-NT (0 mL) and (b) FS-NT (3 mL) for five cycles.
Applsci 14 05205 g008
Figure 9. Linear fitting of first-order kinetics.
Figure 9. Linear fitting of first-order kinetics.
Applsci 14 05205 g009
Figure 10. Schematic diagram of the photocatalytic degradation mechanism of MO in pure MO solution.
Figure 10. Schematic diagram of the photocatalytic degradation mechanism of MO in pure MO solution.
Applsci 14 05205 g010
Figure 11. Schematic diagram of the photocatalytic degradation mechanism of MO in MO-Na2SO4 mixed solution.
Figure 11. Schematic diagram of the photocatalytic degradation mechanism of MO in MO-Na2SO4 mixed solution.
Applsci 14 05205 g011
Table 1. The photocatalytic activity of related materials [36,37,38].
Table 1. The photocatalytic activity of related materials [36,37,38].
MaterialInitial Concentration of MO (mg/L)Light SourceAmount of Catalyst per mL (mg)Degradation Time (min) and Degradation RateCyclic Degradation Rate
ATT-SnO2-TiO220High-pressure mercury lamp130 (99%)non
ZnO/TiO216High-pressure mercury lamp0.3100 (98%)non
Pt-TiO2/zeolite20High-pressure mercury lamp290 (99%)non
FS-NT (3 mL)10High-pressure mercury lamp232 (95.53% ± 0.53%)92%
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

Sun, L.; Ouyang, X.; Li, Z.; Yuan, Z.; Gong, W.; Chen, Z.; Mei, S.; Liu, Y.; Zhou, Q. Preparation of Fe3O4@SiO2@N-TiO2 and Its Application for Photocatalytic Degradation of Methyl Orange in Na2SO4 Solution. Appl. Sci. 2024, 14, 5205. https://doi.org/10.3390/app14125205

AMA Style

Sun L, Ouyang X, Li Z, Yuan Z, Gong W, Chen Z, Mei S, Liu Y, Zhou Q. Preparation of Fe3O4@SiO2@N-TiO2 and Its Application for Photocatalytic Degradation of Methyl Orange in Na2SO4 Solution. Applied Sciences. 2024; 14(12):5205. https://doi.org/10.3390/app14125205

Chicago/Turabian Style

Sun, Li, Xingyu Ouyang, Zilong Li, Zhigang Yuan, Wenbang Gong, Zhen Chen, Shunqi Mei, Ying Liu, and Quan Zhou. 2024. "Preparation of Fe3O4@SiO2@N-TiO2 and Its Application for Photocatalytic Degradation of Methyl Orange in Na2SO4 Solution" Applied Sciences 14, no. 12: 5205. https://doi.org/10.3390/app14125205

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