Next Article in Journal
β-Mannosidase from Cellulomonas fimi: Immobilization Study and Application in the β-Mannoside Synthesis
Next Article in Special Issue
Impact of Copper(II)-Imidazole Complex Modification on Polycrystalline TiO2: Insights into Formation, Characterization, and Photocatalytic Performance
Previous Article in Journal
Sn-Doped Hematite Films as Photoanodes for Photoelectrochemical Alcohol Oxidation
Previous Article in Special Issue
Visible-Light-Driven Peroxymonosulfate Activation for Accelerating Tetracycline Removal Using Co-TiO2 Nanospheres
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Photocatalytic Degradation of Tetracycline Hydrochloride by Mn/g-C3N4/BiPO4 and Ti/g-C3N4/BiPO4 Composites: Reactivity and Mechanism

1
College of Resources and Environment, Zhongkai University of Agriculture and Engineering, Guangzhou 510225, China
2
School of Environment, Jinan University, Guangzhou 510632, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Catalysts 2023, 13(11), 1398; https://doi.org/10.3390/catal13111398
Submission received: 16 August 2023 / Revised: 20 September 2023 / Accepted: 22 September 2023 / Published: 26 October 2023

Abstract

:
The environmental pollution caused by antibiotics is becoming more serious. In this study, the Mn/BiPO4/g-C3N4 composite (Mn-BPC) and the Ti/g-C3N4/BiPO4 composite (Ti-BPC) were prepared by hydrothermal reaction method and solvent method, respectively, and applied to the degradation of tetracycline (TC) in an aqueous environment. The XRD and HRTEM results showed that these materials had the crystalline rod-like structure of BiPO4 and abundant carbon, nitrogen and carbon–oxygen surface functional groups. The degradation of TC by Ti-BPC and Mn-BPC were nearly 92% and 79%, respectively. The degradation processes of TC were well consistent with the pseudo-second-order kinetics model and R2 values were closer to 1. The trapping experiment showed that electron holes (h+) were the main reactive species for the degradation of tetracycline, OH· and O2 also have certain effects. Also, the possible photocatalytic degradation mechanism of Ti-BPC and Mn-BPC composites was thereby proposed. TC was firstly adsorbed on the surface of catalysts, and subsequently degraded by reactive species such as h+, OH· and O2 generated under visible light excitation. This study shows that the Ti-BPC and Mn-BPC photocatalysts have great potential in antibiotic degradation and can provide new ideas for antibiotic removal in aqueous environments.

1. Introduction

Tetracycline (TC) compounds have been increasingly promoted due to their broad spectrum, low cost and easy availability, accounting for approximately 26% of all antibiotic use. The first-generation TCs (tetracycline, oxytetracycline, and chlortetracycline) are the most used TCs [1]. China consumes approximately 6950 tons of TCs every year, and the total TC consumption in Europe has reached 2957 tons [2]. In recent years, TCs have been widely found in the environment, with the highest levels of TC pollution in water environments [3]. In the Pearl River basin of China, the concentration of antibiotics is as high as 3384 ng/L. The concentration of antibiotics in the medical wastewater discharged by some enterprises exceeds standards. The highest concentration of medical wastewater has been tested, and the concentration of tetracycline antibiotics is 53.688 μg/L [4,5]. At present, the techniques for removal of TCs from wastewater mainly include physical treatment, biological treatment, and chemical oxidation. However, traditional biological treatment methods have a low degradation rate and high cost; physical treatment may cause secondary pollution; and chemical oxidation can reduce the difficulty of subsequent treatment of pollutants by deactivating them. In contrast, the photocatalytic technique achieves thorough treatment through a simple, nontoxic, and environmentally friendly process; thus, this approach has attracted the attention of researchers [6,7,8,9].
In recent years, BiPO4 has been used to degrade organic pollutants due to its high electron–hole separation rate, high chemical stability, simple preparation method, and photocatalytic ability under solar irradiation [10]. For example, the composite material g-C3N4/BiPO4 can be obtained by hydrothermal reaction of BiPO4 and graphitic carbon nitride (g-C3N4). BiPO4 has a wide band gap and a high electron–hole rate, while g-C3N4 can improve the stability of photogenerated electrons and holes in BiPO4, thus improving the antibiotic degradation efficiency of BiPO4 without causing secondary pollution [11]. The degradation rate of antibiotics by a g-C3N4/BiPO4 composite photocatalyst was 1.5 times higher than that by pure BiPO4 under UV irradiation and 2.5 times than that by g-C3N4 under visible light irradiation [12,13]. However, g-C3N4/BiPO4 materials still cannot absorb visible light with wavelengths exceeding 420 nm, and effective utilization methods need to be sought to maximize the excitation efficiency of g-C3N4/BiPO4 materials [14,15]. Some researchers have prepared different g-C3N4/BiPO4 based composite photocatalysts, such as loaded CDs. And the optical properties of the composite material, such as diffusion reflection UV characterization and Tauc plot analysis, were used to determine whether the composite catalyst successfully expanded the absorption range of the target pollutant. The results showed that the addition of CD reduced the energy band gap and led to a wider light absorption range. Qian et al.’s study showed that the degradation rate of tetracycline in water by loaded CDs composite catalysts is 75.5% [16]. Related studies have shown that metal ions can modify g-C3N4/BiPO4 through complexation, effectively improving the catalytic activity and efficiency of g-C3N4/BiPO4 in degrading antibiotics [17]. Darkwah [18] used TiO2 and g-C3N4/BiPO4 to prepare a ternary composite catalyst, and Li [19] used AgBr and g-C3N4/BiPO4 to prepare ternary composite catalysts; these ternary composite catalysts are more effective than g-C3N4/BiPO4 in degrading antibiotics.
To search for loading materials that can more efficiently stimulate the degradation effect of g-C3N4/BiPO4 materials. We prepared two different composite photocatalysts based on the precursor g-C3N4/BiPO4 and supported metal ions, namely, TiO2/g-C3N4/BiPO4 (hereinafter referred to as Ti-BPC) and Mn/g-C3N4/BiPO4 (hereinafter referred to as Mn-BPC). This article compares the improvement effect of two material loads on the g-C3N4/BiPO4 material. The physical and chemical characteristics of the two composite catalysts were analyzed to explore the effects of different influencing factors on the ability of the composite photocatalysts to degrade antibiotics, and the underlying mechanisms were analyzed based on radical trapping experiments; accordingly, the findings of this study provide a reference for the treatment of antibiotics in wastewater.

2. Results

2.1. Effect of the Mass Fractions of TiO2 and MnCl2 on the Properties of Composites

Based on the data collected from the experimental results, the degradation efficiency map was drawn using Origin software, and the time when the xenon lamp was turned on was used as the zero point in time (i.e., the start of the photoreaction).
The composite photocatalysts were prepared by adding different mass fractions of TiO2 and MnCl2, and their measured properties are shown in Figure 1. The figure shows that in the adsorption reaction stage, both Ti-BPC and Mn-BPC had limited adsorption efficiency, and their adsorption removal rates were both in the range of 8% to 20%. In the photocatalytic reaction stage (The determination of adsorption time can be found in Supplementary Figure S1), Ti-BPC samples prepared with different TiO2 mass fractions all had a satisfactory degradation effect on TC. The best photocatalytic effect occurred at TiO2 mass fractions of 12.5 wt% and 10 wt%, at which the degradation rate of TC reached 79.3% in 180 min. Considering the preparation cost, a TiO2 mass fraction of 10 wt% was used in the subsequent preparation of Ti-BPC. When the mass fraction of MnCl2 increased from 5 wt% to 25 wt%, the degradation rate of TC decreased. When the mass fraction of Mn was 30 wt%, the degradation rate of TC reached 84%; when the mass fraction of Mn was 5 wt%, the degradation rate was 85%. Based on comprehensive consideration, a MnCl2 mass fraction of 5 wt% was used to prepare Mn-BPC.

2.2. Analysis of the Physicochemical Properties of the Photocatalytic Materials

The morphology and crystal structure of the two composites were determined by transmission electron microscopy (TEM) and high-resolution transmission electron microscopy (HRTEM), and the results are shown in Figure 2. Figure 2a shows that the constituent molecules of the composite photocatalyst Ti-BPC formed a compact cluster structure with tightly packed rod-shaped and spherical crystals [20]. The BiPO4 nanorods shown in Figure 2b are approximately 60–75 nm in diameter with tiny spherical crystals around the edges. Figure 2c shows tiny spherical crystal structures uniformly attached to the crystal rods, which are presumed to be TiO2 crystals [21]. The HRTEM image in Figure 2d shows that the crystal spacing is 0.2669 nm, corresponding to the (210) plane of elemental Bi. Figure 2e,f shows that the composite photocatalyst Mn-BPC is composed of agglomerates of flower-shaped or granular particles and overlapping layered nanosheet structures. The morphological characteristics of Mn-BPC are the same as those reported by Chen, and the layered graphite structure is g-C3N4. The lattice spacings measured in Figure 2g,h are 0.283 nm and 0.312 nm, corresponding to the (111) plane of BiPO4 and the (012) plane of elemental Bi [22].
Figure 3 shows the X-ray photoelectron spectroscopy (XPS) spectra of the two catalysts. The XPS spectrum of Ti-BPC shows that the composite catalyst contains C, N, O, P, Bi, and Ti with different molecular weights. For the Bi spectrum, Bi4f has the highest intensity, the smallest peak width (158–162 eV) and the best symmetry, and it is the main spectral line of Bi. The Mn-BPC spectral analysis results show that the peak at 133.08 eV is the P2p peak, the peak at 160.60 eV is the Bi4f peak, with the highest peak value, and the peak at 531.08 eV is the characteristic peak of O1s, with the second highest peak value. The peaks of C1s, N1s, and Mn2p appear at 285.05 eV, 339.08 eV, and 642.08 eV, respectively.
From the information in Table 1, it can be seen that the molar ratio of each element of Ti-BPC is 23.89% for C1s, 5.44% for N1s, 43.12% for O1s, 7.17% for P2p, 4.25% for Bi4f, and 16.12% for Ti2p. The content of oxygen is the highest in the composite catalyst, which is consistent with the chemical formula of the catalyst [23]. The element with the lowest molar ratio is bismuth, with a content of only 4.25%, which is lower than the content of both carbon and titanium. The XPS peak information of the Mn-BPC catalyst shows that the molar ratio of each element is 32.41% for C1s, 12.74% for N1s, 36.31% for O1s, 2.89% for P2p, 8.96% for Bi4f, and 6.69% for Ti2p. The results showed that the content of oxygen was the highest, followed by carbon and nitrogen, and the content of manganese was the lowest. For the catalyst material, the manganese content is relatively low.
Figure 4 shows the XPS peak position, peak height, and peak width of each element in Ti-BPC. The peak areas of Bi and O are the largest; there are obvious peaks at the binding energies corresponding to P=O and P-O; obvious main peaks corresponding to P, Bi, and Ti binding to O are visible; and the main peaks of C and N correspond to the binding energies of C-C and C=N, respectively. The larger the peak area is, the higher the crystal phase content; the narrower the peak is, the larger the crystal grains. The XPS analysis of Mn in Figure 4g shows that the signal peaks at 641.28 eV and 653.58 eV correspond to Mn3+ and Mn2+, respectively, indicating the presence of both trivalent and divalent manganese ions in the catalyst. Figure 4h analyzes the P element. In the Mn BPC system, the P element has a characteristic peak at 132.88 eV, which is close to the binding energy of P2p. Figure 4j shows the analysis of Bi, with two peaks at 159.5 eV and 164.9 eV, corresponding to the binding energies of Bi4f5/2 and Bi4f7/2, indicating that the Bi element of BiPO4 exists in the +3 valence state in the composite photocatalyst. Figure 4k shows the analytical spectrum of C, where the two peaks of 284.6 eV and 288.4 eV correspond to N-C=N and C=C bonds. The analysis spectra of O and N correspond to Figure 4i,l, respectively. The peak photoelectron analysis spectra of O are at 531.3 eV, indicating the presence of Bi-O bonds in the catalyst. The analysis spectra of N correspond to C=N-C, N-(C)3, C-N-H at 398.4 eV, 399.5 eV, and 400.1 eV, respectively. The increase in peak area of C=N-C bonds indicates the chemical interaction between g-C3N4, BiPO4, and manganese ions [24], This leads to an increase in the strength of carbon nitrogen conjugated π bonds, as shown in Figure 4.
The XRD pattern of Ti-BPC in Figure 5 shows diffraction peaks at 2θ of 25.281°, 34.946°, 37.800°, 38.575°, and 48.049°, which are the characteristic peaks of TiO2, corresponding to the (101), (103), (004), (112), and (200) planes of the JCPDS standard card PDF#21-1281, respectively. The diffraction peaks at 2θ of 29.033°, 31.316°, 37.850°, 41.866° and 52.001° are characteristic peaks of BiPO4, corresponding to the (111), (102), (112), (211), and (203) planes of the JCPDS standard card PDF#15-0766, respectively [25,26]. The diffraction peaks at 2θ of 42.738° and 49.268° are characteristic peaks of g-C3N4, corresponding to the (111) and (102) planes of the JCPDS standard card PDF#0848, respectively. Mn-BPC has diffraction peaks at 2θ of 19.00°, 36.78°, 38.64°, and 44.96°, which are the characteristic peaks of MnO2, corresponding to the (111), (311), (222), and (400) planes of MnO2 in the JCPDS standard card PDF#44-0992 [27], respectively. Since BiPO4 was prepared in this study by the hydrothermal method, which is the same hydrothermal method used by Jin [28] to prepare BiPO4 (BiPO4 was monoclinic in both studies), the seven diffraction peaks at 2θ of 21.72°. The characteristic peaks of BiPO4 are 29.08°, 31.18°, 34.44°, 46.28° and 52.84°, which correspond to the (−111), (120), (012), (−202), (212), and (132) planes of the JCPDS standard card PDF#80-0209. The diffraction peaks of BiPO4 exhibit an increased background and width, which is mainly caused by a reduction in particle size. A weak peak and a strong peak appear at 2θ of 18.58° and 27.12°, respectively. By comparison with the XRD results of Azam and Xu, it is concluded that the two peaks correspond to the (100) and (002) planes, respectively, where the (002) plane corresponds to the interlayer stacking of conjugated aromatics. The other characteristic peaks of g-C3N4 correspond to the (220) and (310) planes.
Figure 6 shows that the relatively broad peak of Ti-BPC between 3200 cm−1 and 3000 cm−1 is caused by C-H bending; there are several sharp peaks near 1600 cm−1 and 1240 cm−1, all of which are caused by the stretching vibration of C=N, corresponding to C-N and C=N bonds activated by infrared light [29]; the absorption peak at 1400 cm−1 is caused by the bending vibration of C-O bonds; the peaks between 1100 cm−1 and 1000 cm−1 correspond to the strong stretching of O-P-O; the peak near 800 cm−1 corresponds to the bending of O-P-O; and the peaks between 500 cm−1 and 750 cm−1 [30] correspond to the stretching and bending vibration of Ti-O bonds. These results confirm the XRD analysis results. The peaks of Mn-BPC in the range of 529 to 553 cm−1 represent the bending vibration of BiPO4, while the peak near 553 cm−1 corresponds to O-P-O [31]. There is a weak peak near 1250 cm−1, which corresponds to the stretching vibration of C-N, and there is a sharp peak near 1100 cm−1 formed by the C-O stretching vibration, with a steep slope of approximately 6.9%. The absorption peak at 3100 cm−1 corresponds to the N-H stretching vibration. The peak near 1500 cm−1 is the characteristic absorption peak of Mn ions, and the characteristic peak of C=N stretching vibrations is attributed to the asymmetric stretching vibration of polypyrrole rings [32]. The peak near 1700 cm−1 has a small width and a low peak height, so it is attributed to the stretching vibration of C=O in carboxyl groups on the surface of Mn-BPC. Mn-BPC has similar peaks to BiPO4 and g-C3N4, which indicates that the catalyst was successfully prepared.

2.3. Degradation of Tetracycline Performance Analysis

The experimental data in Figure 7a,b show that Ti-BPC had the best TC degradation performance when the pH was weakly acidic (pH = 5), with the degradation rate reaching 58.71% in 120 min. The photocatalytic TC degradation efficiency of Mn-BPC gradually decreased with increasing pH. At pH = 3, the TC degradation rate reached 76.47%. In general, acidic conditions improve the TC degradation efficiency of composite catalysts. The photocatalytic degradation activity of composite catalysts is closely related to their surface charge. A change in pH changes the properties of the photocatalyst surface and solvent molecules, as well as the electrostatic interaction between the substance to be degraded and OH, thus affecting the photodegradation rate [33].
According to the data in Figure 7c,d, as the dosage of composite catalyst increases, the degradation rate also increases. When the dosages of Ti-BPC and Mn-BPC were 1 g/L, the corresponding degradation rates of TC were 91.54% and 79%, respectively. The main reason is that the greater the catalyst dosage is, the more active sites are available for contact between the catalyst and TC, and the better the catalytic degradation performance is. Therefore, the larger the amount of catalyst added, the higher the degradation rate, which is similar to the conclusions of previous research [34,35].
Figure 7e,f shows that the photocatalytic degradation efficiency of Ti-BPC and Mn-BPC decreased with increasing TC concentration. One reason for this is that when the dosage of the photocatalyst is fixed, the number of active sites in the system is constant, so the composite photocatalyst has a limited ability to degrade TC in high-concentration TC solution. Another reason is that as the photocatalytic reaction progresses, many degradation intermediates generated in high-concentration TC solution will compete with TC molecules for active photocatalytic sites on the catalyst surface, hindering the formation of photocatalytically active species [36].
In summary, the effects of various influencing factors on the two composite catalysts are basically the same. Under the optimal conditions, the best TC degradation efficiencies of Ti-BPC and Mn-BPC were 91.54% and 79%, respectively. From this, it can be concluded that Ti-BPC may have better degradation effects.

3. Discussion

3.1. Photocatalytic Kinetic Analysis

Figure 8 and Table 2 and Table 3 show that the degradation of TC by Ti-BPC can be well simulated by first- and second-order kinetics. The minimum R2 values of the pseudo-first-order and pseudo-second-order models were 0.9576 and 0.97465, respectively. The half-life of TC increased with increasing concentration. The half-life values of TC at 4 mg/L and 12 mg/L were 25.23 min and 126.84 min, respectively. The strong photocatalysis effect of Mn-BPC was better simulated by pseudo-second-order kinetics. At low concentrations, the half-life of TC degradation under photocatalysis by Mn-BPC was shorter. When the initial TC concentration was 12 mg/L, the half-life of TC was 138.89 mg/L. Overall, the two catalysts induced higher reaction rates and better TC removal effects at lower initial TC concentrations.

3.2. Analysis of the Capture Experiment of the Active Object Capture Agent

Figure 9a shows that the TC degradation efficiency of Ti-BPC in the control group reached 60% in 2 h. After different radical trapping agents were added, the catalytic effect in each test tube changed to different degrees, and the degradation rate of TC was only 12% 2 h after adding EDTA to trap holes (h+). The results show that the photocatalytic degradation effect of the catalyst was significantly reduced after the addition of EDTA, whereas the addition of tert-butanol or BQ as the radical trapping agent only slightly altered the photocatalytic degradation effect. Figure 9b shows that the photocatalytic TC degradation efficiency of Mn-BPC in the control group was 82.11%, while the addition of tert-butanol or BQ as a radical trapping agent had a certain effect on the degradation efficiency; these results demonstrate that hydroxyl radicals (OH·) and electrons (e) have a certain effect on the degradation efficiency but are not the main effective substances. When EDTA-2Na was added as a hole (h+) trapping agent, the degradation effect was greatly reduced, with a degradation rate of only 20.33%. The results indicate that holes and hydroxyl radicals are the main active species for the degradation of TC by the two catalysts and that electrons also have certain effects on the degradation efficiency.

3.3. Analysis of Adsorption Mechanism

Figure 10 shows the possible photocatalytic degradation mechanism of TiO2-BPC and Mn-BPC [37]. The semiconductor BiPO4 is combined with g-C3N4 to form a heterogeneous structure with Ti and Mn attached to the surface and interface. The semiconductor material BiPO4 can directly absorb photons with wavelengths less than 320 nm in sunlight, and g-C3N4 can directly absorb photons with wavelengths less than 420 nm [38]. These two materials can absorb visible-light photons with a wavelength greater than 500 nm through the Ti and Mn particles attached to their surfaces to improve their effectiveness. When the semiconductors BiPO4 and g-C3N4 are in contact with each other, g-C3N4 tends to lose e and thus carries a net positive charge, and the semiconducting BiPO4 tends to accept e and thus carries a net negative charge. Charge transfer continues until the Fermi energy levels are balanced and an internal electric field is formed at the interface, which accelerates the migration and separation of photogenerated e-h+ [39]. After the two materials absorb solar energy, the electrons in the valence band (VB) are excited into the conduction band (CB), thereby generating holes in the VB. Holes can directly degrade TC into the intermediate products CO2 and H2O and can also react with OH- or H2O to generate -OH. Hydroxyl radicals can directly act on TC degradation. The hopping electrons combine with O2 to produce O2−. O2− and superoxide radicals can catalyze TC degradation.

4. Materials and Methods

4.1. Materials

The TC used in this experiment was purchased from Hefei Ge’en Technology Co., Ltd. (Hefei, China) and stored at 4 °C; melamine was purchased from the Wuxi Yuanwang Chemical Reagent Co., Ltd. (Wuxi, China); Tert-Butyl alcohol, disodium Ethylenediaminetetraacetic acid, anhydrous ethanol (99.7% by mass), and other reagents were purchased from the Aladdin Reagent Company; all chemical substances used in the experiment are analytical pure.

4.2. Preparation of Composite Photocatalyst

The g-C3N4 was prepared by calcining 20 g of melamine with high temperature in a muffle furnace at 550 °C; (1) Weighing a fixed ratio of TiO2, g-C3N4, Bi(NO3)3-5H2O, NaH2PO4-2H2O, where the ratio of g-C3N4 and BiPO4 masses was a constant 1:5, changing only the mass fraction ratio of TiO2 in the catalyst [40]. Deionized water was added for ultrasonic stirring, followed by washing the solution in small amounts into the Teflon bottle (the solution in the bottle did not exceed 80% of the total capacity of the Teflon bottle [41]). Subsequently it was placed in a hydrothermal reactor for 8 h at 160°; after extraction, drying, and grinding to obtain 5 wt%, 7.5 wt%, 10 wt%, 12.5 wt%, 15 wt%, and 17.5 wt% Ti-BPC. (2) Weigh an appropriate amount of g-C3N4 and add an appropriate amount of Bi(NO3)3-5H2O with NaH2PO4-2H2O with ultrasonic stirring to form suspension; subsequently placed in a hydrothermal reactor at 160° for 6 h [42]; the precursor complex g-C3N4/BiPO4 was obtained as a white solid, an appropriate amount of MnCl2 dissolved in deionized water was added, the mixed solution was poured into the autoclave, sealed and compacted in an oven at 160° for 8 h; 5 wt%, 10 wt%, 15 wt%, 20 wt%, 25 wt%, and 30 wt% of Mn-BPC, respectively.

4.3. Analysis and Testing Methods

To perform a more detailed analysis of the semiconductor composites Mn-BPC and Ti-BPC, X-ray photoelectron spectroscopy (XPS), X-ray diffraction (XRD; model: D8 Advance, Bruker, Germany), and high-resolution transmission electron microscopy (X-ray photoelectron spectroscopy) were used. The material was characterized by HRTEM and Fourier transform infrared spectroscopy (FTIR; model: Nicolet iS10, Thermo Fisher, USA), and the physical and chemical properties of the material were obtained from the data. XRD was used to characterize the crystal form and grain size of the composite photocatalyst, and the scanning range was 2θ = 10–90°. FTIR was used to explore the functional groups of the composite photocatalyst, and the measurement range was 4000–400 cm−1. The morphological characteristics, particle size, and lattice width of the material could be observed and compared with the standard crystal plane spacing d of the crystal, and the crystal plane to which the material belongs could be determined.

4.4. Degradation Experiment

4.4.1. Tube Experiment

The experiment explored the effect of four different influencing factors, namely preparation ratio, catalyst dosage, initial concentration of TC, and ph value of solution, on the degradation effect. One of the conditions was changed, respectively, and the other three influencing factors were maintained unchanged. Except for the TC initial concentration as an influencing factor on the experiment, all other three groups were set with an initial concentration of 5 mg/L for tetracycline wastewater. The experiment was carried out in a 10 mL quartz tube, and the photocatalytic reactor was opened to start agitation. The first 1 h was used to remove the adsorption. After 1 h, the absorbance of the solution in the test tube was measured by sampling, and the data were recorded. The photocatalytic reaction lasted for a total of 90 min. The Xenon lamp was turned on and adjusted to 1000 w (the maximum setting), stirring and rotating were started. Samples were taken every 20 min within the first 1 h of the reaction, and then every 30 min, the absorbance was measured and the data were recorded. The removal rate of tetracycline in wastewater (Formula (1)) was used to represent the effect, and the adsorption rates of the composite photocatalyst were calculated, respectively.
The removal efficiency of TC is expressed by the following formula:
η = C t C 0 × 100 %
In the formula: C 0 —the concentration of TC at the beginning of catalysis, mg/L; C t —concentration of TC-HCl at time t in the catalytic process, mg/L.

4.4.2. Active Substance Capture Experiment

By adding different trapping agents in the degradation process, the different degradation performance is obtained, and then the degradation rate determines the active substance. In this experiment, tert-butanol, EDTA-2Na and BQ. were added during the degradation process as trapping agents of ·OH, Vacancy (h+) and e, respectively.

4.5. Study on Photocatalytic Kinetics

In order to further explore whether the photocatalytic degradation of TC solution by composite materials conforms to the kinetic reaction order, this experiment takes a TC solution with different initial concentrations as the condition to measure its degradation effect. According to the first order (Formula (2)) [43] and the second-order-reaction kinetic equation (Formula (3)) [44], the simulation study was carried out:
l n C 0 C t = k 1 t
1 C t 1 C 0 = k 2 t
The half-life (t1/2) of the quasi-primary kinetic and quasi-secondary kinetic models is calculated by the equation (Equations (4) and (5))
t 1 / 2 = l n ( 2 ) 1 K 1
t 1 / 2 = 1 C 0 K 2
In the formula: C0—the concentration of TC-HCl at the beginning of catalysis, mg/L; Ct—concentration of TC-HCl at time t in the catalytic process, mg/L. k1, k2-kinetic model rate constants; t-Reaction time.

5. Conclusions

In this study, the Mn/BiPO4/g-C3N4 composite (Mn-BPC) and the Ti/g-C3N4/BiPO4 composite (Ti- BPC) photocatalysts were successfully prepared and applied to the degradation of TC in water. The experimental results showed that the reactivity of Ti-BPC and Mn-BPC composite were higher than that of the original binary g-C3N4/BiPO4. Meanwhile, the main factors affecting the photocatalytic process and degradation efficiency of TC were investigated, and the degradation efficiencies of Ti-BPC and Mn-BPC for TC were about 92% and 79% under the optimal conditions, respectively. The data fitting of reaction kinetics showed that TC degradation was well consistent with pseudo-second-order kinetics model. Several reactive species such as h+, OH· and O2 were generated in this photocatalytic system, and the possible reaction mechanism of TC degradation was proposed. In the experiment for exploring catalytic performance, the optimal conditions for the degradation of antibiotics by the catalyst were obtained. The materials used in the preparation process are common and will not cause secondary pollution to the water body during the application process. In subsequent applications, powdered biochar will be prepared into spherical carbon for easy recovery and utilization during the application process. However, in practical applications, the types of wastewater are complex and diverse, and the degradation mechanisms of different modified photocatalysts are also different. Currently, it is difficult to apply them on a large scale. Subsequent research can be conducted in different wastewater to develop highly selective catalysts, and to seek more energy-saving and low-cost stable preparation processes.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal13111398/s1, Figure S1: Effect of Mass Fractions of TiO2 and MnCl2 in Catalysts on the Adsorption Efficiency of TC-HCl ((a): Ti-BPC; (b): Mn-BPC).

Author Contributions

Conceptualization, Z.D. and M.L.; methodology, Z.D.; software, Y.F.; validation, H.L., Y.D., Y.L. and Y.Z.; formal analysis, W.Q. and Y.F.; writing—original draft preparation, W.Q. and Y.F.; writing—review and editing, Z.D.; supervision, M.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Key Realm Research and Development Program of Guangdong Province (No. 2020B1111350002), the Special project in key areas of Guangdong Province Ordinary Universities (Nos. 2020ZDZX1003 and 2021ZDJS007), the National Natural Science Foundation of China (No. 21407155), the Guangdong Provincial Special Fund for Modern Agriculture Industry Technology Innovation Teams (No. 2019KJ140) and the Zhongkai Graduate Science and Technology Innovation Fund Project (No. KJCX2023023). This work was funded by the Project of Educational Commission of Guangdong Province of China (No. 2019KTSCX067 and 2017KQNCX102), the Guangdong Special Commissioners in Agricultural Science and Technology (No. KTP20190016), the Key Realm R&D Program of Guangdong Province (No. 2022032011800001), the Key Realm R&D Program of Guangdong Province (No. 2020B0202080002).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ma, J.; Li, B.; Zhou, L.; Zhu, Y.; Li, J.; Qiu, Y. Simple Urea Immersion Enhanced Removal of Tetracycline from Water by Polystyrene. Microspheres Int. J. Environ. Res. Public Health 2018, 15, 1524. [Google Scholar] [CrossRef] [PubMed]
  2. Xie, X.J.; Zhou, Q.X.; He, Z.C.; Bao, Y.Y. Physiological and potential genetic toxicity of chlortetracycline as an emerging pollutant in wheat (Triticum aestivum L.). Environ. Toxicol. Chem. 2010, 29, 922–928. [Google Scholar] [CrossRef] [PubMed]
  3. Lu, P.; Fang, Y.; Barvor, J.B.; Neth, N.L.K.; Fan, N.; Li, Z.; Cheng, J. Review of Antibiotic Pollution in the Seven Watersheds in China. Pol. J. Environ. Stud. 2019, 28, 4045–4055. [Google Scholar] [CrossRef] [PubMed]
  4. Kumar, M.; Jaiswal, S.; Sodhi, K.K.; Shree, P.; Singh, D.K.; Agrawal, P.K.; Shukla, P. Antibiotics bioremediation: Perspectives on its ecotoxicity and resistance. Environ. Int. 2019, 124, 448–461. [Google Scholar] [CrossRef] [PubMed]
  5. Zhang, Q.Q.; Ying, G.G.; Pan, C.G.; Liu, Y.S.; Zhao, J.L. Comprehensive evaluation of antibiotics emission and fate in the river basins of China: Source analysis, multimedia modeling, and linkage to bacterial resistance. Environ. Sci. Technol. 2015, 49, 6772–6782. [Google Scholar] [CrossRef]
  6. Rafael, A.H.; Carlos, G.D.; María, A.I.; Gabriel, G.R.; Jaime, C.; Enrique, E. Comparative adsorption of tetracyclines on biochars and stevensite: Looking for the most effective adsorbent. Appl. Clay Sci. 2018, 160, 162–172. [Google Scholar]
  7. Javid, A.; Mesdaghinia, A.; Nasseri, S.; Mahvi, A.H.; Alimohammadi, M.; Gharibi, H. Assessment of tetracycline contamination in surface and groundwater resources proximal to animal farming houses in Tehran, Iran. J. Environ. Health Sci. Eng. 2016, 14, 4. [Google Scholar] [CrossRef]
  8. Rizzo, L.; Manaia, C.; Merlin, C.; Schwartz, T.; Dagot, C.; Ploy, M.C.; Michael, I.; Fatta-Kassinos, D. Urban wastewater treatment plants as hotspots for antibiotic resistant bacteria and genes spread into the environment: A review. Sci. Total Environ. 2013, 447, 345–360. [Google Scholar] [CrossRef]
  9. Thai, H.L.; Charmaine, N.; Ngoc, H.T.; Hongjie, C.; Karina, Y.G. Removal of antibiotic residues, antibiotic resistant bacteria and antibiotic resistance genes in municipal wastewater by membrane bioreactor systems. Water Res. 2018, 145, 498–508. [Google Scholar]
  10. Huang, S.; Wang, G.; Liu, J.; Du, C.; Su, Y. A novel CuBi2O4/BiOBr direct Z-scheme photocatalyst for efficient antibiotics removal: Synergy of adsorption and photocatalysis on degradation kinetics and mechanism insight. ChemCatChem 2020, 12, 4431–4445. [Google Scholar] [CrossRef]
  11. Yang, F.; Yu, X.J.; Liu, Z.B.; Wang, K.; Liang, Y.D.; Qiu, J.; Zhang, T.; Niu, J.F.; Zhao, J.; Yao, B.H. The effects of pH on reactive species formation in moxifloxacin photocatalytic degradation via BiPO4 prismatic. Opt. Mater. 2022, 133, 112892. [Google Scholar] [CrossRef]
  12. Hu, S.; Ma, L.; You, J.; Li, F.; Fan, Z.; Wang, F.; Liu, D.; Gui, J. A simple and efficient method to prepare a phosphorus modified g-C3N4 visible light photocatalyst. RSC Adv. 2014, 4, 21657–21663. [Google Scholar] [CrossRef]
  13. Das, S.; Ahn, Y.H. Synthesis and application of CdS nanorods for LED-based photocatalytic degradation of tetracycline antibiotic. Chemosphere 2022, 291, 132870. [Google Scholar] [CrossRef]
  14. Ahmed, F.; Ahmed, A.; Tamoor, T.; Hassan, T. Comment on Dual-Band Perfect Metamaterial Absorber Based on an Asymmetric H-Shaped Structure for Terahertz Waves. Materials 2018, 12, 2193. [Google Scholar] [CrossRef]
  15. Alasri, T.M.; Ali, S.L.; Salama, R.S.; Alshorifi, F.T. Band-Structure Engineering of TiO2 Photocatalyst by AuSe Quantum Dots for Efficient Degradation of Malachite Green and Phenol Journal of Inorganic and Organometallic. Polym. Mater. 2023, 33, 1729–1740. [Google Scholar]
  16. Qian, W.; Hu, W.T.; Jiang, Z.F.; Wu, Y.Y.; Li, Z.H.; Diao, Z.H.; Li, M.Y. Degradation of Tetracycline Hydrochloride by a Novel CDs/g-C3N4/BiPO4 under Visible-Light Irradiation: Reactivity and Mechanism. Catalysts 2022, 12, 774. [Google Scholar] [CrossRef]
  17. Chang, X.; Wang, Y.; Zhou, X.; Song, Y.; Zhang, M. ZIF-8-derived carbon-modified g-C3N4 heterostructure with enhanced photocatalytic activity for dye degradation and hydrogen production. Dalton Trans. 2021, 50, 17618–17624. [Google Scholar] [CrossRef]
  18. Darkwah, K.W.; Adormaa, B.B.; Sandrine, C.K.M.; Ao, Y. Modification strategies for enhancing the visible light responsive photocatalytic activity of the BiPO4 nano-based composite photocatalysts. Catal. Sci. Technol. 2019, 9, 546–566. [Google Scholar] [CrossRef]
  19. Li, Y.Z.; Li, Z.; Xia, Y.S.; Li, H.F.; Shi, J.H.; Zhang, A.M.; Huo, H.H.; Tan, S.Y.; Gao, L.Z. Fabrication of ternary AgBr/BiPO4/g-C3N4 heterostructure with dual Z-scheme and its visible light photocatalytic activity for Reactive Blue 19. Environ. Res. 2021, 192, 110260. [Google Scholar] [CrossRef]
  20. Yuan, J.H.; Amano, Y.; Machida, M. Characterization of High-Performance Zinc Chloride Activated Biochar Modified by Thermal Chemical Vapor Deposition (CVD) and Its Removal Mechanism of Aqueous Nitrate Ions. Int. J. Environ. Res. 2022, 16, 57. [Google Scholar] [CrossRef]
  21. Shahabi, N.M.; Sheibani, H. Super-efficient removal of arsenic and mercury ions from wastewater by nanoporous biochar-supported poly 2-aminothiophenol. J. Environ. Chem. Eng. 2022, 10, 107363. [Google Scholar] [CrossRef]
  22. Shi, Z.; Zhang, Y.; Shen, X.; Duoerkun, G.; Zhu, B.; Zhang, L.; Li, M.; Chen, Z. Fabrication of g-C3N4/BiOBr heterojunctions on carbon fibers as weaveable photocatalyst for degrading tetracycline hydrochloride under visible light. Chem. Eng. J. 2020, 386, 124010. [Google Scholar] [CrossRef]
  23. Nouri, A.; Dizaji, B.F.; Kianinejad, N.; Rad, A.J.; Rahimi, S.; Irani, M.; Jazi, F.S. Simultaneous linear release of folic acid and doxorubicin from ethyl cellulose/chitosan/g-C3N4/MoS2 core-shell nanofibers and its anticancer properties. J. Biomed. Mater. Res. Part A 2020, 109, 903–914. [Google Scholar] [CrossRef] [PubMed]
  24. Azzam, A.B.; Djellabi, R.; Sheta, S.M.; El Sheikh, S.M. Ultrafast conversion of carcinogenic 4-nitrophenol into 4-aminophenol in the dark catalyzed by surface interaction on BiPO4/g-C3N4 nanostructures in the presence of NaBH4. RSC Adv. 2021, 11, 18797–18808. [Google Scholar] [CrossRef] [PubMed]
  25. Long, D.; Chen, Z.Q.; Rao, X.; Zhang, Y.P. Sulfur-Doped g-C3N4 and BiPO4 Nanorod Hybrid Architectures for Enhanced Photocatalytic Hydrogen Evolution under Visible Light Irradiation. ACS Appl. Energy Mater. 2020, 3, 5024–5030. [Google Scholar] [CrossRef]
  26. Liu, X.S.; Wang, G.; Zhi, H.; Dong, J.; Hao, J.; Zhang, X.; Wang, J.; Li, D.T.; Liu, B.S. Synthesis of the Porous ZnO Nanosheets and TiO2/ZnO/FTO Composite Films by a Low-Temperature Hydrothermal Method and Their Applications in Photocatalysis and Electrochromism. Coatings 2022, 12, 695. [Google Scholar] [CrossRef]
  27. Muhammad, S.N.; Guorui, Y.; Iqra, A.; Wang, X.J.; Silan, W.; Yan, W. Hybridization of g-C3N4 quantum dots with 1D branched TiO2 fiber for efficient visible light-driven photocatalytic hydrogen generation. Int. J. Hydrogen Energy 2020, 45, 13994–14005. [Google Scholar]
  28. Dai, D.; Wang, P.; Bao, X.; Xu, Y.; Wang, Z.; Guo, Y.H. g-C3N4/ITO/Co-BiVO4 Z-scheme composite for solar overall water splitting. Chem. Eng. J. 2022, 433, 134476. [Google Scholar] [CrossRef]
  29. Liu, N.; Lu, N.; Yu, H.; Chen, S.; Quan, X. Enhanced degradation of organic water pollutants by photocatalytic in situ activation of sulfate based on Z-scheme g-C3N4/BiPO4. Chem. Eng. J. 2022, 428, 132116. [Google Scholar] [CrossRef]
  30. Wang, Q.; Lin, Y.; Li, P.; Ma, M.; Zhang, R. An efficient Z-scheme (Cr, B) codoped g-C3N4/BiVO4 photocatalyst for water splitting: A hybrid DFT study. Int. J. Hydrogen Energy 2020, 46, 247–261. [Google Scholar] [CrossRef]
  31. Qamar, M.A.; Shahid, S.; Javed, M.; Sher, M.; Iqbal, S.; Bahadur, A.; Li, D.X. Fabricated novel g-C3N4/Mn doped ZnO nanocomposite as highly active photocatalyst for the disinfection of pathogens and degradation of the organic pollutants from wastewater under sunlight radiations. Colloids Surf. A Physicochem. Eng. Asp. 2020, 611, 125863. [Google Scholar] [CrossRef]
  32. Xu, X.Q.; Wang, S.M.; Hu, T.J.; Yu, X.F.; Wang, J.P.; Jia, C.J. Fabrication of Mn/O co-doped g-C3N4: Excellent charge separation and transfer for enhancing photocatalytic activity under visible light irradiation. Dye. Pigment. 2020, 175, 108107. [Google Scholar] [CrossRef]
  33. Li, Z.; Ni, G.H.; Wang, H.; Sun, Q.; Wang, G.; Jiang, B.Y.; Zhang, C. Molecular structure characterization of lignite treated with ionic liquid via FTIR and XRD spectroscopy. Fuel 2020, 272, 117705. [Google Scholar]
  34. Yang, N.; Yang, S.; Ma, Q.; Beltran, C.; Guan, Y.; Morsey, M.; Brown, E.; Fernando, S.; Holsen, T.M.; Zhang, W.; et al. Solvent-Free Nonthermal Destruction of PFAS Chemicals and PFAS in Sediment by Piezoelectric Ball Milling. Environ. Sci. Technol. Lett. 2023, 10, 198–203. [Google Scholar] [CrossRef]
  35. Chen, D.; Ray, K.A. Photocatalytic kinetics of phenol and its derivatives over UV irradiated TiO2. Appl. Catal. B Environ. 1999, 23, 143–157. [Google Scholar] [CrossRef]
  36. Suryamani, S.; Rashmi, A. An overview on recent developments in synthesis and molecular level structure of visible-light responsive g-C3N4 photocatalyst towards environmental remediation. Mater. Today Proc. 2020, 35, 150–155. [Google Scholar]
  37. Wang, T.T.; Ren, M.; Bao, S.S.; Cai, Z.S.; Liu, B.; Zheng, Z.H.; Xu, Z.L.; Zheng, L.M. The O-P-O bridged Mn2(salen)2 chains showing coexistence of single chain magnet and metamagnet behaviour. Dalton Trans. 2003, 2015, 44. [Google Scholar] [CrossRef]
  38. Li, L.; Salvador, P.A.; Rohrer, G.S. Photocatalysts with internal electric fields. Nanoscale 2014, 6, 24–42. [Google Scholar] [CrossRef]
  39. Sergio, O.; Zhang, Y.F.; Gerardo, C. Cascade charge separation mechanism by ternary heterostructured BiPO4/TiO2/g-C3N4 photocatalyst. Appl. Catal. B Environ. 2016, 184, 96–103. [Google Scholar]
  40. He, Y.; Li, J.; Sheng, J.; Chen, S.; Dong, F.; Sun, Y. Crystal-Structure Dependent Reaction Pathways in Photocatalytic Formaldehyde Mineralization on BiPO4. J. Hazard. Mater. 2021, 420, 126633. [Google Scholar] [CrossRef]
  41. Zhu, X.; Wang, Y.; Guo, Y.; Sun, C. Environmental-friendly synthesis of heterojunction photocatalysts g-C3N4/BiPO4 with enhanced photocatalytic performance. Appl. Surf. Sci. 2020, 544, 148872. [Google Scholar] [CrossRef]
  42. Nyankson, E.; Yeboah, N.; Jnr, S.O.; Onaja, S.; Mensah, T.; Efavi, J.K. The effect of synthesis route on the photocatalytic performance of Ag-TiO using rhodamine b dyes, pesticides, and pharmaceutical waste as model pollutants. Mater. Res. Express 2022, 9, 094001. [Google Scholar] [CrossRef]
  43. Wang, Y.; Ding, K.; Xu, R.; Yu, D.; Liu, B. Fabrication of BiVO4/BiPO4/GO composite photocatalytic material for the visible light-driven degradation. J. Clean. Prod. 2020, 247, 119108. [Google Scholar] [CrossRef]
  44. Ahmaruzzaman, M.; Mishra, S.R. Photocatalytic performance of g-C3N4 based nanocomposites for effective degradation/removal of dyes from water and wastewater. Mater. Res. Bull. 2021, 143, 111417. [Google Scholar] [CrossRef]
Figure 1. Effect of Mass Fractions of TiO2 and MnCl2 in Catalysts on the Photocatalytic Degradation Efficiency of TC-HCl ((a) Ti-BPC; (b) Mn-BPC).
Figure 1. Effect of Mass Fractions of TiO2 and MnCl2 in Catalysts on the Photocatalytic Degradation Efficiency of TC-HCl ((a) Ti-BPC; (b) Mn-BPC).
Catalysts 13 01398 g001
Figure 2. HRTEM scan of Mn-BPC and Ti-BPC composite catalysts ((ad) Ti-BPC; (eh) Mn-BPC).
Figure 2. HRTEM scan of Mn-BPC and Ti-BPC composite catalysts ((ad) Ti-BPC; (eh) Mn-BPC).
Catalysts 13 01398 g002
Figure 3. Full spectrum of X-ray photoelectron energy of Ti-BPC and Mn-BPC.
Figure 3. Full spectrum of X-ray photoelectron energy of Ti-BPC and Mn-BPC.
Catalysts 13 01398 g003
Figure 4. X-ray photoelectron spectra of Ti-BPC and Mn-BPC elements ((af) Ti-BPC; (gl) Mn-BPC).
Figure 4. X-ray photoelectron spectra of Ti-BPC and Mn-BPC elements ((af) Ti-BPC; (gl) Mn-BPC).
Catalysts 13 01398 g004
Figure 5. X-ray analysis diagram of Ti-BPC and Mn-BPC.
Figure 5. X-ray analysis diagram of Ti-BPC and Mn-BPC.
Catalysts 13 01398 g005
Figure 6. FTIR analysis of Ti-BPC and Mn-BPC.
Figure 6. FTIR analysis of Ti-BPC and Mn-BPC.
Catalysts 13 01398 g006
Figure 7. Effects of different influencing factors on the photocatalytic degradation efficiency of TC-HCl by Ti-BPC and Mn-BPC ((a) effect of pH on Ti-BPC, (b) effect of pH on Mn-BPC; (c) influence of dosage on Ti-BPC; (d) influence of dosage on Mn-BPC; (e) influence of initial concentration of tetracycline on Ti-BPC; (f) influence of initial concentration of tetracycline on Mn-BPC).
Figure 7. Effects of different influencing factors on the photocatalytic degradation efficiency of TC-HCl by Ti-BPC and Mn-BPC ((a) effect of pH on Ti-BPC, (b) effect of pH on Mn-BPC; (c) influence of dosage on Ti-BPC; (d) influence of dosage on Mn-BPC; (e) influence of initial concentration of tetracycline on Ti-BPC; (f) influence of initial concentration of tetracycline on Mn-BPC).
Catalysts 13 01398 g007
Figure 8. Pseudo-first-order kinetics fitting of photocatalytic degradation experiments of Ti-BPC and Mn-BPC ((a) Ti-BPC; (b) Mn-BPC) and pseudo- second-order kinetics fitting ((c) Ti-BPC; (d) Mn-BPC).
Figure 8. Pseudo-first-order kinetics fitting of photocatalytic degradation experiments of Ti-BPC and Mn-BPC ((a) Ti-BPC; (b) Mn-BPC) and pseudo- second-order kinetics fitting ((c) Ti-BPC; (d) Mn-BPC).
Catalysts 13 01398 g008
Figure 9. Capture experiment of Ti-BPC and Mn-BPC composite catalysts ((a) Ti-BPC; (b) Mn-BPC).
Figure 9. Capture experiment of Ti-BPC and Mn-BPC composite catalysts ((a) Ti-BPC; (b) Mn-BPC).
Catalysts 13 01398 g009
Figure 10. Schematic diagram of degradation mechanism ((a) Ti-BPC; (b) Mn-BPC).
Figure 10. Schematic diagram of degradation mechanism ((a) Ti-BPC; (b) Mn-BPC).
Catalysts 13 01398 g010
Table 1. XPS peak information of Ti-BPC and Mn-BPC catalysts.
Table 1. XPS peak information of Ti-BPC and Mn-BPC catalysts.
TiO2-BPCMn-BPC
NameAtomic %Atomic %
C1s23.8932.41
N1s5.4412.74
O1s43.1236.31
P2p7.178.96
Bi4f4.256.69
Ti2p16.12/
Mn2p/2.89
Table 2. Kinetic fitting results of Ti-BPC photocatalytic degradation experiment.
Table 2. Kinetic fitting results of Ti-BPC photocatalytic degradation experiment.
Initial Concentration of Tetracycline Hydrochloride/(mg/L)Pseudo-First Order Kinetic EquationkR2Pseudo-Second Order Kinetic EquationkR2
4y = 0.226x + 1.4780.004630.98554y = 0.00991x − 0.158330.009910.99878
6y = 0.00397x + 0.0840.003970.9925y = 0.00407x − 0.107780.004070.97465
8y = 0.0027x + 0.1020.00270.98722y = 0.00172x − 0.019730.001720.98407
10y = 0.00217x + 0.0640.002170.95762y = 0.00102x − 0.012350.001020.97955
12y = 0.00187 + 0.0480.001870.98616y = 0.000657x − 0.009950.0006570.98529
Table 3. Kinetic fitting results of Mn-BPC photocatalytic degradation experiment.
Table 3. Kinetic fitting results of Mn-BPC photocatalytic degradation experiment.
Initial Concentration of Tetracycline Hydrochloride/(mg/L)Pseudo-First Order Kinetic EquationkR2Pseudo-Second Order Kinetic EquationkR2
3y = 0.0080x + 0.69550.00800.9237y = 0.0156x − 0.15030.01560.8248
4y = 0.0087x + 0.58690.00870.9814y = 0.0096x − 0.05910.00960.9779
5y = 0.0070x + 0.40420.00700.9626y = 0.0048x + 0.01370.00480.9904
7y = 0.0041x + 0.44350.00410.9349y = 0.0014x + 0.06360.00140.9616
9y = 0.0032x + 0.29840.00320.9260y = 0.0006x + 0.03350.00060.9549
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

Qian, W.; Fang, Y.; Liu, H.; Deng, Y.; Li, Y.; Zhang, Y.; Diao, Z.; Li, M. Photocatalytic Degradation of Tetracycline Hydrochloride by Mn/g-C3N4/BiPO4 and Ti/g-C3N4/BiPO4 Composites: Reactivity and Mechanism. Catalysts 2023, 13, 1398. https://doi.org/10.3390/catal13111398

AMA Style

Qian W, Fang Y, Liu H, Deng Y, Li Y, Zhang Y, Diao Z, Li M. Photocatalytic Degradation of Tetracycline Hydrochloride by Mn/g-C3N4/BiPO4 and Ti/g-C3N4/BiPO4 Composites: Reactivity and Mechanism. Catalysts. 2023; 13(11):1398. https://doi.org/10.3390/catal13111398

Chicago/Turabian Style

Qian, Wei, Yi Fang, Hui Liu, Yili Deng, Yingying Li, Yongzheng Zhang, Zenghui Diao, and Mingyu Li. 2023. "Photocatalytic Degradation of Tetracycline Hydrochloride by Mn/g-C3N4/BiPO4 and Ti/g-C3N4/BiPO4 Composites: Reactivity and Mechanism" Catalysts 13, no. 11: 1398. https://doi.org/10.3390/catal13111398

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