Next Article in Journal
Visualization of Phototherapy Evolution by Optical Imaging
Previous Article in Journal
Structural Modeling of Nanobodies: A Benchmark of State-of-the-Art Artificial Intelligence Programs
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Rational Design of Monolithic g-C3N4 with Floating Network Porous-like Sponge Monolithic Structure for Boosting Photocatalytic Degradation of Tetracycline under Simulated and Natural Sunlight Illumination

1
School of Water Resource and Environment, Hebei Province Key Laboratory of Sustained Utilization & Development of Water Recourse, Hebei Province Collaborative Innovation Center for Sustainable Utilization of Water Resources and Optimization of Industrial Structure, Hebei Geo University, Shijiazhuang 050031, China
2
School of Material Science and Engineering, Jiangsu University of Science and Technology, Zhenjiang 212003, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(10), 3989; https://doi.org/10.3390/molecules28103989
Submission received: 26 April 2023 / Revised: 6 May 2023 / Accepted: 8 May 2023 / Published: 9 May 2023
(This article belongs to the Section Green Chemistry)

Abstract

:
In order to solve the problems of powder g-C3N4 catalysts being difficult to recycle and prone to secondary pollution, floating network porous-like sponge monolithic structure g-C3N4 (FSCN) was prepared with a one-step thermal condensation method using melamine sponge, urea, and melamine as raw materials. The phase composition, morphology, size, and chemical elements of the FSCN were studied using XRD, SEM, XPS, and UV–visible spectrophotometry. Under simulated sunlight, the removal rate for 40 mg·L−1 tetracycline (TC) by FSCN reached 76%, which was 1.2 times that of powder g-C3N4. Under natural sunlight illumination, the TC removal rate of FSCN was 70.4%, which was only 5.6% lower than that of a xenon lamp. In addition, after three repeated uses, the removal rates of the FSCN and powder g-C3N4 samples decreased by 1.7% and 2.9%, respectively, indicating that FSCN had better stability and reusability. The excellent photocatalytic activity of FSCN benefits from its three-dimensional-network sponge-like structure and outstanding light absorption properties. Finally, a possible degradation mechanism for the FSCN photocatalyst was proposed. This photocatalyst can be used as a floating catalyst for the treatment of antibiotics and other types of water pollution, providing ideas for the photocatalytic degradation of pollutants in practical applications.

1. Introduction

As a typical antibiotic, tetracycline (TC) ranks second in production and usage globally and has been widely used in medicine and industry. However, TC has a complex structure and is a difficult-to-degrade organic pollutant that can easily accumulate in the environment. Moreover, TC has issues such as ecotoxicity and poor biodegradability, and residual TC in the environment may also increase microbial resistance, posing a serious threat to ecosystems and human health. Therefore, it is necessary to remove TC from the water environment [1,2,3]. The current mainstream methods for water purification include chemical air flotation, advanced oxidation, photocatalytic degradation, adsorption, and microbial treatment [4,5,6]. However, traditional treatment methods, such as chemical, physical, and biological methods, have some limitations, such as easily causing secondary pollution, incomplete degradation, and toxicity for microorganisms [7,8,9]. Photocatalytic technology is an advanced oxidation technology with low energy consumption, high reaction efficiency, and no secondary pollution [10,11,12,13]. Photocatalytic technology utilizes clean solar energy as the driving force to generate free radical active substances, such as superoxide radicals (·O2) and hydroxyl radicals (·OH), by stimulating photo-generated carriers with strong redox ability in semiconductor materials [14,15,16], which thereby react with different pollutants and, ultimately, generate small molecules, such as CO2 and H2O, resulting in the removal of pollutants [17,18,19,20]. Therefore, photocatalytic technology provides a promising method for the removal of TC in aquatic environments.
Among the many photocatalytic materials, carbon nitride (g-C3N4) is favored by many researchers due to its advantages of being metal-free, easy to prepare, and low cost and having high chemical stability; it is thus widely used in research directions such as hydrogen production and pollutant degradation [21,22,23,24,25]. However, due to the limitations of the material properties, g-C3N4 powders are affected by defects such as easy aggregation, a low light response range, and a high photo-generated electron–hole recombination rate [26,27,28,29]. In addition, when added to polluted wastewater, g-C3N4 powders sink to the bottom of the water environment, making recovery difficult and resulting in the powders not being fully utilized [30,31,32]. Fortunately, preparing floating photocatalysts can address these issues, as they can fully utilize light energy and are easy to recover and reuse [33]. For example, a ZnFe2O4@SiO2@TiO2 composite floater prepared according to the sol–gel approach was used for dyestuff degradation by Meng et al. [34]. Under visible light conditions, the removal efficiency could reach 95.1% within 2 h. However, the abovementioned floating photocatalysts are rather complicated to prepare, so there is an urgent need to find a simple method for preparing floating photocatalysts [35]. The three-dimensional (3D) structure of melamine sponge (MS) gives it the advantages of high specific surface area, high nitrogen content, and high elasticity, making it an ideal carrier for preparing floating photocatalytic materials [36,37]. In addition, the carbonization of MS can form a carbon skeleton with a porous mesh structure, which provides sufficient attachment sites for the loading of other materials and effectively enhances the adsorption and catalytic performance of MS. For example, the ZIF-8/carbon-nitrogen foam synthesized by Daeok Kim et al. [38] was used for stratified pore oil capture and chemical fixation of CO2, and the ZIF-8/CN foam was found to reject water while exhibiting a very high affinity for oil.
In this study, g-C3N4 was successfully attached to the surface of MS using a one-step thermal shrinkage method, and an FSCN photocatalyst was prepared and applied to the photocatalytic degradation of TC. The floating network porous-like sponge monolithic structure of FSCN facilitates recycling and makes it possible to avoid the occurrence of secondary pollution. The effects of different dosages and reaction conditions on the degradation performance of FSCN were studied. In addition, the morphology, degradation, and optical properties of the material were studied through various characterization techniques, and the mechanism of FSCN’s photocatalytic degradation of TC was analyzed. The stability of FSCN was verified through cyclic experiments and tested, demonstrating the possibility of its practical application.

2. Results and Discussion

2.1. XRD Analysis

Figure 1a shows the X-ray diffraction (XRD) results for the MS, FSCN, and g-C3N4 powders to illustrate each sample’s crystal structure and phase composition. It can be seen from the figure that the characteristic diffraction peaks of FSCN (37.5°, 43.9°, 64.3°, and 77.3°) [39] were similar to those of the MS, and the peak intensities had some subtle changes, which may have been due to the evolution of the morphology of melamine after calcination at high temperature. In accordance with g-C3N4 powders (JCPDS 87-1526) [40,41], the XRD pattern for FSCN showed a characteristic diffraction peak at 26.7° (002), attributed to the layer stacking of the in-planar repeating units of the continuous heptazine skeleton and the conjugated aromatic structure with a spacing of 0.32 nm [42,43]. The diffraction peak of FSCN was weaker, indicating that the interlayer period relevance length of the tri-s-triazine building block was reduced [44,45]. In summary, FSCN exhibited the characteristics of both MS and g-C3N4, proving that g-C3N4 successfully adhered to the MS with a network porous-like structure.
In order to verify the XRD results and further analyze the composition of FSCN, Fourier-transform infrared spectroscopy (FT-IR) was used to analyze the FSCN and g-C3N4 powders, as shown in Figure 1b. A comparison showed that the FSCN and g-C3N4 powders were basically the same. However, there were some differences in peak strength, which may have been due to the morphological evolution of melamine after high-temperature calcination of FSCN. There was a clear characteristic peak at 500 cm−1–1000 cm−1 found in both materials, which is a typical vibration peak of the tri-s-triazine ring of g-C3N4, once again proving the successful combination of g-C3N4 with MS [46]. Moreover, the FSCN and g-C3N4 powders exhibited a series of dense elastic vibration peaks in the range from 1000 cm−1 to 1700 cm−1, which can be attributed to the C-N and C=N vibration peaks of g-C3N4 [47]. The characteristic peaks of FSCN at 3000 cm−1–3500 cm−1 showed a certain deviation from the findings for the g-C3N4 powders, which may have been influenced by the MS. The characteristic peaks here can be attributed to the N-H and O-H of g-C3N4 [48]. The dense fluctuations at 3500 cm−1–4000 cm−1 may have been caused by potassium bromide doping during the testing process, which did not affect the analysis of the results [49]. In summary, FSCN exhibited all the characteristic peaks of the g-C3N4 material, indicating that g-C3N4 successfully combined with MS to form FSCN with the advantages of both materials, which was also consistent with the XRD results.

2.2. Morphology

As shown in Figure 2a, MS exhibited a simple network of pore-like structures with a very high specific surface area. These structures provide the MS with abundant attachment points, making it an excellent carrier. Zooming-in to the 5 μm scale (Figure 2b), the MS can be seen as having a micron-rod structure with a diameter of 5 μm and a very smooth surface. After soaking with urea and melamine, the network porous-like structure of MS was not destroyed (Figure 2c), and it was also found that the surface of the MS became very rough, which further enhanced its material loading capacity, and urea and melamine particles were clearly visible on the MS (Figure 2d). After high-temperature calcination, it could be clearly seen that the FSCN had a hollow structure (Figure 2e), and the urea and melamine initially wrapped around the three-dimensional skeleton transformed into sheet-like g-C3N4, finally forming FSCN with a network porous-like structure, which was consistent with the XRD and FT-IR results. Moreover, the problem of the easy aggregation of the g-C3N4 monomer was solved. Additionally, as Figure 2f shows, the surface roughness of the material was further enhanced and, together with the flaky g-C3N4, this provided a greater specific surface area, which was more conducive to the adhesion of pollutants and improved the efficiency of contaminant treatment, providing a high reference value for the in-depth study of related materials.

2.3. XPS

As shown in Figure 3, X-ray photoelectron spectroscopy (XPS) was used to observe the surface chemical composition and elemental valence states of the FSCN samples. Figure 3a shows the XPS survey spectrum for FSCN, and it is evident that the FSCN was mainly composed of C and N elements since the primary substance in FSCN was g-C3N4. Figure 3b shows the high-resolution spectrum for C1s. After fitting, the C1s peak of the FSCN sample could be divided into three peaks at 287.5 eV, 286.0 eV, and 284.8 eV, respectively. The characteristic peak at 287.5 eV was formed due to N-C=N, the presence of C-N resulted in a characteristic peak at 286.0 eV, and the characteristic peak at 284.8 eV can be attributed to the existence of C=C [50]. The high-resolution spectrum for N1s is shown in Figure 3c, and the N1s of FSCN was fit to four peaks, with the peaks located at 398.3 eV, 399.1 eV, and 400.7 eV corresponding to C=N-C, N-(C)3, and N-H. The peak at 404.0 eV can be attributed to the excitation of π electrons in the C=N conjugated structure [51,52].

2.4. UV–Visible Spectroscopy

UV–visible diffuse reflectance spectroscopy was used to determine the optical absorption range and energy band gap of the FSCN and g-C3N4 powder samples. As shown in Figure 4a, both the FSCN and g-C3N4 powders had strong absorption capabilities in the UV and visible light regions. FSCN had a black carbonized structure that was more conducive to light absorption and a porous mesh structure that could reflect incident light multiple times, thereby improving the light utilization efficiency. This indicated that the black sponge-like monolithic structure further improved the utilization of light by FSCN, thus enhancing the photocatalytic activity. The forbidden bandwidths for the FSCN and g-C3N4 powders were calculated using the Tauc plot equation, and the results are shown in Figure 4b,c. The band gaps of the FSCN and g-C3N4 powders were 1.62 eV and 2.8 eV, respectively. The narrower band gap of FSCN compared to that of g-C3N4 powder gives it a better ability to utilize visible light, thus improving the photocatalytic performance. Moreover, these findings strongly agree with the previous SEM and XRD characterization results. Subsequently, the flat-band (FB) potentials and semiconductor types of the FSCN and g-C3N4 powders were examined using electrochemical Mott–Schottky analysis [53]. The linear plots of both the FSCN and g-C3N4 powders had positive slopes, which indicated that both the FSCN and g-C3N4 powders were n-type semiconductors. Remarkably, the FB potential was 0.1 V higher than the conduction-band (CB) potential of the n-type semiconductor [54]. The FB potential of the g-C3N4 powder was −0.65 V (−0.45 V vs. NHE), and the FB potential of FSCN was −0.5 V (−0.3 V vs. NHE) according to the Nernst formula: ENHE = EAg/AgCl + 0.197 [55]. Thus, the FSCN and g-C3N4 powders’ CBs were −0.4 eV and −0.55 eV, respectively. The VBs of the FSCN and g-C3N4 powders were 1.22 eV and 2.25 eV, respectively, according to the equation EVB = Eg + ECB.

2.5. Photocatalytic Activities

Figure 5 shows the photocatalytic activity of FSCN, calcination MS, and g-C3N4 powder on TC under xenon lamp irradiation. Figure 5a shows the effects of different FSCN dosages on the performance of degraded TC (40 mg/L). The figure shows that TC did not self-degrade under xenon lamp irradiation without the catalyst. The removal rates for TC achieved with 10 mg, 20 mg, and 30 mg of FSCN were 55%, 66.3%, and 76%, respectively. This indicated that the degradation rate for TC continuously increased with the increase in FSCN dosage and finally reached equilibrium. However, when the FSCN dosage was 40 mg, the removal rate for TC decreased by 5.3% because too much of the photocatalyst led to shielding and scattering of light, thus limiting the photocatalytic activity of the material and reducing the photocatalytic efficiency. Therefore, the optimal catalyst dosage for FSCN was 30 mg. In addition, to make the data more convincing, the photocatalytic activities of FSCN, MS, and g-C3N4 powder were tested for comparison. As shown in Figure 5b, the TC concentration in all three groups of samples gradually decreased with the increase in light time. Among them, the highest TC removal rate achieved with FSCN reached 76%, which was 11.3% higher than that of the g-C3N4 powder (64.7%) and 34.8% higher than that of the MS (41.2%). It is apparent from Figure 5c that the photocatalytic degradation data for all three samples pertained to first-order reaction kinetics. The kinetic constants K of the products were obtained by calculation, and the K values for FSCN (0.00967 min−1) were 1.28 and 2.51 times higher than those for g-C3N4 powder (0.00757 min−1) and calcination MS (0.00385 min−1), respectively. In order to evaluate the photocatalytic stability of the FSCN, three cycle experiments were performed. As shown in Figure 5d, the photocatalytic degradation of TC by FSCN, calcination MS, and g-C3N4 powder decreased by 1.9%, 1.7%, and 2.9%, respectively, after three cycle reactions, which indicated that the material was more stable after the addition of MS. In summary, the FSCN prepared by calcination with MS as the carrier showed an increased specific surface area and provided more active sites due to the sponge’s monolithic three-dimensional-network porous structure, thus effectively promoting the adsorption of TC and improving the photocatalytic degradation rate and photocatalytic activity. Moreover, compared with g-C3N4 powder, FSCN had a higher recycling rate and better photostability.
In order to simulate a practical application scenario, the photocatalytic activities of the FSCN and g-C3N4 powder on TC were tested under natural sunlight illumination. As shown in Figure 6a, the g-C3N4 powder was all deposited in the water. At the same time, the FSCN with a porous network structure could float on the water surface during the photocatalytic reaction, making it easy to recycle. Moreover, a set of blank experiments were set up as a control group to ensure the accuracy of the experimental results. Two sets of experiments were set up under sunny and cloudy conditions to investigate the effect of weather on the photocatalytic performance of the materials. First, in the reaction under sunny conditions, after 6 h, FSCN could remove 70.4% of the TC. The degradation of TC by g-C3N4 powder was only 48.3% (Figure 6b). The photocatalytic performances of the FSCN and g-C3N4 powders were significantly inhibited under cloudy conditions, with removal rates of 47.3% and 25.2%, respectively (Figure 6c). As mentioned above, FSCN maintains excellent photocatalytic activity under sunlight and can be recycled without secondary pollution.

2.6. Electrochemical Test

The electron–hole migration and separation efficiency of the FSCN and g-C3N4 powder samples were analyzed using the photocurrent response and electrochemical impedance under visible light. Figure 7a shows the instantaneous photocurrent responses of the FSCN and g-C3N4 powder samples. The current density of the g-C3N4 powder was significantly lower than that of FSCN, indicating that calcinating FSCN with MS as a carrier could effectively improve the photogenerated carrier separation efficiency of the material compared to g-C3N4 powder. Figure 7b displays the AC impedance spectra for FSCN and the g-C3N4 powder. The radius of curvature of FSCN was smaller than that of g-C3N4, indicating that the resistance of FSCN was lower than that of the g-C3N4 powder, resulting in better conductivity, more significant separation efficiency for photogenerated carriers, and superior photocatalytic performance.

2.7. Radical Trapping and ESR

Reactive radical trapping experiments and electron spin resonance (ESR) spectroscopy were used to verify the photocatalytic mechanism and explore the role of each reactive species during the reaction. VC, TBA, and TOEA were selected as trapping agents and added to the reaction system to trap ∙O2, ∙OH, and h+, respectively, and determine their contributions to the reaction. As shown in Figure 8a, the rate of degradation of TC decreased from 76% to 46.9% after the addition of ∙O2, which indicated that ∙O2 played a more important role in the reaction process. The rate of degradation of TC was most obviously inhibited after the addition of TOEA, decreasing to 34.8%, indicating that h+ was the most critical active factor in the degradation process. In contrast, the simple degradation rate for TC was only reduced by 15.4% after the addition of TBA, which indicated that ∙OH was involved in the reaction but its contribution was low. Notably, the production of ∙OH was inhibited in the process of h+ capture, further indicating the important role of h+ in the degradation process. To further verify the effects of the active factors on degradation, ESR analysis was performed under visible light. Figure 8b shows that no EPR signals for DMPO-∙O2 or DMPO-∙OH were observed under dark conditions. However, a series of distinctive characteristic peaks appeared under light conditions, which indicated that ∙O2 and ∙OH radicals were involved in the photocatalytic reaction and played key roles.

2.8. Photocatalytic Mechanism

Based on the above experiments and analysis, a possible photocatalytic reaction mechanism for the degradation of TC by FSCN under visible light irradiation can be proposed. As shown in Figure 9, under the irradiation of visible light, the electron (e) on the valence band (VB) of the FSCN leapfrogged above the conduction band (CB), leaving the hole (h+) on the VB. Since the CB potential of FSCN was more negative at −0.4 eV compared to −0.33 eV, the e clustered in the CB could react with O2 to form superoxide radicals (∙O2) with strong oxidation properties. Meanwhile, the potential on the VB of the FSCN was 1.22 eV, significantly lower than 1.99 eV, so the h+ on the VB was insufficient to react with H2O and OH to form hydroxyl radicals (∙OH). In summary, in the FSCN photocatalytic system, the main active species that react with TC in wastewater are ∙O2 and h+. The TC undergoes redox reactions with mineralization to produce small molecules, such as CO2 and H2O, which are eventually removed in the water.

3. Materials and Methods

3.1. Chemicals and Materials

Both urea and melamine were purchased from Chengdu Aikeda Chemical Reagent Co. (Chengdu, China). Melamine sponge was supplied by Tianjin Damao Chemical Reagent Factory (Tianjin, China), and tetracycline was purchased from Shanghai Baoman Biotechnology Co. (Shanghai, China). All experimental water was pure water, and all chemical substances used in the experiment were analytical grade and used without further purification.

3.2. Characterization

A D/MAX-2500VL/PC (Rigaku Co., Tokyo, Japan) was used for the X-ray diffraction (XRD) analysis of FSCN, and powder g-C3N4 photocatalysts were analyzed in the range from 20 to 80 in the 2θ. A S-4800 scanning electron microscope (SEM) (Hitachi Co., Tokyo, Japan) was used to characterize and analyze the morphology and size of the FSCN and powder g-C3N4 photocatalysts. An ESCALAB250Xi X-ray photoelectron spectrometer (XPS) (Thermo Co., Waltham, MA, USA) was used to characterize the chemical state and composition of the photocatalysts. A UV-2550 was used for diffuse reflectance spectroscopy (UV–visible spectroscopy) to analyze the optical absorption properties of the photocatalytic materials. A CHI660D electrochemical workstation (Chenhua Co., Ltd., Shanghai, China) was used to study the separation of the photogenerated carriers of the catalyst.

3.3. Synthesis of g-C3N4 Powder and FSCN

For the next step, 6 g of urea and 8 g of melamine were mixed thoroughly and added to a semi-open corundum crucible. The mixture was placed in a tube furnace and heated to 550 °C at a heating rate of 5 °C/min in a N2 environment and then calcined at a constant temperature for 4 h. After the reaction, the light yellow product was cooled to room temperature and ground to obtain g-C3N4 powder.
Then, 16 g of urea was weighed and dissolved into 40 mL of deionized water, after which 8 g of melamine was added and the mixture was stirred magnetically for 3 h at room temperature to obtain a saturated solution. The MS was cut to obtain a rectangle of 4 × 2 × 2 cm3, and its mass was recorded. The MS was immersed in the above solution. After that, the MS was taken out and dried in a freeze dryer for 24 h and then placed in a corundum crucible and heated to 550 °C in a tube furnace with a 5 °C/min heating rate in a N2 environment for 4 h of calcination. After cooling to room temperature, the synthesized product was washed with deionized water and ethanol alternately three times. Then, the samples were dried in a vacuum oven at 80 °C for 10 h and cooled naturally to obtain 2.2 × 1.5 × 1.5 cm3 of black floating network porous-like sponge monolithic structure g-C3N4, which was named FSCN. The preparation principle is shown in Figure 10.

3.4. Measurement of Photocatalytic Activity

3.4.1. Indoor Experiment

First, 30 mg of FSCN, g-C3N4 powder, and calcined MS were respectively weighed and added to 50 mL quartz test tubes. Afterward, 40 mg/L TC solution was added to the quartz test tubes as the target pollutant and a set of quartz test tubes with only TC solution were used as a blank control. Next, the quartz test tubes were placed in a photocatalytic reactor and stirred with magnetic force in a dark environment for 30 min to achieve a dark reaction and ensure that the FSCN, pure calcined sponge, and powdered g-C3N4 photocatalysts achieved adsorption–desorption equilibrium for TC and to reduce experimental errors. A 300 W xenon lamp was turned on as the light source for the photocatalytic reaction. Then, 3 mL of the solution was removed every 20 min and centrifuged for 5 min. After centrifugation, the supernatant was extracted. The wavelength of the UV–visible spectrophotometer was set to 357 nm, the absorbance of the supernatant was measured and recorded, and, ultimately, the degradation rates for TC achieved by the three photocatalysts were determined. The different masses of the FSCN were weighed, keeping the other conditions unchanged, and photocatalytic experiments were conducted to test the effect of dosage on photocatalytic performance.

3.4.2. Outdoor Experiment

First, 60 mg of FSCN and 60 mg of g-C3N4 powder were weighed and poured into beakers filled with 100 mL of TC solution (40 mg/L). Next, a 30 min dark reaction was conducted to ensure adsorption equilibrium. Then, the samples were transferred to a natural sunlight-irradiated environment with a light intensity of 58.28 mW/cm2 for the photocatalytic reaction. The blank control group comprised 100 mL of 40 mg/L TC without a catalyst. In the next step, 3 mL of the solution was removed every hour and centrifuged, the supernatant was extracted, and the TC absorbance was measured and recorded with an ultraviolet–visible spectrophotometer at a wavelength of 357 nm. In addition, the sunny natural environment was replaced with a cloudy natural environment, with other experimental conditions kept the same as above, and the TC absorbance was measured and recorded.

3.5. Reactive Radical Trapping Experiments

Similar to the above photocatalysis experiment, the original experimental process was kept unchanged. Before the reaction, 1 mmol vitamin C (VC), thiobarbituric acid (TBA), and triethanolamine (TEOA) captors were added to the TC solution to explore the contributions of superoxide anion radicals (O2), hydroxyl radicals (OH), and holes (h+) in the experiment on the process of photocatalysis.

4. Conclusions

This study successfully prepared g-C3N4 (FSCN) photocatalysts with a floating network porous-like sponge monolithic structure through a one-step thermal shrinkage method. The experimental results indicated that FSCN exhibited excellent photocatalytic performance and stability for TC degradation with both laboratory light sources and sunlight exposure. After a series of characterization analyses, it can be concluded that the factors affecting the photocatalytic performance of FSCN are the following: (1) the addition of MS effectively solved the problem of easy aggregation of g-C3N4, and the increase in the specific surface area provided more attachment sites for TC; (2) the black structure after MS carbonization was more conducive to light absorption, and the complex porous structure increased the refraction of light in the material and improved light utilization efficiency, thus improving photocatalytic performance. In addition, FSCN also has the advantages of being low cost and easy to prepare, allowing easy recovery, and producing no secondary pollution. This study provides new ideas for the design and preparation of recyclable photocatalysts that can help to reduce the harm from antibiotic pollution on the ecological environment and human health.

Author Contributions

D.C.: Conceptualization, Data curation, Funding acquisition, Writing—original draft. X.W.: Investigation, Formal analysis. H.Z. and D.Y.: Conceptualization, Investigation, Data curation. Z.Y.: Formal analysis, Investigation. Z.L.: Validation, Data curation, Formal analysis. C.L. and F.G.: Conceptualization, Methodology, Supervision, Writing—review and editing, Funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (21906039, 21906072), Hebei Province 333 Talents Project (A202101020), the Science and Technology Project of Hebei Education Department (BJ2021010, QN2021029), the Graduate Student Innovation Ability Training Funding Project of Hebei Province (CXZZSS2023129), the Open Fund for Hebei Province Key Laboratory of Sustained Utilization and Development of Water Recourse (HSZYL2022002), and the Hebei Geo University Student Science and Technology Fund (KAG202306).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data from the study can be provided by the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Xie, Z.J.; Feng, Y.P.; Wang, F.L.; Chen, D.N.; Zhang, Q.X.; Zeng, Y.Q.; Lv, W.Y.; Liu, G.G. Construction of carbon dots modified MoO3/g-C3N4 Z-scheme photocatalyst with enhanced visible-light photocatalytic activity for the degradation of tetracycline. Appl. Catal. B Environ. 2018, 229, 96–104. [Google Scholar] [CrossRef]
  2. Sun, H.R.; Guo, F.; Pan, J.J.; Huang, W.; Wang, K.; Shi, W.L. One-pot thermal polymerization route to prepare N-deficient modified g-C3N4 for the degradation of sstetracycline by the synergistic effect of photocatalysis and persulfate-based advanced oxidation process. Chem. Eng. J. 2021, 406, 126844. [Google Scholar] [CrossRef]
  3. Wang, D.B.; Jia, F.Y.; Wang, H.; Chen, F.; Fang, Y.; Dong, W.B.; Zeng, G.M.; Li, X.M.; Yang, Q.; Yuan, X.Z. Simultaneously efficient adsorption and photocatalytic degradation of tetracycline by Fe-based MOFs. J. Colloid Interface Sci. 2018, 519, 273–284. [Google Scholar] [CrossRef] [PubMed]
  4. Guo, F.; Li, M.Y.; Renb, H.J.; Huang, X.L.; Hou, W.X.; Wang, C.; Shi, W.L.; Lu, C.Y. Fabrication of p-n CuBi2O4/MoS2 heterojunction with nanosheets-on-microrods structure for enhanced photocatalytic activity towards tetracycline degradation. Appl. Surf. Sci. 2019, 491, 88–94. [Google Scholar] [CrossRef]
  5. Guo, F.; Sun, H.R.; Huang, X.L.; Shi, W.L.; Yan, C. Fabrication of TiO2/high-crystalline g-C(3)N(4)composite with enhanced visible-light photocatalytic performance for tetracycline degradation. J. Chem. Technol. Biotechnol. 2020, 95, 2684–2693. [Google Scholar]
  6. Lu, C.Y.; Wang, J.; Cao, D.L.; Guo, F.; Hao, X.L.; Li, D.; Shi, W.L. Synthesis of magnetically recyclable g-C3N4/NiFe2O4 S-scheme heterojunction photocatalyst with promoted visible-light-response photo-Fenton degradation of tetracycline. Mater. Res. Bull. 2023, 158, 112064. [Google Scholar] [CrossRef]
  7. Lu, C.Y.; Wang, L.T.; Yang, D.Q.; Jin, Z.H.; Wang, X.; Xu, J.M.; Li, Z.D.; Shi, W.L.; Guan, W.S.; Huang, W. Boosted tetracycline and Cr(VI) simultaneous cleanup over Z-Scheme BiPO4/CuBi2O4 p-n heterojunction with 0D/1D trepang-like structure under simulated sunlight irradiation. J. Alloys Compd. 2022, 919, 165849. [Google Scholar] [CrossRef]
  8. Zhao, D.M.; Wang, Y.Q.; Dong, C.L.; Huang, Y.C.; Chen, J.; Xue, F.; Shen, S.H.; Guo, L.J. Boron-doped nitrogen-deficient carbon nitride-based Z-scheme heterostructures for photocatalytic overall water splitting. Nat. Energy 2021, 6, 388–397. [Google Scholar] [CrossRef]
  9. He, F.; Meng, A.Y.; Cheng, B.; Ho, W.K.; Yu, J.G. Enhanced photocatalytic H-2-production activity of WO3/TiO2 step-scheme heterojunction by graphene modification. Chin. J. Catal. 2020, 41, 9–20. [Google Scholar] [CrossRef]
  10. Lu, C.Y.; Yang, D.Q.; Wang, L.T.; Wen, S.J.; Cao, D.L.; Tu, C.Q.; Gao, L.N.; Li, Y.L.; Zhou, Y.H.; Huang, W. Facile construction of CoO/Bi2WO6 p-n heterojunction with following Z-Scheme pathways for simultaneous elimination of tetracycline and Cr(VI) under visible light irradiation. J. Alloys Compd. 2022, 904, 164046. [Google Scholar] [CrossRef]
  11. Li, S.J.; Chen, J.L.; Hu, S.W.; Wang, H.L.; Jiang, W.; Chen, X.B. Facile construction of novel Bi2WO6/Ta3N5 Z-scheme heterojunction nanofibers for efficient degradation of harmful pharmaceutical pollutants. Chem. Eng. J. 2020, 402, 126165. [Google Scholar] [CrossRef]
  12. Lu, C.Y.; Guo, F.; Yan, Q.Z.; Zhang, Z.J.; Li, D.; Wang, L.P.; Zhou, Y.H. Hydrothermal synthesis of type II ZnIn2S4/BiPO4 heterojunction photocatalyst with dandelion-like microflower structure for enhanced photocatalytic degradation of tetracycline under simulated solar light. J. Alloys Compd. 2019, 811, 151976. [Google Scholar] [CrossRef]
  13. Zhang, F.Z.; Chen, J.; Wallace, G.G.; Yang, J.P. Engineering electrocatalytic fiber architectures. Prog. Mater. Sci. 2023, 133, 101069. [Google Scholar] [CrossRef]
  14. Liang, S.J.; Zhou, Z.M.; Wu, X.W.; Zhu, S.Y.; Bi, J.H.; Zhou, L.M.; Liu, M.H.; Wu, L. Constructing a MoS2 QDs/CdS Core/Shell Flower-like Nanosphere Hierarchical Heterostructure for the Enhanced Stability and Photocatalytic Activity. Molecules 2016, 21, 213. [Google Scholar] [CrossRef]
  15. Bai, J.X.; Shen, R.C.; Jiang, Z.M.; Zhang, P.; Li, Y.J.; Li, X. Integration of 2D layered CdS/WO3 S-scheme heterojunctions and metallic Ti3C2 MXene-based Ohmic junctions for effective photocatalytic H-2 generation. Chin. J. Catal. 2022, 43, 359–369. [Google Scholar] [CrossRef]
  16. Wu, X.F.; Fang, S.; Zheng, Y.; Sun, J.; Lv, K.L. Thiourea-Modified TiO2 Nanorods with Enhanced Photocatalytic Activity. Molecules 2016, 21, 181. [Google Scholar] [CrossRef]
  17. Bose, R.; Manna, G.; Jana, S.; Pradhan, N. Ag2S-AgInS2: P-n junction heteronanostructures with quasi type-II band alignment. Chem. Commun. 2014, 50, 3074–3077. [Google Scholar] [CrossRef]
  18. Nishad, K.K.; Joseph, J.; Tiwari, N.; Kurchania, R.; Pandey, R.K. Investigation on Size Dependent Elemental Binding Energies and Structural Properties of ZnO Nanoparticles and Their Correlation with Observed Photo-Luminescence Behavior. Sci. Adv. Mater. 2015, 7, 1368–1378. [Google Scholar] [CrossRef]
  19. Elshypany, R.; Selim, H.; Zakaria, K.; Moustafa, A.H.; Sadeek, S.A.; Sharaa, S.I.; Raynaud, P.; Nada, A.A. Magnetic ZnO Crystal Nanoparticle Growth on Reduced Graphene Oxide for Enhanced Photocatalytic Performance under Visible Light Irradiation. Molecules 2021, 26, 2269. [Google Scholar] [CrossRef]
  20. Tantraviwat, D.; Nattestad, A.; Chen, J.; Inceesungvorn, B. Enhanced photoactivity and selectivity over BiOI-decorated Bi2WO6 microflower for selective oxidation of benzylamine: Role of BiOI and mechanism. J. Colloid Interface Sci. 2023, 629, 854–863. [Google Scholar] [CrossRef]
  21. Guo, F.; Huang, X.L.; Chen, Z.H.; Cao, L.W.; Cheng, X.F.; Chen, L.Z.; Shi, W.L. Construction of Cu3P-ZnSnO3-g-C3N4 p-n-n heterojunction with multiple built-in electric fields for effectively boosting visible-light photocatalytic degradation of broad-spectrum antibiotics. Sep. Purif. Technol. 2021, 265, 118477. [Google Scholar] [CrossRef]
  22. Bavani, T.; Madhavan, J.; Preeyanghaa, M.; Neppolian, B.; Murugesan, S. Construction of direct Z-scheme g-C3N4/Bi2WO6 heterojunction photocatalyst with enhanced visible light activity towards the degradation of methylene blue. Environ. Sci. Pollut. Res. 2023, 30, 10179–10190. [Google Scholar] [CrossRef]
  23. Kumar, A.; Sharma, G.; Kumari, A.; Guo, C.S.; Naushad, M.; Vo, D.V.N.; Iqbal, J.; Stadler, F.J. Construction of dual Z-scheme g-C3N4/Bi4Ti3O12/Bi4O5I2 heterojunction for visible and solar powered coupled photocatalytic antibiotic degradation and hydrogen production: Boosting via I/I3 and Bi3+/Bi5+ redox mediators. Appl. Catal. B Environ. 2021, 284, 119808. [Google Scholar] [CrossRef]
  24. Guo, F.; Shi, W.L.; Zhu, C.; Li, H.; Kang, Z.H. CoO and g-C3N4 complement each other for highly efficient overall water splitting under visible light. Appl. Catal. B Environ. 2018, 226, 412–420. [Google Scholar] [CrossRef]
  25. Wei, X.Q.; Wang, X.; Pu, Y.; Liu, A.N.; Chen, C.; Zou, W.X.; Zheng, Y.L.; Huang, J.S.; Zhang, Y.; Yang, Y.C.; et al. Facile ball-milling synthesis of CeO2/g-C3N4 Z-scheme heterojunction for synergistic adsorption and photodegradation of methylene blue: Characteristics, kinetics, models, and mechanisms. Chem. Eng. J. 2021, 420, 127719. [Google Scholar] [CrossRef]
  26. Szyman, J. Multiple Steady States in the Photocatalytic Reactor for Colored Compounds Degradation. Molecules 2021, 26, 3804. [Google Scholar] [CrossRef]
  27. Duan, S.F.; Tao, C.L.; Geng, Y.Y.; Yao, X.Q.; Kang, X.W.; Su, J.Z.; Rodriguez-Gutierrez, I.; Kan, M.; Romero, M.; Sun, Y.; et al. Phosphorus-doped Isotype g-C3N4/g-C3N4: An Efficient Charge Transfer System for Photoelectrochemical Water Oxidation. ChemCatChem 2019, 11, 729–736. [Google Scholar] [CrossRef]
  28. Ho, W.C.J.; Tay, Q.L.; Qi, H.; Huang, Z.H.; Li, J.; Chen, Z. Photocatalytic and Adsorption Performances of Faceted Cuprous Oxide (Cu2O) Particles for the Removal of Methyl Orange (MO) from Aqueous Media. Molecules 2017, 22, 677. [Google Scholar] [CrossRef]
  29. Bharagav, U.; Reddy, N.R.; Rao, V.N.K.; Ravi, P.; Sathish, M.; Rangappa, D.; Prathap, K.; Chakra, C.S.; Shankar, M.V.; Appels, L.; et al. Bifunctional g-C3N4/carbon nanotubes/WO3 ternary nanohybrids for photocatalytic energy and environmental applications. Chemosphere 2023, 311, 137030. [Google Scholar] [CrossRef]
  30. Kavil, J.; Anjana, P.M.; Joshy, D.; Babu, A.; Raj, G.; Periyat, P.; Rakhi, R.B. g-C3N4/CuO and g-C3N4/Co3O4 nanohybrid structures as efficient electrode materials in symmetric supercapacitors. RSC Adv. 2019, 9, 38430–38437. [Google Scholar] [CrossRef]
  31. Zhou, X.; Zhao, C.H.; Chen, J.H.; Chen, L.Y. Influence of B, Zn, and B-Zn doping on electronic structure and optical properties of g-C3N4 photocatalyst: A first-principles study. Results Phys. 2021, 26, 104338. [Google Scholar] [CrossRef]
  32. Karimi-Nazarabad, M.; Ahmadzadeh, H.; Goharshadi, E.K. Porous perovskite-lanthanum cobaltite as an efficient cocatalyst in photoelectrocatalytic water oxidation by bismuth doped g-C3N4. Sol. Energy 2021, 227, 426–437. [Google Scholar] [CrossRef]
  33. Tang, J.L.; Wang, J.J.; Tang, L.; Feng, C.Y.; Zhu, X.; Yi, Y.Y.; Feng, H.P.; Yu, J.F.; Ren, X.Y. Preparation of floating porous g-C3N4 photocatalyst via a facile one-pot method for efficient photocatalytic elimination of tetracycline under visible light irradiation. Chem. Eng. J. 2022, 430, 132669. [Google Scholar] [CrossRef]
  34. Meng, X.F.; Zhuang, Y.; Tang, H.; Lu, C.H. Hierarchical structured ZnFe2O4@SiO2@TiO2 composite for enhanced visible-light photocatalytic activity. J. Alloys Compd. 2018, 761, 15–23. [Google Scholar] [CrossRef]
  35. Jiang, W.J.; Zhu, Y.F.; Zhu, G.X.; Zhang, Z.J.; Chen, X.J.; Yao, W.Q. Three-dimensional photocatalysts with a network structure. J. Mater. Chem. A 2017, 5, 5661–5679. [Google Scholar] [CrossRef]
  36. Sevastaki, M.; Suchea, M.P.; Kenanakis, G. 3D Printed Fully Recycled TiO2-Polystyrene Nanocomposite Photocatalysts for Use against Drug Residues. Nanomaterials 2020, 10, 2144. [Google Scholar] [CrossRef]
  37. Li, X.B.; Xiong, J.; Gao, X.M.; Huang, J.T.; Feng, Z.J.; Chen, Z.; Zhu, Y.F. Recent advances in 3D g-C3N4 composite photocatalysts for photocatalytic water splitting, degradation of pollutants and CO2 reduction. J. Alloys Compd. 2019, 802, 196–209. [Google Scholar] [CrossRef]
  38. Kim, D.; Kim, D.W.; Buyukcakir, O.; Kim, M.K.; Polychronopoulou, K.; Coskun, A. Highly Hydrophobic ZIF-8/Carbon Nitride Foam with Hierarchical Porosity for Oil Capture and Chemical Fixation of CO2. Adv. Funct. Mater. 2017, 27, 1700706. [Google Scholar] [CrossRef]
  39. Zhang, M.X.; Du, H.X.; Ji, J.; Li, F.F.; Lin, Y.C.; Qin, C.W.; Zhang, Z.; Shen, Y. Highly Efficient Ag3PO4/g-C3N4 Z-Scheme Photocatalyst for Its Enhanced Photocatalytic Performance in Degradation of Rhodamine B and Phenol. Molecules 2021, 26, 2062. [Google Scholar] [CrossRef]
  40. Kumaresan, N.; Sinthiya, M.M.A.; Kumar, M.P.; Ravichandran, S.; Babu, R.R.; Sethurman, K.; Ramamurthi, K. Investigation on the g-C3N4 encapsulated ZnO nanorods heterojunction coupled with GO for effective photocatalytic activity under visible light irradiation. Arabian J. Chem. 2020, 13, 2826–2843. [Google Scholar] [CrossRef]
  41. Chen, Y.; Li, A.; Fu, X.L.; Peng, Z.J. One-Step Calcination to Gain Exfoliated g-C3N4/MoO2 Composites for High-Performance Photocatalytic Hydrogen Evolution. Molecules 2022, 27, 7178. [Google Scholar] [CrossRef]
  42. Liu, Y.Y.; Wang, X.J.; Sun, Q.N.; Yuan, M.; Sun, Z.H.; Xia, S.Q.; Zhao, J.F. Enhanced visible light photo-Fenton-like degradation of tetracyclines by expanded perlite supported FeMo3Ox/g-C3N4 floating Z-scheme catalyst. J. Hazard. Mater. 2022, 424, 127387. [Google Scholar] [CrossRef]
  43. Yousefi, M.; Eshghi, H.; Karimi-Nazarabad, M.; Farhadipour, A. P5W30/g-C3N4 heterojunction thin film with improved photoelectrochemical performance for solar water splitting. New J. Chem. 2020, 44, 20470–20478. [Google Scholar] [CrossRef]
  44. Zheng, Y.; Lin, L.H.; Ye, X.J.; Guo, F.S.; Wang, X.C. Helical Graphitic Carbon Nitrides with Photocatalytic and Optical Activities. Angew. Chem. Int. Ed. 2014, 53, 11926–11930. [Google Scholar] [CrossRef]
  45. Mao, Z.Y.; Chen, J.J.; Yang, Y.F.; Wang, D.J.; Bie, L.J.; Fahlman, B.D. Novel g-C3N4/CoO Nanocomposites with Significantly Enhanced Visible-Light Photocatalytic Activity for H-2 Evolution. ACS Appl. Mater. Interface 2017, 9, 12427–12435. [Google Scholar] [CrossRef]
  46. Tan, S.Y.; Xing, Z.P.; Zhang, J.Q.; Li, Z.Z.; Wu, X.Y.; Cui, J.Y.; Kuang, J.Y.; Yin, J.W.; Zhou, W. Meso-g-C3N4/g-C3N4 nanosheets laminated homojunctions as efficient visible-light-driven photocatalysts. Int. J. Hydrogen Energy 2017, 42, 25969–25979. [Google Scholar] [CrossRef]
  47. Chen, Y.F.; Huang, W.X.; He, D.L.; Yue, S.T.; Huang, H. Construction of Heterostructured g-C3N4/Ag/TiO2 Microspheres with Enhanced Photocatalysis Performance under Visible-Light Irradiation. ACS Appl. Mater. Interfaces 2014, 6, 14405–14414. [Google Scholar] [CrossRef]
  48. Sun, J.W.; Yang, S.R.; Liang, Z.Q.; Liu, X.; Qiu, P.Y.; Cui, H.Z.; Tian, J. Two-dimensional/one-dimensional molybdenum sulfide (MoS2) nanoflake/graphitic carbon nitride (g-C3N4) hollow nanotube photocatalyst for enhanced photocatalytic hydrogen production activity. J. Colloid Interface Sci. 2020, 567, 300–307. [Google Scholar] [CrossRef]
  49. Qian, X.F.; Wu, Y.W.; Kan, M.; Fang, M.Y.; Yue, D.T.; Zeng, J.; Zhao, Y.X. FeOOH quantum dots coupled g-C3N4 for visible light driving photo- Fenton degradation of organic pollutants. Appl. Catal. B Environ. 2018, 237, 513–520. [Google Scholar] [CrossRef]
  50. Yang, Y.; Zhang, C.; Huang, D.L.; Zeng, G.M.; Huang, J.H.; Lai, C.; Zhou, C.Y.; Wang, W.J.; Guo, H.; Xue, W.J.; et al. Boron nitride quantum dots decorated ultrathin porous g-C3N4: Intensified exciton dissociation and charge transfer for promoting visible-light-driven molecular oxygen activation. Appl. Catal. B Environ. 2019, 245, 87–99. [Google Scholar] [CrossRef]
  51. Mishra, A.; Mehta, A.; Basu, S.; Shetti, N.P.; Reddy, K.R.; Aminabhavi, T.M. Graphitic carbon nitride (g-C3N4)-based metal-free photocatalysts for water splitting: A review. Carbon 2019, 149, 693–721. [Google Scholar] [CrossRef]
  52. Fu, J.W.; Xu, Q.L.; Low, J.X.; Jiang, C.J.; Yu, J.G. Ultrathin 2D/2D WO3/g-C3N4 step-scheme H-2-production photocatalyst. Appl. Catal. B Environ. 2019, 243, 556–565. [Google Scholar] [CrossRef]
  53. Kianipour, S.; Razavi, F.S.; Hajizadeh-Oghaz, M.; Abdulsahib, W.K.; Mahdi, M.A.; Jasim, L.S.; Salavati-Niasari, M. The synthesis of the P/N-type NdCoO3/g-C3N4 nano-heterojunction as a high-performance photocatalyst for the enhanced photocatalytic degradation of pollutants under visible-light irradiation. Arabian J. Chem. 2022, 15, 103840. [Google Scholar] [CrossRef]
  54. Li, S.J.; Wang, C.C.; Cai, M.J.; Yang, F.; Liu, Y.P.; Chen, J.L.; Zhang, P.; Li, X.; Chen, X.B. Facile fabrication of TaON/Bi2MoO6 core-shell S-scheme heterojunction nanofibers for boosting visible-light catalytic levofloxacin degradation and Cr(VI) reduction. Chem. Eng. J. 2022, 428, 131158. [Google Scholar] [CrossRef]
  55. Huang, L.Y.; Zhang, R.X.; Sun, X.J.; Cheng, X.N. Synthesis and Characterization of g-C3N4/Alpha-Fe2O3 Composites with Enhanced Photocatalytic Activity. In Proceedings of the 2nd International Congress on Advanced Materials (ICAM), Zhenjiang, China, 16–19 May 2013; Jiangsu University: Zhenjiang, China, 2014; pp. 225–228. [Google Scholar]
Figure 1. (a) XRD patterns for MS, FSCN, and g-C3N4 powders; (b) FT-IR patterns for FSCN and g-C3N4 powders.
Figure 1. (a) XRD patterns for MS, FSCN, and g-C3N4 powders; (b) FT-IR patterns for FSCN and g-C3N4 powders.
Molecules 28 03989 g001
Figure 2. Scanning electron microscope results: (a,b) MS, (c,d) sponge soaked in urea and melamine, and (eg) FSCN.
Figure 2. Scanning electron microscope results: (a,b) MS, (c,d) sponge soaked in urea and melamine, and (eg) FSCN.
Molecules 28 03989 g002
Figure 3. The XPS spectrum of the FSCN sample: (a) survey spectrum, (b) C1s, (c) N1s.
Figure 3. The XPS spectrum of the FSCN sample: (a) survey spectrum, (b) C1s, (c) N1s.
Molecules 28 03989 g003
Figure 4. (a) UV–visible spectroscopy results for FSCN and g-C3N4 powders. Mott–Schottky test results and Eg values: (b) FSCN, (c) g-C3N4 powder.
Figure 4. (a) UV–visible spectroscopy results for FSCN and g-C3N4 powders. Mott–Schottky test results and Eg values: (b) FSCN, (c) g-C3N4 powder.
Molecules 28 03989 g004
Figure 5. (a) Degradation of TC by FSCN; (b) degradation of TC by FSCN, g-C3N4 powder, and calcination MS; (c) first-order reaction kinetics of TC degradation with FSCN, g-C3N4 powder, and calcination MS; (d) cycle experiment results for FSCN, g-C3N4 powder, and calcination MS material.
Figure 5. (a) Degradation of TC by FSCN; (b) degradation of TC by FSCN, g-C3N4 powder, and calcination MS; (c) first-order reaction kinetics of TC degradation with FSCN, g-C3N4 powder, and calcination MS; (d) cycle experiment results for FSCN, g-C3N4 powder, and calcination MS material.
Molecules 28 03989 g005
Figure 6. (a) Photocatalytic removal of TC under natural light conditions. Photocatalytic removal performance of FSCN and g-C3N4 powder with TC: (b) sunny day (c) cloudy day.
Figure 6. (a) Photocatalytic removal of TC under natural light conditions. Photocatalytic removal performance of FSCN and g-C3N4 powder with TC: (b) sunny day (c) cloudy day.
Molecules 28 03989 g006
Figure 7. FSCN and g-C3N4 powder: (a) transient photocurrent diagram; (b) electrochemical impedance spectrum.
Figure 7. FSCN and g-C3N4 powder: (a) transient photocurrent diagram; (b) electrochemical impedance spectrum.
Molecules 28 03989 g007
Figure 8. (a) Active-species trapping experiments for visible photocatalytic degradation of TC by FSCN and (b) ESR spectra of FSCN under visible light and dark conditions.
Figure 8. (a) Active-species trapping experiments for visible photocatalytic degradation of TC by FSCN and (b) ESR spectra of FSCN under visible light and dark conditions.
Molecules 28 03989 g008
Figure 9. Diagram of the mechanism of TC degradation by FSCN photocatalytic material under visible light.
Figure 9. Diagram of the mechanism of TC degradation by FSCN photocatalytic material under visible light.
Molecules 28 03989 g009
Figure 10. Diagram of the preparation of FSCN photocatalytic materials.
Figure 10. Diagram of the preparation of FSCN photocatalytic materials.
Molecules 28 03989 g010
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

Cao, D.; Wang, X.; Zhang, H.; Yang, D.; Yin, Z.; Liu, Z.; Lu, C.; Guo, F. Rational Design of Monolithic g-C3N4 with Floating Network Porous-like Sponge Monolithic Structure for Boosting Photocatalytic Degradation of Tetracycline under Simulated and Natural Sunlight Illumination. Molecules 2023, 28, 3989. https://doi.org/10.3390/molecules28103989

AMA Style

Cao D, Wang X, Zhang H, Yang D, Yin Z, Liu Z, Lu C, Guo F. Rational Design of Monolithic g-C3N4 with Floating Network Porous-like Sponge Monolithic Structure for Boosting Photocatalytic Degradation of Tetracycline under Simulated and Natural Sunlight Illumination. Molecules. 2023; 28(10):3989. https://doi.org/10.3390/molecules28103989

Chicago/Turabian Style

Cao, Delu, Xueying Wang, Hefan Zhang, Daiqiong Yang, Ze Yin, Zhuo Liu, Changyu Lu, and Feng Guo. 2023. "Rational Design of Monolithic g-C3N4 with Floating Network Porous-like Sponge Monolithic Structure for Boosting Photocatalytic Degradation of Tetracycline under Simulated and Natural Sunlight Illumination" Molecules 28, no. 10: 3989. https://doi.org/10.3390/molecules28103989

Article Metrics

Back to TopTop