Next Article in Journal
Tetracyclic Bis-Piperidine Alkaloids: Structures, Bioinspired Synthesis, Synthesis, and Bioactivities
Previous Article in Journal
Towards Understanding the Basis of Brucella Antigen–Antibody Specificity
Previous Article in Special Issue
Efficient Degradation of Methylene Blue in Industrial Wastewater and High Cycling Stability of Nano ZnO
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Highly Efficient Photocatalytic Degradation of Tetracycline Antibiotics by BiPO4/g-C3N4: A Novel Heterojunction Nanocomposite with Nanorod/Stacked-like Nanosheets Structure

1
State Environmental Protection Key Laboratory of Soil Environmental Management and Pollution Control, Nanjing Institute of Environmental Sciences, Ministry of Ecology and Environment of China, Nanjing 210042, China
2
State Key Laboratory of Pollution Control and Resource Reuse, School of the Environment, Nanjing University, Nanjing 210023, China
*
Author to whom correspondence should be addressed.
Molecules 2025, 30(14), 2905; https://doi.org/10.3390/molecules30142905
Submission received: 1 December 2024 / Revised: 30 January 2025 / Accepted: 2 February 2025 / Published: 9 July 2025
(This article belongs to the Special Issue Advances in Photocatalytic Degradation of Organic Pollutants)

Abstract

The use of semiconductors for photocatalytic degradation of organic pollutants has garnered considerable attention as a promising solution to environmental challenges. Compared to TiO2, BiPO4 exhibits superior photocatalytic activity. However, its large band gap restricts its light absorption to the UV region. One effective technique for extending BiPO4’s absorption wavelength into the visible spectrum is the construction of the heterostructure. This study aimed to synthesize monodisperse BiPO4 nanorods via a solvothermal approach and fabricate BiPO4/g-C3N4 heterojunctions with varying loadings through in situ deposition. Tetracyclines were employed as the target pollutant to evaluate the photocatalytic performance and stability of the prepared materials. The results indicated that 5 wt% of composite exhibited better photocatalytic performance than single catalysts, which showed the highest photodegradation efficiency of approximately 98% for tetracyclines. The prepared bi-photocatalyst presented favorable stability under sunlight irradiation, the photocatalytic activity of which remained almost unchanged after four cycles. The enhanced photocatalytic activity was attributed to the synergistic effect. Additionally, the possible degradation mechanism was elucidated utilizing the semiconductor energy band theory. Overall, this work presents new perspectives on synthesizing innovative and efficient visible-light-driven photocatalysts. It also offers a mechanistic analysis approach by integrating theoretical calculations with experimental observations.

1. Introduction

With the rapid advancement of science and technology, the escalating use of antibiotics in medical treatment, livestock, and aquaculture has led to their widespread detection in natural water bodies and other environmental media [1,2,3]. Tetracycline (TC), as a basic broad-spectrum bacteriostatic agent among the most commonly used tetracycline family of antibiotics in the breeding industry, has been extensively applied worldwide [4]. TC boasts strong water solubility, a long half-life, and structural stability under acidic conditions, which can persist for a long time in natural water bodies and soils while participating in the material circulation process [5]. However, a variety of antibiotic resistance bacteria may be induced by the residual TC in the environment, affecting the normal life activities of native organisms and potentially threatening humans and the ecosystem through the transmission and enrichment of the food chain [6]. In response, extensive research and efforts have been devoted to the rapid degradation of residual tetracycline antibiotics in the environment.
Initially, TC in wastewater is expected to be removed via conventional sewage treatment processes such as physical adsorption and biological treatment. However, the removal efficiency is severely hindered by the complex structure and stable physicochemical properties of TC antibiotics. Additionally, this challenge is compounded by potential risks of secondary pollution [7,8]. In recent years, photocatalytic oxidation technology—characterized by its green chemistry principles, high efficiency, and absence of secondary pollution—has emerged as a promising advanced treatment for TC from wastewater, attracting significance research attention. Muller et al. [9] and Fujishima et al. [10] discovered that the photocatalytic oxidation process had the ability to efficiently destroy the structure of organic pollutants in wastewater and to achieve removal of pollutants in the 1970s. Since then, the photocatalytic oxidation process has been extensively studied in the field of antibiotic pollution treatment in wastewater, the main research fields of which focus on optimizing the photocatalytic degradation process and synthesizing innovative and efficient photocatalysts. Chen Yan et al. [11] prepared N-doped TiO2/diatomite-integrated photocatalytic particles (N-IPP), which achieved approximately 85% tetracycline removal efficiency after 5 reuse cycles. Recently, Tang Tao et al. [12] reported a new type of p-n heterojunction photocatalyst Bi2O3/Ti3+-TiO2. The photocatalytic degradation efficiency of tetracyclines under visible light irradiation was enhanced by 15% compared with traditional TiO2. However, traditional TiO2-based photocatalysts still faced several limitations, including low utilization efficiency of sunlight, low photocatalytic quantum efficiency, and insufficient photocatalytic reaction driving force.
Extensive research has explored the photocatalytic potential of various novel semiconductor materials to develop alternatives for TiO2 with superior photocatalytic performance [13,14,15,16]. Due to their characteristics of non-toxic properties, strong physicochemical stability, enhanced visible light response, and suitable energy bandwidth, bismuth-based semiconductors have attracted extensive attention as novel semiconductor photocatalysts for the degradation of antibiotics [4,17,18,19,20]. As initially reported by Pan’s research team [21],bismuth phosphate (BiPO4) showed excellent performance in degrading organic pollutants as a photocatalyst. This efficacy was primarily attributed to the efficient electron–hole separation due to the inherent inductive effect of the highly negatively charged PO43− groups [22,23]. However, the band bap of BiPO4 was appropriately 3.85 eV, and the relatively high band gap resulted in the limitation of the light absorption range to the short-wavelength ultraviolet region. Concurrently, the inherent structural defect severely limited its photocatalytic efficiency under visible light.
Heterojunction catalysts constructed by doping non-metallic semiconductors in bismuth-based materials generally exhibit higher photocatalytic efficiency over single semiconductors. The energy level difference between the compound semiconductors would aid the separation of photogenerated electron pairs and reduce their recombination efficiency, thereby enhancing the photocatalytic activity of heterojunction photocatalysts [24,25]. Graphite-like carbon nitride (g-C3N4) was considered as highly promising for enhancing catalytic degradation performance of metal–semiconductor photocatalysts toward tetracycline, owing to its high chemical and thermal stability, non-toxicity, and strong visible-light catalytic activity even at non-nanoscale [26]. Hyung Jun Kong et al. [27] prepared a composite photocatalyst with bismuth vanadate (BiVO4) and sulfur-doped g-C3N4 for water oxidation. Under identical reaction conditions, the oxygen release rate was demonstrated to be twice that of BiVO4, while its photon efficiency also increased by 19%. Jielin Yuan et al. [28] synthesized a nanoscale phosphorus-doped g-C3N4 and BiPO4 composite catalyst with a hydrolysis rate of up to 1110 μmol h−1 g−1. However, limited research has been conducted on the enhancement effect of heterojunction photocatalyst BiPO4/g-C3N4 on the photocatalytic oxidation of tetracyclines in wastewater. Additionally, the specific contributions and underlying mechanisms driving the improvement in photocatalytic performance remain underexplored.
In this study, the BiPO4/g-C3N4 heterojunction photocatalysts with varying loadings were synthesized using the in situ precipitation method. The morphology, structure, and photoelectrochemical properties were characterized, and the photocatalytic performance and stability of the prepared heterojunction catalysts for tetracyclines were investigated. In addition, the experimental observations were integrated with theoretical calculations to further delve into the mechanism of the formation of the heterojunction and the effective separation of the charge carriers at the heterojunction interface.

2. Results and Discussion

2.1. Crystal Structure and Morphology

The crystal structures and purity of BiPO4, g-C3N4, and composite g-C3N4/BiPO4 (CN/BiP) were analyzed by X-ray diffraction (Figure 1). The characteristic peaks of g-C3N4 at 13.0° and 27.4° corresponded to the crystalline planes of repetitive structure (100) and the crystalline planes of lamellar stacking (002) in the same plane of g-C3N4, respectively (JCPDS-50-0367). The characteristic peak at 13° was caused by the in-plane structural stacking of the aromatic ring system, corresponding to a lattice spacing of 0.679 nm. The strongest peak at 27.4° was the typical interlayer stacking peak of the aromatic ring system, with the lattice spacing value reflecting the intercalation spacing. For the prepared BiPO4, the peaks at 19.0°, 21.3°, 25.3°, 27.2°, 29.1°, 31.2°, 34.5°, and 36.9° corresponded to the (011), (-111), (111), (200), (120), (012), (-202), and (-212) crystal planes, respectively (JCPDS 80-0209). With the increasing content of g-C3N4, diffraction peaks corresponding to BiPO4 were detectable in all the CN/BiP samples, indicating that the introduction of g-C3N4 did not alter the crystal structure of BiPO4. However, the existence of g-C3N4 could not be determined by the XRD across different scales of composite CN/BiP samples due to the close proximity of the BiPO4 characteristic peak (200) and the g-C3N4 characteristic peak (002), with the latter being masked by the former.
The X-ray photoelectron spectroscopy data revealed the surface elemental composition and chemical valence states of BiPO4, g-C3N4, and composite CN/BiP (Figure 2). The XPS spectra clearly showed the coexistence of C, N, Bi, P, and O elements in 5%-CN/BiP, confirming the successful synthesis of CN/BiP composites (Figure 2a). The high-resolution XPS spectra of Bi 4f uncovered two characteristic peaks Bi 4f7/2 and Bi 4f5/2 of BiPO4 with binding energies of 159.9 eV and 165.2 eV, respectively [29] (Figure 2b). Moreover, a negative migration of 0.2 eV in the binding energy of Bi 4f was observed in the BiP/CN system compared to pure BiPO4. The N 1s spectrum of g-C3N4 was divided into four peaks at a binding energy of 398.5 eV, 399.8 eV, 400.9 eV, and 404.2 eV, respectively (Figure 2c). The characteristic peaks at binding energies of 398.5 eV and 399.8 eV corresponded to pyridine nitrogen (N- sp2 C) and pyrrole nitrogen (N- sp3 C) for the triazine units (C-N=C), respectively. The latter two were assigned to C-N-H and N-N configurations, consistent with previous literature reports [29]. In contrast, the N 1s peak of the 5%-CN/BiP sample shifted towards the higher binding energy. Generally, a negative correlation existed between the binding energy and the surface electron density [30]. The binding energy shifts of Bi 4f and N 1s in the XPS spectra indicated that the BiPO4 and g-C3N4 were not merely physically mixed but formed a heterojunction between the two semiconductors. Despite both being n-type semiconductors, g-C3N4 acted as the electron donor in the composite, and electrons were transferred from g-C3N4 to BiPO4 upon heterojunction formation.
Photocatalyst performance is closely related to its particle size and morphological structure [31]. Herein, during the photocatalytic processes, the time for randomly generated photogenerated electron–hole pairs to diffuse from the interior of the photocatalyst to the surface could be obtained from Equation (1) [32].
τ = r22D
where r is the particle size radius, and D is the diffusion coefficient of the photogenerated carriers. Smaller crystal sizes of photocatalysts enable faster diffusion of charge carriers to the interface, where they react with free radicals or target pollutants.
SEM images of BiPO4 indicated that the BiPO4 nanorods synthesized in this study using the solvothermal method featured a smooth surface and homogeneous morphology with an average diameter and length of approximately 50–70 nm and 600 nm, respectively (Figure 3a). One-dimensional (1D) nanostructured materials (e.g., nanowires, nanorods, nanotubes) facilitated efficient separation of photogenerated electron–hole pairs. This stemmed from the ability of charge carriers to undergo ballistic charge transport along the axial direction [33]. On the other hand, g-C3N4 was a typical irregularly layered stacking structure, consisting of two-dimensional (2D) lamellae stacked (Figure 3b). The composite of 1D BiPO4 and 2D g-C3N4 flakes would provide more active sites, thereby enabling the diffusion of photogenerated carriers to the interface of the composite and improving the photocatalytic performance [34]. Morphological analyses of CN/BiP materials with varying g-C3N4 loading ratios revealed that the surface of BiP nanorods exhibited an increasing coverage of g-C3N4 nanoflakes as the loading of g-C3N4 increased (Figure 3c–g). When the loading mass ratio of g-C3N4 exceeded 5%, the nanoflake-like g-C3N4 tended to agglomerate and accumulate, compromising the dispersion of the samples. Therefore, an optimal loading mass ratio between g-C3N4 and BiPO4 might exist.
TEM and HRTEM characterization of the samples’ microscopic morphology revealed that the structure of BiPO4 nanorods and the stacked flake morphology of g-C3N4 were consistent with the above SEM observations (Figure 4a,b). The microscopic morphology of the samples characterized by TEM showed that the structure of BiPO4 nanorods, g-C3N4 stacked flakes, and the dispersion of BiPO4 nanorods in g-C3N4 flakes aligned with the SEM results (Figure 4a–c). The HRTEM image of the 5%-CN/BiP sample suggested a visible interface region between the two semiconductors, which had been marked with red dotted line. The lattice stripe with a spacing of 0.244 nm corresponded to the (112) crystalline plane of BiPO4, while the amorphous region was attributed to g-C3N4 (Figure 4d). The results were consistent with the XRD data, confirming the successful construction of heterojunctions between the two photocatalysts.
The FESEM-EDS analysis showed that the synthesized 5%-CN/BiP sample contained five elements, including Bi, P, O, C, and N (Figure 4e). These elements were uniformly distributed in the composite catalyst, further proving the successful synthesis of the CN/BiP sample (Figure 5).

2.2. Photoelectrochemical Properties

The photoelectrochemical properties of the samples were characterized using DRS, PL, photocurrent, and EIS impedance. The optical absorption properties of single BiPO4 samples and CN/BiP samples with varying mass ratios were analyzed via DRS. The results indicated that BiPO4 could only respond to UV light, with the absorption band edge at roughly 300 nm (Figure 6a). The absorption spectra of the CN/BiP composites with different mass ratios extended from the short UV wavelength to the visible region with absorption edges close to 450 nm. The addition of g-C3N4 not only expanded the light absorption range of the catalyst to the visible region, but also improved its light response in the absorption region of 260–450 nm. This indicated that the CN/BiP composite had high visible light utilization. The 5%-CN/BiP exhibited the best optical absorption performance in the visible region. The energy bands of the semiconductor were calculated according to the Kubelka–Munk function Equation (2)
αhν = A(hν − Eg)n/2
where α, ν, A, and Eg are the optical absorption coefficient, optical frequency, constant, and forbidden band width, respectively. The value of n is 1 when the bandgap type of the semiconductor is direct bandgap (BiPO4) and 4 for indirect bandgap (g-C3N4) [35]. Herein, the forbidden band widths of the BiPO4, g-C3N4 and 5%-CN/BiP samples were 3.63 eV, 2.71 eV, and 2.58 eV, respectively (Figure 6b). This indicated that the CN/BiP composites exhibited lower band gap energy compared to the single-component samples, a characteristic favorable for the generation of photogenerated electron–hole pairs.
The radiative complexation of photogenerated carriers significantly influences the activity of photocatalysts. Fluorescence emission spectra (PL) are frequently employed to investigate the luminescence properties of synthetic materials. Generally, lower PL intensity indicates reduced recombination efficiency of photogenerated carriers, implying higher separation efficiency and superior catalytic performance of the material. Herein, the BiPO4 sample demonstrated a strong emission peak at 440 nm with higher intensity compared to all the materials compounded with g-C3N4 (Figure 7). This indicated that the addition of g-C3N4 effectively promoted the separation of carriers while reducing the complexation rate, thus improving the activity of the photocatalyst. Under the optimal conditions (5% mass ratio), the photoluminescence intensity was substantially reduced, and the photocatalytic performance was improved, consistent with the DRS results.
The transient photocurrent responses of the electrodes of BiPO4, g-C3N4, and CN/BiP materials were tested to visually elucidate the separation and migration properties of photogenerated electron–hole pairs on the synthesized materials. Figure 8a presents the electrochemical impedance diagrams (EIS) of the corresponding samples. The results indicated that the photocurrent density of all samples remained stable and electron regenerative after seven light–dark intermittent cycles of experiments. The CN/BiP composites showed higher photocurrent intensity than single BiPO4, g-C3N4 materials, indicating that the composite heterojunctions had more efficient electron–hole separation and migration properties [36]. The smaller electrochemical impedance radius of the 5%-CN/BiP sample implied lower charge transfer resistance and higher charge transfer efficiency, reflecting more effective separation of photogenerated carriers (Figure 8b). The results revealed that the heterojunction of CN/BiP composites was favorable to suppressing the compounding process of electrons and holes, while promoting the migration and separation of charge carriers.

2.3. Photocatalytic Degradation Properties

The degradation experiments on TC indicated that the photocatalytic activity of the synthetic catalysts was significantly enhanced compared to the single catalyst. Without the addition of a photocatalyst, the TC concentration remained almost unchanged after 90 min of Xe lamp irradiation, demonstrating its negligible photolytic effect. When BiPO4 and g-C3N4 were employed as photocatalysts, the degradation rates of TC were 12.8% and 48.0%, respectively, after 90 min of irradiation (Figure 9a). The loading of g-C3N4 had a more pronounced influence on the photocatalytic activity of the composites, among which the 5%-CN/BiP sample demonstrated the optimal photocatalytic activity, with the degradation rate of TC close to 100% at the conclusion of the experiment (Figure 9a).
The photodegradation rate constant of TC could be calculated according to the following quasi-level kinetic Equation (3) of the Langmuir–Hinshelwood model [37]:
ln(C0/C) = kappt
where the slope kapp denotes the apparent rate constant; C is the TC concentration at different reaction times; C0 is the initial TC concentration; and t is the time of illumination. The fitted curves of ln(C0/C) versus time t exhibited a good linear relationship, indicating that the reaction of sample degradation of TC was in accordance with pseudo-first-order kinetics (Figure 9b). The apparent rate constants of g-C3N4/BiPO4 composites were significantly higher than those of single photocatalysts (Table 1). This enhancement was likely attributed to the formation of a heterojunction between g-C3N4 and BiPO4 in the composite, which substantially improved the material’s photocatalytic performance [38]. In addition, the photocatalytic activity of the synthesized catalyst showed a trend of increasing and then decreasing with the growth of g-C3N4 loading, and the optimal photocatalytic degradation performance was achieved when the loading of g-C3N4 was 5 wt%. This phenomenon might arise because excessiveg-C3N4 wrapped on the surface of the synthesized catalyst restricted the light absorption of the heterojunction and weakened the catalyst’s surface adsorption capacity on TC, thereby reducing the photocatalytic degradation efficiency [36].
Figure 9c,d show the degradation efficiency and kinetics curves of OTC. Similarly to TC, 5%-CN/BiP also demonstrated the best photocatalytic activity. Concurrently, the photodegradation rate constants of BiPO4, g-C3N4 and 5%-CN/BiP samples for OTC were 1.37 × 10−3, 7.42 × 10−3, and 2.77 × 10−2, respectively (Table 1). The above results confirmed that the combination of BiPO4 and g-C3N4 was advantageous to the separation of photogenerated carriers of the photocatalyst, thereby improving the catalytic degradation activity of the sample. Insufficient loading of g-C3N4 prevented full contact between BiPO4 and g-C3N4, thus hindering heterojunction formation. Excessive loading of g-C3N4, while ensuring charge transfer, might hinder the absorption of visible light by the material and decelerate the generation rate of photogenerated electrons. Consequently, the 5%-CN/BiP photocatalyst exhibited the best photocatalytic activity, which was also consistent with the previous characterization results.
From the perspective of practical application, in addition to photocatalytic efficiency, the stability of the photocatalyst was another key factor reflecting its performance. In order to investigate the stability and reusability of 5%-CN/BiP, the cyclic degradation experiments of two tetracycline antibiotics by photocatalysts under sunlight were carried out. As depicted in Figure 10, after four cycles, the photocatalytic degradation efficiency of TC and OTC by 5%-CN/BiP showed no significant loss, with only a slight reduction observed. This indicated the excellent stability and repeatability of 5%-CN/BiP for degradation under sunlight irradiation.
The photodegradation efficiency of 5%-CN/BiP was compared with that of other photocatalysts under visible light reported in previous literature. The detailed information is listed in Table 2. Among them, the as-prepared 5%-CN/BiP in this study demonstrated superior photocatalytic activity for TC degradation.

2.4. Identification of Active Species

The addition of quenching agents decreased the degradation efficiency of TC by 5%-CN/BiP (Figure 11). The performance of the samples was significantly inhibited upon the addition of TEMPOL, indicating that the •O2 species played a dominant role in the reaction system. The introduction of AO also negatively affected the degradation rate of TC. This suggested that despite the involvement of h+ in the photodegradation of TC, the effect on the reaction was not as significant as that of •O2. Conversely, the addition of IPA had a negligible effect on the degradation rate, revealing that •OH was not the primary active species.
The EPR spectroscopy of BiPO4 and 5%-CN/BiP samples revealed that the material recombination altered the catalytic mechanism of the reaction (Figure 12). In the absence of light, no characteristic peaks were observed in all samples. However, •O2 radicals were detected in all samples under sunlight, and the signal intensity of 5%-CN/BiP composites were stronger than that of BiPO4. In a single BiPO4 catalytic reaction system, a reduced intensity of •OH signal was obtained. Notably, almost no •OH signals were detected in the 5%-CN/BiP samples, indicating a change in the active species within the composite catalyst reaction system.

2.5. Mechanism of the Photocatalytic Performance Improvement

The photocatalyst energy band structure and density of states calculated using density functional theory revealed the mechanism of the formed heterojunction on the catalyst photocatalytic activity enhancement (Figure 13). The results demonstrated both BiPO4 and g-C3N4 as n-type semiconductors with band gap widths of 3.60 eV and 2.69 eV, respectively. The band gap of g-C3N4 with amorphous structures would be lower than that for the crystalline structure [43]. The conduction band minimum of BiPO4 was composed of Bi 6p orbital, while its valence band maximum primarily included O 2p, Bi 6p, Bi 6s, and P 3p orbitals. For g-C3N4, the conduction band minimum mainly consisted of C 2p and N 2p orbitals, and the valence band maximum was dominated by N 2p orbitals (Figure 13a,b). The low orbital hybridization of O, P, and Bi atoms at the valence band maximum of BiPO4 led to a high probability of electron occurrence, whereas the orbital hybridization density of the atoms at the conduction band minimum was low. Thus, the electrons in the conduction band were easily excited to the surface of BiPO4. In comparison to BiPO4, g-C3N4 exhibited a higher degree of hybridization and electron hybridization density in the bottom atomic orbitals at the conduction band minimum, while the atomic orbitals at the valence band maximum presented a lower degree of hybridization. To confirm the electron transfer path for the built-in-electric field at the heterojunction interface, the work functions of BiPO4 and g-C3N4 were further calculated. As illustrated in Figure 13e,f, the work functions of BiPO4 and g-C3N4 were determined to be 4.378 and 4.754 eV, being in the sequence BiPO4 < g-C3N4. This indicated that the electrons in the g-C3N4 nanoparticles would incline to transfer to BiPO4. Therefore, when BiPO4 coupled with g-C3N4 to form an n-n heterojunction, the electrons were transferred from the conduction band of g-C3N4 with higher energy to the conduction band of BiPO4. Conversely, the holes in the valence band usually tended to transfer from BiPO4 to g-C3N4. Eventually, the formation of heterojunctions effectively suppressed the compounding of photogenerated electron–hole pairs.
XPS analysis of the heterojunction material revealed the energy band positions of BiPO4 and g-C3N4. Upon sunlight absorption by the photocatalytic material, electrons in the valence band were excited to the conduction band, generating holes in the valence band (Figure 13c,d). The formation of the heterojunction contributed to the transfer of conduction band electrons from g-C3N4 to BiPO4, which showed a greater conduction band potential than the reduction potential of O2/•O2 (−0.33 eV). Therefore, the g-C3N4/BiPO4 would reduce O2 to •O2 [44]. Moreover, the photogenerated electrons on the surface of BiPO4 could trap molecular oxygen to generate •O2, and the holes in the valence band of g-C3N4 reacted directly with the target pollutant. The valence band potential of BiPO4 (3.18 eV) was higher than the redox potentials of •OH/H2O (2.27 eV) and •OH/OH (2.38 eV) [45]. Following the addition of g-C3N4, the holes in the BiPO4 valence band were transferred to the valence band of g-C3N4, a valence band potential (1.65 eV) of which was lower than that of •OH/H2O and •OH/OH. Consequently, the g-C3N4/BiPO4 composite could not oxidize H2O/OH to •OH. EPR experiments supported this conclusion: the •OH signal was detected in the reaction system using single BiPO4, but absent in the 5%-CN/BiP sample. Therefore, the main active species involved in the photocatalytic degradation process was probably •O2, while h+ and •OH also played supplementary roles (Figure 14).

3. Materials and Methods

3.1. Materials

Tetracycline (C22H24N2O8), oxytetracycline (C22H24N2O9), tetracycline hydrochloride (C22H24N2O9·HCl), bismuth nitrate (Bi(NO3)3·5H2O), sodium dihydrogen phosphate (NaH2PO4·2H2O), urea (CH4N2O), and other chemicals used in the experiments were purchased from Nanjing Chemical Reagents Co., Ltd. (Nanjing, China) with analytically pure grade (99%), without further purification. Solutions were prepared utilizing deionized water.

3.2. Synthesis of BiPO4/g-C3N4 Heterojunction Photocatalysts

Graphitic carbon nitride (g-C3N4) was prepared using the calcination method via two stages. Specifically, 10 g urea powder was placed into an alumina crucible and calcined at 500 °C at a heating rate of 5 °C/min, followed by raising the temperature to 550 °C at 10 °C/min and maintained for 4 h. Upon cooling to room temperature, the obtained nanosheet g-C3N4 was milled into powders.
BiPO4 photocatalyst was synthesized using the solvothermal method. Briefly, 1 mmol Bi (NO3)3·5H2O and 1 mmol NaH2PO4·2H2O were added to 60 mL solution with deionized water and glycerol (v/v, 1:3). After vigorously stirring for 30 min, the reaction was performed in a stainless-steel autoclave at 180 °C for 24 h. Subsequently, the precipitate was collected by centrifugation and dried in an oven at 60 °C overnight.
Different mass proportions of g-C3N4 (1 wt%, 2 wt%, 5 wt%, 10 wt%, 15 wt%) were added to a mixture of 60 mL deionized water and glycerin (1:3, v/v) and dispersed evenly. Following that, 1 mmol Bi(NO3)3·5H2O and 1 mmol NaH2PO4·2H2O were continuously added and stirred vigorously for 10 min. Subsequently, the suspension was put into a stainless steel autoclave, calcined at 180 °C for 24 h, then centrifuged, collected, and dried overnight in an oven at 60 °C. The g-C3N4/BiPO4 composite photocatalysts with varying g-C3N4 loadings were ultimately obtained. The sample name was abbreviated as X%-CN/BiP, where X represents the mass ratio of g-C3N4, CN stands for g-C3N4, and signifies BiPO4.

3.3. Characterization of Synthetic Catalysts

X-ray diffraction (XRD) patterns of the samples were determined using an XRD-6000 X-ray powder diffractometer (Shimadzu, Kyoto, Japan) with monochromatic Cu-Kα radiation at a setting of 40 kV and 30 mA. The FT-IR spectra of the samples were determined on a Nicolet iS10 FT-IR instrument within the IR range (500–4000 cm−1). X-ray photoelectron spectroscopy (XPS) was performed on a PHI5000 Versa Probe spectrometer (ULVAC-PHI, Maozaki, Japan). Scanning electron microscopy (SEM) images were acquired using a QUANTA FEG 250 (FEI, Hillsboro, OR, USA), while transmission electron micrographs (TEMs) were obtained via a JEM-200CX instrument (JEOL, Kyoto, Japan).
UV–vis diffuse reflectance spectroscopy (DRS) was performed using a UV-3600 spectrophotometer (Shimadzu, Kyoto, Japan) equipped with an integrating sphere attachment, covering a wavelength range of 200–800 nm. PL were recorded on a Horiba Fluorolog 3-22 fluorescence spectrophotometer (Horiba, Irvine, CA, USA) at an excitation wavelength of 365 nm, with measurements taken over the 200–800 nm wavelength range.
Photoelectrochemical characterization was performed on a CHI760E electrochemical workstation (Shanghai, China) with a standard three-electrode system. The samples were loaded onto an ITO electrode (1 cm × 2 cm squares) and served as the working electrode. Pt plate and Ag/AgCl electrode were utilized as the counter and reference electrode, respectively. The electrolyte was 0.2 M Na2SO4 aqueous solution, and a 500 W xenon lamp (NBeT, Beijing, China) was employed to provide the light source.

3.4. Photocatalytic Degradation Experimental Set Up

The photocatalytic activities of as-prepared photocatalysts were evaluated by catalytic degradation of TC under visible light using 1000 W Xe lamp irradiation (Xujiang, Nanjing, China). In a typical experiment, the reactant and the catalysts were placed in a quartz tube. 20 mg of the photocatalyst was inspired into 50 mL of TC aqueous solution (20 mg/L). Prior to the light illumination, the suspension was magnetically stirred for 60 min in dark to reach an adsorption-desorption equilibrium. Following the equilibrium, the samples (volume of each is 4 mL) were taken at given time intervals, centrifuged at 7000 rpm for 10 min, and filtered through a 0.22 µm Millipore filter to remove the particles. Following that, the concentration of TC in the solution was analyzed using an Agilent 1200 high-performance liquid chromatography (HPLC) system. Ammonium oxalate (AO), isopropanol (IPA), and 4-hydroxy-2,2,6,6-tetramethylpiperidine-1-oxyl radical (TEMPOL) were added to the original reaction system as quenchers for vacancies (h+), hydroxyl radicals (•OH), and superoxide radicals (•O2), respectively. The main reactive species involved in the reaction and their contributions were compared under the identical experimental conditions. Additionally, electron paramagnetic resonance (EPR) spectroscopy with 5,5-dimethyl-1-pyrroline-N-oxide (DMPO) as a trapping agent was employed to characterize radicals generated during the photocatalytic reaction. Superoxide radicals (•O2) were detected in the methanol system, while hydroxyl radicals (•OH) were analyzed in the aqueous system. To ensure result reproducibility, duplicate experiments were performed for each condition, and average data were recorded. Blank experiments without catalysts were also performed.

3.5. Analytical and Computational Simulation Methods

The concentration of tetracyclines was determined using an Agilent 1200 HPLC system (Agilent, Santa Clara, CA, USA) with the UV–vis spectrophotometer 2450 and a 4.6 mm × 150 mm × 5 μm Zorbax Eclipse XDB-C18 at 35 °C. The wavelength of the ultraviolet (UV) detector was set at 355 nm. The mixture of methanol-water (10:90, v/v, 0.1% formic acid contained in water) was employed as the mobile phase with a flow rate of 1.0 mL   min 1 . The photocatalytic degradation efficiency was calculated by the following Equation (4):
η = C 0     C t   C 0   ×   100 %
where η is the photocatalytic efficiency; C0 is the concentration of reactant before illumination; and Ct is the concentration of reactant following illumination of t hours.
The generalized gradient approximation (GGA) in the Perdew–Burke–Ernzerh (PBE) functional form was applied along with the ultrasoft pseudopotentials. Moreover, density functional (DFT) calculations were performed utilizing the plane-wave pseudopotential method in the CASTEP code [46]. The GGA with the PBE correction was also adopted to estimate the work function using CASTEP code.

4. Conclusions

Herein, g-C3N4/BiPO4 heterojunction photocatalysts were successfully prepared using the solvothermal method. The morphologies of the prepared photocatalysts were characterized, with their photoelectrochemical properties, photocatalytic activities, and photocatalytic mechanisms investigated. The results showed that the heterojunction material exhibited better photocatalytic performance compared to the single catalyst for the degradation of tetracycline antibiotics. The photocatalytic performance of the catalysts with varying g-C3N4 loadings revealed that 5 wt% of the heterojunction material exhibited the highest photodegradation efficiency for tetracyclines. The prepared heterojunction catalysts demonstrated excellent stability under sunlight irradiation, and the photocatalytic activity of the heterojunction catalysts remained almost unchanged after four cycles of recycling. Furthermore, mechanistic analysis of the photocatalytic degradation process indicated that •O2 served as the primary active species, with h+ and •OH acting as supporting participants. Overall, this study offers new perspectives for the preparation of novel, environmentally friendly, and efficient visible-light-driven photocatalysts.

Author Contributions

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

Funding

This research was funded by the Program of Nanjing Institute of Environmental Sciences, Ministry of Ecology and Environment: Integration of Ecological and Environmental Protection in Yangtze River Delta, grant number ZX2022QT047; the National Key R&D Program of China, grant number 2021YFC1809205; the Innovative team project of Nanjing Institute of Environmental Sciences, grant number ZXQT202301002; and the Fundamental Research Funds for the Central Public Welfare Research Institutes, grant number GYZX220202.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Data are contained within this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Chang, P.-H.; Li, Z.; Yu, T.-L.; Munkhbayer, S.; Kuo, T.-H.; Hung, Y.-C.; Jean, J.-S.; Lin, K.-H. Sorptive removal of tetracycline from water by palygorskite. J. Hazard. Mater. 2009, 165, 148–155. [Google Scholar] [CrossRef] [PubMed]
  2. Hou, L.; Zhang, H.; Wang, L.; Chen, L. Ultrasound-enhanced magnetite catalytic ozonation of tetracycline in water. Chem. Eng. J. 2013, 229, 577–584. [Google Scholar] [CrossRef]
  3. Hou, J.; Wan, W.; Mao, D.; Wang, C.; Mu, Q.; Qin, S.; Luo, Y. Occurrence and distribution of sulfonamides, tetracyclines, quinolones, macrolides, and nitrofurans in livestock manure and amended soils of Northern China. Environ. Sci. Pollut. Res. 2015, 22, 4545–4554. [Google Scholar] [CrossRef]
  4. Zhu, X.; Wang, Y.; Guo, Y.; Wan, J.; Yan, Y.; Zhou, Y.; Sun, C. Environmental-friendly synthesis of heterojunction photocatalysts g-C3N4/BiPO4 with enhanced photocatalytic performance. Appl. Surf. Sci. 2021, 544, 148872. [Google Scholar] [CrossRef]
  5. Wang, D.; Jia, F.; Wang, H.; Chen, F.; Fang, Y.; Dong, W.; Zeng, G.; Li, X.; Yang, Q.; Yuan, X. Simultaneously efficient adsorption and photocatalytic degradation of tetracycline by Fe-based MOFs. J. Colloid Interface Sci. 2018, 519, 273–284. [Google Scholar] [CrossRef]
  6. Mao, D.; Yu, S.; Rysz, M.; Luo, Y.; Yang, F.; Li, F.; Hou, J.; Mu, Q.; Alvarez, P.J.J. Prevalence and proliferation of antibiotic resistance genes in two municipal wastewater treatment plants. Water Res. 2015, 85, 458–466. [Google Scholar] [CrossRef]
  7. Phoon, B.L.; Ong, C.C.; Saheed, M.S.M.; Show, P.-L.; Chang, J.-S.; Ling, T.C.; Lam, S.S.; Juan, J.C. Conventional and emerging technologies for removal of antibiotics from wastewater. J. Hazard. Mater. 2020, 400, 122961. [Google Scholar]
  8. Wu, G.; Qin, W.; Sun, L.; Yuan, X.; Xia, D. Role of peroxymonosulfate on enhancing ozonation for micropollutant degradation: Performance evaluation, mechanism insight and kinetics study. Chem. Eng. J. 2019, 360, 115–123. [Google Scholar] [CrossRef]
  9. Muller, H.D.; Steinbac, F. Decomposition of Isopropyl Alcohol photosensitized by Zinc Oxide. Nature 1970, 225, 728–729. [Google Scholar] [CrossRef]
  10. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  11. Chen, Y.; Liu, K.R. Preparation and characterization of nitrogen-doped TiO2/diatomite integrated photocatalytic pellet for the adsorption-degradation of tetracycline hydrochloride using visible light. Chem. Eng. J. 2016, 302, 682–696. [Google Scholar] [CrossRef]
  12. Tang, T.; Yin, Z.L.; Chen, J.R.; Zhang, S.; Sheng, W.C.; Wei, W.X.; Xiao, Y.G.; Shi, Q.Y.; Cao, S.S. Novel p-n heterojunction Bi2O3/Ti3+-TiO2 photocatalyst enables the complete removal of tetracyclines under visible light. Chem. Eng. J. 2021, 417, 128058. [Google Scholar] [CrossRef]
  13. Liu, J.; Liu, Y.; Liu, N.; Han, Y.; Zhang, X.; Huang, H.; Lifshitz, Y.; Lee, S.-T.; Zhong, J.; Kang, Z. Metal-free efficient photocatalyst for stable visible water splitting via a two-electron pathway. Science 2015, 347, 970–974. [Google Scholar] [PubMed]
  14. Wu, A.; Tian, C.; Jiao, Y.; Yan, Q.; Yang, G.; Fu, H. Sequential two-step hydrothermal growth of MoS2/CdS core-shell heterojunctions for efficient visible light-driven photocatalytic H2 evolution. Appl. Catal. B Environ. 2017, 203, 955–963. [Google Scholar]
  15. Liu, T.; Zhang, X.; Zhao, F.; Wang, Y. Targeting inside charge carriers transfer of photocatalyst: Selective deposition of Ag2O on BiVO4 with enhanced UV–vis-NIR photocatalytic oxidation activity. Appl. Catal. B Environ. 2019, 251, 220–228. [Google Scholar]
  16. Bi, Y.; Hu, H.; Ouyang, S.; Jiao, Z.; Lu, G.; Ye, J. Selective growth of Ag3PO4 submicro-cubes on Ag nanowires to fabricate necklace-like heterostructures for photocatalytic applications. J. Mater. Chem. 2012, 22, 14847–14850. [Google Scholar]
  17. Ye, L.; Su, Y.; Jin, X.; Xie, H.; Zhang, C. Recent advances in BiOX (X = Cl, Br and I) photocatalysts: Synthesis, modification, facet effects and mechanisms. Environ. Sci. Nano 2014, 1, 90–112. [Google Scholar]
  18. Wang, P.; Ao, Y.; Wang, C.; Hou, J.; Qian, J. A one-pot method for the preparation of graphene-Bi2MoO6 hybrid photocatalysts that are responsive to visible-light and have excellent photocatalytic activity in the degradation of organic pollutants. Carbon 2012, 50, 5256–5264. [Google Scholar] [CrossRef]
  19. Rao, P.M.; Cai, L.; Cho, I.S.; Lee, C.H.; Weisse, J.M.; Zheng, X.; Liu, C.; Yang, P. Simultaneously efficient light absorption and charge separation in WO3/BiVO4 Core/shell nanowire photoanode for photoelectrochemical water oxidation. Nano Lett. 2014, 14, 1099–1105. [Google Scholar] [CrossRef]
  20. Gui, M.S.; Zhang, W.D.; Su, Q.X.; Chen, C.H. Preparation and visible light photocatalytic activity of Bi 2O3/Bi2WO6 heterojunction photocatalysts. J. Solid State Chem. 2011, 184, 1977–1982. [Google Scholar] [CrossRef]
  21. Pan, C.; Zhu, Y. New Type of BiPO4 Oxy-Acid Salt Photocatalyst with High Photocatalytic Activity on Degradation of Dye. Environ. Sci. Technol. 2010, 44, 5570–5574. [Google Scholar] [CrossRef] [PubMed]
  22. Nithya, V.D.; Kalai Selvan, R.; Vasylechko, L. Hexamethylenetetramine assisted hydrothermal synthesis of BiPO4 and its electrochemical properties for supercapacitors. J. Phys. Chem. Solids 2015, 86, 11–18. [Google Scholar] [CrossRef]
  23. Pan, C.; Xu, J.; Chen, Y.; Zhu, Y. Influence of OH-related defects on the performances of BiPO4 photocatalyst for the degradation of rhodamine B. Appl. Catal. B Environ. 2012, 115–116, 314–319. [Google Scholar] [CrossRef]
  24. Marschall, R. Semiconductor composites: Strategies for enhancing charge carrier separation to improve photocatalytic activity. Adv. Funct. Mater. 2014, 24, 2421–2440. [Google Scholar]
  25. Moniz, S.J.; Shevlin, S.A.; Martin, D.J.; Guo, Z.-X.; Tang, J. Visible-light driven heterojunction photocatalysts for water splitting–a critical review. Energy Environ. Sci. 2015, 8, 731–759. [Google Scholar]
  26. Song, Y.; Tian, J.; Gao, S.; Shao, P.; Qi, J.; Cui, F. Photodegradation of sulfonamides by g-C3N4 under visible light irradiation: Effectiveness, mechanism and pathways. Appl. Catal. B Environ. 2017, 210, 88–96. [Google Scholar]
  27. Kong, H.J.; Won, D.H.; Kim, J.; Woo, S.I. Sulfur-Doped g-C3N4/BiVO4 Composite Photocatalyst for Water Oxidation under Visible Light. Chem. Mater. 2016, 28, 1318–1324. [Google Scholar] [CrossRef]
  28. Yuan, J.; Gao, Q.; Li, X.; Liu, Y.; Fang, Y.; Yang, S.; Peng, F.; Zhou, X. Novel 3-D nanoporous graphitic-C3N4 nanosheets with heterostructured modification for efficient visible-light photocatalytic hydrogen production. RSC Adv. 2014, 4, 52332–52337. [Google Scholar] [CrossRef]
  29. Tan, G.Q.; She, L.N.; Liu, T.; Xu, C.; Ren, H.J.; Xia, A. Ultrasonic chemical synthesis of hybrid mpg-C3N4/BiPO4 heterostructured photocatalysts with improved visible light photocatalytic activity. Appl. Catal. B-Environ. 2017, 207, 120–133. [Google Scholar] [CrossRef]
  30. Zhuang, Z.; Li, Y.; Li, Z.; Lv, F.; Lang, Z.; Zhao, K.; Zhou, L.; Moskaleva, L.; Guo, S.; Mai, L. MoB/g-C3N4 interface materials as a schottky catalyst to boost hydrogen evolution. Angew. Chem. 2018, 130, 505–509. [Google Scholar]
  31. Ge, L.; Chen, J.; Wei, X.; Zhang, S.; Qiao, X.; Cai, X.; Xie, Q. Aquatic photochemistry of fluoroquinolone antibiotics: Kinetics, pathways, and multivariate effects of main water constituents. Environ. Sci. Technol. 2010, 44, 2400–2405. [Google Scholar] [PubMed]
  32. Xu, Y.; Shen, M. Fabrication of anatase-type TiO2 films by reactive pulsed laser deposition for photocatalyst application. J. Mater. Process. Technol. 2008, 202, 301–306. [Google Scholar]
  33. Shi, H.; Zhou, C.; Zhang, C. Silver vanadate nanowires: Photocatalytic properties and theoretical calculations. Res. Chem. Intermed. 2015, 41, 7725–7737. [Google Scholar]
  34. He, R.; Cheng, K.; Wei, Z.; Zhang, S.; Xu, D. Room-temperature in situ fabrication and enhanced photocatalytic activity of direct Z-scheme BiOI/g-C3N4 photocatalyst. Appl. Surf. Sci. 2019, 465, 964–972. [Google Scholar]
  35. He, R.; Zhou, J.; Fu, H.; Zhang, S.; Jiang, C. Room-temperature in situ fabrication of Bi2O3/g-C3N4 direct Z-scheme photocatalyst with enhanced photocatalytic activity. Appl. Surf. Sci. 2018, 430, 273–282. [Google Scholar]
  36. Zheng, X.; Wang, J.; Liu, J.; Wang, Z.; Chen, S.; Fu, X. Photocatalytic degradation of benzene over different morphology BiPO4: Revealing the significant contribution of high–energy facets and oxygen vacancies. Appl. Catal. B Environ. 2019, 243, 780–789. [Google Scholar]
  37. Vincent, M.J.; Gonzalez, R.D. A Langmuir–Hinshelwood model for a hydrogen transfer mechanism in the selective hydrogenation of acetylene over a Pd/γ-Al2O3 catalyst prepared by the sol–gel method. Appl. Catal. A Gen. 2001, 217, 143–156. [Google Scholar]
  38. Long, B.; Huang, J.; Wang, X. Photocatalytic degradation of benzene in gas phase by nanostructured BiPO4 catalysts. Prog. Nat. Sci. Mater. Int. 2012, 22, 644–653. [Google Scholar] [CrossRef]
  39. Liu, Y.; Kong, J.; Yuan, J.; Zhao, W.; Zhu, X.; Sun, C.; Xie, J. Enhanced photocatalytic activity over flower-like sphere Ag/Ag2CO3/BiVO4 plasmonic heterojunction photocatalyst for tetracycline degradation. Chem. Eng. J. 2018, 331, 242–254. [Google Scholar] [CrossRef]
  40. Chen, Z.-J.; Guo, H.; Liu, H.-Y.; Niu, C.-G.; Huang, D.-W.; Yang, Y.-Y.; Liang, C.; Li, L.; Li, J.-C. Construction of dual S-scheme Ag2CO3/Bi4O5I2/g-C3N4 heterostructure photocatalyst with enhanced visible-light photocatalytic degradation for tetracycline. Chem. Eng. J. 2022, 438, 135471. [Google Scholar] [CrossRef]
  41. Chen, W.; Huang, J.; He, Z.-C.; Ji, X.; Zhang, Y.-F.; Sun, H.-L.; Wang, K.; Su, Z.-W. Accelerated photocatalytic degradation of tetracycline hydrochloride over CuAl2O4/g-C3N4 p-n heterojunctions under visible light irradiation. Sep. Purif. Technol. 2021, 277, 119461. [Google Scholar] [CrossRef]
  42. Shao, B.; Liu, Z.; Zeng, G.; Wu, Z.; Liu, Y.; Cheng, M.; Chen, M.; Liu, Y.; Zhang, W.; Feng, H. Nitrogen-Doped Hollow Mesoporous Carbon Spheres Modified g-C3N4/Bi2O3 Direct Dual Semiconductor Photocatalytic System with Enhanced Antibiotics Degradation under Visible Light. ACS Sustain. Chem. Eng. 2018, 6, 16424–16436. [Google Scholar] [CrossRef]
  43. Durandurdu, M. Amorphous carbon nitride (C3N4). J. Non-Cryst. Solids 2024, 631, 122916. [Google Scholar] [CrossRef]
  44. Zhao, W.; Liu, Y.; Wei, Z.; Yang, S.; He, H.; Sun, C. Fabrication of a novel p–n heterojunction photocatalyst n-BiVO4@p-MoS2 with core–shell structure and its excellent visible-light photocatalytic reduction and oxidation activities. Appl. Catal. B Environ. 2016, 185, 242–252. [Google Scholar]
  45. Guo, Y.; Dai, Y.; Zhao, W.; Li, H.; Xu, B.; Sun, C. Highly efficient photocatalytic degradation of naphthalene by Co3O4/Bi2O2CO3 under visible light: A novel p–n heterojunction nanocomposite with nanocrystals/lotus-leaf-like nanosheets structure. Appl. Catal. B Environ. 2018, 237, 273–287. [Google Scholar] [CrossRef]
  46. Segall, M.D.; Lindan, P.J.D.; Probert, M.J.; Pickard, C.J.; Hasnip, P.J.; Clark, S.J.; Payne, M.C. First-principles simulation: Ideas, illustrations and the CASTEP code. J. Phys.-Condens. Matter 2002, 14, 2717–2744. [Google Scholar] [CrossRef]
Figure 1. XRD spectra of BiPO4, g-C3N4, and composite CN/BiP with varying mass ratios.
Figure 1. XRD spectra of BiPO4, g-C3N4, and composite CN/BiP with varying mass ratios.
Molecules 30 02905 g001
Figure 2. (a) XPS survey spectra of BiPO4, g-C3N4 and 5%-CN/BiP; High-resolution XPS spectra, (b) Bi 4f of BiPO4 and 5%-CN/BiP, and (c) N1s of g-C3N4 and 5%-CN/BiP.
Figure 2. (a) XPS survey spectra of BiPO4, g-C3N4 and 5%-CN/BiP; High-resolution XPS spectra, (b) Bi 4f of BiPO4 and 5%-CN/BiP, and (c) N1s of g-C3N4 and 5%-CN/BiP.
Molecules 30 02905 g002
Figure 3. SEM images for (a) BiPO4, (b) g-C3N4, (c) 1%-CN/BiP, (d) 2%-CN/BiP, (e) 5%-CN/BiP, (f) 10%-CN/BiP, and (g) 15%-CN/BiP.
Figure 3. SEM images for (a) BiPO4, (b) g-C3N4, (c) 1%-CN/BiP, (d) 2%-CN/BiP, (e) 5%-CN/BiP, (f) 10%-CN/BiP, and (g) 15%-CN/BiP.
Molecules 30 02905 g003
Figure 4. TEM images for (a) BiPO4, (b) g-C3N4, (c) 5%-CN/BiP, (d) HRTEM images, and (e) EDS line map of the sample 5%-CN/BiP.
Figure 4. TEM images for (a) BiPO4, (b) g-C3N4, (c) 5%-CN/BiP, (d) HRTEM images, and (e) EDS line map of the sample 5%-CN/BiP.
Molecules 30 02905 g004
Figure 5. FESEM-EDS mapping of 5%-CN/BiP.
Figure 5. FESEM-EDS mapping of 5%-CN/BiP.
Molecules 30 02905 g005
Figure 6. (a) DRS spectra of single BiPO4, g-C3N4, and composite CN/BiP with different mass ratios; and (b) (αhv)2/(αhv)1/2versus photon energy (hv) of the as prepared BiPO4, g-C3N4, and 5%-CN/BiP.
Figure 6. (a) DRS spectra of single BiPO4, g-C3N4, and composite CN/BiP with different mass ratios; and (b) (αhv)2/(αhv)1/2versus photon energy (hv) of the as prepared BiPO4, g-C3N4, and 5%-CN/BiP.
Molecules 30 02905 g006
Figure 7. Fluorescence emission spectra of BiPO4 and composite CN/BiP with different mass ratios.
Figure 7. Fluorescence emission spectra of BiPO4 and composite CN/BiP with different mass ratios.
Molecules 30 02905 g007
Figure 8. (a) Transient photocurrent response; and (b) electrochemical impedance spectroscopy of BiPO4, g-C3N4, and CN/BiP composite.
Figure 8. (a) Transient photocurrent response; and (b) electrochemical impedance spectroscopy of BiPO4, g-C3N4, and CN/BiP composite.
Molecules 30 02905 g008
Figure 9. (a) Degradation and (b) pseudo-first-order kinetics curves of TC over g-C3N4/BiPO4 samples with different molar ratios; as well as (c) degradation and (d) pseudo-first-order kinetics curves of OTC.
Figure 9. (a) Degradation and (b) pseudo-first-order kinetics curves of TC over g-C3N4/BiPO4 samples with different molar ratios; as well as (c) degradation and (d) pseudo-first-order kinetics curves of OTC.
Molecules 30 02905 g009
Figure 10. Recycling runs of 5%-CN/BiP for (a) TC and (b) OTC photodegradation.
Figure 10. Recycling runs of 5%-CN/BiP for (a) TC and (b) OTC photodegradation.
Molecules 30 02905 g010
Figure 11. Effects of different quenchers on the photodegradation of TC by 5%-CN/BiP under sunlight irradiation.
Figure 11. Effects of different quenchers on the photodegradation of TC by 5%-CN/BiP under sunlight irradiation.
Molecules 30 02905 g011
Figure 12. DMPO trapping EPR spectra of BiPO4 and 5%-CN/BiP composite in (a) aqueous and (b) methanol dispersion.
Figure 12. DMPO trapping EPR spectra of BiPO4 and 5%-CN/BiP composite in (a) aqueous and (b) methanol dispersion.
Molecules 30 02905 g012
Figure 13. Crystal structures, calculated band structures, and density of states of (a) BiPO4 and (b) g-C3N4; VB values of (c) BiPO4 and (d) g-C3N4 calculated by XPS; Work functions of (e) g-C3N4 and (f) BiPO4.
Figure 13. Crystal structures, calculated band structures, and density of states of (a) BiPO4 and (b) g-C3N4; VB values of (c) BiPO4 and (d) g-C3N4 calculated by XPS; Work functions of (e) g-C3N4 and (f) BiPO4.
Molecules 30 02905 g013
Figure 14. Schematic illustrations of the energy band structures upon the formation of g-C3N4 and BiPO4 heterojunction.
Figure 14. Schematic illustrations of the energy band structures upon the formation of g-C3N4 and BiPO4 heterojunction.
Molecules 30 02905 g014
Table 1. Photodegradation rate constants k of BiPO4, g-C3N4, and 5%-CN/BiP samples for TC and OTC.
Table 1. Photodegradation rate constants k of BiPO4, g-C3N4, and 5%-CN/BiP samples for TC and OTC.
Target Pollutantsg-C3N4BiPO45%-CN/BiP
TC7.12 × 10−37.13 × 10−43.71 × 10−2
OTC7.42 × 10−31.37 × 10−32.77 × 10−2
Table 2. Summary of photodegradation efficiency of various types of photocatalysts on TC.
Table 2. Summary of photodegradation efficiency of various types of photocatalysts on TC.
Type of PhotocatalystDosage
(g/L)
C0 of NA
(mg/L)
Light SourcesTime
(min)
Removal Efficiency (%)References
g-C3N4/BiPO4120λ > 420 nm9099.4%This study
Ag/Ag2CO3/BiVO4120λ > 40015094.9%[39]
Ag2CO3/Bi4O5I2/g-C3N42/320λ > 4006082.16%[40]
CuAl2O4/g-C3N40.2100λ > 4006089.6%[41]
g-C3N4/Bi2O3@N-HMCs210λ > 4206090.06%[42]
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

Zhu, X.; Luo, M.; Sun, C.; Jiang, J.; Guo, Y. Highly Efficient Photocatalytic Degradation of Tetracycline Antibiotics by BiPO4/g-C3N4: A Novel Heterojunction Nanocomposite with Nanorod/Stacked-like Nanosheets Structure. Molecules 2025, 30, 2905. https://doi.org/10.3390/molecules30142905

AMA Style

Zhu X, Luo M, Sun C, Jiang J, Guo Y. Highly Efficient Photocatalytic Degradation of Tetracycline Antibiotics by BiPO4/g-C3N4: A Novel Heterojunction Nanocomposite with Nanorod/Stacked-like Nanosheets Structure. Molecules. 2025; 30(14):2905. https://doi.org/10.3390/molecules30142905

Chicago/Turabian Style

Zhu, Xin, Moye Luo, Cheng Sun, Jinlin Jiang, and Yang Guo. 2025. "Highly Efficient Photocatalytic Degradation of Tetracycline Antibiotics by BiPO4/g-C3N4: A Novel Heterojunction Nanocomposite with Nanorod/Stacked-like Nanosheets Structure" Molecules 30, no. 14: 2905. https://doi.org/10.3390/molecules30142905

APA Style

Zhu, X., Luo, M., Sun, C., Jiang, J., & Guo, Y. (2025). Highly Efficient Photocatalytic Degradation of Tetracycline Antibiotics by BiPO4/g-C3N4: A Novel Heterojunction Nanocomposite with Nanorod/Stacked-like Nanosheets Structure. Molecules, 30(14), 2905. https://doi.org/10.3390/molecules30142905

Article Metrics

Back to TopTop