Solar-light-driven photocatalytic degradation and detoxification of ciprofloxacin using sodium niobate nanocubes decorated g-C3N4 with built-in electric field

Hui Wang Haodong Ji Dandan Zhang Xudong Yang Hanchun Chen Chunqian Jiang Weiliang Sun Jun Duan Wen Liu

Citation:  Hui Wang, Haodong Ji, Dandan Zhang, Xudong Yang, Hanchun Chen, Chunqian Jiang, Weiliang Sun, Jun Duan, Wen Liu. Solar-light-driven photocatalytic degradation and detoxification of ciprofloxacin using sodium niobate nanocubes decorated g-C3N4 with built-in electric field[J]. Chinese Chemical Letters, 2025, 36(5): 110200. doi: 10.1016/j.cclet.2024.110200 shu

Solar-light-driven photocatalytic degradation and detoxification of ciprofloxacin using sodium niobate nanocubes decorated g-C3N4 with built-in electric field

English

  • Broad-spectrum antibiotics, particularly second-generation fluoroquinolones, have been monitored worldwide in natural water matrices because of their high demand and extensive application in the treatment of diseases caused by bacterial infections [1]. Ciprofloxacin (CIP) is a typical fluoroquinolone antibiotic, which is frequently detected in the effluents of pharmaceutical manufacturing. The trace-level CIP with high ecotoxicity will severely threaten human health and the safety of the natural ecosystem [2]. Therefore, efficient treatment strategies and technologies are urgently required.

    Considering the deficiencies in the biodegradability and mineralization rate of antibiotics, conventional treatment processes using activated sludge cannot efficiently remove CIP, and transformation products (TPs) may have higher ecotoxicity [3-6]. Photocatalysis is a cost-efficient and environmentally friendly technology that has gained increasing attention for the treatment of antibiotics such as CIP [7]. Compared to other technologies, photocatalysis has specific advantages such as solar light utilization, moderate reaction conditions and low secondary pollution [8]. More importantly, photocatalysis can effectively mineralize various antibiotics to non-toxic final products such as H2O and CO2 [9]. Recently, metal-free graphitic carbon nitride (g-C3N4), synthesized via a facile thermal polymerization method, has been proposed as a promising photocatalyst owing to its medium band gap (~2.65 eV), high photogenerated carrier transfer efficiency, and good stability. Thus, g-C3N4 has been widely applied for photocatalytic degradation of various organic contaminants [10,11]. However, the practical application of g-C3N4 is limited by its deficiencies in light absorption and conversion in the visible light range, electron transfer efficiency and recoverability [12]. More importantly, the photoinduced electron-hole pairs generally possess a higher recombination ratio in g-C3N4, leading to a low utilization rate of photogenerated carriers [11,12].

    Various strategies have been developed to improve the photocatalytic activity of neat g-C3N4, In particular, heterostructure/heterojunction construction has been widely proven as an adequate approach for enhancing its photocatalytic activity [13]. Compared with other heterojunctions (type-Ⅱ, Z-scheme and S-scheme), the type-Ⅰ heterojunction facilitates the efficient separation of electron-hole pairs and controls the hole transfer direction and accumulation, resulting in the construction of a stable photocatalyst and the formation of a built-in electric field [14]. However, few studies reported type-Ⅰ heterojunctions because of the special band energy levels of the two semiconductors. Thus, it is necessary to precisely regulate type-Ⅰ heterojunctions coupled with g-C3N4.

    Sodium niobate (NaNbO3), which has thermoelectric, piezoelectric, and ferroelectric properties [15,16], is a potential candidate for environmental photocatalysis applications. In addition, NaNbO3 with a high valence band level exhibits strong oxidation ability owing to the generation of rich reactive species, such as photoinduced holes (h+) and hydroxyl radicals (OH), which are suitable for micropollutant degradation. In addition, incorporation of NaNbO3, as a new electron donor level, can increase the electrons density and transfer efficiency, thus the introduction of NaNbO3 with high photocatalytic ability in composite material is a potential method for improving photocatalytic performance [17]. Therefore, modification of g-C3N4 with NaNbO3 to construct type-Ⅰ heterojunctions will boost electron transfer, assuming the positive effect of “1 + 1 > 2” during the organic pollutant elimination process. In addition, the morphology and structure of NaNbO3 could be controlled by varying the synthesis conditions. Therefore, the in-situ coordinated 3-D cubic NaNbO3 from Na2Nb2O6·H2O shows better photocatalytic activities because of the crystal plane heterogeneity, thus facilitating photoexcited electron transfer in the type-Ⅰ heterojunction of g-C3N4 [18,19].

    In this study, an NaNbO3 nanocube-decorated g-C3N4 nanohybrid (NbNC/g-C3N4) was synthesized and used for the photocatalytic degradation of CIP under simulated solar light. Energy-band structures were designed to achieve efficient electron transfer. The photocatalytic detoxification efficiency of CIP was evaluated. The mechanism of reactive species production was revealed through material characterization and radical identification. In addition, the reaction mechanism of CIP attack by radicals was investigated via integrated experimental analysis and computational calculations.

    All chemicals and reagents used in the experiments were of analytical grade or higher without further purification, as provided in Text S1(Supporting information). NbNC/g-C3N4 was synthesized via a two-step route, i.e., Na2Nb2O6·H2O (precursor of NaNbO3) and g-C3N4 were prepared first, followed by a calcination process to achieve in-situ transformation and decoration of NaNbO3 onto g-C3N4. Detailed synthesis methods are shown in Text S2 (Supporting information). The synthesized materials were characterized by field-emission scanning electron microscopy (SEM), high-resolution transmission electron microscopy (HRTEM), X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), Fourier transform infrared spectroscopy. In addition, UV–vis spectrophotometry, photocurrent, electrochemical impedance spectroscopy (EIS), photoluminescence spectra (PL), linear sweep voltammetry (LSV) and Brunauer-Emmett-Teller (BET) nitrogen adsorption were also analyzed to obtain the diffuse reflectance spectra (DRS), photoelectrochemical properties, surface area, and pore properties of the catalysts. The detailed material characterization methods are described in Text S3 (Supporting information).

    Batch experiments on the photocatalytic performance were performed in a quartz reactor (volume = 250 mL) with a water-cooling system to maintain the reaction temperature at 25 ℃ (Fig. S1 in Supporting information). A Microsolar 300 solar-light simulator (Beijing Perfectlight, Beijing, China) coupled with a Xe lamp (300 W) was used and operated at a light irradiation of 100 ± 0.5 mW/cm2 in the AM 1.5 G mode. The typical photocatalytic test was conducted with 10 µmol/L CIP (100 mL) and 0.1 g/L catalysts in the reactor, and the pH was adjusted to 7.0 ± 0.1 by NaOH or HClO4 (0.1 mol/L). The CIP concentration was detected using a Dionex UltiMate 3000 high-performance liquid chromatograph (HPLC, Thermo Fisher Scientific). The TPs and degradation intermediates were determined using an ultra-HPLC equipped with an electrospray ionization source and a triple quadrupole mass spectrometer (Dionex UltiMate 3000 Series; MS, Thermo Scientific, USA) [20]. The detailed analytical methods are described in Text S4 (Supporting information).

    Density functional theory (DFT) calculations on CIP are performed using Gaussian 16 software (Version C.01) [21]. B3LYP combined with 6–31+G(d, p) is applied for structure optimization [22,23]. Details of the calculation method for the Fukui index are described in Text S5 (Supporting information). To investigate the theoretical ecotoxicity of CIP and its TPs [3]. The Cambridge Sequential Total Energy Package in Materials Studio 2020 was used to calculate the band gap, density of state (DOS), projected DOS (PDOS), work function, electron density, and electron density difference (EDD) of the materials (Text S6 in Supporting information) [24-26]. g-C3N4 exhibits a smooth surface and 2-D layered structure (Figs. 1a-g), while pristine Na2Nb2O6·H2O exhibited as 1-D nanorods (Figs. 1b and e), and the lattice distance of 0.15 nm displayed in the HRTEM image was indexed to the (200) plane of Na2Nb2O6·H2O (Fig. 1h) [27,28]. After calcination treatment for the in-situ formation and decoration of NaNbO3 onto g-C3N4, the Na2Nb2O6·H2O 1-D nanowires were transformed to NaNbO3 3-D nanocubes via dehydration (Fig. 1f) due to the more thermodynamically stability than NaNbO3 nanowires [27]. In addition, the surface of NbNC/g-C3N4–1 was coarse owing to the niobate decoration (Fig. 1c). The measured lattice distance of 0.39 nm in NbNC/g-C3N4–1 was assigned to the (110) plane of orthorhombic NaNbO3 (JCPDS 01–073–0803) (Fig. 1i), indicating the different crystalline structures for pristine Na2Nb2O6·H2O and NaNbO3 phase in NbNC/g-C3N4–1. The successful hybridization of the NaNbO3 nanocubes and g-C3N4 can enhance the utilization of solar light, thus facilitating electron transfer to achieve better photocatalytic effectiveness for CIP degradation. In addition, an interface connection between g-C3N4 and Na2Nb2O6 phases is clearly observed in Fig. 1i, allowing the electrons transfer between these two components and beneficial to the built-in electric field construction [29].

    Figure 1

    Figure 1.  (a-c) Field emission scanning electron microscopy (FESEM), (d-f) TEM and (g-i) HRTEM images of g-C3N4, Na2Nb2O6·H2O and NbNC/g-C3N4–1.

    In the XRD patterns (Fig. 2a), the significant peaks of pristine g-C3N4 at 13.0° and 27.5° were assigned to the (100) and (002) crystalline planes (JCPDS 01–087–1526), respectively, which can be characterized as the tri-s-triazine system in the interlayer structure and stacking of aromatic units [30]. For pristine Na2Nb2O6·H2O, the 2θ peaks at 11.4°, 12.7°, 29.1°, 30.4°, and 31.6° corresponded to monoclinic crystal Na2Nb2O6 with a C2/c space group (a = 17.114 Å, b = 5.0527 Å, and c = 16.5587 Å), comprising [NbO6] octahedra lattice [31]. For pristine NaNbO3 formed after thermal dehydration from Na2Nb2O6·H2O, new distinct peaks at 22.8°, 32.5°, 46.5°, 52.4°, and 57.9° appeared, assigned to the (110), (114), (220), (118), and (028) planes of the orthorhombic crystal structure (JCPDS 01–073–0803) (Fig. S2 in Supporting information), respectively, with the Pbcm space group (a = 5.506 Å, b = 5.566 Å, and c = 15.520 Å). An “octahedral tilting” [NbO6] octahedra corner-sharing network was displayed, indicating that structure fabrication along with phase transformation occurred during dehydration of Na2Nb2O6·H2O, i.e., the [NbO6] layers and chemical bonding (Nb–O–Nb and Nb–O–Na) were atomically rearranged to [NbO6] octahedra via the thermal excitation due to the minimization of Gibbs free energy, resulting in the formation of NaNbO3 nanocubes (Figs. 1e and f). For all NbNC/g-C3N4 nanohybrids, all the Na2Nb2O6·H2O peaks disappeared while the diffractions of NaNbO3 emerged, indicating the in-situ crystalline phase transformation via calcination.

    Figure 2

    Figure 2.  (a) XRD patterns of NbNC/g-C3N4 with different component ratios. The XPS spectra of NaNbO3, g-C3N4 and NbNC/g-C3N4–1: (b) The survey, (c) Nb 2p, (d) O 1s, (e) C 1s and (f) N 1s.

    Figs. 2b-f show the XPS spectra of materials. In the high-resolution Nb 3d spectrum (Fig. 2c), the two significant split peaks at 206.1 eV and 208.9 eV for pristine NaNbO3 are assigned to the Nb 3d3/2 and 3d5/2 orbitals, respectively. The spin-orbit separation energy of 2.8 eV corresponded to the [NbO6] octahedra [32]. There was a positive shift (+0.5 eV) for the two Nb orbitals for NbNC/g-C3N4–1 (Table S1 in Supporting information), indicating electron transfer from NaNbO3 to g-C3N4 and the formation of chemical bonds for Nb. In the high-resolution spectra of O 1s (Fig. 2d), the peaks at 529.6, 530.9, and 533.6 eV for pristine NaNbO3 indexed to the lattice O derived from the [NbO6] octahedra, Nb–OH, and chemisorbed O/H2O, respectively. For NbNC/g-C3N4–1, the atomic percentage of lattice O increased from 63.07 at% (NaNbO3) to 70.39 at% (NbNC/g-C3N4–1) (Table S1), indicating the formation of Nb–O–C or Nb–O–N, except the main chemical bond Nb–O–Nb in the [NbO6] octahedra [33]. In the high-resolution spectra of C 1s (Fig. 2e), the peaks at 284.8 eV and 288.3 eV in pristine g-C3N4 corresponded to C–C/C=C and N–C=N of the triazine rings, respectively [4]. The N–C=N peak decreased dramatically from 85.16 at% for g-C3N4 to 9.82 at% for NbNC/g-C3N4–1 (Table S1), indicating new bond formation for C. Moreover, a new peak at ca. 285.9 eV assigned to C–O (accounting for 38.33 at% of C) increased in NbNC/g-C3N4–1, which was in good agreement with the formation of Nb–O–C.

    Fig. S3a (Supporting information) shows the photocatalytic effectiveness of various photocatalysts for CIP degradation. The adsorption of CIP was first assessed for all materials (Figs. S3a and S4 in Supporting information). Neat g-C3N4 barely adsorbed CIP in 30 min, while neat NaNbO3 displayed a CIP adsorption efficiency of 39.0%. In addition, the NbNC/g-C3N4 composites showed CIP adsorption efficiencies of 10.0%-16.9%, indicating efficient interaction between CIP and NbNC/g-C3N4 materials. Then, a pseudo-first order model was used to interpret the kinetic results (Table S2 in Supporting information) [34,35]. 64.1% and 84.5% of CIP was degraded by pristine g-C3N4 and NaNbO3 with degradation kinetic rate constants (k1) of 0.052 and 0.076 min−1, respectively. Enhanced photocatalytic effectiveness of CIP degradation was observed using all NbNC/g-C3N4 nanohybrids. Specifically, NbNC/g-C3N4–1 (Na2Nb2O6·H2O: g-C3N4 mass ratio of 1:1) exhibited the highest k1 value (0.173 min−1), which was 3.3 and 2.3 times that of pristine g-C3N4 and NaNbO3, respectively. Fig. S3b (Supporting information) shows that the optimum NbNC/g-C3N4–1 dosage was 0.10 g/L, with a high CIP removal efficiency of 96.8%; further increasing the dosage to 0.20 g/L achieved a slight increase in the CIP removal efficiency to 99.3%.

    After scavenger quenching tests, Fig. 3a shows that the addition of Tiron and potassium iodide (KI) inhibited CIP degradation by 56.5% and 36.3% at 20 min, respectively, indicating that O2 played the most important role, whereas h+ contributed less (Table S3 in Supporting information). However, the addition of tert-butyl alcohol (tBA) slightly retarded the photocatalytic degradation of CIP by 2.6%, suggesting that OH plays a negligible role. The electron spin resonance (ESR) spectra further confirmed the generation of OH and O2 in the NbNC/g-C3N4–1 photocatalytic system under solar irradiation (Figs. 3b and c).

    Figure 3

    Figure 3.  (a) Photocatalytic degradation of CIP by NbNC/g-C3N4–1 in the presence of various scavengers (Initial CIP concentration = 10 µmol/L, catalyst dosage = 0.1 g/L, temperature = 25 ± 0.2 ℃, solution pH = 7.0 ± 0.1, scavenger dosage = 10 mmol/L). ESR spectra of (b) DMPO-OH and (c) DMPO-O2. (d) Eg values calculated from Kubelka-Munk method. (e) XPS-VB spectra and (f) photocurrent spectra for pristine NaNbO3, g-C3N4, and NbNC/g-C3N4–1. (g-i) EIS, PL and LSV curves of pristine NaNbO3, g-C3N4 and NbNC/g-C3N4–1.

    Based on the UV–vis DRS spectra (Fig. S5 in Supporting information), the measured Eg for NaNbO3, g-C3N4, and NbNC/g-C3N4–1 were 4.34, 2.81, and 3.75 eV, respectively (Fig. 3d). The synthesized NaNbO3 in this study had a markedly wider band gap energy; therefore, the highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) were introduced to describe its energy levels instead of the valance band (VB) and conduction band (CB). The light absorption edge of NbNC/g-C3N4–1 was red-shifted, with a narrower band gap (3.75 eV) than that of pristine NaNbO3 (286 nm, 4.34 eV), indicating the enhanced light utilization efficiency of the new material. Although g-C3N4 exhibited an absorption edge in the visible light region at 441 nm and an Eg of 2.81 eV, NbNC/g-C3N4–1 showed higher photocatalytic activity owing to efficient charge migration and electron-hole separation. In addition, the calculated EVB and ECB were 2.07 and −1.68 eV for NbNC/g-C3N4–1, respectively (Fig. 3e), and negative ECB values led to higher electron reactivity on CB, facilitating the production of O2. The photocurrent spectra clearly show that the photoelectric conversion efficiency of NbNC/g-C3N4–1 was better than that of pristine NaNbO3 and g-C3N4 (Fig. 3f). Besides, EIS spectra in Fig. 3g display that NbNC/g-C3N4–1 with a smaller arc radius possessed a lower resistance for charge transfer and higher charge transfer efficiency than pristine g-C3N4 and NaNbO3. Then, PL spectra in Fig. 3h also present that NbNC/g-C3N4–1 with a lower steady-state PL emission intensity than g-C3N4 and NaNbO3, suggesting NbNC/g-C3N4–1 had a lower recombination rate of photoinduced electron-hole pairs than the other two materials. Finally, LSV curves in Fig. 3i show that NbNC/g-C3N4–1 had a stronger current density and redox capacity than pristine g-C3N4 and NaNbO3, indicating the electrical conductivity and photocatalytic performance has been obviously enhanced after the heterojunction construction [36].

    To explore the mechanism on enhanced photocatalytic activity of NbNC/g-C3N4, the band structure, DOS/PDOS spectra, electron density, work function, and EDD were studied via DFT calculations (Fig. 4). The structural models of the catalysts are presented in Figs. S6a and b (Supporting information). The calculated band gap of pure g-C3N4 (Fig. 4a) and NaNbO3 (Fig. S6c in Supporting information) were 1.281 and 2.373 eV, respectively. The obtained band gap values are lower than the experimental values because of the known limitations of the plain DFT method [36]. Fig. 4d shows that NbNC/g-C3N4 possesses a more negative CB than pure g-C3N4, indicating that the CB of NbNC/g-C3N4 has a high electronic reactivity. In the PDOS spectra for NbNC/g-C3N4 (Fig. 4e), the N 2p and O 2p orbitals primarily contributed to the VB, whereas the C 2p and Nb 3d orbitals contributed to the CB. Hence, the electron transfer path in NbNC/g-C3N4 followed: N 2p → O 2p → Nb 3d → C 2p based on the PDOS distributions. Meanwhile, Fig. 4e shows that the PDOS of N 2p moves to the left for NbNC/g-C3N4 compared to that of pure g-C3N4, indicating that the electron density of N 2p increased after the introduction of NaNbO3 and more electrons were transferred along the above path under light irradiation. Moreover, the d-band center of Nb 3d in NbNC/g-C3N4 shifts to 0.40 eV compared with 0.44 eV of NaNbO3 (Fig. 4e and Fig. S6d in Supporting information), which can be attributed to the compositing with g-C3N4. Based on the PDOS of Nb 3d in Fig. 4e, an antibonding state is formed at nearly 2 eV compared to pure NaNbO3 (Fig. S6d), suggesting that dx2-y2 orbitals with higher coupling ability combine with p orbitals during the bond formation of Nb–O–N/Nb–O–C. Accordingly, the introduction of g-C3N4 changes the electronic structure of Nb 3d orbitals, further leading to the alteration of d band center [37]. Therefore, the d-band center of NbNC/g-C3N4 was markedly closer to the Fermi level, suggesting that the electrons in Nb 3d were more reactive and could be more effectively transferred to C 2p with higher catalytic activity [38]. The electron density images (Figs. 4c and f) show higher electron densities in the N 2p and C 2p spectra of NbNC/g-C3N4 than those of g-C3N4, which benefits the photocatalytic reaction. The work function (Fig. 4g) indicates that NaNbO3 with a low value (4.488 eV) transfers electrons to g-C3N4 with a high value (4.689 eV) when the NbNC/g-C3N4 heterojunction is constructed. EDD analysis indicates that electrons are more likely to transfer from NaNbO3 to g-C3N4 (Fig. 4h); thus, holes accumulated on NaNbO3 and electrons gathers on g-C3N4 [38,39]. The EDD result clearly shows a 0.37 e transfer from NaNbO3 to g-C3N4 based on the Mulliken population analysis of the built-in electric field intensity. Furthermore, because the e transfer direction is from NaNbO3 to g-C3N4 based on the EDD result in NaNbO3/g-C3N4 composite, the e density is greatly increased in g-C3N4, and the built-in electric field is also constructed because of their electrons density difference. Thus, g-C3N4 can be likened to a switch with a large amount of e, and more e transfer occurs along N 2p → O 2p → Nb 3d → C 2p chain under light irradiation, leading to a high e utilization rate and activity [40]. Thus, the electron density increased in g-C3N4 and the formed built-in electric field significantly boost the separation rate of charge carriers, further promoting photocatalytic activity [39].

    Figure 4

    Figure 4.  (a) The band structure, (b) DOS spectra and (c) electron density of pure g-C3N4. (d) The band structure, (e) DOS spectra and (f) electron density of NbNC(110)/g-C3N4. (g) The work function of g-C3N4 and NaNbO3. (h) EDD of NbNC(110)/g-C3N4 (yellow region: electrons depletion; cyan region: electrons accumulation).

    Fig. S7 (Supporting information) shows a schematic of efficient charge migration at the interface of NbNC/g-C3N4–1 nanohybrid during photocatalysis. Based on the energy levels of NaNbO3 and g-C3N4, the NbNC/g-C3N4–1 structure is a type-Ⅰ heterojunction. Then, under light irradiation, the photogenerated electrons and holes transfer to the same side [14,41]. Specifically, the LUMO potential of NaNbO3 (−1.93 V) is more negative than the CB of g-C3N4 (−0.84 V; Fig. S7), which helps the movement of photogenerated electrons from the LUMO of NaNbO3 to the CB of g-C3N4. In addition, the photogenerated holes cannot move from the VB of g-C3N4 (+1.97 V) to the HOMO of NaNbO3 (+2.41 V) because of the greater negativity of the VB of g-C3N4. The transformation of these photogenerated electrons by heterojunctions and the generation of holes in the HOMO and VB can reduce the charge recombination rate and provide more electrons/holes on the surface of the photocatalyst [14,41]. The photocurrent intensity in Fig. 3f and the DFT calculations in Fig. 4 prove the efficient separation of electron-hole pairs compared with neat g-C3N4 and NaNbO3; thus, an internal electric field can also be generated during electron transfer from NaNbO3 to g-C3N4.

    To accurately evaluate the reactive sites of the CIP molecule, DFT calculations of CIP were conducted based on the natural population analysis charge distribution (Fig. 5). Fig. 5b shows the HOMO is placed on the piperazine ring, and the C atom of the benzene ring is bonded with the F atom, which are the most reactive sites for electron loss [42]. Furthermore, the piperazine ring side of CIP± with a high positive ESP at neutral pH preferred to attach on the negative-charged surface of NbNC/g-C3N4–1 (Fig. 5c). Fig. 5d indicates that the most reactive sites of CIP are C5, C8, N10, O17, N19, and N24, with high Fukui indices representing electrophilic attack, i.e., f = 0.0417, 0.1008, 0.0719, 0.0990, 0.0831, and 0.1472, respectively. O2 is the dominant ROS in this photocatalysis system [43]. Therefore, electrophilic attack for CIP± oxidation is focused.

    Figure 5

    Figure 5.  (a) CIP± chemical structure. (b) HOMO and LUMO of CIP±. (c) ESP of CIP±. (d) NPA charge distribution and Fukui index (f ) of CIP±. (e) Proposed pathways on photocatalytic degradation of CIP by NbNC/g-C3N4–1 under solar light. (Initial CIP concentration = 10 µmol/L, catalyst dosage = 0.1 g/L, temperature = 25 ± 0.2 ℃, solution pH = 7.0 ± 0.1). The white, gray, blue, red, and cyan balls denote H, C, N, O and F atoms.

    Table S4 and Fig. S8 (Supporting information) present the TPs of CIP during photocatalytic degradation, and the four CIP degradation pathways are shown in Fig. 5e. Pathway Ⅰ involves the photoinduced substitution of fluorine by the hydroxyl moiety for the formation of TP330 (m/z 330.06). The reaction occurred with the breakage of the C–F bond by attack at the C5 site (f = 0.0417). Pathway Ⅱ involved defluorination and oxidation of the quinolone ring, leading to the formation of TP346 (m/z 346.06). TP330 and TP346 were generated via attack on the C5 (f = 0.0417) and C8 (f = 0.1008) sites. Pathway Ⅲ involved the cleavage of the piperazine moiety via oxidation, which was attributed to the attack on N19 (f = 0.0831) and N24 (f = 0.1472) sites with a high Fukui index. TP362 (m/z 336.03) was a key intermediate in this pathway. Pathway Ⅳ involved deep defluorination after dealkylation of the piperazine moiety in Pathway Ⅲ. TP344 (m/z 344.04) was a key intermediate formed upon the ring opening of the piperazine moiety and defluorination of CIP or TP362. Finally, mineralization of these TPs led to formation of inorganic small molecules/ions such as CO2, H2O, NH4+ and F [44].

    Five consecutive CIP degradation tests were performed to evaluate the reusability and stability of NbNC/g-C3N4–1 (Fig. S9 in Supporting information), and the CIP removal efficiency just slightly decreased from 97.1% to 90.1%, indicating the good reusability and high stability of NbNC/g-C3N4–1. In addition, the physicochemical properties of NbNC/g-C3N4–1 after reaction were analyzed by XRD and XPS (Fig. S10 in Supporting information), which also indicate slight changes on crystal phase and composition of the material after photocatalysis [20]. In addition, negligible Na and Nb leached during the photocatalytic reaction based on ICP-OES analysis, indicating good reusability and high stability of NbNC/g-C3N4–1.

    In this study, NbNC/g-C3N4 nanohybrid was synthesized through in-situ transformation of Na2Nb2O6·H2O at the interface of g-C3N4, which exhibited as type-Ⅰ heterojunction. The optimized NbNC/g-C3N4–1 showed high photocatalytic activity for efficient photocatalytic degradation and detoxification of CIP under simulated solar light. In the NbNC/g-C3N4 nanohybrid, the electron transfer path was as follows: N 2p → O 2p → Nb 3d → C 2p. Electrons can be effectively transferred along the Nb–O–N/Nb–O–C bonds, which greatly facilitates the migration of photoexcited electrons/holes. Meanwhile, the electron density increased in g-C3N4 to form a built-in electric field, because 0.37 e transferred from NaNbO3 to g-C3N4 based on Mulliken population analysis, which can significantly boost the separation rate of charge carriers. Upon solar light irradiation, simultaneous degradation and detoxification of CIP by the NbNC/g-C3N4 nanohybrid was achieved, implying its considerable application potential in practical wastewater treatment.

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

    Hui Wang: Writing – original draft, Validation, Resources, Methodology, Investigation, Funding acquisition, Conceptualization. Haodong Ji: Writing – review & editing, Validation, Supervision, Methodology, Funding acquisition, Data curation. Dandan Zhang: Validation, Investigation, Formal analysis, Data curation. Xudong Yang: Validation, Software. Hanchun Chen: Validation, Investigation. Chunqian Jiang: Writing – review & editing, Supervision, Funding acquisition, Conceptualization. Weiliang Sun: Validation, Methodology, Investigation, Data curation. Jun Duan: Validation, Methodology, Data curation. Wen Liu: Writing – review & editing, Validation, Supervision, Project administration, Methodology, Funding acquisition, Data curation, Conceptualization.

    Financial supports from the National Key Research and Development Program of China (Nos. 2021YFA1202500 and 2022YFF1303004), Shenzhen Science and Technology Program (No. JCYJ20220531093205013), the National Natural Science Foundation of China (NSFC) (Nos. 52100069, 52270053 and 52200084), the Beijing Natural Science Foundation (No. 8232035), the Beijing Nova Program (No. 20220484215), the Beijing National Laboratory for Molecular Sciences (No. BNLMS2023011) and Emerging Engineering Interdisciplinary-Young Scholars Project (Peking University), the Fundamental Research Funds for the Central Universities are greatly acknowledged. DFT calculations supported by the High-Performance Computing Platform of Peking University and the National Key Scientific and Technological Infrastructure project “Earth System Numerical Simulation Facility” (EarthLab) are also acknowledged. The work is also supported by the program of “Research on Advanced Treatment Technology of New Pollutants in Domestic Sewage of Residential District”.

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2024.110200.


    1. [1]

      B. Zhou, L. Chen, F. Li, et al., Chin. Chem. Lett. 34 (2023) 108558. doi: 10.1016/j.cclet.2023.108558

    2. [2]

      F. Sun, X. Yang, F. Shao, et al., Chin. Chem. Lett. 34 (2023) 108563. doi: 10.1016/j.cclet.2023.108563

    3. [3]

      C. Dang, F. Sun, H. Jiang, et al., J. Hazard. Mater. 400 (2020) 123225. doi: 10.1016/j.jhazmat.2020.123225

    4. [4]

      W. Liu, Y. Li, F. Liu, et al., Water Res. 151 (2019) 8–19. doi: 10.15302/j-laf-1-010002

    5. [5]

      J. Richard, A. Boergers, C. vom Eyser, K. Bester, J. Tuerk, Int. J. Hyg. Environ. Health. 217 (2014) 506–514. doi: 10.1016/j.ijheh.2013.09.007

    6. [6]

      H. Ji, T. Wang, T. Huang, B. Lai, W. Liu, J. Cleaner Prod. 278 (2021) 123924. doi: 10.1016/j.jclepro.2020.123924

    7. [7]

      S. Li, T. Huang, P. Du, W. Liu, J. Hu, Water Res. 185 (2020) 116286. doi: 10.1016/j.watres.2020.116286

    8. [8]

      J. Xu, P. Liang, X. Shen, et al., Sep. Purif. Technol. 339 (2024) 126734. doi: 10.1016/j.seppur.2024.126734

    9. [9]

      O. Baaloudj, I. Assadi, N. Nasrallah, et al., J. Water Process Eng. 42 (2021) 102089. doi: 10.1016/j.jwpe.2021.102089

    10. [10]

      Y. Liu, L. Chen, X. Liu, et al., Chin. Chem. Lett. 33 (2022) 1385–1389. doi: 10.3390/cryst12101385

    11. [11]

      A. Kumar, P. Raizada, A. Hosseini-Bandegharaei, et al., J. Mater. Chem. A 9 (2021) 111–153. doi: 10.1039/d0ta08384d

    12. [12]

      A. Balakrishnan, E.S. Kunnel, R. Sasidharan, M. Chinthala, A. Kumar, Chem. Eng. J. 475 (2023) 146163. doi: 10.1016/j.cej.2023.146163

    13. [13]

      D. He, Z. Zhang, Y. Xing, et al., Chem. Eng. J. 384 (2020) 123258. doi: 10.1016/j.cej.2019.123258

    14. [14]

      T. Muhmood, M. Xia, W. Lei, F. Wang, Appl. Catal. B: Environ. 238 (2018) 568–575. doi: 10.1016/j.apcatb.2018.07.029

    15. [15]

      Y.F. Xu, M.Z. Yang, B.X. Chen, et al., J. Am. Chem. Soc. 139 (2017) 5660–5663. doi: 10.1021/jacs.7b00489

    16. [16]

      E. Grabowska, Appl. Catal. B: Environ. 186 (2016) 97–126. doi: 10.1016/j.apcatb.2015.12.035

    17. [17]

      X. Liu, P. Du, W. Pan, et al., Appl. Catal. B: Environ. 231 (2018) 11–22. doi: 10.1016/j.apcatb.2018.02.062

    18. [18]

      P. Li, H. Xu, L. Liu, et al., J. Mater. Chem. A 2 (2014) 5606–5609. doi: 10.1039/C4TA00105B

    19. [19]

      P. Li, S. Ouyang, Y. Zhang, T. Kako, J. Ye, J. Mater. Chem. A 1 (2013) 1185–1191. doi: 10.1039/C2TA00260D

    20. [20]

      D. Zhang, J. Qi, H. Ji, et al., Chem. Eng. J. 400 (2020) 125918. doi: 10.1016/j.cej.2020.125918

    21. [21]

      M.J. Frisch, G.W. Trucks, H.B. Schlegel, et al., Gaussian 16 Rev. C.01, 2016.

    22. [22]

      X. Zeng, J. Zhu, G. Zhang, et al., Chem. Eng. J. 468 (2023) 143536. doi: 10.1016/j.cej.2023.143536

    23. [23]

      S. He, E. Wu, M. Shen, ACS ES&T Eng. 3 (2023) 651–660. doi: 10.1021/acsestengg.2c00373

    24. [24]

      M.D. Segall, J.D.L. Philip, M.J. Probert, et al., J. Phy. Condens. Matter 14 (2002) 2717. doi: 10.1088/0953-8984/14/11/301

    25. [25]

      S. Gao, H. Ji, P. Yang, et al., Small 19 (2023) 2206114. doi: 10.1002/smll.202206114

    26. [26]

      S. Gao, R. Wu, M. Sun, et al., Appl. Catal. B: Environ. 324 (2023) 122260. doi: 10.1016/j.apcatb.2022.122260

    27. [27]

      A. Yu, J. Qian, L. Liu, H. Pan, X. Zhou, Appl. Surf. Sci. 258 (2012) 3490–3496. doi: 10.1016/j.apsusc.2011.08.133

    28. [28]

      M. Nyman, A. Tripathi, J.B. Parise, R.S. Maxwell, T.M. Nenoff, J. Am. Chem. Soc. 124 (2002) 1704–1713. doi: 10.1021/ja017081z

    29. [29]

      Y. Yang, S. Zhao, F. Bi, et al., Appl. Catal. B: Environ. 315 (2022) 121550. doi: 10.1016/j.apcatb.2022.121550

    30. [30]

      A. Kumar, P. Raizada, P. Singh, R.V. Saini, A.K. Saini, Chem. Eng. J. 391 (2020) 123496. doi: 10.1016/j.cej.2019.123496

    31. [31]

      J.H. Jung, C.Y. Chen, W.W. Wu, et al., J. Phys. Chem. C 116 (2012) 22261–22265. doi: 10.1021/jp308289r

    32. [32]

      L. Yan, T. Zhang, W. Lei, et al., Catal. Today 224 (2014) 140–146. doi: 10.1016/j.cattod.2013.11.033

    33. [33]

      R. Radha, Y. Ravi Kumar, M. Sakar, K. Rohith Vinod, S. Balakumar, Appl. Catal. B: Environ. 225 (2018) 386–396. doi: 10.1016/j.apcatb.2017.12.004

    34. [34]

      H. Ji, Y. Gong, J. Duan, D. Zhao, W. Liu, Mar. Pollut. Bull. 135 (2018) 427–440. doi: 10.1016/j.marpolbul.2018.07.047

    35. [35]

      H. Ji, Y. Zhu, W. Liu, et al., Colloids Surf. A: Physicochem. Eng. Asp. 561 (2019) 373–380. doi: 10.1016/j.colsurfa.2018.10.048

    36. [36]

      X. Yang, F. Li, W. Liu, et al., Appl. Catal. B: Environ. 324 (2023) 122202. doi: 10.1016/j.apcatb.2022.122202

    37. [37]

      J. Qi, X. Yang, P.Y. Pan, et al., Environ. Sci. Technol. 56 (2022) 5200–5212. doi: 10.1021/acs.est.1c08806

    38. [38]

      H. Shi, G. Chen, C. Zhang, Z. Zou, ACS Catal. 4 (2014) 3637–3643. doi: 10.1021/cs500848f

    39. [39]

      X. Yang, J. Duan, J. Qi, et al., J. Hazard. Mater. 445 (2023) 130576. doi: 10.1016/j.jhazmat.2022.130576

    40. [40]

      H. Zhang, Y. Wang, S. Zuo, et al., J. Am. Chem. Soc. 143 (2021) 2173–2177. doi: 10.1021/jacs.0c08409

    41. [41]

      J. Low, J. Yu, M. Jaroniec, S. Wageh, A.A. Al-Ghamdi, Adv. Mater. 29 (2017) 1601694. doi: 10.1002/adma.201601694

    42. [42]

      J. Duan, L. Chen, H. Ji, et al., Chin. Chem. Lett. 33 (2022) 3172–3176. doi: 10.1016/j.cclet.2021.11.072

    43. [43]

      F. De Vleeschouwer, V. Van Speybroeck, M. Waroquier, P. Geerlings, F. De Proft, Org. Lett. 9 (2007) 2721–2724. doi: 10.1021/ol071038k

    44. [44]

      Y. Wang, X. Sun, J. Ma, et al., Sep. Purif. Technol. 337 (2024) 126392. doi: 10.1016/j.seppur.2024.126392

  • Figure 1  (a-c) Field emission scanning electron microscopy (FESEM), (d-f) TEM and (g-i) HRTEM images of g-C3N4, Na2Nb2O6·H2O and NbNC/g-C3N4–1.

    Figure 2  (a) XRD patterns of NbNC/g-C3N4 with different component ratios. The XPS spectra of NaNbO3, g-C3N4 and NbNC/g-C3N4–1: (b) The survey, (c) Nb 2p, (d) O 1s, (e) C 1s and (f) N 1s.

    Figure 3  (a) Photocatalytic degradation of CIP by NbNC/g-C3N4–1 in the presence of various scavengers (Initial CIP concentration = 10 µmol/L, catalyst dosage = 0.1 g/L, temperature = 25 ± 0.2 ℃, solution pH = 7.0 ± 0.1, scavenger dosage = 10 mmol/L). ESR spectra of (b) DMPO-OH and (c) DMPO-O2. (d) Eg values calculated from Kubelka-Munk method. (e) XPS-VB spectra and (f) photocurrent spectra for pristine NaNbO3, g-C3N4, and NbNC/g-C3N4–1. (g-i) EIS, PL and LSV curves of pristine NaNbO3, g-C3N4 and NbNC/g-C3N4–1.

    Figure 4  (a) The band structure, (b) DOS spectra and (c) electron density of pure g-C3N4. (d) The band structure, (e) DOS spectra and (f) electron density of NbNC(110)/g-C3N4. (g) The work function of g-C3N4 and NaNbO3. (h) EDD of NbNC(110)/g-C3N4 (yellow region: electrons depletion; cyan region: electrons accumulation).

    Figure 5  (a) CIP± chemical structure. (b) HOMO and LUMO of CIP±. (c) ESP of CIP±. (d) NPA charge distribution and Fukui index (f ) of CIP±. (e) Proposed pathways on photocatalytic degradation of CIP by NbNC/g-C3N4–1 under solar light. (Initial CIP concentration = 10 µmol/L, catalyst dosage = 0.1 g/L, temperature = 25 ± 0.2 ℃, solution pH = 7.0 ± 0.1). The white, gray, blue, red, and cyan balls denote H, C, N, O and F atoms.

  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  163
  • HTML全文浏览量:  10
文章相关
  • 发布日期:  2025-05-15
  • 收稿日期:  2024-03-12
  • 接受日期:  2024-07-01
  • 修回日期:  2024-06-14
  • 网络出版日期:  2024-07-01
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章