Synthesis of carbon quantum dots decorating Bi2MoO6 microspherical heterostructure and its efficient photocatalytic degradation of antibiotic norfloxacin

Meijuan Chen Liyun Zhao Xianjin Shi Wei Wang Yu Huang Lijuan Fu Lijun Ma

Citation:  Meijuan Chen, Liyun Zhao, Xianjin Shi, Wei Wang, Yu Huang, Lijuan Fu, Lijun Ma. Synthesis of carbon quantum dots decorating Bi2MoO6 microspherical heterostructure and its efficient photocatalytic degradation of antibiotic norfloxacin[J]. Chinese Chemical Letters, 2024, 35(8): 109336. doi: 10.1016/j.cclet.2023.109336 shu

Synthesis of carbon quantum dots decorating Bi2MoO6 microspherical heterostructure and its efficient photocatalytic degradation of antibiotic norfloxacin

English

  • Recently, antibiotic contamination has attracted a lot of attention [1]. As one of them, norfloxacin is widely used worldwide as a therapeutic drug and growth promoter due to its excellent properties of low price and stable efficacy, but its massive utilization brings a lot of environmental problems [2]. Moreover, the inefficient operation of conventional wastewater treatment plants in eliminating norfloxacin results in the discharge of norfloxacin into various water sources and causes water pollution [1]. Relevant investigations showed that 960 ng/L and 862 ng/L of norfloxacin were detected in a sewage treatment plant in Finland and Beijing (China), respectively, 2366 ng/L of norfloxacin was detected in sewage in Spain, 2200 ng/L in surface water in Kenya, 1070 ng/L in sewage in Hong Kong, China, and 989 ng/L in rivers in Shanghai, China [1,3]. The input of norfloxacin in the aquatic environment has a negative impact on aquatic organisms, resulting in high ecological and health risks [4]. In an attempt to address concerns about norfloxacin contamination, various technologies have been developed, including adsorption, biodegradation, and photocatalysis [57]. Among them, photocatalysis is of great interest owing to inexpensive and stable and is now widely used in water treatment research [8]. The key to the photocatalytic removal of norfloxacin in water is the development of suitable photocatalysts that can absorb natural light to achieve efficient mineralization of norfloxacin in complex environments.

    Many recent studies have been devoted to the development of novel antibiotic-degrading materials, but they all have suffered from some limitations, such as the OCN-MnFe2O4 photocatalyst synthesized by Zeng et al. [9] which could only achieve 42.7% ciprofloxacin degradation efficiency at 1 h; MIL-88A(Fe)/cotton fibers (MC) photocatalyst synthesized by Wang et al. [10] which could only achieve about 30% tetracycline degradation efficiency at 8 min and which must be accompanied by the addition of the oxidant peroxynitrite to achieve a better degradation efficiency; The Co3O4 NPs@N-PC catalyst synthesized by Mohtasham et al. [11] also required the addition of oxidant peroxymonosulfate to achieve better degradation efficiency for sulfamethoxazole; The g-C3N4/Ni0.5Zn0.5Fe2O4 catalyst prepared by Dhiman et al. [12] was only able to achieve a better antibiotic removal efficiency when the reaction was up to 60 min long. In view of this situation, the most basic Aurivillius oxide, Bi2MoO6 displays promising photocatalytic performance towards the breakdown over organic contaminants [13]. Nevertheless, the photocatalytic performance of pure Bi2MoO6 is restricted due to its underutilization of visible light (Bi2MoO6 band gap of 2.60–2.90 eV) and the fast recombination of photogenerated electrons and holes [13]. Several publications have reported that the Bi2MoO6-based photocatalysts, including those combined with different semiconductors to fabricate composites, such as Bi2MoO6/MoO3 [14], Bi2MoO6/WO3 [15], and Bi2MoO6/ZnO [16], showed superior photocatalyst performance because of its fast electron-hole separation. But they still suffered from the limitation of the insufficient visible light response.

    Carbon quantum dots (CQDs) has attracted a great deal of interest due to their high visible light absorption intensity, photostability, biocompatibility, and low cytotoxicity [17]. Recently, the composites by coupling CQDs with Bi2MoO6 have been proposed as a promising approach to enhance electron-hole separation [18]. For example, Di et al. [19] prepared CQDs/Bi2MoO6 composites and found that it had 5 times higher photocatalytic activity of pure Bi2MoO6. The degradation efficiency on ciprofloxacin reached 88% within 120 min. In the CQDs/Bi2MoO6 synthesized by Sun et al. [20], although the introduction of CQDs increased the light absorption intensity, it showed ignorable effect on the extending of absorption edge because all the CQDs/Bi2MoO6 samples showed the same absorption edge of 479 nm. Therefore, while modifying Bi2MoO6 with CQDs to improve photoelectron-hole charge separation, it is also necessary to further improve the visible light absorption performance. Furthermore, the size of the catalyst surface area largely determines the adsorption and conversion efficiency of pollutants by catalysts [19]. However, the reported synthesized CQDs/Bi2MoO6 catalyst have relatively small specific surface areas. Considering the relatively low specific surface area and narrow light absorption edge in previous reports, it is expected that the photocatalytic performance of CQDs/Bi2MoO6 could be further improved by enhancing these two properties. Therefore, this study focused on exploring feasible preparation methods to synthesize CQDs/Bi2MoO6 with a wider range of light absorption and larger specific surface area.

    In this study, an alcohol-thermal method was adopted to synthesize three-dimensional microspherical CQDs/Bi2MoO6 heterostructures. The chemical composition, microscopic morphology, pore structure, chemical valence, and optical properties of the synthesized CQDs/Bi2MoO6 were investigated. The photocatalytic degradation and mineralization performance on antibiotic norfloxacin were examined, from which the optimal loading amount of CQDs was determined. The photocatalytic mechanism was proposed accordingly. Moreover, in order to evaluate the application potential of the synthesized CQDs/Bi2MoO6 in real aquatic environment, optimization experiments including norfloxacin concentration, catalyst dosage, pH, and coexisting anion, were further investigated.

    The three-dimensional microspherical CQDs/Bi2MoO6 heterostructures were synthesized using a simple alcohol-thermal method. CQDs/Bi2MoO6 with different CQDs loading amount were synthesized, in which CQDs were mainly added by taking the supernatant obtained from the hydrothermal method as a stock solution, and the CQDs loading was adjusted by controlling the volume of added CQDs stock solution, and the samples synthesized with the addition of 0, 100, 200, and 300 µL of the CQDs stock solution were named Bi2MoO6, 100-CQDs/Bi2MoO6, 200-CQDs/Bi2MoO6, and 300-CQDs/Bi2MoO6, respectively. Detailed materials, synthesis, characterization and photocatalytic activity measurement was shown in Supporting information. With X-ray diffraction (XRD), the phase structures of Bi2MoO6 and CQDs/Bi2MoO6 composites were measured. Fig. 1a demonstrated that the pure Bi2MoO6 showed obvious and sharp characteristic peaks at 2θ of 28.2°, 32.5°, 46.7°, 55.4°, and 58.5°, which coincided with the orthogonal phase Bi2MoO6 (JCPDS No. 21–0102) planes (131), (200), (202), (331), and (262). No other peak was observed in the Bi2MoO6 sample, suggesting no impurity in the synthesized Bi2MoO6. In the CQDs/Bi2MoO6 composites, all peaks of Bi2MoO6 were determined and showed a shift to a lower angle, which was probably due to the existence of strong interfacial interactions between the CQDs and Bi2MoO6 resulting in a slight distortion of the Bi2MoO6 lattice [21]. In addition, with the introduction of CQDs, the peaks of Bi2MoO6 showed slight weakening and broadening, indicating that its crystallinity deteriorated. Notably, no characteristic peak of CQDs was identified, which might be attributed to the CQDs/Bi2MoO6 composites' low concentration and high dispersion of CQDs [22]. The same observations were obtained from other similar systems [20].

    Figure 1

    Figure 1.  Phase structure and chemical composition of the as-prepared catalysts. (a) The XRD patterns. (b) The FTIR spectra. (c) Mo 3d spectra. (d) Bi 4f spectra. (e) O 1s spectra. (f) C 1s spectra.

    Fig. 1b showed the Fourier transform infrared (FT-IR) spectra of different samples, from which it is clear that the absorption peaks of the Bi2MoO6 sample appeared at 1620, 840, 729, 557, and 441 cm−1. Among the five absorption peaks, the adsorbed water molecule's H-O bending vibration was at 1620 cm−1. The Mo-O bond stretching vibration induced a peak of 840 cm−1 [8], there was an asymmetric stretching pattern of the equatorial oxygen atoms in the MoO66− layer resulting in a peak at 729 cm−1 [23], on 557 cm−1 the absorption peak was originated from the MoO66− based bending vibration [24], while the 441 cm−1 absorption peak was responsible for stretching vibrations of Bi-O bonds in Bi2MoO6. In all CQDs/Bi2MoO6 composites, the five absorption peaks of Bi2MoO6 were clearly observed, as well as a new absorption peak at 1390 cm−1 appeared. A 1390 cm−1 peak was found to correlate with the C-O-C tension vibration, indicating that CQDs was present in all CQDs/Bi2MoO6 composites [25].

    The high-resolution X-ray photoelectron spectroscopy (XPS) spectra for Bi2MoO6 as well as 200-CQDs/Bi2MoO6 were evident that both of them contained binding energy peaks for four elements: O, Mo, Bi and C (Fig. S2 in Supporting information), where C in Bi2MoO6 was mainly derived from contaminated carbon in the environment and instruments [26]. In Fig. 1c, the 232.4 eV and 235.5 eV signature peaks correspond to Mo 3d5/2 and Mo 3d3/2, indicating that Mo6+ was present in Bi2MoO6. The 159.2 eV and 164.5 eV characteristic peaks of Fig. 1d were induced by Bi 4f7/2 and Bi 4f5/2, demonstrating the existence of Bi as Bi3+ [13]. In Fig. 1e, the distinctive peaks at 530.0 eV, 530.6 eV, and 531.4 eV, respectively, were the Bi-O, Mo-O, and C = O bonds [27,28]. The separation of the C 1s peaks in Fig. 1f occurred at energies corresponding to the C-C sp2-hybridized carbon, the C-O-C bond, and the C=O bond of CQDs, respectively, at 284.8 eV, 286.2 eV, and 288.4 eV [29], proving that the composite catalyst was effectively infused with CQDs, which was consistent with the FTIR analysis. It was observed that after the decorating with CQDs, the characteristic peak of Mo 3d5/2 shifted from 232.4 eV to 232.3 eV, whereas the Mo 3d3/2 shifted from 235.5 eV to 235.4 eV. Simultaneously, the characteristic peak of Mo-O bonds in O 1s shifted from 530.6 eV to 530.9 eV. Obviously, these shifts were attributed to the presence of CQDs, suggesting the formation of Mo-O-C bonds [26,30].

    The scanning electron microscopy (SEM) image of obtained sample suggested that Bi2MoO6 was a three-dimensional microsphere composed of nanorods (Fig. 2a). With its rough surface, there should be plenty of active sites on the catalyst surface according to the Bi2MoO6's huge specific surface area [23]. The prepared CQD's image was shown using high-resolution transmission electron microscopy (HRTEM) in Fig. 2b. The 0.33 nm lattice stripe corresponded with (002) crystal plane on CQDs [19], suggesting that the hydrothermal approach had been successfully used to create CQDs. Moreover, as shown, the CQDs were made up of uniformly small particles with a diameter of around 6 nm. Fig. 2c displayed the TEM picture of the 200-CQDs/Bi2MoO6 composite, which can be clearly seen as a rod-like structure. From 200-CQDs/Bi2MoO6 HRTEM images shown as Fig. 2d, 0.23 and 0.28 nm lattice spacing respectively matched the (100) crystal plane of CQDs [31] and that of Bi2MoO6 (200) crystal plane. The obvious interface between the lattice of CQDs and Bi2MoO6 demonstrated that the synthesized CQDs/Bi2MoO6 is a heterojunction structure, where the heterojunction structure was reported to be able to facilitate electron-hole pairs separation and lengthen photogenerated electron and hole lifetimes [19].

    Figure 2

    Figure 2.  Morphology, pore structure, optical absorption and electrochemical properties of the as-prepared catalysts. (a) SEM images of Bi2MoO6. (b) HRTEM images of CQDs. (c) TEM image of the 200-CQDs/Bi2MoO6. (d) HRTEM images of 200-CQDs/Bi2MoO6. (e, f) Nitrogen adsorption isotherm of the Bi2MoO6 and 200-CQDs/Bi2MoO6. (g) The UV–vis absorption spectra of the Bi2MoO6 and CQDs/Bi2MoO6. (h) The plots of (αhν)2 versus energy () of the Bi2MoO6 and CQDs/Bi2MoO6. (i) The transient photocurrent response of Bi2MoO6 and 200-CQDs/Bi2MoO6.

    Figs. 2e and f illustrated the information about the pore size and nitrogen (N2) adsorption isotherms on the prepared Bi2MoO6 and 200-CQDs/Bi2MoO6. The nitrogen adsorption-desorption isotherms of both Bi2MoO6 and 200-CQDs/Bi2MoO6 corresponded to type Ⅳ of the BDDT (Brunauer-Deming-Deming-Teller) classification, which suggested that Bi2MoO6 and 200-CQDs/Bi2MoO6 were mesoporous materials [32]. The specific surface area (SSA) of the catalysts was mainly calculated based on the Brunauer-Emmett-Teller (BET) test method and the results were displayed in Table 1. The SSA of Bi2MoO6, 100-CQDs/Bi2MoO6, 200-CQDs/Bi2MoO6, and 300-CQDs/Bi2MoO6 were 54.4, 54.9, 56.0 and 57.6 m2/g respectively, which are at a relatively high level compared to the existing reports (Table S1 in Supporting information). From this, the introduction of CQDs could slightly increase the specific surface area of the catalyst, which could facilitate the contact between the contaminants and the catalyst [18]. However, the different loading amount of CQDs showed similar effects on the specific surface area. This observation is the agreement with the previous literature, where in Di's study [19], the SSA of the CQDs/Bi2MoO6 composite slightly increased from 6.08 m2/g to 8.68 m2/g after loading CQDs. In Sun's study [20], with the increment in loading of CQDs reagent to 0.1 g from 0.05 g, the SSA rose to 16.83 m2/g from 16.27 m2/g. It is worth noting that our prepared CQDs/Bi2MoO6 composite showed a much bigger SSA, which ascribed to our preparation method is beneficial to produce CQDs/Bi2MoO6 with high specific surface area. The larger the SSA, the more active sites available in the photocatalytic process and better the degradation performance [22].

    Table 1

    Table 1.  The specific surface area of the Bi2MoO6 and CQDs/Bi2MoO6 composites.
    DownLoad: CSV

    The UV–vis diffuse reflectance spectra (DRS) from Bi2MoO6 and CQDs/Bi2MoO6 composites were displayed in Fig. 2g. It can be seen that each sample exhibited absorption edges at 480–496 nm, indicating that an excellent visible light absorption ability of the synthesized catalysts. Compared with pure Bi2MoO6, as mentioned in the introduction the loading of CQDs could significantly enhance the visible light absorption intensity of CQDs/Bi2MoO6, in addition, our synthesized CQDs/Bi2MoO6 also exhibited broadened absorption edges, among which 200-CQDs/Bi2MoO6 exhibited the largest absorption edge (broadening the absorption edge of the pure Bi2MoO6 from 480 nm to 496 nm) and the strongest visible light absorption intensity. Fig. 2h displayed for the prepared sample the band gap. It was evident that Bi2MoO6 had a band gap of 2.78 eV. Band gaps of CQDs/Bi2MoO6 composites were tuned with the CQDs content, among which 200-CQDs/Bi2MoO6 owned the lowest band gap of 2.71 eV. From above, it is evident that adding CQDs to Bi2MoO6 considerably increased the catalyst's visible light usage, in which 200-CQDs/Bi2MoO6 showed the optimal visible light absorption performance [24]. This broader absorption edge and lower band gap may be attributed to the role of CQDs, it functioned by enhancing the intensity of visible light absorption [22], which would absorb more visible light thus causing CQDs/Bi2MoO6 to exhibit enhanced visible light absorption. Besides, the upconversion feature of CQDs may be responsible for the absorption edge's expansion [33], where the CQDs could emit light at shorter wavelengths of 300–530 nm when they were excited with light at 700–1000 nm [34]. So as the loading amount of CQDs increased from 0 µL to 200 µL, the bandgap of CQDs/Bi2MoO6 gradually lowered. While continuing to increase the loading of CQDs to 300 µL, the bandgap of 300-CQDs/Bi2MoO6 became higher instead, which might be due to the overloading of 300 µL CQDs resulting in the accumulation of CQDs on the surface of Bi2MoO6, it would hinder light absorption of Bi2MoO6 and produced an internal filtering effect on light [35], which resulted in the poorer light absorption performance.

    The transient photocurrent response curves of Bi2MoO6 and CQDs/Bi2MoO6 were tested to study charge separation (Fig. 2i). The CQDs/Bi2MoO6 composite showed a larger photocurrent intensity than the pure Bi2MoO6, indicating its better electron-hole separation efficiency and longer electron lifetime. The increase in photocurrent intensity may be related to CQDs' ability to transport electrons, which served as an electron collector, capturing electrons spawned from the Bi2MoO6 semiconductor to prolong the carrier lifetime [36]. And the photocurrent intensity increased with the increase of CQDs loading amount from 100 µL to 200 µL, further increasing the CQDs loading amount to 300 µL, the photocurrent intensity decreased instead.

    The degradation of norfloxacin by Bi2MoO6 and CQDs/Bi2MoO6 composites was tested under visible light in Fig. 3a. It showed that the photodegradation rate of 200-CQDs/Bi2MoO6 reached 99% within 30 min, which were notably greater than Bi2MoO6, 100-CQDs/Bi2MoO6, and 300-CQDs/Bi2MoO6. As shown in Table S2 (Supporting information), the CQDs/Bi2MoO6 photocatalyst synthesized in this study exhibited obviously superior performance compared to the reported photocatalysts for the degradation of norfloxacin by visible light irradiation. The inset in Fig. 3a showed the results of the fitting using the quasi-first order kinetic model, in which k (where the absolute value of the slope of the line represented the k value) presented at first an increasing tendency followed by a decreasing tendency with an increase in CQDs amount, where the 200-CQDs/Bi2MoO6 composite exhibited a maximum reaction rate constant as 0.2066 min−1. Combined with the photocurrent test results, it is speculated that it may be due to the addition of CQDs, where the CQDs acted as the electron trapping agent and the construction of the CQDs/Bi2MoO6 heterojunction benefited the transferring of electrons from Bi2MoO6 to CQDs, which improved the efficiency of photogenerated electron-hole separation to prolong electron as well as hole longevity in order to degrade pollutants [29]. Meanwhile, combined with the DRS results, it could be known that the presence of CQDs broadened the spectral response range of the catalyst. Moreover, the upconversion luminescence property of CQDs resulted in its ability to emit shorter wavelengths of light in stimulation of the photocatalyst after absorbing near-infrared light [34], thereby enhancing the photocatalytic behavior. The performance was considerably impacted by the CQDs loading amount, namely, CQDs/Bi2MoO6 composites demonstrated an increment in photocatalytic activity in response to the growth of CQDs loading, where 200-CQDs/Bi2MoO6 reached the best performance. This is mainly because 200-CQDs/Bi2MoO6 exhibited the largest absorption edge (broadening the absorption edge of the pure Bi2MoO6 from 480 nm to 496 nm), the strongest visible light absorption intensity and the best electron-hole separation efficiency.

    Figure 3

    Figure 3.  (a) Degradation of norfloxacin on Bi2MoO6 and CQDs/Bi2MoO6. (b) The photocatalytic degradation and TOC removal curves of norfloxacin on 200-CQDs/Bi2MoO6 catalyst. (c) The possible photocatalytic mechanism for norfloxacin degradation over CQDs/Bi2MoO6 catalyst ([norfloxacin] = 20 mg/L, [catalysts] = 0.8 g/L, pH 7).

    For further evaluation of mineralization ability of the 200-CQDs/Bi2MoO6 photocatalyst, the change of TOC was monitored during the norfloxacin degradation. The TOC removal rate was approximately 20% in 30 min, as illustrated in Fig. 3b. This result indicated that the CQDs/Bi2MoO6 photocatalyst can effectively mineralize norfloxacin to CO2 when exposed to visible light [6]. It merits our attention that the norfloxacin decay was fast in the first 15 min when the TOC removal was almost inert. This observation should be ascribed to the photocatalytic reaction in the earlier stage mainly degraded the norfloxacin to small organic molecules rather than CO2. Prolonging the reaction time to 30 min, the norfloxacin decay leveled off since the norfloxacin almost completely degraded while the TOC removal was accelerated, which might be because the small organic molecules started to convert into CO2 right after the destruction of the target norfloxacin molecules. Therefore, it can be deduced that if the experimental time is extended, the TOC can still maintain a certain rate of removal [37].

    In light of aforementioned findings, Fig. 3c presented mechanisms for norfloxacin's degradation under this system. When the CQDs/Bi2MoO6 photocatalyst was illuminated by light, the catalyst generated hole-electron (h+ and e) pairs under the stimulation of light, and then the excited electrons skipped to the conduction band (CB) from the valence band (VB) to generate holes (hVB+) and electrons (eCB). The CQDs acting as electron carriers could trap the photogenerated e, thus effectively preventing the combination of h+ and e. As known in the reported literature, the EVB of Bi2MoO6 was +2.34 eV [38], which was more positive than that of H2O/HO (+2.27 eV vs. NHE) [39], so the h+ produced in Bi2MoO6 could oxidize H2O to HO, and the ECB of Bi2MoO6 was −0.34 eV, which was more negative than that of O2/O2•− (−0.28 eV vs. NHE), so the e+ generated in Bi2MoO6 could reduce O2 to O2•− [18,40,41]. In the CQDs/Bi2MoO6 system, the norfloxacin can be degraded by the hvb+, HO and O2•−. The detailed reaction mechanism is shown in Eqs. (1)-(4).

    (1)

    (2)

    (3)

    (4)

    The impact of operating factors, such as the catalyst dose, norfloxacin's starting concentration, as well as pH level, on the degradation of norfloxacin was depicted in Fig. S3 (Supporting information). Figs. S3a and b displayed the effect of different dosages of catalyst on the degradation of norfloxacin. In Fig. S3a, the degradation of norfloxacin by light alone was also investigated, i.e., at the catalyst dosage of 0 g/L, it was clear from the figure that there was almost no degradation of norfloxacin by light alone. And there was a continuous rise in k up to 0.2066 min−1 from 0.0749 min−1 if the dosage of CQD/Bi2MoO6 was raised to 0.8 g/L from 0.3 g/L and then a decrease as the dosage of CQD/Bi2MoO6 increased unceasingly. The creation of a significant amount of photo-generated h+ and e from the higher catalyst dosage contributed to the rise in k, which in turn generated abundant radicals to degrade norfloxacin [13]. Nevertheless, once CQDs/Bi2MoO6 was dosed at concentrations exceeding 0.8 g/L, the catalyst should gradually weaken the light transmission of the solution and produce a physical masking effect [42], causing the catalyst to absorb less light and generate fewer radicals. Furthermore, if the catalyst overdosed, the excess radicals would self-react with deactivation, as described in Eq. 5 [43]. Therefore, the best amount of catalyst to use was considered as 0.8 g/L in this study by considering the degradation rate of norfloxacin.

    (5)

    It was investigated how norfloxacin's initial concentrations affected the rate of degradation. Figs. S3c and d displayed that there was a drop in norfloxacin degradation efficiency in response to an increment in the norfloxacin concentration, i.e., there was a continuous decrease in k to 0.0934 min−1 from 0.3019 min−1 if norfloxacin initial concentration was raised to 25 mg/L from 10 mg/L. It is presumed that the active site's utilization of the catalyst reached saturation at higher norfloxacin concentrations, and the same amount of catalyst cannot degrade more norfloxacin at a certain time; in other words, the greater initial norfloxacin concentration solution was deficient in reactive free radicals [13].

    In addition, higher concentrations of norfloxacin absorbed more photons, leaving relatively fewer photons available for the active catalyst, and the shortage of photons on the catalyst surface somewhat delayed the initial stage in the photocatalytic reaction, namely the activation to catalyst by absorbing photons, thus decreasing the degradation rate of norfloxacin [35]. Besides, there were more intermediates generated in the degradation at higher norfloxacin concentrations, which would compete with norfloxacin for the active radicals, rendering the degradation rate of norfloxacin declined [44].

    Photocatalytic degradation tests were performed in solutions with different pH values for the purpose of understanding how the pH affected the degradation of norfloxacin. Figs. S3e and f revealed that norfloxacin degradation was inhibited in the strong acid (pH 2.2) and strong base (pH 12.0), and the degradation efficiency of norfloxacin was only 11.48% and 70.22% after 30 min of reaction, respectively, with k values of 0.0064 min−1 and 0.0918 min−1. This is mainly related to the interaction of the catalyst's surface charge with the contaminant. Bi2MoO6 is purported to have a zero point charge of 6.1 [45,46], therefore, as pH < 6.1, there are positive charges on the catalyst surface, and as pH > 6.1, there are negative charges. The carboxyl and piperazine groups, which have pKa values of 6.2 and 8.5, respectively, are the norfloxacin's two proton-binding sites. When the pH is lower than 6.2, there are positive charges of the norfloxacin molecule. With pH values ranging from 6.2 to 8.5, the norfloxacin molecule is electrically neutral and has both positive and negative charges; and when the pH is greater than 8.5, the norfloxacin molecule is negatively charged [47]. The catalyst's surface was strongly positively and negatively charged when pH 2.2 or 12.0, respectively, which strongly repulse the same charged norfloxacin, weakening the binding affinity between the norfloxacin and CQDs/Bi2MoO6 composite and thus inhibiting the norfloxacin decay. In contrast, norfloxacin breakdown was favored by an intermediate pH scope (pH 4.4 to 10.8), with the pH adjusted from 4.4 to 10.8, the degradation efficiency all exceeded 94% in 30 min. In detail, the k increased with increasing pH between pH 4.4 and 10.8, and the optimum pH level was determined at pH 10.8 with the biggest k of 0.5278 min−1. That is because a higher pH level provided more hydroxyl ions, which resulted in more hydroxyl radicals during the photocatalytic process and efficient norfloxacin decay. As a result, the photocatalyst CQDs/Bi2MoO6 operated well at a wide pH range, which is a great advantage in practice.

    It was necessary by analyzing CQDs/Bi2MoO6 degradation properties within an interfering substance presence to simulate its application potential in real water. Separate experiments on norfloxacin degradation in the presence of HCO3, Cl, NO3, and SO42− were conducted. Fig. 4 demonstrated that HCO3 aided norfloxacin decomposition, whereas other anions such as Cl, NO3, and SO42− inhibited norfloxacin degradation at a certain degree.

    Figure 4

    Figure 4.  Effect of coexisting ions on norfloxacin degradation.

    With the addition of HCO3, degradation performance enhanced as the k having gone up to 0.2974 min−1 from 0.2066 min−1. It is assumed that the CO3•− could be generated by the reaction between HCO3 and HO through Eq. 6. Despite the formed CO3•− being less active than HO [48], CO3•− is reported to be a more selective active species than HO, which can rapidly react with the substances containing the aromatic aniline functional group [49]. Considering the target norfloxacin is one of the substances containing aromatic aniline, therefore, it can be easily oxidized and degraded in the presence of CO3•− [7]. Moreover, it was monitored that the addition of HCO3 made the solution alkaline, with a pH around 8.4. The effect of pH on the system from the previous section indicates that the reaction is more suitable for weakly alkaline conditions, which is also one of the reasons why the presence of HCO3 exhibited a facilitative effect on norfloxacin photocatalytic degradation.

    (6)

    The addition of Cl, NO3, and SO42− all inhibited norfloxacin degradation to some extent. It is mainly because Cl, NO3, and SO42− can all act as scavengers of HO, through the Eqs. 7-9, giving rise to less active species such as CIHO•−, NO3, and NO2 [13,48,50]. Because freshly created radicals had a lower oxidizing capacity than consumed HO, decomposition of norfloxacin had been suppressed when Cl, NO3, and SO42− existed.

    (7)

    (8)

    (9)

    In this study, three-dimensional microspheres of CQDs/Bi2MoO6 in heterojunction structure were synthesized by an alcohol-thermal method. The 200-CQDs/Bi2MoO6 showed the best performance among the catalysts with different CQDs amounts, which showed a large specific surface area of 56.0 m2/g and the extending light-absorption edge from 480 nm to 496 nm. The intimate contact between Bi2MoO6 and CQDs resulted in the formation of Mo-O-C bonds and the improved interfacial charge transfer. The optimal degradation effect of 20 mg/L norfloxacin was obtained when the catalyst dosage = 0.8 g/L and pH 4.4–10.8, i.e., 99% norfloxacin degradation efficiency and 20% TOC removal were achieved in 30 min. Moreover, the coexistence of anions such as HCO3, Cl, NO3, and SO42− did not significantly impact the photocatalytic performance. In summary, the CQDs/Bi2MoO6 composite is an effective photocatalyst for the removal and mineralization of norfloxacin in real aqueous environments.

    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.

    This work was funded by the National Science Foundation of China (Nos. 42377452 and 41877481), the Strategic Priority Research Program of the Chinese Academy of Sciences, China (Nos. XDA23010300 and XDA23010000), the Opening Fund of Key Laboratory of Aerosol Chemistry and Physics, Institute of Earth Environment, CAS (No. KLACP2002). Thanks to Miss Chenyu Liang from the Instrumental Analysis Center and Ms. Xiaojing Zhang from School of Physics at Xi'an Jiaotong University for their assistance.

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2023.109336.


    1. [1]

      M.S. de Ilurdoz, J.J. Sadhwani, J.V. Reboso, J. Water Process. Eng. 45 (2022) 102474. doi: 10.1016/j.jwpe.2021.102474

    2. [2]

      L. Catteau, L. Zhu, F. Van Bambeke, et al., Phytochem. Rev. 17 (2018) 1129–1163. doi: 10.1007/s11101-018-9564-2

    3. [3]

      A. Jia, Y. Wan, Y. Xiao, et al., Water Res. 46 (2012) 387–394. doi: 10.1016/j.watres.2011.10.055

    4. [4]

      W. Deng, N. Li, H. Zheng, et al., Ecotoxicol. Environ. Saf. 125 (2016) 121–127. doi: 10.1016/j.ecoenv.2015.12.002

    5. [5]

      M. Chen, Y. Huang, W. Chu, Chin. J. Catal. 40 (2019) 673–680. doi: 10.1016/S1872-2067(19)63285-7

    6. [6]

      W. Wang, Z. Zeng, G. Zeng, et al., Chem. Eng. J. 378 (2019) 122132. doi: 10.1016/j.cej.2019.122132

    7. [7]

      F. Zhu, Y. Wu, Y. Liang, et al., Chem. Eng. J. 389 (2020) 124276. doi: 10.1016/j.cej.2020.124276

    8. [8]

      T. Chankhanittha, V. Somaudon, J. Watcharakitti, et al., J. Mater. Sci. : Mater. Electron. 32 (2021) 1977–1991. doi: 10.1007/s10854-020-04965-5

    9. [9]

      H. Yi, C. Lai, X. Huo, et al., Environ. Sci. : Nano 9 (2022) 815–826. doi: 10.1039/D1EN00869B

    10. [10]

      J.S. Wang, X.H. Yi, X. Xu, et al., Chem. Eng. J. 431 (2022) 133213. doi: 10.1016/j.cej.2021.133213

    11. [11]

      H. Mohtasham, M. Rostami, B. Gholipour, et al., Chemosphere 310 (2023) 136625. doi: 10.1016/j.chemosphere.2022.136625

    12. [12]

      P. Dhiman, G. Rana, R.A. Alshgari, et al., Environ. Res. 216 (2023) 114665. doi: 10.1016/j.envres.2022.114665

    13. [13]

      Z. Shen, H. Zhou, Z. Pan, et al., J. Hazard. Mater. 400 (2020) 123187. doi: 10.1016/j.jhazmat.2020.123187

    14. [14]

      H. Ma, Y. He, X. Li, et al., Appl. Catal. B 292 (2021) 120159. doi: 10.1016/j.apcatb.2021.120159

    15. [15]

      S. Li, S. Hu, W. Jiang, et al., J. Colloid Interface Sci. 556 (2019) 335–344. doi: 10.1016/j.jcis.2019.08.077

    16. [16]

      G. Zhang, D. Chen, N. Li, et al., Appl. Catal. B 250 (2019) 313–324. doi: 10.1016/j.apcatb.2019.03.055

    17. [17]

      Q. Zhao, Z. Zhang, T. Yan, et al., RSC Adv. 11 (2021) 28674–28684. doi: 10.1039/D1RA05164D

    18. [18]

      Z. Zhang, T. Zheng, J. Xu, et al., J. Photochem. Photobiol. A 346 (2017) 24–31. doi: 10.1016/j.jphotochem.2017.05.029

    19. [19]

      J. Di, J. Xia, M. Ji, et al., Nanoscale 7 (2015) 11433–11443. doi: 10.1039/C5NR01350J

    20. [20]

      C. Sun, Q. Xu, Y. Xie, et al., J. Alloys Compd. 723 (2017) 333–344. doi: 10.1016/j.jallcom.2017.06.130

    21. [21]

      X. Miao, Z. Ji, J. Wu, et al., J. Colloid Interface Sci. 502 (2017) 24–32. doi: 10.1016/j.jcis.2017.04.087

    22. [22]

      J. Wang, L. Tang, G. Zeng, et al., Appl. Catal. B 222 (2018) 115–123. doi: 10.1016/j.apcatb.2017.10.014

    23. [23]

      A. Phuruangrat, S. Buapoon, T. Bunluesak, et al., J. Aust. Ceram. Soc. 58 (2022) 999–1008. doi: 10.1007/s41779-022-00765-8

    24. [24]

      X. Lü, J. Shen, D. Fan, et al., Res. Chem. Intermed. 41 (2015) 9629–9642. doi: 10.1007/s11164-015-1953-1

    25. [25]

      J. Wang, F. Han, Y. Rao, et al., Ind. Eng. Chem. Res. 57 (2018) 10226–10233. doi: 10.1021/acs.iecr.8b01731

    26. [26]

      J. Zhao, Y. Yang, W. Yu, et al., J. Mater. Sci. Mater. Electron. 28 (2017) 543–552. doi: 10.1007/s10854-016-5557-3

    27. [27]

      L. Zhang, Z. Wang, C. Hu, et al., Appl. Catal. B 257 (2019) 117785. doi: 10.1016/j.apcatb.2019.117785

    28. [28]

      T. Zhou, J. Hu, J. Li, Appl. Catal. B 110 (2011) 221–230. doi: 10.1016/j.apcatb.2011.09.004

    29. [29]

      X. Sun, W. He, T. Yang, et al., Chem. Eng. J. 412 (2021) 128679. doi: 10.1016/j.cej.2021.128679

    30. [30]

      W. Dai, W. Xiong, J. Yu, et al., ACS Appl. Mater. Interfaces 12 (2020) 25861–25874. doi: 10.1021/acsami.0c04730

    31. [31]

      Y. Zhang, L. Wang, X. Ma, et al., J. Colloid Interface Sci. 531 (2018) 47–55. doi: 10.1016/j.jcis.2018.07.041

    32. [32]

      R. Abazari, A.R. Mahjoub, J. Shariati, et al., J. Cleaner Prod. 221 (2019) 582–586. doi: 10.1016/j.jclepro.2019.03.008

    33. [33]

      H. Zhang, H. Huang, H. Ming, et al., J. Mater. Chem. 22 (2012) 10501–10506. doi: 10.1039/c2jm30703k

    34. [34]

      Y. Su, M. Xie, X. Lu, et al., RSC Adv. 4 (2014) 4839–4842. doi: 10.1039/c3ra45453c

    35. [35]

      C. Guo, J. Xu, S. Wang, et al., Catal. Sci. Technol. 3 (2013) 1603–1611. doi: 10.1039/c3cy20811g

    36. [36]

      M. Ji, Z. Zhang, J. Xia, et al., Chin. Chem. Lett. 29 (2018) 805–810. doi: 10.1016/j.cclet.2018.05.002

    37. [37]

      S. Li, S. Hu, W. Jiang, et al., J. Colloid Interface Sci. 530 (2018) 171–178. doi: 10.1016/j.jcis.2018.06.084

    38. [38]

      X. Gao, W. Du, X. Gong, et al., Opt. Mater. 105 (2020) 109828. doi: 10.1016/j.optmat.2020.109828

    39. [39]

      P. Peng, Z. Chen, X. Li, et al., Sep. Purif. Technol. 291 (2022) 120901. doi: 10.1016/j.seppur.2022.120901

    40. [40]

      X. Gao, X. Ji, T.T. Nguyen, et al., Vacuum 164 (2019) 256–264. doi: 10.1016/j.vacuum.2019.03.032

    41. [41]

      Y. Qu, X. Li, H. Zhang, et al., J. Hazard. Mater. 429 (2022) 128310. doi: 10.1016/j.jhazmat.2022.128310

    42. [42]

      Z. Chen, X. Li, Q. Xu, et al., Chem. Eng. J. 390 (2020) 124454. doi: 10.1016/j.cej.2020.124454

    43. [43]

      Y. Liu, X. He, Y. Fu, et al., Chem. Eng. J. 284 (2016) 1317–1327. doi: 10.1016/j.cej.2015.09.034

    44. [44]

      Y. Wang, D. Tian, W. Chu, et al., Sep. Purif. Technol. 212 (2019) 536–544. doi: 10.1016/j.seppur.2018.11.051

    45. [45]

      M. Long, W. Cai, H. Kisch, Chem. Phys. Lett. 461 (2008) 102–105. doi: 10.1016/j.cplett.2008.06.081

    46. [46]

      J. Zhang, T. Wang, X. Chang, et al., Chem. Sci. 7 (2016) 6381–6386. doi: 10.1039/C6SC01803C

    47. [47]

      W. Yang, Y. Lu, F. Zheng, et al., Chem. Eng. J. 179 (2012) 112–118. doi: 10.1016/j.cej.2011.10.068

    48. [48]

      J. Li, M. Xu, G. Yao, et al., Chem. Eng. J. 348 (2018) 1012–1024. doi: 10.1016/j.cej.2018.05.032

    49. [49]

      X.H. Wang, A.Y.C. Lin, Environ. Sci. Technol. 46 (2012) 12417–12426. doi: 10.1021/es301929e

    50. [50]

      X. Gao, J. Niu, Y. Wang, et al., J. Hazard. Mater. 403 (2021) 123860. doi: 10.1016/j.jhazmat.2020.123860

  • Figure 1  Phase structure and chemical composition of the as-prepared catalysts. (a) The XRD patterns. (b) The FTIR spectra. (c) Mo 3d spectra. (d) Bi 4f spectra. (e) O 1s spectra. (f) C 1s spectra.

    Figure 2  Morphology, pore structure, optical absorption and electrochemical properties of the as-prepared catalysts. (a) SEM images of Bi2MoO6. (b) HRTEM images of CQDs. (c) TEM image of the 200-CQDs/Bi2MoO6. (d) HRTEM images of 200-CQDs/Bi2MoO6. (e, f) Nitrogen adsorption isotherm of the Bi2MoO6 and 200-CQDs/Bi2MoO6. (g) The UV–vis absorption spectra of the Bi2MoO6 and CQDs/Bi2MoO6. (h) The plots of (αhν)2 versus energy () of the Bi2MoO6 and CQDs/Bi2MoO6. (i) The transient photocurrent response of Bi2MoO6 and 200-CQDs/Bi2MoO6.

    Figure 3  (a) Degradation of norfloxacin on Bi2MoO6 and CQDs/Bi2MoO6. (b) The photocatalytic degradation and TOC removal curves of norfloxacin on 200-CQDs/Bi2MoO6 catalyst. (c) The possible photocatalytic mechanism for norfloxacin degradation over CQDs/Bi2MoO6 catalyst ([norfloxacin] = 20 mg/L, [catalysts] = 0.8 g/L, pH 7).

    Figure 4  Effect of coexisting ions on norfloxacin degradation.

    Table 1.  The specific surface area of the Bi2MoO6 and CQDs/Bi2MoO6 composites.

    下载: 导出CSV
  • 加载中
计量
  • PDF下载量:  1
  • 文章访问数:  285
  • HTML全文浏览量:  2
文章相关
  • 发布日期:  2024-08-15
  • 收稿日期:  2023-08-31
  • 接受日期:  2023-11-22
  • 修回日期:  2023-11-13
  • 网络出版日期:  2023-11-30
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