Bo Zhanga,
Chaoqi Lia,
Shasha Liua,
Lixuan Zhuanga,
Weiqi Zhanga,
Limei Huanga,
Zhenzhen Jiab,
Dongdong Chen*a and
Xiang Li*a
aSchool of Environmental and Chemical Engineering, Guangdong Provincial Key Laboratory of Eco-environmental Studies and Low-carbon Agriculture in Peri-urban Areas, Zhaoqing University, 526061 Zhaoqing, China
bSchool of Environmental Science and Engineering, Sun Yat-Sen University, 510006 Guangzhou, China
First published on 12th November 2024
A CaMnO3/g-C3N4 heterostructure, demonstrating promising photocatalytic performance for tetracycline (TC), was successfully synthesized using a straightforward calcination approach in this study. A series of characterization methods were employed to assess the physicochemical properties and visible-light responsiveness of the synthesized photocatalysts. The photocatalytic degradation rates of TC for the three prepared samples were evaluated. The results indicate that the CaMnO3/g-C3N4 composite exhibits the highest photocatalytic activity under visible-light irradiation, surpassing that of the individual components. Specifically, the degradation rate of CaMnO3/g-C3N4 is 0.031 min−1, which is 2.07 and 2.82 times greater than that of the pristine g-C3N4 and CaMnO3, respectively. Our findings highlight the significant potential of eco-friendly perovskites in developing visible-light-activated g-C3N4-based heterostructures for practical photocatalytic applications.
Perovskite oxides with an ABO3 structure are among the most promising heterogeneous catalysts for environmental remediation due to their tunable crystal structures, optical and chemical stability, adjustable band gaps, long charge carrier lifetimes, and sufficient oxygen vacancies.23–26 Additionally, the transition metals within these perovskite oxides exhibit variable oxidation states, which facilitate electron movement, making them valuable as electron transport materials in the design of electrocatalysts and photocatalysts.27,28 In the structure of perovskite oxides, the A sites are typically occupied by larger alkaline or rare earth metals (e.g., Ca, La, Be, Ba), while the B sites are filled by transition metals (e.g., Mn, Co, Sr, Ti, Ni), with A representing a divalent cation and B a tetravalent cation.29,30 Among them, Mn-based perovskites are noted for their oxidative stability, temperature resistance, and good conductivity for both oxygen and electrons, making them widely used in catalytic reactions.30 Calcium manganite (i.e. CaMnO3) is a significant member of this group, characterized by a framework of MnO6 octahedrons with Ca located in the interstitial spaces.31–33 The raw materials for synthesizing CaMnO3 are environmentally friendly, non-toxic, earth-abundant, and cost-effective, thus leading to increased demand for its application across various fields.31–34 Furthermore, CaMnO3 possesses an extremely narrow band gap, allowing it to effectively absorb visible light in the photocatalytic reactions.34,35 Therefore, it is reasonable to propose that forming a heterojunction between g-C3N4 and narrow-gap CaMnO3 could enhance both the separation efficiency of photogenerated electron–hole pairs and the ability to harvest visible light. Notably, the hybridization of g-C3N4 with CaMnO3 perovskite oxides has been relatively underexplored in existing literature, warranting further investigation.
In this study, a novel CaMnO3/g-C3N4 heterojunction was created through a straightforward calcination process. The morphology and structure of the catalysts were characterized using field emission scanning electron microscopy (FE-SEM), transmission electron microscopy (TEM), and X-ray diffraction (XRD). Optical properties were assessed through UV-vis Diffuse reflectance spectroscopy (UV-vis DRS) and Fourier transform infrared spectroscopy (FT-IR), while X-ray photoelectron spectroscopy (XPS) was employed to determine the chemical composition of the samples. The separation efficiency of photo-generated charge carriers including electrons (e−) and holes (h+) was evaluated via the photoluminescence (PL) and Photoelectrochemical measurements. The photocatalytic performance of the as-prepared samples under the irradiation of visible light was investigated using TC as a model pollutant. The results demonstrated that the inclusion of CaMnO3 significantly improved the photocatalytic activity of the CaMnO3/g-C3N4 composite compared to the bare g-C3N4. The findings indicate that the notable photocatalytic performance of CaMnO3/g-C3N4 is attributed to its effective visible-light harvesting ability and higher separation efficiency of photo-generated carriers. This study explores a new idea for the development of highly efficient and robust heterostructures for use in environmental remediation.
The FT-IR method was employed to further elucidate the chemistry states on the surface of CaMnO3, g-C3N4 and CaMnO3/g-C3N4 samples, as depicted in Fig. 2. In the FT-IR spectrum of g-C3N4, the prominent peak at 811 cm−1 is attributed to the characteristic breathing modes of the triazine units, while the absorption bands ranging from 1000 to 1700 cm−1 are assigned to the typical stretching vibration modes of C–N and CN (i.e., N–CN heterocycles) within the “melon” framework.7,10,18,23 Meanwhile, the FT-IR spectrum of CaMnO3 displays wide bands between 560 and 580 cm−1, corresponding to the stretching vibration of the Mn–O or Mn–O–Mn bonds, which is indicative of the typical perovskite structure formation.37,38 Additionally, the absorption band at 435 cm−1 is attributed to the bending mode resulting from changes in the Mn–O–Mn bond angle.37,38 The FT-IR spectra of the CaMnO3/g-C3N4 composite reveal the presence of all characteristic peaks of g-C3N4 and CaMnO3, indicating the successful construction of the heterostructure.
The light absorption capability is a crucial property that significantly influences the photocatalytic performance of semiconductors. To evaluate this, the light absorption properties of the synthesized samples were examined using UV-vis DRS measurements, with the results presented in Fig. 3. As shown in Fig. 3a, the light absorption edge of the pristine g-C3N4 is approximately 455 nm, indicating strong absorption in the UV region.6,9 In contrast, the bare CaMnO3 exhibits light absorption across both the UV and visible regions, demonstrating its efficient photoelectrical properties, similar to those observed in LaCoO3 perovskite oxide.39 When g-C3N4 is combined with CaMnO3, the light absorption edge experiences a red shift, which corresponds to an enhanced ability to absorb visible light. The bandgap energies of the photocatalysts were calculated using the Kubelka–Munk formula:
α(hν) = A(hν − Eg)n/2 |
Fig. 3 (a) UV-vis diffuse reflectance spectra of g-C3N4, CaMnO3 and CaMnO3/g-C3N4 samples; (b) (αhν)2 − hν plots of g-C3N4 and CaMnO3/g-C3N4 samples determined by Kubelka–Munk formula. |
The BET specific surface area and pore size distributions of the materials were analyzed using N2 adsorption–desorption isotherms, as depicted in Fig. 4. It is evident that all the samples exhibit type IV isotherms with type H3 hysteresis loops (Fig. 4a), indicating the presence of a mesoporous structure. Specifically, the BET specific surface areas of CaMnO3, g-C3N4, and CaMnO3/g-C3N4 are calculated to be 4.07, 23.07, and 25.72 m2 g−1, respectively. Following the introduction of CaMnO3, the BET specific surface area of CaMnO3/g-C3N4 shows a slight improvement compared to the bare g-C3N4, while the pore volume decreases from 0.134 to 0.105 cm3 g−1 (Table S1†). Fig. 3b presents the pore size distributions of the prepared samples calculated based on the Barrett–Joyner–Halenda (BJH) method, with the average pore sizes of CaMnO3, g-C3N4, and CaMnO3/g-C3N4 determined to be 8.06, 23.19, and 16.29 nm, respectively. Consequently, the reduced pore volume in CaMnO3/g-C3N4 is attributed to the formation of a greater number of smaller-sized pores during the calcination process of the composite. These results indicate that the introduction of CaMnO3 perovskite oxide into the g-C3N4 system provides the composite with more reactive sites and an enhanced ability to absorb smaller molecules, which is favorable for photocatalytic reactions.
Fig. 4 (a) N2-adsorption and desorption isotherms and (b) Barrett–Joyner–Halenda (BJH) pore size distribution curves of g-C3N4, CaMnO3 and CaMnO3/g-C3N4 samples. |
The surface morphologies of the as-prepared catalysts were analyzed using SEM and TEM measurements, with the results presented in Fig. 5, 6, S1 and S2.† The representative SEM images in Fig. 5a and b reveal that the bare g-C3N4 displays a compact block structure assembled by irregular lamellar architectures. Meanwhile, the pure CaMnO3 perovskite oxide exhibits a porous microspherical structure composed of aggregated nanoparticles (Fig. 5c and S1†). The SEM image of CaMnO3/g-C3N4 composite (Fig. 5d and S2†) illustrates the interweaving of lamellar architectures with porous microspheres when CaMnO3 perovskite oxide is mixed with g-C3N4. And it can be clearly seen that a significant quantity of small pores is formed in the structure of CaMnO3/g-C3N4 as compared to the pristine g-C3N4, this finding is consistent with the outcomes obtained from N2 adsorption–desorption measurements. Additionally, the close contact between CaMnO3 and g-C3N4 is observed, which can facilitate the mass transfer and charge transport processes of the CaMnO3/g-C3N4 composite in the photocatalytic reactions. Further investigation of the specific micro-structure of the CaMnO3/g-C3N4 composite was conducted using TEM method, as indicated in Fig. 6. The images at three different magnifications (Fig. 6a–c) reveal that a large amount of CaMnO3 nanoparticles is deposited on the surface of g-C3N4 nanosheets, indicating successful hybridization of CaMnO3 with g-C3N4. Fig. 6d presents the high-resolution TEM (i.e. HRTEM) image of the CaMnO3/g-C3N4 catalyst, showing a fringe spacing of 0.26 nm corresponding to the typical inter-planar spacing of the (121) plane of CaMnO3 perovskite material,36 while the related lattice spacing of g-C3N4 is lost due to its extremely low crystallinity. In summary, the results of SEM and TEM further confirm the successful construction of a binary heterojunction between CaMnO3 and g-C3N4, which aligns with the results of XRD and FT-IR measurements.
Fig. 5 The representative SEM image of pristine g-C3N4 (a and b), (c) CaMnO3 and (d) CaMnO3/g-C3N4 composite. |
The XPS technology was utilized to analyze the chemical compositions and binding states of the as-obtained samples, as depicted in Fig. 7. The survey spectrum in Fig. 7a reveals the presence of C, O, N, Ca, and Mn elements in CaMnO3/g-C3N4, whereas the other two catalysts (CaMnO3 and g-C3N4) only contain a subset of these elements (Fig. S3†), indicating the successful synthesis of the binary composites. In the high-resolution XP C 1s spectrum (Fig. 7b), three representative peaks at 289.5, 285.9 and 284.8 eV corresponding to sp2-bonded carbon (N–CN), C–N bond, and sp2 carbon (CC or C–C bond) are observed.40,41 The deconvoluted XP N 1s spectrum (Fig. 7c) reveals four distinct characteristic peaks at 403.8, 400.3, 399.1 and 398.3 eV corresponding to graphitic N, surface uncondensed amino groups bonded with an H atom (C–N–H), tertiary nitrogen bond (N–(C)3), and sp2 hybridized N (CN–C).40,42 Additionally, the XP O 1s spectrum (Fig. 7d) enables the differentiation of absorbed oxygen-containing species at approximately 532.1, 531.2, and 529.8 eV, respectively.43,44 The former is corresponding to CO, the middle is ascribed to the adsorbed oxygen species, and the latter arises from the metal oxygen bond (i.e. M−O).41,44 The curve-fitting of XP Ca 2p spectrum (Fig. 7e) displays two peaks (with an energy gap of 3.5 eV) corresponding to Ca 2p1/2 and Ca 2p3/2 orbitals centered at 350.7 and 347.2 eV, indicating that that Ca in the composite is the divalent oxide form.29,30 Fig. 7f exhibits the XP Mn 2p spectrum, and the binding energy gap between Mn 2p1/2 (653.5 eV) and Mn 2p3/2 (641.9 eV) spin–orbit splitting is about 11.6 eV, which is consistent with the previous works.30,45 The doublet peaks can be further deconvoluted by six peaks centered at 656.5, 653.9, 652.9, 644.9, 642.3 and 641.3 eV, respectively. The main peaks located at 653.9 and 642.3 eV are attributed to Mn4+ ions, while the two peaks located at 652.9 and 641.3 eV are assigned to Mn3+ ions.29,30 The other two bands at 656.5 and 644.9 eV are corresponding to the satellite peaks.29,30 The results indicate that Mn4+ is dominant on the surface of CaMnO3. The XPS measurements also demonstrate the successful preparation of CaMnO3/g-C3N4.
Fig. 7 The XPS survey spectra (a), high-resolution XP C 1s (b), N 1s (c), O 1s (d), Ca 2p (e) and Mn 2p (f) spectra for CaMnO3/g-C3N4 photocatalysts. |
The photocatalytic performance of a semiconductor is closely tied to the quantity of photo-generated charge carriers (e− and h+).7,46 To assess the separation efficiency and the transport ability of photogenerated e− and h+, PL measurements, transient photocurrent response measurements and EIS tests were conducted, with the results presented in Fig. 8. Fig. 8a illustrates that the PL intensity of CaMnO3/g-C3N4 sample significantly decreases compared to the pristine g-C3N4, indicating a substantial suppression of the recombination process of the charge carriers after CaMnO3 loading. Moreover, no obvious emission peak is observed for the bare CaMnO3, possibly due to its extremely narrow band gap.47,48 The transient photocurrent spectra in Fig. 8b demonstrate that the CaMnO3/g-C3N4 composite exhibits the highest photocurrent intensity compared to the pristine g-C3N4 and CaMnO3. These results suggest that the introduction of CaMnO3 enhances the separation efficiency of photogenerated electrons and holes. In summary, the PL and transient photocurrent analyses indicate that the hybridization of CaMnO3 and g-C3N4 significantly promotes the interfacial transfer process of the photogenerated charge carriers. Moreover, EIS measurements were also conducted to study the interfacial charge transfer processes, with the results shown in Fig. 8c. The equivalent circuit used in the measurements is exhibited in the inset of Fig. 8c, it contains the solution resistance (i.e. Rs), the electron-transfer resistance (i.e. Rct) as well as the constant phase element (CPE). Rs means the total resistance from the wire, electrolyte and electrode. Rs is mainly decided by the semiconductors due to fact that the measurement system is nearly identical except the electrode material.44,49 Rct is associated with the interfacial charge transfer between the electrode and electrolyte. Generally, a lower Rct value corresponds to a faster interfacial charge transfer process.44,49 In other words, a smaller semicircle radius in the Nyquist plots suggests a lower charge transfer resistance across the electrolyte/electrode interface, corresponding to the higher separation efficiency of charge carriers.44,49 It can be seen that CaMnO3/g-C3N4 demonstrates the smallest arc radius in the three samples, suggesting that the composite possesses the highest separation efficiency of charge carriers. Based on the results of PL and photoelectrochemical measurements, it can be concluded that the recombination of the charge carriers is effectively inhibited because of the strong interaction between CaMnO3 and g-C3N4.
Fig. 8 (a) PL spectra, (b) transient photocurrent responses and (c) EIS Nyquist plots of the as-prepared photocatalysts. |
To evaluate the photocatalytic performances of the catalysts, TC, a widely prevalent antibiotic in wastewater, was selected as the model pollutant.50,51 The photocatalytic degradation activities of TC over the obtained catalysts were specifically tested under visible light irradiation, as depicted in Fig. 9. Each experiment involved a dark reaction period of approximately 30 minutes before light irradiation to establish a balance of TC adsorption and desorption processes over the photocatalysts. The results in Fig. 9a reveal that the photocatalytic performances of the three catalysts for TC degradation under visible-light irradiation follow the order: CaMnO3/g-C3N4 > g-C3N4 > CaMnO3, consistent with the photoelectrochemical measurements. Additionally, a blank experiment (without photocatalyst) was also conducted, the results indicate that TC cannot be degraded in the absence of photocatalysts (Fig. S4†). Clearly, among the catalysts, the CaMnO3/g-C3N4 catalyst demonstrates the highest photocatalytic activity under visible light irradiation, achieving approximately 95% within 90 minutes, while CaMnO3 and g-C3N4 achieve 66% and 74%, respectively. The performance of CaMnO3/g-C3N4 sample in this work in comparison with the systems reported in the previous publications is provided in Table S2.† The CaMnO3/g-C3N4 sample exhibits outstanding photocatalytic activity as compared to the reported samples in terms of catalyst dosage, pollutant concentration and degradation efficiency. Furthermore, the reaction kinetics in the TC degradation processes were analyzed by fitting all the degradation rates based on the typical pseudo-first-order correlation: ln(C0/C) = kt, where C0 and C represent the initial TC concentration and the real-time TC concentration, k is the apparent pseudo-first-order rate constant (min−1), and t indicates the reaction time. As shown in Fig. 9b, the apparent rate constants of CaMnO3/g-C3N4, g-C3N4, and CaMnO3 are calculated based on the above pseudo-first-order correlation, yielding values of 0.031, 0.015, and 0.011 min−1, respectively. Obviously, among these, CaMnO3/g-C3N4 exhibits the highest reaction rate constant, surpassing that of g-C3N4 and CaMnO3 by 2.07 and 2.82 times, respectively.
Fig. 9 (a) Photocatalytic performances and (b) pseudo-first-order kinetic plots of the as-obtained catalysts for degradation of TC under the irradiation of visible light. |
In addition to its impressive performance, the stability and reusability of the photocatalyst is vital considerations for the practical applications. Consequently, the typical cycling experiments were performed under visible light irradiation to assess the stability of the CaMnO3/g-C3N4 catalyst, with the results showcased in Fig. 10a. Clearly, only a marginal decrease (approximately 3%) in performance of degrading TC was observed for the CaMnO3/g-C3N4 photocatalyst after 5 cycling tests. This slight reduction could be attributed to the inevitable loss of the photocatalyst during the cycling processes. The outcomes of the cycling experiments indicate that the CaMnO3/g-C3N4 photocatalyst exhibits excellent stability and can be effectively reused for the photocatalytic degradation of TC in wastewater under visible light irradiation. In addition, XRD tests were further conducted to investigate the crystal structure of CaMnO3/g-C3N4 before and after photocatalytic reaction, the results are shown in Fig. 10b. It indicates that the composite remains stable after the photo-degradation reaction, further indicating that CaMnO3/g-C3N4 exhibits the commendable stability.
Fig. 10 (a) Cycling tests of CaMnO3/g-C3N4 sample for the photocatalytic degradation of TC; (b) XRD spectra of CaMnO3/g-C3N4 before and after the photocatalytic reaction. |
To uncover the reaction mechanism, the primary active species involved in the photocatalytic degradation process of TC over CaMnO3/g-C3N4 were identified. It is well-established that ˙OH and ˙O2− radicals, along with the photogenerated h+, are key reactive species in the degradation reactions.7,52 In order to investigate the role of the radicals in the photocatalytic degradation process of TC, the radical scavenging experiments were conducted. Herein, benzoquinone (i.e. BQ), tert-butyl alcohol (i.e. TBA) and ammonium oxalate (i.e. (NH4)2C2O4) are applied as the scavenging reagents of ˙O2−, ˙OH and h+, respectively.7 The results, presented in Fig. 11, reveal that the photocatalytic performances for TC degradation using the CaMnO3/g-C3N4 sample significantly decreases upon the addition of (NH4)2C2O4 scavengers, while TBA and BQ have minimal impact on the catalyst's activity, indicating that h+ radicals are the primary active species. Specifically, the photocatalytic efficiency for TC degradation over CaMnO3/g-C3N4 decreases to 28.9%, 62.4% and 78.6% after the addition of (NH4)2C2O4, BQ and TBA, respectively. Additionally, ESR tests were conducted to further explore the roles of free radicals in the reactions. In the ESR tests, DMPO was employed as a spin-trapping reagent, with results depicted in Fig. 12a and b. The ESR signals, with intensity ratios of 1:2:2:1 and 1:1:1:1, correspond to the DMPO-˙OH and DMPO-˙O2− adducts, respectively, indicating the presence of both ˙O2− and ˙OH radical species in the photocatalytic process. These findings align well with the radical trapping test results. Overall, the importance of radical species in the photocatalytic degradation of TC using CaMnO3/g-C3N4 follows the order: h+ > ˙O2− >˙OH.
Fig. 11 Photocatalytic performance of CaMnO3/g-C3N4 for TC degradation with different radical scavengers. |
Based on the aforementioned observations, the potential degradation mechanism for TC is proposed, as depicted in Fig. 13. The increased surface area (25.72 m2 g−1) of CaMnO3/g-C3N4, along with its pore structures as indicated in Table S1,† facilitate TC adsorption process. Furthermore, the introduction of CaMnO3 enhances the ability of g-C3N4 to capture more visible light. Under visible-light irradiation, g-C3N4 is excited to generate electrons (e−) on the conduction band (CB) and h+ on the valence band (VB). Due to the excellent conductivity of CaMnO3, the e− on the CB of g-C3N4 could easily transfer to CaMnO3, thereby preventing the recombination process of e− and h+ in g-C3N4. According to the VB-XPS results (Fig. S5†), the VB value of CaMnO3/g-C3N4 composite (1.51 eV) is slightly smaller than that of g-C3N4 (1.66 eV). Integrating the results presented in Fig. 3, the calculated CB potentials can be determined to be 0.21 eV and −1.07 eV, respectively. Evidently, the CB potential of CaMnO3/g-C3N4 is not negative enough to reduce O2 in water to produce ˙O2− (O2/˙O2− = −0.33 eV), and the composite also cannot produce ˙OH (˙OH/OH− = 1.99 eV).53,54 Thus, h+ in the composite should be responsible for the degradation of TC. Meanwhile, e− remaining on the CB of g-C3N4 (−1.07 eV) can react with O2 to produce ˙O2−, and a portion of ˙O2− can further react with H2O to produce ˙OH. In a whole, both h+, ˙OH and ˙O2− participate in the degradation process of TC. However, h+ plays the major role in degrading TC into smaller harmless molecules, such as CO2 and H2O.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ra06407k |
This journal is © The Royal Society of Chemistry 2024 |