2. Wenzhou Aids to Navigation Division, Donghai Navigation Safety Administration (DNSA), Wenzhou 035000, China;
3. School of Chemical Engineering and Technology, Taiyuan University of Science and Technology, Taiyuan 030024, China;
4. College of Chemical Engineering and Technology, Taiyuan University of Technology, Taiyuan 030024, China;
5. School of Environment and Resources, Taiyuan University of Science and Technology, Taiyuan 030024, China
2. 交通运输部东海航海保障中心温州航标处, 浙江 温州 325000;
3. 太原科技大学 化学工程与技术学院, 山西 太原030024;
4. 太原理工大学 化学工程与技术学院, 山西 太原030024;
5. 太原科技大学 环境与资源学院, 山西 太原030024
Phenol, as a vital industrial chemical product, is well known for its bio-treatment of recalcitrant and acute toxicity. The ubiquitous presence of phenol in wastewater and its associated environmental risks have raised significant public health concerns[1]. Phenol is continuously discharged into aquatic environments through various anthropogenic activities. Its presence in storm and wastewater effluent poses a major obstacle to the widespread adoption of water recycling[2−3]. Moreover, it can directly harm ecosystems and endanger human health through the contamination of drinking water sources, including surface and groundwater[4]. Consequently, developing effective strategies for the removal of persistent phenol from wastewater effluent is crucial to minimize pollution risks.
Currently, numerous efforts have been devoted to developing suitable methods for removing phenol from wastewater effluent[5−12]. In recent years, photocatalytic degradation of phenol has emerged as a prominent research area. As an advanced oxidation technology, photocatalytic degradation has garnered significant attention due to its ability to efficiently degrade organic compounds in wastewater without generating secondary pollution[7−11]. However, conventional photocatalysts like TiO2 and ZnO are unable to fully utilize visible light, which constitutes approximately 45% of the solar spectrum, and thus fail to meet the rapidly increasing energy demand. This limitation has spurred extensive research into the development and exploration of efficient visible-light-driven photocatalysts to utilize solar energy more effectively.
Perovskite oxide semiconductors (ABO3) have shown great potential as photocatalysts, due to their excellent absorption of solar energy, particularly in the visible light spectrum, and their high quantum efficiency[13−16]. The term “perovskite” refers to the crystal structure of calcium titanate (CaTiO3), discovered by Gustav Rose in 1839 and named in honor of Lev Perovski[17]. In perovskite structures, the A-site typically accommodates rare earth or alkali metals with larger ionic radii, while the B-site is occupied by transition metals with smaller ionic radii.
According to the literature, doping, as a modification method, has been shown to enhance the catalytic properties of perovskites. By introducing metal dopants at either the A or B sites, the lattice structure can be modified, indirectly influencing and improving the activity and mobility of lattice oxygen, which in turn enhances catalytic performance[18−24]. For example, Wang et al.[18] synthesized BiCoxFe1−xO3 by doping Fe into BiCoO3, demonstrating a significant increase in magnetization due to Co substitution for Fe ions, while DTA confirmed the preservation of ferroelectric properties. Fan et al.[19] prepared La1−xCexMnO3 perovskites with varying Ce doping ratios for catalytic ozonation of phenol and found that La0.9Ce0.1MnO3 exhibited superior catalytic performance and good reusability. Huang et al.[20] reported the synthesis of LaCuxFe1−xO3 with different Fe doping ratios for the removal of organic pollutants, achieving optimal performance with 0.4 Fe doping at a gelation temperature of 80 ℃ and calcination temperature of 700 ℃, resulting in a 91.64% COD removal rate in cellulosic ethanol wastewater. Bouchal et al.[21] prepared via a sol-gel method with different concentrations of Bi nitrate and examined as a photocatalyst for RhB degradation under sunlight, and its antioxidant and antibacterial activities were examined. The results showed that the highest degradation rate was exhibited by 15% Bi-doped BaBiO3 for the degradation of RhB under solar radiation. According to the antibacterial activity study, the addition of Bi enhanced the antibacterial activity of the resulting material against both Gram-positive and Gram-negative microorganisms.
Based on the previous experimental findings, the results indicated LaCoO3 perovskite catalyst exhibited a certain photocatalytic activity for phenol, and the degradation rate reached 52.56% with 4 h irradiation. To further improve the performance of LaCoO3 catalysts, this study focused on doping Ba into the A-site of LaCoO3 to synthesize La1−xBaxCoO3 catalysts. The physical and chemical properties of these A-site doped materials were characterized and analyzed to understand the effect of doping on their catalytic behavior. Furthermore, the photocatalytic degradation of phenol under visible light irradiation was investigated to evaluate the efficacy of the doped perovskites.
1 Experimental section 1.1 MaterialsLanthanum nitrate (La(NO3)3·6H2O, AR, 99%), cobalt nitrate (Co(NO3)2·6H2O, AR, 99%), barium nitrate (Ba(NO3)2, AR, 99%), phenol (AR), citric acid (AR, ≥99.5%), ammonia (AR, 25%), ethanol (AR, ≥99.7%). The phenol solution was prepared into distilled water.
1.2 Catalyst preparationThe specific preparation method of modified LaCoO3 catalyst is citric acid complexation-hydrothermal synthesis combined method and as follows: (ⅰ) Solution preparation: 80 mL of deionized water was added to a beaker. La(NO3)3·6H2O, Ba(NO3)2, Co(NO3)2·6H2O, and citric acid were accurately weighed and dissolved in the water, maintaining the stoichiometric ratio of La1−xBaxCoO3 (x = 0.2, 0.3, 0.4, 0.5, 0.6). The molar ratio of n(La3++Ba2+)∶ n(Co2+)∶ n(citric acid) was kept at 1∶1∶1; (ⅱ) pH adjustment and complexation: The solution was stirred continuously at a uniform speed on a magnetic stirrer for 1 h to ensure complete complexation of the metal ions with citric acid. Meanwhile ammonia solution was added dropwise to the stirred solution to adjust the pH to approximately 8.5. This alkaline environment facilitated the complete precipitation of the metal ions as hydroxides; (ⅲ) Hydrothermal reaction: The mixture was poured into a teflon stainless steel container and reacted in an oven at 150 ℃ for 20 h; (ⅳ) Washing and drying: After the hydrothermal reaction, the resulting precipitate was removed from the teflon stainless steel container, centrifuged to separate it from the liquid, and washed thoroughly with deionized water to remove any residual reactants. The precipitate was then dried in an oven at 120 ℃; (ⅴ) Calcination: The dried precipitate was ground into a fine powder and calcined at 750 ℃ for 2 h. This high-temperature treatment removed any remaining organic matter and facilitated the formation of the final perovskite crystal structure; (ⅵ) Labeling: The synthesized catalysts were labeled as Ba-0.2, Ba-0.3, Ba-0.4, Ba-0.5 and Ba-0.6, corresponding to their respective Ba doping ratios (x values in La1−xBaxCoO3). The specific synthesis flowchart was shown in Fig.1.
![]() |
Fig.1 La1−xBaxCoO3 synthesization flowchart |
X-ray powder diffraction (XRD): Crystalline phases and crystal structure of the synthesized catalysts were determined using a Rigaku D/max-2500 X-ray diffractometer. The diffractometer was operated with Cu Kα radiation (λ =
The photocatalytic activity of the synthesized La1−xBaₓCoO3 catalysts was evaluated by monitoring the degradation of phenol in aqueous solution (100 mL, 10 mg·L−1) under visible light irradiation. All experiments were conducted at a constant temperature of 20 ℃ using a jacketed beaker with continuous water circulation to maintain the temperature. The solution was continuously stirred magnetically and ventilated throughout the experiment. A PLS-SXE300 xenon lamp equipped with a cutoff filter was used as the visible light source.
Experimental procedure: (Ⅰ) Dark adsorption: 100 mg of the synthesized photocatalyst was dispersed in 100 mL of the phenol solution in the jacketed beaker. The dispersion was stirred in the dark for 40 min to ensure adsorption-desorption equilibrium between phenol and the catalyst surface; (Ⅱ) Photocatalytic reaction: The dispersion was then irradiated with visible light. At predetermined time intervals of 30 min, 2 mL of the specified dispersion was withdrawn from the reaction mixture and filtered through a 0.22 µm syringe filter to remove the catalyst particles; (Ⅲ) UV-Vis spectrophotometry: The concentration of phenol in the filtered aliquots was determined by measuring the absorbance at 270 nm using a UV-2550 spectrophotometer. A calibration curve was constructed using standard phenol solutions to relate absorbance to concentration; (Ⅳ) Control experiments: Control experiments were performed under identical conditions but in the absence of visible light irradiation and without the addition of the photocatalyst; (Ⅴ) Degradation efficiency calculation: The photocatalytic degradation efficiency (D) was calculated using the following equation (1):
$ \begin{array}{c}D=\dfrac{\left({C}_{0}-{C}_{t}\right)}{{C}_{0}}\times 100\text{%}\end{array} $ | (1) |
where C0 is the initial concentration of phenol, and Ct is the concentration at time t.
2 Results and discussion 2.1 Structure characterizationThe phase structure of prepared samples were analyzed by XRD, and the XRD pattern of catalysts with different Ba doping ratios were compared with the standard spectra (in Fig.2). It can be found that the characteristic diffraction peaks of LaCoO3 in the 2θ of 23.79°, 26.18°, 32.90°, 40.54°, 47.41°, 58.77°, 69.07° and 78.78° were attributed to (110), (111), (200), (211), (220), (222), (400) and (420) lattice planes (JPCDS No.32-0480)[25], respectively. XRD results showed that LaCoO3 was successfully synthesized by citric acid complexation-hydrothermal synthesis combined method. Notably, no additional peaks corresponding to impurities or secondary phases were observed, confirming the phase purity of the synthesized catalysts. Furthermore, Fig.2 demonstrated a gradual increase in the intensity of diffraction peaks with increasing Ba doping ratio, and the Ba-0.6 showed the strongest diffraction peaks. This observation suggested that Ba successfully substituted La in the perovskite lattice. Due to the larger ionic radius of Ba compared to La, the lattice parameters expanded, leading to a slight shift of the XRD pattern towards lower angles.
![]() |
Fig.2 XRD patterns of the synthesized A-site Ba doped LaCoO3 catalysts |
The morphology and microstructure of modified LaCoO3 samples were studied by SEM. The distribution of elements within the samples was further analyzed using EDS spectroscopy. SEM images of the Ba-0.2, Ba-0.3, Ba-0.4, Ba-0.5, Ba-0.6 and LaCoO3 catalysts were presented in Fig.3(a–f), respectively. These images revealed a general trend of increasing average particle size with increasing Ba doping ratio. Notably, the Ba-0.6 catalyst exhibited the largest particle size, ranging from 20 to 50 nm, with a certain degree of agglomeration observed between the particles, which further illustrated that the Ba doping could increase the catalyst particle size. The smaller size was beneficial to the transfer of photo-generated electrons and enhancement of the photocatalytic performance[26]. The EDS was employed to analyze the elemental composition of the modified LaCoO3 catalysts. The EDS spectrum of the Ba-0.5 catalyst in Fig.4, revealed the presence of La, Ba, Co, O and C. The atomic percentages of these elements were determined to confirm the successful incorporation of Ba into the perovskite lattice. Table 1 presented the elemental ratios, which closely approximate the expected stoichiometry of (La+Ba)∶Co∶O = 1∶1∶3 for the perovskite ABO3 structure. This finding confirms the successful doping of Ba into the LaCoO3 lattice. Notably, the Ba-0.5 catalyst exhibited an elemental ratio closest to the ideal 1∶1∶3 stoichiometry among the five synthesized catalysts, suggesting a higher degree of purity. This enhanced purity may contribute to the superior photocatalytic performance observed for the Ba-0.5 catalyst.
![]() |
Fig.3 SEM of modified LaCoO3 catalysts (a) Ba-0.2, (b) Ba-0.3, (c) Ba-0.4, (d) Ba-0.5, (e) Ba-0.6, (f) LaCoO3 |
![]() |
Fig.4 EDS of Ba-0.5 catalyst |
![]() |
Table 1 Atomic content (mole fraction) of LaCoO3 and modified LaCoO3 catalysts |
The FT-IR spectroscopy was employed to identify the functional groups in the catalysts. The FT-IR spectra of the catalysts with different Ba doping ratios were shown in Fig.5, and the corresponding peak assignments were listed in Table 2. The peak at approximately 417 cm−1 corresponded to the bending vibration of O—Co—O in the perovskite ABO3 structure, while the peak at about 550 cm−1 was attributed to the stretching vibration of Co—O[27]. The strong absorption peak at 602 cm−1 further confirmed the presence of Co—O stretching vibrations in the ABO3 phase[28]. A weak absorption peak observed at about 850 cm−1 suggests the presence of carbonate species[29], likely originating from residual carbon and oxygen impurities introduced during the synthesis process using citric acid as the complexing agent. Notably, the Co—O stretching vibration peak at 550 cm−1 gradually shifted to lower wavenumbers with increasing Ba doping ratio, indicating a weakening of the Co—O bond covalency and an increase in ionic character upon Ba incorporation[30]. This observation can be attributed to lattice strain induced by lattice distortion due to the larger ionic radius of Ba compared to La. The FT-IR results, in conjunction with XRD and other characterization data, provided further evidence for the successful doping of Ba into the LaCoO3 crystal structure.
![]() |
Fig.5 Infrared spectrogram attribution analysis of modified LaCoO3 catalysts |
![]() |
Table 2 Infrared spectrum attribution analysis of modified LaCoO3 catalysts |
The textural properties of the modified LaCoO3 catalysts, as determined by N2 adsorption-desorption isotherms, were summarized in Table 3. The specific surface areas of the synthesized samples were relatively low, likely due to the presence of citric acid during synthesis. As shown in Table 3, the specific surface area of the catalysts decreased with increasing Ba doping ratio. Among the prepared catalysts, Ba-0.2 exhibited the highest specific surface area (23.95 m2·g−1), along with the largest pore volume and pore size. This decrease in textural properties with increasing Ba content may be attributed to the increased energy required for interdiffusion between grains in the presence of Ba(NO3)2[31, 32]. The presence of Ba ions could hinder the direct contact between crystal particles and impede crystal growth, leading to a reduction in both specific surface area and pore size[33]. There were research findings indicating that specific surface area of photocatalyst was related with its photocatalytic performance. A larger specific surface area was beneficial for the contact between the catalyst and the reaction solution, thereby improving the photocatalytic performance of the catalyst.
![]() |
Table 3 Pore structure parameters of modified LaCoO3 catalysts |
The N2 adsorption-desorption isotherms and pore size distributions of the LaCoO3 catalysts with different Ba doping ratios were presented in Fig.6. According to the IUPAC classification, all catalysts exhibited Ⅳ type isotherms, characteristic of mesoporous materials. The presence of H3 hysteresis loops in the catalysts, with closure points in the p/p0 between 0.1 and 0.2, indicated the presence of slit-shaped pores or plate-like particles. Furthermore, the pore size distributions revealed that all the catalysts possessed mesopores with sizes ranging from 2 to 50 nm. The gradual decrease in adsorption volume with increasing Ba doping ratio suggested a corresponding decrease in the specific surface area of the catalysts. This observation was consistent with the SEM analysis, which showed an increase in particle size and agglomeration with higher Ba content, leading to a reduction in the available surface area.
![]() |
Fig.6 Adsorption desorption curve and BJH pore size distribution of La1−xBaxCoO3 catalysts (x = 0.2, 0.3, 0.4, 0.5, 0.6) |
XPS was employed to investigate the elemental composition and valence states of the La1−xBaxCoO3 (x=0.2, 0.3, 0.4, 0.5, 0.6) catalysts. The XPS survey spectra (Fig.7(a)) confirmed the presence of La, Co, Ba, O and C, consistent with the EDS analysis. The carbon signal (C 1s) was used as a reference for binding energy calibration.
![]() |
Fig.7 XPS analysis energy spectrum of modified LaCoO3 catalysts (a) survey, (b) La 3d, (c) Ba 3d, (d) Co 2p, (e) O 1s |
The La 3d spectra in Fig.7(b) exhibited two characteristic peaks, La 3d5/2 and La 3d3/2 separated by approximately 4 eV due to spin-orbit coupling. Each of these peaks was further resolved into two components, arising from the transfer of electrons from the oxygen valence band to the empty La 4f orbital during the ionization process[34]. The La 3d XPS spectra confirmed the presence of La in the trivalent oxidation state[35], indicating that it did not participate in redox reactions as an active site. The Ba 3d spectra (Fig.7(c)) displayed two peaks, the binding energy centered at 795.1 and 779.65 eV ascribed to Ba 3d3/2 and Ba 3d5/2 respectively, with a separation of 15.4 eV, consistent with literature values[36]. Each peak was further resolved into two components, separated by approximately 3 eV, corresponding to Ba in the bivalent oxidation state[37−40]. The Co 2p spectra (Fig.7(d)) exhibited four peaks at 778.9, 781.6, 793.9, and 796.8 eV, attributed to Co3+ and Co2+ species, respectively. The peak at 790.0 eV was assigned to the satellite peak of Co2+, while the peak at 805.1 eV was ascribed to the satellite peak of Co species in Co3O4[41−42].
Finally, the O 1s spectra (Fig.7(e)) was deconvoluted into three peaks located at 528.9, 531.3 and 532.8 eV, corresponding to lattice oxygen, oxygen vacancies, and surface-adsorbed oxygen, respectively[43]. The presence of oxygen vacancies was noteworthy as it could significantly influence the catalytic activity of the perovskite materials.
The magnetic properties of the La1−xBaxCoO3 catalysts were investigated by measuring their magnetic hysteresis loops at room temperature (Fig.8). As La and Co were known to exhibit magnetic behavior, the observed strong linear M-H curves passing through the origin are indicative of paramagnetic behavior in these catalysts[44−46]. The modified LaCoO3 catalyst displayed a relatively strong paramagnetism, with an increasing trend observed as the Ba doping ratio increased. Notably, the Ba-0.5 and Ba-0.6 catalysts exhibited the highest paramagnetic susceptibility among the samples, suggesting a potential correlation between Ba content and the density of unpaired electrons responsible for paramagnetism. While the observed paramagnetism may be associated with enhanced catalytic activity, it was not the sole determining factor, as other factors such as surface area and crystal structure also play crucial roles. Nevertheless, the strong paramagnetic nature of the catalysts presents an opportunity for their efficient recovery and recycling using magnetic separation techniques, thereby reducing costs and promoting the sustainability of the photocatalytic process.
![]() |
Fig.8 MPMS diagram of modified LaCoO3 catalysts |
The photocatalytic activity of LaCoO3 and La1−xBaxCoO3 catalysts with different Ba doping ratios were assessed by degradation of phenol under visible light irradiation. As shown in Fig.9(a), the degradation rate of phenol initially increased and then decreased as the Ba doping ratio increased, indicating an optimal Ba content beneficial for photocatalytic performance. The Ba-0.5 catalyst exhibited the highest photocatalytic activity, achieving the phenol degradation rate of 83.50% after 6 h irradiation at 20 ℃. In contrast, the Ba-0.2 catalyst showed the lowest degradation rate (28.09%), while the Ba-0.3, Ba-0.4 and Ba-0.6 catalysts achieved degradation rates of 36.92%, 54.07% and 63.40%, respectively, under the same conditions. To elucidate the role of the Ba-0.5 photocatalyst in phenol degradation and to differentiate the contributions of adsorption and photocatalysis, a comparative experiment was conducted with and without visible light irradiation. As depicted in Fig.9(b), in the absence of light, the Ba-0.5 catalyst adsorbed approximately 10% of phenol, and this adsorption capacity had no change over 6 h. Moreover, the degradation rate of phenol significantly increased to over 80% under visible light irradiation. The results demonstrated that the Ba-0.5 catalyst exhibited high photocatalytic activity towards phenol, and the adsorption played a minor role in the overall degradation process.
![]() |
Fig.9 (a) Degradation rate curves of different Ba doped LaCoO3 catalysts for photocatalytic degradation of phenol; (b) Degradation of phenol under different conditions of Ba-0.5 |
To further corroborate the observed photocatalytic activities, the degradation kinetics of phenol by the La1−xBaxCoO3 catalysts were investigated. The linear relationship between ln(C0/C) and irradiation time (t) was given in Fig.10(a), indicated that the photodegradation of phenol follows pseudo-first-order kinetics. This kinetic model suggested that the rate of phenol degradation was directly proportional to its concentration, with the photocatalyst acting as a constant factor. The rate constant (k) for each catalyst could be determined from the slope of the linear fit in Fig.10(a)[47].
![]() |
Fig.10 The corresponding kinetic plots of phenol degradation over different catalysts (a) linear relationship between ln(C0/C) and t, (b) rate constants corresponding to different catalysts |
$ \begin{array}{c}r=k{C}_{A}\end{array} $ | (2) |
$ \begin{array}{c}r=-\dfrac{d\left({C}_{A}\right)}{dt}\end{array} $ | (3) |
$ \begin{array}{c}-\dfrac{d\left({C}_{A}\right)}{dt}=k{C}_{A}\end{array} $ | (4) |
$ \begin{array}{c}-{\displaystyle\int }_{{C}_{A,0}}^{{C}_{A}}\dfrac{1}{{C}_{A}}={\displaystyle\int }_{{t}_{0}}^{t}k\cdot dt\end{array} $ | (5) |
$ \therefore \text{ln}\left({C}_{0}\text{/}C\right)={kt} $ | (6) |
The rate constants (k) for phenol degradation by the different La1−xBaxCoO3 catalysts, derived from the slopes of the linear fits in Fig.10(a), were presented in Fig. 10(b). The Ba-0.5 catalyst exhibited the highest rate constant (
The reusability of the catalyst was a critical factor in assessing its overall effectiveness. To this end, the stability of the Ba-0.5 catalyst was evaluated by recyclability test. The photocatalyst was recycled by filtration, washing and drying for further evaluation. The recovered catalyst (100 mg) was introduced into a 100 mL phenol solution (10 mg·L−1), and its degradation rate was measured using absorbance spectroscopy. As depicted in Fig.11(a), although the degradation rate of phenol displayed a slightly decrease over five cycles, it consistently remained above 70%. This result demonstrates the Ba-0.5 catalyst maintains substantial photocatalytic activity after repeated use. XRD characterization and comparative analysis were then performed on Ba-0.5 before and after five repeated experiments, as shown in Fig.11(b). There was no obvious change could be found in the XRD patterns compared to the used and fresh Ba-0.5 catalyst, suggesting that the crystal structure of Ba-0.5 was well-preserved after five cycles.
![]() |
Fig.11 (a) Reusability of the spent Ba-0.5 catalyst and (b) XRD diagram of Ba-0.5 before and after 5 cycles |
To elucidate the photocatalytic mechanism of the modified Ba-0.5 catalyst in phenol degradation[6, 8, 48−49], the roles of various active radicals in the phenol decomposition process were investigated. Four trapping agents isopropanol (IPA), p-benzoquinone (BQ), ammonium oxalate (AO), and silver nitrate (AgNO3) were introduced into the phenol solution under consistent experimental conditions, the agents served as scavengers for hydroxyl radicals (•OH), superoxide radicals (
![]() |
Fig.12 Effect of scavengers on the degradation of phenol |
The perovskite-type LaCoO3 photocatalyst was known to exhibit catalytic activity under visible light. Upon irradiation with a xenon lamp, the surface of the LaCoO3 catalyst absorbs photons with energy exceeding its bandgap. This absorption led to the formation of electrons (e−) in the conduction band and holes (h+) in the valence band. These charge carriers subsequently reacted with water and oxygen molecules in the system, generating highly reactive species such as hydroxyl radicals (•OH) and superoxide radicals (
$ \mathrm{\ LaCoO}_{\mathrm{3}}\mathrm{+}\mathit{h\nu\to}\mathrm{e}^-\mathrm{(LaCoO}_{\mathrm{3}}\mathrm{)+h}^{\mathrm{+}}\mathrm{(LaCoO}_{\mathrm{3}}\mathrm{)} $ |
$ \mathrm{\ h}^{\mathrm{+}}\mathrm{+H}_{\mathrm{2}}\mathrm{O}\mathit{\to}\mathrm{\text{•}OH+H}^{\mathrm{+}} $ |
$ \mathrm{\ h}^{\mathrm{+}}\mathrm{+OH}^-\mathit{\to}\mathrm{\text{•}OH} $ |
$ \mathrm{e}^-\mathrm{+O}_{\mathrm{2}}\mathit{\to}\mathrm{\text{•}O}_{\mathrm{2}}^- $ |
$ \mathrm{H}_{\mathrm{2}}\mathrm{O+\text{•}O}_{\mathrm{2}}^-\mathit{\to}\mathrm{HO}_{\mathrm{2}}\mathrm{\text{•}+OH}^- $ |
The results demonstrated that after 6 h of reaction, the phenol degradation rate was 83.5% in the system without a trapping agent. In contrast, the degradation rates were 26.4%, 61.3%, 65.7% and 80.9% in the systems containing IPA, BQ, AO and AgNO3, respectively. This indicated that photo-generated electrons (e−) had a negligible effect on phenol degradation, while hydroxyl radicals (•OH), superoxide radicals (
![]() |
Fig.13 Three pathways of photocatalytic phenol degradation (a), and La1−xBaxCoO3 mechanism diagram (b) |
The La1−xBaxCoO3 catalyst was prepared by the citric acid complexation-hydrothermal combined method, and the synthesis mechanism was elucidated through a series of assumptions and rigorous deductions.
Complexation of citric acid with metal ions:
$ \mathrm{C}_{ \mathrm{6}} \mathrm{H}_{ \mathrm{8}} \mathrm{O}_{ \mathrm{7}} \mathrm{+La}^{ \mathrm{3+}} \mathrm{\to LaC}_{ \mathrm{6}} \mathrm{H}_{ \mathrm{5}} \mathrm{O}_{ \mathrm{7}} \mathrm{+3H}^{ +} $ |
$ \mathrm{C}_{ \mathrm{6}} \mathrm{H}_{ \mathrm{8}} \mathrm{O}_{ \mathrm{7}} \mathrm{+Co}^{ \mathrm{2+}} \mathrm{\to CoC}_{ \mathrm{6}} \mathrm{H}_{ \mathrm{6}} \mathrm{O}_{ \mathrm{7}} \mathrm{+2H}^{ \mathrm+} $ |
$ \mathrm{C}_{ \mathrm{6}} \mathrm{H}_{ \mathrm{8}} \mathrm{O}_{ \mathrm{7}} \mathrm{+Ba}^{ \mathrm{2+}} \mathrm{\to BaC}_{ \mathrm{6}} \mathrm{H}_{ \mathrm{6}} \mathrm{O}_{ \mathrm{7}} \mathrm{+2H}^{ \mathrm+} $ |
Hydrolysis reaction:
$ \mathrm{LaC}_{ \mathrm{6}} \mathrm{H}_{ \mathrm{5}} \mathrm{O}_{ \mathrm{7}} \mathrm{\to LaC}_{ \mathrm{6}} \mathrm{H}_{ \mathrm{3}} \mathrm{O}_{ \mathrm{6}} \mathrm{+H}_{ \mathrm{2}} \mathrm{O} $ |
$ \mathrm{2LaC}_{ \mathrm{6}} \mathrm{H}_{ \mathrm{3}} \mathrm{O}_{ \mathrm{6}} \mathrm{\to La}_{ \mathrm{2}} \mathrm{C}_{ \mathrm{12}} \mathrm{H}_{ \mathrm{6}} \mathrm{O}_{ \mathrm{12}} $ |
$ \mathrm{CoC}_{ \mathrm{6}} \mathrm{H}_{ \mathrm{6}} \mathrm{O}_{ \mathrm{7}} \mathrm{\to CoC}_{ \mathrm{6}} \mathrm{H}_{ \mathrm{4}} \mathrm{O}_{ \mathrm{6}} \mathrm{+H}_{ \mathrm{2}} \mathrm{O} $ |
$ \mathrm{2CoC}_{ \mathrm{6}} \mathrm{H}_{ \mathrm{4}} \mathrm{O}_{ \mathrm{6}} \mathrm{\to Co}_{ \mathrm{2}} \mathrm{C}_{ \mathrm{12}} \mathrm{H}_{ \mathrm{8}} \mathrm{O}_{ \mathrm{12}} $ |
$ \mathrm{BaC}_{ \mathrm{6}} \mathrm{H}_{ \mathrm{6}} \mathrm{O}_{ \mathrm{7}} \mathrm{\to BaC}_{ \mathrm{6}} \mathrm{H}_{ \mathrm{4}} \mathrm{O}_{ \mathrm{6}} \mathrm{+H}_{ \mathrm{2}} \mathrm{O} $ |
$ \mathrm{2BaC}_{ \mathrm{6}} \mathrm{H}_{ \mathrm{4}} \mathrm{O}_{ \mathrm{6}} \mathrm{\to Ba}_{ \mathrm{2}} \mathrm{C}_{ \mathrm{12}} \mathrm{H}_{ \mathrm{8}} $ |
Oxidation-reduction creaction:
$ \mathrm{La}_{ \mathrm{2}} \mathrm{C}_{ \mathrm{12}} \mathrm{H}_{ \mathrm{6}} \mathrm{O}_{ \mathrm{12}} \mathrm{+9O}_{ \mathrm{2}} \mathrm{\to La}_{ \mathrm{2}} \mathrm{O}_{ \mathrm{2}} \mathrm{CO}_{ \mathrm{3}} \mathrm{+CO}_{ \mathrm{2}} \mathrm{+H}_{ \mathrm{2}} \mathrm{O} $ |
$ \mathrm{La}_{ \mathrm{2}} \mathrm{O}_{ \mathrm{2}} \mathrm{CO}_{ \mathrm{3}} \mathrm{\to La}_{ \mathrm{2}} \mathrm{O}_{ \mathrm{3}} \mathrm{+CO}_{ \mathrm{2}} $ |
$ \mathrm{3Co}_{ \mathrm{2}} \mathrm{C}_{ \mathrm{12}} \mathrm{H}_{ \mathrm{8}} \mathrm{O}_{ \mathrm{12}} \mathrm{+28O}_{ \mathrm{2}} \mathrm{\to 2Co}_{ \mathrm{3}} \mathrm{O}_{ \mathrm{4}} \mathrm{+36CO}_{ \mathrm{2}} \mathrm{+12H}_{ \mathrm{2}} \mathrm{O} $ |
$ \mathrm{Ba}_{ \mathrm{2}} \mathrm{C}_{ \mathrm{12}} \mathrm{H}_{ \mathrm{8}} \mathrm{O}_{ \mathrm{12}} \mathrm{+9O}_{ \mathrm{2}} \mathrm{\to 2BaO+12CO}_{ \mathrm{2}} \mathrm{+4H}_{ \mathrm{2}} \mathrm{O} $ |
Decomposition reaction:
$ \mathrm{La}_{ \mathrm{2}} \mathrm{O}_{ \mathrm{3}} \mathrm{+Co}_{ \mathrm{3}} \mathrm{O}_{ \mathrm{4}} \mathrm{+BaO\to LaBaCoO}_{ \mathrm{3}} \mathrm{+CoO} $ |
A-site Ba-doped La1−xBaxCoO3 materials were successfully synthesized using a citric acid complexation-hydrothermal synthesis combined method. The physicochemical structures and properties of the resulting samples were characterized using XRD, SEM, BET and XPS. Among the synthesized materials, La0.5Ba0.5CoO3 exhibited the highest phenol degradation rate (83.5%), following pseudo-first-order kinetics.
Various characterization techniques confirmed the successful incorporation of Ba into the LaCoO3 crystal structure. While the morphology remained largely unchanged, a slight degree of agglomeration was observed in the modified catalyst. The specific surface area of the catalysts was relatively small and decreased with increasing Ba doping. Magnetization, as assessed by M-H curves, appeared to correlate with catalytic activity, with stronger magnetic intensity associated with better phenol photodegradation. The strong paramagnetism of the catalyst suggested good recyclability and potential cost savings. Notably, the Ba-0.5 catalyst retained high photocatalytic activity after five cycles, with XRD analysis confirming minimal structural changes. The XPS characterization results showed that Ba-0.5 had the highest oxygen content, and its oxygen vacancies made it positively charged, leading to an extension of the positron lifetime, reducing the recombination of electron-hole pairs, and enhancing the photocatalytic activity. Additionally, the magnetic properties of the photo-catalyst were also related to its photocatalytic activity. A photocatalyst with certain magnetic properties could better adsorb the light source required for photocatalytic reactions, thereby improving photocatalytic activity. Furthermore, magnetism could also help in the recovery and reuse of the photocatalyst, reducing catalyst loss, and ultimately improving the economic and environmental friendliness of photocatalysis. The MPMS characterization results showed that Ba-0.5 has high magnetism, which helped to enhance its catalytic activity. In conclusion, Ba-0.5 had the highest photocatalytic activity.
The photocatalytic mechanism study suggested that electron-hole pairs convert surface-adsorbed oxygen into highly active free radicals (•OH and
[1] |
Risk assessment of xenobiotics in stormwater discharged to Harrestrup Å, Denmark[J]. Desalination, 2007, 215(1/3): 187–197.
DOI:10.1016/j.desal.2006.12.008 |
[2] |
Photocatalytic degradation of agricultural N-heterocyclic organic pollutants using immobilized nanoparticles of titania[J]. J Hazard Mater, 2007, 145(1/2): 65–71.
DOI:10.1016/j.jhazmat.2006.10.089 |
[3] |
Comparison of different advanced oxidation processes for phenol degradation[J]. Water Res, 2002, 36(4): 1034–1042.
DOI:10.1016/S0043-1354(01)00301-3 |
[4] |
Synergy effect in the photocatalytic degradation of phenol on a suspended mixture of titania and activated carbon[J]. Appl Catal B-Environ, 1998, 18(3/4): 281–291.
DOI:10.1016/S0926-3373(98)00051-4 |
[5] |
Phenol degradation on the surface of mesoporous TiO2 particles via ligand-to-metal charge transfer under visible light irradiation[J]. Chem Phys Lett, 2024, 840: 141162.
DOI:10.1016/j.cplett.2024.141162 |
[6] |
ZnO hierarchical structures with tunable oxygen vacancies for high performance in photocatalytic degradation of phenol[J]. J Mol Struct, 2024, 1304: 137656.
DOI:10.1016/j.molstruc.2024.137656 |
[7] |
Nano polyoxometalate-based copper complex with double catalytic sites as heterogeneous catalyst for high-efficient degradation of phenol[J]. Inorg Chem Commun, 2024, 161: 112123.
DOI:10.1016/j.inoche.2024.112123 |
[8] |
Preparation of Co-doped TiO2 activated carbon nanocomposite and its photocatalytic degradation of phenol wastewater[J]. Inorg Chem Commun, 2024, 166: 112644.
DOI:10.1016/j.inoche.2024.112644 |
[9] |
Heterogeneous catalytic degradation of phenol by CuFe2O4/Bi12O15Cl6 photocatalyst activated peroxymonosulfate[J]. Mater Res Bull, 2023, 167: 112435.
DOI:10.1016/j.materresbull.2023.112435 |
[10] |
Effect of pore-forming agent on degradation of phenol by iron tailings based porous ceramics[J]. Ceram Int, 2024, 50(18): 33791–33801.
DOI:10.1016/j.ceramint.2024.06.197 |
[11] |
Novel dual Z-scheme Bi/BiOI-Bi2O3-C3N4 heterojunctions with synergistic boosted photocatalytic degradation of phenol[J]. Sep Purif Technol, 2023, 307: 122733.
DOI:10.1016/j.seppur.2022.122733 |
[12] |
Adsorption and catalytic degradation of phenol in water by a Mn, N co-doped biochar via a non-radical oxidation process[J]. Sep Purif Technol, 2024, 330: 125267.
DOI:10.1016/j.seppur.2023.125267 |
[13] |
Perovskite oxides for semiconductor-based gas sensors[J]. Sens Actuator B-Chem, 2007, 123(2): 1169–1179.
DOI:10.1016/j.snb.2006.10.051 |
[14] |
Machine learning-aided discovery of bismuth-based transition metal oxide double perovskites for solar cell applications[J]. Sol Energy, 2024, 267: 112209.
DOI:10.1016/j.solener.2023.112209 |
[15] |
Fabrication, characterization, metal ions sensor and photocatalytic activity of visible light harvesting lead free perovskite oxide[J]. Mater Chem Phys, 2024, 325: 129738.
DOI:10.1016/j.matchemphys.2024.129738 |
[16] |
Strategies for improving oxygen ionic conducting in perovskite oxides and their practical applications[J]. Energy Rev, 2024, 3(4): 100085.
DOI:10.1016/j.enrev.2024.100085 |
[17] |
Contributed papers in specimen mineralogy: 38th Rochester mineralogical symposium abstracts: Part 3[J]. Rocks Miner, 2012, 87(5): 444–454.
DOI:10.1080/00357529.2012.716338 |
[18] |
Enhanced ferromagnetic properties of multiferroic BiCoxFe1−xO3 synthesized by hydrothermal method[J]. Mater Lett, 2008, 62(23): 3806–3808.
DOI:10.1016/j.matlet.2008.04.062 |
[19] |
Ce-doped LaMnO3-δ perovskite as an efficient ozonation catalyst for the degradation of high-concentration phenol wastewater[J]. J Environ Chem Eng, 2024, 12(3): 112761.
DOI:10.1016/j.jece.2024.112761 |
[20] |
Fe element B-site doped lanthanum-copper perovskite type material for cellulosic ethanol wastewater treatment[J]. Mater Chem Phys, 2024, 313: 128751.
DOI:10.1016/j.matchemphys.2023.128751 |
[21] |
Bi-doped BaBiO3 (x = 0%, 5%, 10%, 15%, and 20%) perovskite oxides by a sol–gel method: Comprehensive biological assessment and RhB photodegradation[J]. RSC Adv, 2014, 14(11): 7359–7370.
DOI:10.1039/d3ra06354b |
[22] |
Cu-doped CaFeO3 perovskite oxide as oxygen reduction catalyst in air cathode microbial fuel cells[J]. Eviron Res, 2022, 214: 113968.
DOI:10.1016/j.envres.2022.113968 |
[23] |
Exploiting the Bi-doping effect on the properties of NaNbO3 perovskite-type materials[J]. Chem Phys, 2024, 579: 112203.
DOI:10.1016/j.chemphys.2024.112203 |
[24] |
The effect of Al3+ doping on the infrared radiation and thermophysical properties of SrZrO3 perovskites as potential low thermal infrared material[J]. Acta Mater, 2021, 209: 116795.
DOI:10.1016/j.actamat.2021.116795 |
[25] | |
[26] |
Photocatalytic performances of BiFeO3 particles with the average size in nanometer, submicrometer, and micrometer[J]. Mater Res Bull, 2014, 50: 369–373.
DOI:10.1016/j.materresbull.2013.11.039 |
[27] |
Characterization of cobalt oxides studied by FT-IR, Raman, TPR and TG-MS[J]. Thermochim Acta, 2008, 473(1/2): 68–73.
DOI:10.1016/j.tca.2008.04.015 |
[28] |
Crystal structures, hhirshfeld surfaces analysis, Infrared and Raman studies of organic-inorganic hybrid perovskite salts NH3(CH2)nNH3MnCl4 (n=5, 6)[J]. J Solid State Chem, 2022, 314: 123401.
DOI:10.1016/j.jssc.2022.123401 |
[29] |
On the formation mechanism of BaTiO3-BaZrO3 solid solution through solid-oxide reaction[J]. Mater Lett, 2005, 59(1): 135–138.
DOI:10.1016/j.matlet.2004.07.053 |
[30] |
Perovskite catalysts with different dimensionalities for environmental and energy applications: A review[J]. Sep Purif Technol, 2023, 307: 122716.
DOI:10.1016/j.seppur.2022.122716 |
[31] | |
[32] |
Preparation and characterization of iodine-doped mesoporous TiO2 by hydrothermal method[J]. Appl Surf Sci, 2011, 257(8): 3688–3696.
DOI:10.1016/j.apsusc.2010.11.108 |
[33] |
Crystallite size, surface properties and photocatalytic activity of Gd3+-doped TiO2 nano-powder[J]. J Chem Ind Eng (China), 2006, 57(9): 2194–2200.
DOI:10.3321/j.issn:0438-1157.2006.09.035 |
[34] |
Effect of Pd or Ag additive on the activity and stability of monolithic LaCoO3 perovskites for catalytic combustion of methane[J]. Catal Today, 2004, 90(1/2): 121–126.
DOI:10.1016/j.cattod.2004.04.016 |
[35] |
XPS study on La0.67Ca0.33Mn1−xCoxO3 compounds[J]. J Mol Struct, 2014, 1073: 66–70.
DOI:10.1016/j.molstruc.2014.04.065 |
[36] |
X-Ray photoelectron spectroscopy on zeolites and related materials[J]. Microporous Mat, 1996, 6(5/6): 235–257.
DOI:10.1016/0927-6513(96)00034-X |
[37] |
Preparation of Ba1−xSrxWO4 and Ba1−xCaxWO4 films on tungsten plate by mechanically assisted solution reaction at room temperature[J]. Mater Chem Phys, 2008, 109(2/3): 217–223.
DOI:10.1016/j.matchemphys.2007.11.029 |
[38] |
Sn-CeO2 thin films prepared by rf magnetron sputtering: XPS and SIMS study[J]. Appl Surf Sci, 2009, 255(13/14): 6656–6660.
DOI:10.1016/j.apsusc.2009.02.080 |
[39] |
Preparation and dielectric properties of compositionally graded (Ba, Sr)TiO3 thin film by sol-gel technique[J]. Trans Nonferrous Met Soc China, 2006, 16: S119–S122.
DOI:10.1016/S1003-6326(06)60157-X |
[40] |
Multifunctional characterization of multiferroic 0.5(NdFeO3)-0.5 (Ba0.5Sr0.5TiO3) for storage and photovoltaics applications[J]. Inorg Chem Commun, 2023, 151: 110588.
DOI:10.1016/j.inoche.2023.110588 |
[41] |
Co3O4/LaCoO3 nanocomposites derived from MOFs as anodes for high-performance lithium-ion batteries[J]. Inorg Chem Commun, 2022, 104: 109447.
DOI:10.1016/j.inoche.2022.109447 |
[42] |
One-step impregnation method to prepare direct Z-scheme LaCoO3/g-C3N4 heterojunction photocatalysts for phenol degradation under visible light[J]. Appl Surf Sci, 2019, 491: 432–442.
DOI:10.1016/j.apsusc.2019.06.143 |
[43] |
Low temperature and controllable formation of oxygen vacancy SrTiO3-x by loading Pt for enhanced photocatalytic hydrogen evolution[J]. Energy Technol, 2018, 6(11): 2166–2171.
DOI:10.1002/ente.201800181 |
[44] |
Room-temperature ferromagnetism of Co-doped Na0.5Bi0.5TiO3: Diluted magnetic ferroelectrics[J]. J Alloy Compd, 2019, 475(1/20): L25–L30.
DOI:10.1016/j.jallcom.2008.07.073 |
[45] |
Analysis of cubic perovskite oxide AgNbO3 under stress conditions: Potential for photovoltaic and high-temperature applications - A DFT study[J]. Mater Sci Semicond Process, 2024, 182: 108711.
DOI:10.1016/j.mssp.2024.108711 |
[46] | |
[47] |
The effect of calcination temperature on the photocatalytic performance of Bi2MoO6 for the degradation of phenol under visible light[J]. Optik, 2019, 199: 163319.
DOI:10.1016/j.ijleo.2019.163319 |
[48] |
Effect of MoS2 loading on photocatalytic degradation efficiency of phenol by MoS2/TiO2 and its mechanism[J]. J Fuel Chem Technol, 2017, 45(8): 1001–1008.
|
[49] |
Preparation of Ti0.7W0.3O2@BiVO4 p-n heterogeneous interfacial photocatalyst and mechanism of photodegradation of phenol by visible light[J]. Acta Sci Circumstantiae, 2020, 40(5): 1674–1691.
DOI:10.13671/j.hjkxxb.2020.0002 |
[50] |
Preparation and photocatalytic properties of AgI-BiOCl composites[J]. Appl Chem Ind, 2021, 7(50): 1751–1756.
DOI:10.16581/j.cnki.issn1671-3206.20210419.014 |
[51] |
Multifunctional strontium titanate perovskitebased composite photocatalysts for energy conversion and other applications[J]. Int J Hydrogen Energ, 2023, 48(98): 38634–38654.
DOI:10.1016/j.ijhydene.2023.06.168 |
[52] |
Preparation of LaCoO3 perovskite catalyst and its photocatalytic degradation of acid fuchsin and basic fuchsin[J]. J Mol Catal (China), 2024, 38(1): 71–80.
DOI:10.16084/j.issn1001-3555.2024.01.009 |
[53] |
Construction of TiO2@C/g-C3N4 and its photocatalytic degradation performance of tetra-cycline[J]. J Mol Catal (China), 2024, 38(4): 342–351.
DOI:10.16084/j.issn1001-3555.2024.04.005 |
[54] |
Synthesis and properties of two-dimensional composite photocatalyst[J]. J Mol Catal (China), 2024, 38(2): 99–103.
DOI:10.16084/j.issn1001-3555.2024.02.001 |
[55] |
Iron-doped bismuth sulfide PhotoFenton degrades organic pollutants[J]. J Mol Catal (China), 2024, 38(1): 17–25.
DOI:10.16084/j.issn1001-3555.2024.01.003 |
[56] |
Preparation of Z-scheme g-C3N4/Bi/BiOBr heterojunction photocatalyst and its visible light degradation of formaldehyde gas[J]. J Mol Catal (China), 2023, 37(1): 53–62.
DOI:10.16084/j.issn1001-3555.2023.01.006 |
[57] |
Preparation and photocatalytic properties of K/Cl doped g-C3N4[J]. J Mol Catal (China), 2022, 36(6): 561–570.
DOI:10.16084/j.issn1001-3555.2022.06.005 |
[58] |
Preparation of perovskite LaFexCu(1−x)O3 heterogeneous Fenton catalyst and its degradation of methylene blue[J]. J Mol Catal (China), 2021, 35(6): 503–517.
DOI:10.16084/j.issn1001-3555.2021.06.002 |
[59] |
Study on the electronic structure modulation and photocatalytic performance of bismuth oxychloride photocatalysts[J]. J Mol Catal (China), 2022, 36(5): 433–445.
DOI:10.16084/j.issn1001-3555.2022.05.003 |