Abstract
This study proposes a novel LaCoO3‑δ/H2O2 Fenton-like process for the efficient generation of nonradical singlet oxygen (1O2) and hydroxyl radicals (•OH), enabling the deep oxidation of nitric oxide (NO), as well as simultaneous denitrification and desulfurization. A LaCoO3‑δ perovskite material enriched with Co3+ ions and oxygen vacancies (OV) was successfully synthesized by using a one-step citric acid sol–gel method. OV enhances electron mobility and expedites the redox cycling of surface Co2+/Co3+ pairs, thereby promoting the generation of •OH and •O2 –. The high Co3+ content further facilitates the transformation of •O2 – into 1O2. Moreover, electron-rich centers and highly reactive lattice oxygen (O2–) induced by abundant OV also contribute to the activation of H2O2, enhancing the production of •OH and 1O2. When combined with a comprehensive Na2SO3/NaOH absorption system, simultaneous removal efficiencies of 91.2% for NO and 100% for SO2 were achieved under optimal experimental conditions while ensuring good operational stability. This innovative noniron-based Fenton-like system not only rivals the performance of typical iron-based Fenton-like systems but also provides new insights into the development of multipollutant deep oxidation removal technologies for application in coal-fired power plants.


1. Introduction
Excessive emissions of nitric oxide (NO) and sulfur dioxide (SO2) from coal-fired flue gas can cause severe environmental pollution problems. − Conventional NH3-selective catalytic reduction (NH3–SCR) denitrification and limestone-based wet flue gas (Ca-WFGD) desulfurization , technologies are often unsuitable for green and low-energy development owing to high operational costs and potential secondary environmental contamination. − In recent years, heterogeneous Fenton-like oxidation combined with absorption processes for simultaneously removing multiair pollutants has attracted attention owing to its low cost and minimal environmental impact. − In this process, cost-effective and environmentally friendly hydrogen peroxide (H2O2) is activated by iron-based heterogeneous catalysts to generate hydroxyl radicals (•OH) and superoxide radicals (•O2 –). , These reactive oxygen species (ROS) can oxidize water-insoluble NO into absorbable, higher-valence nitrogen-containing compounds. Subsequently, these compounds, along with SO2, are simultaneously absorbed by the absorption device. However, the ROS generation process is typically restricted due to the relatively slow kinetics of the Fe2+/Fe3+ redox couple, the scarcity of active sites, and self-quenching effects of ROS, leading to a low NO oxidation efficiency and excessive consumption of H2O2. , Therefore, it is highly significant yet challenging to achieve the deep oxidation of NO by promoting the efficient generation and utilization of ROS in a heterogeneous Fenton-like system.
To enhance H2O2 utilization and ROS yield, a series of Fe-based materials, including Fe0, Fe2O3, Fe–Al, Fe/TiO2, La1–x Ca x FeO3, Fe2(SO4)3, Fe2(MoO4)3, Fe/ZSM5, and Fe3O4/Fe0/Fe3C, have been comprehensively designed and developed. These studies demonstrated the efficient oxidation ability of •OH by accelerating the circulation of surface Fe2+/Fe3+ and increasing the number of active sites for H2O2 activation. Nevertheless, •OH can inevitably be consumed by certain undesirable side reactions, leading to the formation of H2O2 or •O2 –. , On the one hand, the contribution of H2O2/•O2 – to NO oxidation is less significant compared to that of •OH. On the other hand, •OH can further react with •O2 – to generate H2O and O2. Consequently, the excessive consumption of •OH and the negative effects caused by •O2 – can result in an inevitable inefficiency in H2O2 utilization and deep removal of NO.
Research indicates that •O2 – can be converted into nonradical singlet oxygen (1O2) through electron transfer during the activation of peroxymonosulfate (PMS). , Moreover, 1O2 has been demonstrated to play a crucial role in the degradation of persistent organic pollutants in water , and in the photocatalytic oxidation of NO to NO2, which is attributed to its relatively long lifetime (3.5 μs), excellent oxidation capacity, and high selectivity. Consequently, it may be feasible to improve the utilization rate of H2O2 and promote the deep oxidation of NO by transforming •O2 – into 1O2 within a Fenton-like process. Studies have shown that high-valent metal catalysts can provide an effective pathway for converting •O2 – into 1O2. Qian et al. successfully synthesized U2.5-LaFeO3 perovskite materials with high-valent iron (Fe4+) using an ammonia-mediated method, which significantly boosted the production of 1O2 through electron transfer from •O2 – to Fe4+, indicating perovskite catalysts demonstrate significant potential for the generation of 1O2 in a Fenton-like system. Besides iron-based materials, compounds containing Co4+ (CoS2‑x) and Mo6+ also facilitate the conversion of •O2 – to 1O2, providing notable advantages for the selective elimination of water pollutants. However, in these catalytic systems, the generation of •OH is restricted to maximize 1O2 production, which is not beneficial for NO oxidation. Therefore, it is essential to investigate new catalytic materials or strategies to achieve balanced regulation of 1O2 and •OH and to develop a catalytic oxidation system cooperated by nonradical (1O2) and radical (•OH) species.
Cobalt-based heterogeneous catalysts exhibit considerable potential in Fenton-like catalysis, with the rapid cycling of Co2+/Co3+ being crucial for the activation of H2O2 to generate •OH and •O2 –. , Notably, the standard reduction potential of cobalt (Co3+/Co2+ (E θ = 1.82 V)) surpasses that of iron (Fe3+/Fe2+ (E θ = 0.77 V)), indicating that Co3+ has a greater tendency to accept electron from •O2 – compared to Fe3+. Meanwhile, the electron-accepting capability of Co3+ is comparatively lower than that of Co4+. Consequently, it is speculated that Co3+ with a moderate electron-accepting ability (Fe3+ < Co3+ < Co4+) may effectively convert •O2 – into 1O2 without significantly affecting the generation of •OH. In addition, previous studies have suggested that oxygen vacancies (OV) not only facilitate the mobility of lattice oxygen but also create electron-rich centers that enhance electron transfer capabilities, consequently improving the catalytic performance of catalysts. , Therefore, materials with a high abundance of Co3+ and OV could serve as effective candidates for converting H2O2 into 1O2 and •OH. In Lanthanum-based perovskite materials (LaBO3), lanthanum predominantly exists in its trivalent state (La3+), and OV can be easily formed and controlled. , Incorporating Co into the B-site to form a OV-rich LaCoO3 perovskite may yield a promising material capable of effectively catalyzing H2O2 to generate abundant 1O2 and •OH. Recently, LaCoO3 perovskite has been reported to activate peracetic acid (PAA) and H2O2 to facilitate the generation of ROS for the degradation of organic pollutants in liquid-phase systems. This further indicates that the LaCoO3‑δ/H2O2 process holds potential for gas-phase applications such as deep oxidation of NO and efficient simultaneous denitrification and desulfurization. In this study, LaCoO3‑δ materials rich in Co3+ and OV were successfully prepared by using a one-step citric acid (CA) sol–gel method. Comprehensive characterization techniques, including X-ray diffraction (XRD), Fourier transform infrared spectroscopy (FTIR), inductively coupled plasma-optimal emission spectroscopy (ICP-OES), energy-dispersive spectroscopy (EDS), elemental analysis (EA), electron paramagnetic resonance (EPR), Brunauer–Emmett–Teller (BET), scanning electron microscopy (SEM), transmission electron microscopy (TEM), and X-ray photoelectron spectroscopy (XPS), were utilized to systematically investigate the physicochemical properties of LaCoO3‑δ. Activity assessments and SO2 resistance tests were performed to evaluate the catalytic performance and anti-SO2 capability of LaCoO3‑δ. Moreover, systematic experiments on sodium-based absorbent selection and operational parameter optimization were conducted to enhance the LaCoO3‑δ/H2O2 oxidation and absorption process to achieve deep oxidation removal of NO and SO2. Subsequently, the stability of this process was assessed through cyclic and continuous operation tests. Finally, the potential mechanisms of ROS generation and the removal pathways of NO and SO2 were elucidated based on ROS determination, quenching experiments, and product analyses.
2. Experimental Section
2.1. Materials Synthesis
The LaCoO3‑δ sample was synthesized by using the CA sol–gel method (Figure a). The detailed synthesis process of catalysts in this study is described as follows: (a) Preparation of the precursor solution: a molar ratio of 1:1:2 La(NO3)3·6H2O (Guaranteed Reagent, Macklin Reagent Company), Co(NO3)2·6H2O (Guaranteed Reagent, Macklin Reagent Company), and CA (C6H8O7·H2O, Analytical Reagent, Macklin Reagent Company) was dissolved in 80 mL of high-purity water to prepare a mixed precursor solution; (b) wet sol formation process: the precursor solution was placed in an open beaker and subjected to magnetic stirring at 300 rpm under water bath heating at 90 °C to evaporate the majority of the solvent water, resulting in the formation of a viscous wet sol; and (c) dry gel formation process: the wet sol was placed in an air-drying oven at 110 °C for 12 h to remove the residual water, resulting in the formation of the dry gel (note: the material was exposed to air in this process); (d) calcination process of dry gel: the dry gel was placed in an open crucible within a conventional muffle furnace at a heating rate of 5 °C/min, followed by calcination at 700 °C for 6 h (note: the material was exposed to static air in this process); and (e) sample grinding and sieving: following calcination, the material was ground using a ceramic mortar and subsequently sieved through a 200-mesh screen. Additionally, two other samples (La2O3 and Co3O4) were prepared by using La(NO3)3·6H2O or Co(NO3)2·6H2O as precursors through the same procedure.
1.
(a) Schematic diagram for the synthesis of the LaCoO3‑δ catalyst and (b) flow diagram of the experimental apparatus. 1N2, 2O2, 3NO, 4SO2, 5CO2, 6pressure relief valve, 7mass flow controller, 8buffer bottle, 9evaporation device, 10thermal control electric heater, 11peristaltic pump, 12H2O2 solution, 13electrical furnace, 14catalytic reactor, 15midget impinger, 16spiral condenser, 17thermostat water bath, 18Na2SO3/NaOH absorption solution, 19dry tower, and 20flue gas analyzer.
2.2. Characterization Methods
XRD (Bruker D8 Advance) was employed to analyze the crystal structure of the dry gels prior to calcination and as-prepared samples within a scan range of 10°–80° (2θ) at a scan rate of 5°/min. FTIR spectroscopy (Nicolet iZ10) was performed to identify and distinguish the characteristic species present in LaCoO3‑δ and La–Co–CA dry gel. EA (Vario EL Cube) and ICP-OES (Agilent 700 Series) analyses were conducted to determine the elemental composition and quantitative content of LaCoO3‑δ. EPR spectroscopy using a Bruker A-300 instrument was employed to investigate OV in the materials at 150 °C, as well as to identify ROS generated during the catalytic process at room temperature. Specifically, 5,5-Dimethyl-1-pyrroline (DMPO) served as the spin-trapping agent for detecting •OH and •O2 –, whereas 1O2 was identified using 2,2,6,6-tetramethyl-4-piperidone (TEMP). In quenching experiments, furfuryl alcohol (FFA) was used to scavenge both •OH (k• OH/FFA = (1.8 ± 0.2)×1010 M–1 s–1, pH 1–11) and 1O2 (k1 O2/FFA = (9.4 ± 0.1)×107 M–1 s–1, pH 3–12), tert-butanol (TBA) was employed to quench •OH (k• OH/TBA = (4.8 ± 0.3)×108 M–1 s–1, pH 3–11), and p-benzoquinone (BQ) was used to quench •O2 – (k• O2 -/FFA = 3–5 × 108 M–1 s–1, pH < 7). SEM (Hitachi S-4800) and TEM (JEM-2100F, JEM (Japan Electron Optics Laboratory electron microscope)) were conducted to examine the morphology and microstructure of LaCoO3‑δ. The BET (ASAP2020) method was applied to evaluate the specific surface area and pore structure of the LaCoO3‑δ sample. EDS (Bruker quantax400) and EDS mapping were employed to analyze the distribution of the elements. XPS (ESCALAB 250Xi) was performed to determine the chemical states and relative contents of the specific elements. Additionally, ion chromatography (IC) (Thermo ICS-5000+) was used to analyze the sulfur- and nitrogen-containing species in the spiral condenser and absorbent solution.
2.3. Experimental Procedure and Activity Tests
As illustrated in Figure b, activity tests of the as-prepared catalysts were conducted in a fixed-bed apparatus. The total flow rate of the simulated flue gas was set at 1.5 L/min and was precisely controlled using a mass flowmeter. The simulated flue gas compositions included N2, NO (100–1100 mg/m3), SO2 (500–3000 mg/m3), O2 (0–15%, v/v), and CO2 (0–20%, v/v). A peristaltic pump delivered the hydrogen peroxide solution (pH = 3–8) at a constant rate of 4.5 mL/h into a custom-designed vaporization device. Subsequently, the H2O2 vapor was thoroughly mixed with the simulated flue gas and directed to the catalyst bed (catalyst dosage: 0–0.25 g, catalytic reaction temperature: 110–210 °C), where the catalytic oxidation reaction occurred within a quartz tube measuring 150 mm in length and 10 mm in diameter. Quartz wool was employed to ensure a uniform distribution of the catalyst throughout the tube, thereby maximizing the contact between the simulated flue gas containing H2O2 vapor and the catalyst. The oxidation products were collected using an ice-bath condenser and an absorbent solution (0.1 mol/L Na2SO3 + 0.1 mol/L NaOH, 500 mL) located at the tail end of the system. Finally, the concentrations of pollutants (NO, NO2, and SO2) in the flue gas were measured by using a flue gas analyzer (Ecom-J2KN, Germany), and the removal efficiencies of NO and SO2 were calculated using the following equations:
| 1 |
| 2 |
| 3 |
where “C” denotes the pollutant concentration, mg/m3.
3. Results and Discussion
3.1. Characterization of the LaCoO3‑δ Sample
The crystal structures of the three presynthesized materials were characterized by XRD, and the results are presented in Figure a. The diffraction peaks observed at 2θ = 23.2°, 32.9°, 33.3°, 39.0°, 40.7°, and 79.5° correspond to the characteristic peaks of LaCoO3 (JCPDS PDF#84-0848), which exhibits a rhombohedral crystal structure. Furthermore, the use of lanthanum nitrate as the sole precursor resulted in the formation of La2O3 (JCPDS PDF#74-2430), whereas the use of cobalt nitrate as the sole precursor led to the formation of Co3O4 (JCPDS PDF#74-2120). However, the three precursor dry gels showed no distinct XRD characteristic peaks (Figure S1) and exhibited an amorphous structure. The average crystallite sizes of the three catalysts calculated using the Scherrer equation were 38.0 nm (LaCoO3‑δ), 18.8 nm (La2O3), and 28.6 nm (Co3O4), respectively (Table S2). In the XRD pattern of LaCoO3, the absence of characteristic peaks corresponding to La2O3 and cobalt oxides confirms the successful synthesis of high-purity LaCoO3 samples via the CA sol–gel method, which is further supported by the FTIR spectra of LaCoO3 and its precursor dry gel (Figure S2).
2.
(a) XRD patterns of LaCoO3, Co3O4, and La2O3 and (b) EPR spectra of the detected OV in LaCoO3‑δ, Co3O4, and La2O3.
CNH/O element analysis and ICP-OES tests were performed on the LaCoO3 sample to further determine the atomic ratios of each element (C, N, H, O, La, and Co), and the results are summarized in Table . The absence of C, N, and H elements indicates that the precursors (CA and NO3 –) have been completely calcined in an air atmosphere, with no other species remaining. The La/Co atomic ratio was close to the stoichiometric ratio of 1 (1.01–1.03). However, the O/La(Co) atomic ratio (2.28–2.40) was lower than the theoretical value of 3. This ratio was denoted as 3-δ (0.60 ≤ δ ≤ 0.72), and the LaCoO3 compound was denoted as LaCoO3‑δ. These findings suggest the presence of oxygen defects in the LaCoO3‑δ material. Consequently, EPR measurements were carried out on LaCoO3‑δ to quantify the OV. As depicted in Figure b, the results confirm that the LaCoO3‑δ samples exhibit a substantial number of OV, which is consistent with the results of element analysis. Additionally, EPR tests were conducted on the La2O3 and Co3O4 samples, and the oxygen vacancy concentrations in these two samples were significantly lower than those observed in LaCoO3‑δ. This difference may be attributed to the lattice defects formed during the crystallization of the perovskite structure, as a portion of the lattice oxygen in LaCoO3‑δ is derived from atmosphere oxygen (with another part originates from metal nitrate) that enters the La–Co perovskite lattice structure via bulk diffusion during the air calcination process. , Assuming that the catalyst without OV possesses an ideal chemical formula of LaCoO3, the content of OV of three samples was 24.0%, 22.3%, and 20% (relative error: 8.6%, 0.9%, 9.5%), respectively, suggesting a relatively stable OV concentration, which demonstrates that the incorporation of Co into the B-site of the LaBO3 perovskite structure enables the stable synthesis of the OV-rich LaCoO3‑δ material via the CA sol–gel method in the present study.
1. Analysis of the Chemical Elemental Composition of the LaCoO3‑δ Material.
| LaCoO3-δ sample | La (wt %) | Co (wt %) | O (wt %) | C/N/H (wt %) | atomic ratios (La/Co/O) | chemical composition | OV content (%) |
|---|---|---|---|---|---|---|---|
| 1 | 58.2 | 25.3 | 15.3 | not detected | 1:1.03:2.28 | LaCo1.03O2.28 | 24.0 |
| 2 | 58.6 | 25.0 | 15.8 | not detected | 1:1.01:2.33 | LaCo1.01O2.33 | 22.3 |
| 3 | 58.0 | 24.8 | 16.0 | not detected | 1:1.02:2.40 | LaCo1.02O2.40 | 20.0 |
| average | 58.3 | 25.0 | 15.7 | not detected | 1:1.02:2.34 | LaCo1.02O2.34 | 22.1 |
Data obtained from ICP-OES tests.
Data obtained from CNH/O EA.
Figure a–c presents the SEM and low-resolution TEM images of the LaCoO3‑δ sample, respectively. It is clearly observed that LaCoO3‑δ exhibits a uniformly spherical morphology. The average grain size was determined to be 87.0 nm by using Nano Measure 1.2 software, which is larger than the average crystallite size (38.5 nm), indicating that the LaCoO3‑δ catalyst exists as a polycrystalline aggregate material. Additionally, a slight degree of particle agglomeration may be attributed to intergranular electrostatic attraction. The BET test of LaCoO3‑δ revealed a specific surface area of 7.1 m2/g. The relatively low specific surface area was primarily attributed to the high-temperature sintering process during synthesis. The adsorption–desorption isotherms are presented in Figure S3, the average pore diameter was 16.4 nm, and the total pore volume was 0.03 cm3/g. These mesoporous and macroporous structures were formed due to the substantial release of CO2 and NO2 during the pyrolysis of CA and nitrates, which are expected to positively influence gas–solid heterogeneous catalysis reactions.
3.
(a,b) SEM images of LaCoO3‑δ; (c) low-resolution TEM image of LaCoO3‑δ; (d) high-resolution TEM images of LaCoO3‑δ; and (e) EDS elemental mapping images of LaCoO3‑δ.
Figure d shows the high-resolution TEM image of LaCoO3‑δ, where the crystalline interplanar spacings of 0.27 and 0.22 nm correspond to the (−1 1 0) and (0 2 0) planes of rhombohedral LaCoO3‑δ, respectively. These results are consistent with the XRD data and further confirm the successful synthesis of LaCoO3‑δ. EDS mapping was performed to investigate the elemental distribution across the LaCoO3‑δ sample, as shown in Figure e. The results revealed that La, Co, and O were uniformly dispersed on the surface of the sample, which facilitated the exposure and dispersion of active sites and significantly enhanced the catalytic performance.
3.2. Oxidation Removal Performances of NO and SO2
3.2.1. Evaluation of Catalytic Activity and SO2 Resistance
The catalytic activities of the three samples were evaluated using the NO conversion efficiency as an indicator. The results are shown in Figure a. The catalytic activity followed the order LaCoO3‑δ > Co3O4 > La2O3. Notably, La2O3 exhibited almost negligible catalytic activity, while LaCoO3‑δ demonstrated the highest catalytic performance. This suggests that the LaCoO3‑δ/H2O2 system generated the largest amount of ROS, thereby achieving the highest NO conversion efficiency. Subsequently, 1000 mg/m3 of SO2 was introduced into the catalytic reaction system to evaluate its impact on the catalytic performances of Co3O4 and LaCoO3‑δ. The experimental results are presented in Figure b. Prior to the addition of SO2, both systems maintained stable NO conversion rates (91.0% for LaCoO3‑δ and 74.0% for Co3O4) and NO2 production levels (287 mg/m3 for LaCoO3‑δ and 114 mg/m3 for Co3O4), with a higher NO conversion rate and NO2 production observed in the LaCoO3‑δ system compared to the Co3O4 system. Upon the introduction of SO2 (t = 30 min), the NO conversion rate of the LaCoO3‑δ system only decreased by 2.0%, from 91.0% to 89.0%, but subsequently stabilized. In contrast, the NO conversion rate in the Co3O4 system continuously declined by 23%, from 74.0% to 51.0%, between 30 and 210 min before stabilizing. Moreover, the variation trend of the NO2 production was consistent with that of the NO conversion rate in both systems. These results indicate that the LaCoO3‑δ system exhibits strong resistance to SO2 interference, whereas the Co3O4 system demonstrates relatively weak tolerance to SO2. When the supply of SO2 was terminated (t = 270 min), the NO conversion rate of the LaCoO3‑δ system returned to its initial value, whereas that of the Co3O4 system increased by approximately 1.5% over the subsequent 3 h, further suggesting that LaCoO3‑δ exhibits greater stability than Co3O4 under SO2 exposure conditions.
4.
(a) Catalytic activity tests of LaCoO3‑δ, Co3O4, and La2O3 and (b) SO2 tolerance tests of LaCoO3‑δ and Co3O4.
3.2.2. Evaluation of Sodium-Based Absorbent
To achieve more efficient oxidation of NO and simultaneous absorption of NO2 and SO2, the performance of sodium-based absorbents (Na2SO3 and NaOH), which have demonstrated excellent effectiveness in previous studies, , was evaluated in combination with the LaCoO3‑δ/H2O2 system, and the results are shown in Figure . Using Na2SO3 alone exhibited relatively high absorption efficiency for NO2 and SO2; however, the desulfurization efficiency of this system gradually decreased over time, indicating limited stability. Using NaOH alone achieved 100% desulfurization efficiency but demonstrated inferior absorption performance for NO2 compared to that for Na2SO3. In contrast, the combination of Na2SO3 and NaOH maintained a relatively high level of simultaneous desulfurization and denitrification efficiency over an extended period, suggesting the superior absorption capacity and stability of the combined absorbent system.
5.

Oxidation–absorption properties of NO/SO2/NO2 in the LaCoO3‑δ/H2O2 + sodium-based absorbents (A): 0.2 mol/L Na2SO3 solution, (B): 0.2 mol/L NaOH solution, and (C): 0.1 mol/L NaOH + 0.1 mol/L Na2SO3 mixed solution systems.
Based on the above experimental results, it can be concluded that the LaCoO3‑δ/H2O2 catalytic oxidation system effectively promotes the oxidation of NO, while the Na2SO3–NaOH absorption system achieves highly efficient simultaneous absorption of NO2 and SO2. Therefore, the deep oxidation-removal of NO and SO2 was accomplished through the synergistic action of the combined LaCoO3‑δ/H2O2 and Na2SO3–NaOH system.
3.2.3. Effects of Experimental Parameters
Given the outstanding catalytic performance of the LaCoO3‑δ catalyst and its potential for simultaneous desulfurization and denitrification, various experimental conditions, including H2O2 concentration (molar ratio of H2O2 to NO), catalyst dosage (gas hourly space velocity, GHSV), catalytic temperature, pH of the H2O2 solution, and gas compositions (NO, SO2, O2, and CO2), were systematically investigated, and the results are shown in Figure .
6.
(a) Effect of H2O2 concentration (molar ratio of H2O2 to NO) on simultaneous removal of NO and SO2 in the LaCoO3‑δ/H2O2 system; (b) effects of LaCoO3‑δ dosage and GHSV on simultaneous removal of NO and SO2; (c) effect of reaction temperature on simultaneous removal of NO and SO2; (d) effect of H2O2 solution pH on simultaneous removal of NO and SO2; (e) effect of NO concentration on simultaneous removal of NO and SO2; (f) effect of SO2 concentration on simultaneous removal of NO and SO2; (g) effect of O2 concentration on simultaneous removal of NO and SO2; (h) effect of CO2 concentration on simultaneous removal of NO and SO2; and (i) cyclic and continuous running experiments for the simultaneous removal of NO and SO2.
It is evident that under all operating conditions, SO2 was entirely removed by the absorption liquid, and its removal efficiency showed minimal sensitivity to variations in the experimental conditions. This clearly highlights the exceptional desulfurization performance of this system. In contrast, the NO removal efficiency was considerably influenced by the various experimental parameters. Notably, the concentration of H2O2 played a critical role in NO removal (Figure a). As the concentration of H2O2 increased from 0.2 to 1.0 mol/L and the molar ratio rise from 0.6 to 3, the NO removal efficiency progressively improved. This improvement can be attributed to the increased generation of ROS resulting from the reaction between LaCoO3‑δ and higher concentrations of H2O2, which enhances the oxidation rate of NO and, consequently, boosts its removal efficiency. However, as the concentration of H2O2 continued to increase beyond this range, the rate of improvement in the NO removal efficiency gradually diminished. This phenomenon was primarily due to the saturation of the active sites of the catalyst and radical self-quenching. , Therefore, taking into account both removal efficiency and economic feasibility, as indicated by the red circular annotations, we have determined 1 mol/L H2O2 (with a molar ratio of H2O2 to NO of 3) as the optimal experimental condition. By increasing the dosage of LaCoO3‑δ (Figure b), the number of active sites for the catalytic reaction was directly enhanced, and the GHSV of the system decreased. This prolongs the contact time between H2O2 and LaCoO3‑δ, thereby promoting the complete reaction of H2O2 with the catalyst and facilitating ROS generation. Considering overall cost and treatment capacity, an optimal LaCoO3‑δ dosage of 0.15 g (corresponding to a GHSV of 150,000 h–1) was determined and highlighted by red circles in Figure b. As shown in Figure c, as the temperature increased, the NO removal efficiency initially exhibited a significant enhancement, followed by a slight decline. On the one hand, an increase in temperature enhanced the reaction rate constant and facilitated the atomization and diffusion of H2O2, thereby accelerating the catalytic process. On the other hand, H2O2 tends to decompose into H2O and O2 at excessively high temperatures, leading to a reduction in the formation of ROS, which negatively influences the removal efficiency of NO. , Consequently, as indicated by the red circle in Figure c, 150 °C represents the optimal temperature condition for the experiment. In Figure d, under acidic conditions, the removal efficiency of NO was maintained at a high and stable level. In contrast, in neutral or alkaline environments, the ineffective decomposition of H2O2 leads to a marked decrease in the NO removal efficiency. As the pH of the 1 mol/L H2O2 solution was 5.3, no additional pH adjustment of the solution was required, and pH 5.3 was the optimal parameter in this study, as highlighted by the red circle in Figure d.
It is important to examine the effects of coexisting gases on NO removal efficiency because the load of industrial boilers typically influences the concentration of pollutants and the composition of gas components; the results are presented in Figure e–h. As shown in Figure e, as the NO concentration increased, the removal efficiency of NO decreased owing to the limited availability of ROS for oxidizing excessive NO. In Figure f, the NO removal efficiency exhibited an initial increase, followed by a decrease as the SO2 concentration increased. An appropriate level of SO2 can enhance the conversion and absorption of NO and NO2; however, excessive SO2 competes with NO for the limited oxidant, leading to a reduction in NO removal efficiency. As shown in Figure g, the presence of O2 accelerates the absorption of NO and NO2, improving the NO removal efficiency to some extent. Furthermore, CO2 had a relatively minor impact on the NO removal efficiency (Figure h).
Based on the above experimental results, the optimal reaction parameters for the catalytic system are determined as follows: H2O2 concentration: 1 mol/L (molar ratio of H2O2/NO: 3), LaCoO3‑δ dosage: 0.15 g, reaction temperature: 150 °C, pH of H2O2 solution: 5.3, NO concentration: 500 mg/m3, SO2 concentration: 2000 mg/m3, O2 concentration: 6% (v/v), and CO2 concentration 8% (v/v). Subsequently, a systematic test of the stability of LaCoO3‑δ was conducted under optimal experimental conditions, and the experimental results are presented in Figure i. The optimal simultaneous removal efficiencies of NO and SO2 were 91.2% and 100%, respectively. After undergoing ten experimental cycles, the catalytic performance of the catalyst did not show a significant decline. Furthermore, the removal efficiency of NO decreased by only 0.9% after continuous operation for 18 h. Additionally, no new diffraction peaks were observed in the XRD spectrum of the spent LaCoO3‑δ sample (Figure S4), which indicates the excellent stability of LaCoO3‑δ. The energy consumption rates in this study were calculated to be 7.61 × 10–3 kWh/g (NO) and 1.74 × 10–3 kWh/g (SO2) (calculation process provided in the Supporting Information), suggesting good potential for industrial application.
3.3. ROS Generation Mechanisms
EPR spectroscopy was employed to investigate the generation of ROS (1O2, •OH, and •O2 –) in the LaCoO3‑δ/H2O2 and Co3O4/H2O2 systems. In Figure a–c, it is observed that the signal intensities of the three ROS in the H2O2 system were nearly negligible, indicating that H2O2 scarcely generated ROS in the absence of a catalyst. In the LaCoO3‑δ/H2O2 system, the relative signal intensities followed the order 1O2 (1.1× 1017 spins/mL) > •OH (6.1× 1016 spins/mL) ≈ •O2 – (6.0 × 1016 spins/mL), whereas in the Co3O4/H2O2 system, the order was •O2 – (8.9× 1016 spins/mL) > •OH (5.7 × 1016 spins/mL) > 1O2 (5.2 × 1017 spins/mL). As shown in Figure c, although the generation of •O2 – in the Co3O4/H2O2 system exceeded that in the LaCoO3‑δ/H2O2 system, the catalytic activity of Co3O4 was nonetheless lower than that of LaCoO3‑δ, indicating that •O2 – played a relatively minor role in the NO oxidation process. Additionally, the LaCoO3‑δ/H2O2 system exhibited a higher production of 1O2 than the Co3O4/H2O2 system (Figure a), while both systems demonstrated similar levels of •OH generation (Figure b). These findings confirm that 1O2 is primarily responsible for the superior catalytic performance of LaCoO3‑δ over Co3O4. To further investigate the contributions of various ROS to the NO oxidation process, quenching experiments were conducted in the LaCoO3‑δ/H2O2 system, and the results are presented in Figure d. In the absence of quenchers, the NO removal efficiency was 90.3%. Upon the addition of FFA, TBA, and BQ, the NO removal efficiency decreased by 39.8% (20.9% for 1O2), 18.9% (•OH), and 15.7% (•O2 –), respectively. The contribution of •O2 – (15.7%) was deemed unreliable owing to its relatively low oxidation capacity, and it was hypothesized that the formation of 1O2 was restricted because •O2 –, acting as an intermediate in the generation of 1O2, being quenched by BQ. Consequently, an EPR test was conducted to measure 1O2 in the LaCoO3‑δ/H2O2/BQ system to further verify this hypothesis. As shown in Figure e, the signal intensity of 1O2 diminished upon the introduction of BQ, indicating that the quenching of •O2 – resulted in a reduction in the generation of 1O2, thereby confirming that •O2 – served as an intermediate in the formation of 1O2. Therefore, the cooperation of 1O2 and •OH in the LaCoO3‑δ/H2O2 system played a significant role in the oxidation of NO. Furthermore, it is important to investigate the SO2 oxidation process. Radical quenching experiments during SO2 oxidation in the LaCoO3‑δ/H2O2 system were conducted under both NO-containing and NO-free conditions. As shown in Figure f, under NO-free conditions, the addition of FFA and TBA led to a reduction in the SO2 oxidation efficiency by 9.9% and 9.2%, respectively. Under NO-containing conditions, the corresponding reductions were 6.4% and 6.7%, respectively. In addition, the introduction of BQ exerts a negligible influence on the oxidation rate of SO2. These results indicate that 1O2 and •O2 – play a negligible role in the oxidation of SO2, whereas •OH and H2O2 are identified as the predominant oxidative species responsible for SO2 oxidation.
7.
(a) EPR spectra of the detected TEMP-1O2 in H2O2, Co3O4/H2O2, and LaCoO3‑δ/H2O2 systems; (b) EPR spectra of the detected DMPO–OH in H2O2, Co3O4/H2O2, and LaCoO3‑δ/H2O2 systems; (c) EPR spectra of the detected DMPO-O2 – in H2O2, Co3O4/H2O2, and LaCoO3‑δ/H2O2 systems; (d) radical quenching experiments during NO removal for the LaCoO3‑δ/H2O2 system; (e) EPR spectra of the detected TEMP-1O2 in H2O2, LaCoO3‑δ/H2O2, and LaCoO3‑δ/H2O2/BQ systems; (f) radical quenching experiments during SO2 oxidation for the LaCoO3‑δ/H2O2 system; (g) the La 3d XPS spectra of fresh and spent LaCoO3‑δ; (h) the Co 2p XPS spectra of fresh and spent LaCoO3‑δ; and (i) the O 1s XPS spectra of fresh and spent LaCoO3‑δ.
To investigate the generation mechanisms of ROS and the catalytic reactions occurring in the LaCoO3‑δ/H2O2 system, XPS analysis was performed on the La 3d, Co 2p, and O 1s peaks for both fresh and spent LaCoO3‑δ samples. In Figure g, La 3d5/2 and La 3d3/2 peaks were observed at binding energies of 834.0–838.3 eV and 850.8–855.1 eV, respectively. The spin–orbit splitting of the La 3d level was measured at 16.8 eV, which is consistent with the values of pure La2O3, demonstrating the presence of stable La3+ in the fresh and spent LaCoO3‑δ samples, indicating La3+ was not involved in the catalytic reaction. As shown in Figure h, the Co 2p spectrum of fresh LaCoO3‑δ exhibits two cobalt surface cobalt species. The binding energies of peaks at 779.8 and 795.0 eV were corresponded to Co3+, while those at 782.0 and 797.2 eV were assigned to Co2+. Quantitative analysis revealed that the molar ratio of surface Co3+/Co2+ was 67.6%/32.4%, suggesting a relatively high Co3+ content in the fresh LaCoO3‑δ samples. The formation of Co2+ in the perovskite structure is primarily attributed to the generation of OV to ensure charge balance. , Following the reaction between LaCoO3‑δ and H2O2, the Co3+ content decreased, while the Co2+ content increased, as evidenced by the binding energy of the Co3+ peak shifting to a higher value (from 779.8 to 780.3 eV for Co 2p3/2) resulting from a decrease of electron cloud density on the Co atom. These findings indicate that the Co3+/Co2+ redox pair plays a critical role in the catalytic activation of H2O2. , In Figure i, the O 1s spectra of both fresh and spent LaCoO3‑δ samples displayed three predominant oxygen species: lattice oxygen Olat (O2–, 528.9 or 529.3 eV), absorbed H2O or surface hydroxyl (OI, 532.9 eV), and surface-chemisorbed oxygen (OII, 531.2 eV) originating from OV. ,, In fresh LaCoO3‑δ, the relatively low content of OI (8%) indicated a limited presence of water molecules or surface hydroxyl groups, whereas the high level of OII (51%) suggested an abundance of electron-donating centers attributed to substantial OV. After the reaction, the contents of OII and Olat decreased by 13% and 7%, respectively. Moreover, the binding energy of Olat shifted to a higher value, suggesting a decrease in electron density on the O atom. In contrast, the OI content increased by 20%, indicating that electron-rich structures (originating from OV) and lattice oxygen (O2–) are involved in the catalytic decomposition of H2O2 and form surface hydroxyl groups or water molecules.
Thus, the potential ROS generation mechanisms are proposed as follows (Figure a): the cyclic redox pair of surface Co2+/Co3+ can catalyze H2O2 to generate •OH (eq ) and •O2 – (eq ). ,, The abundant surface Co3+ species can further react with intermediate •O2 – to form 1O2 (eq ). The presence of electron-rich OV (OV(e–)) enhanced the electron transfer capability and facilitated the transformation between surface Co3+ and Co2+ (eq ). , Additionally, OV reduces the stability of O2– and promote its reactivity, enabling it to react with •O2 – to generate 1O2 (eq ). Moreover, the electron-rich centers on the surface can donate electrons to H2O2, thereby generating more •OH (eq ).
| 4 |
| 5 |
| 6 |
| 7 |
| 8 |
| 9 |
8.
(a) Potential ROS generation mechanism in the LaCoO3‑δ/H2O2 system and (b) products analysis and N/S balances.
3.4. Deep Oxidation Removal Pathways of NO and SO2
To investigate the reaction pathways of NO and SO2 in both the oxidation and absorption systems, the oxidation products from the LaCoO3‑δ/H2O2 catalytic system and the absorption products from the Na2SO3/NaOH absorption system were analyzed by using exhaust gas detection and IC technology, respectively. Additionally, mass balance analysis was performed for N- and S-containing species, as presented in Figure b. The results demonstrated that the primary oxidation products of NO were NO2 (44.7%), NO2 – (12.2%), and NO3 – (43.1%), demonstrating a higher ratio of (NO2 + NO3 –) compared to previous studies which reported that •OH was the predominant ROS in the Fenton-like system, , indicating that the synergistic effect of 1O2 and •OH can lead to a more extensive oxidation of NO compared to •OH acting alone. Following the absorption treatment, the predominant absorption products were NO3 – (62.3%) and NO2 – (37.7%). The relative errors in the N balance during the oxidation and absorption processes were 9.8% and 6.8%, respectively. Regarding SO2, approximately 17.6% was oxidized to SO4 2– during the oxidation process, while the remaining 83.4% was fully removed and converted to SO4 2– during the absorption process. The relative errors in the S-element balance for the two processes were 7.7% and 6.6%, respectively. These findings suggest that within this oxidation–absorption reaction system, NO and SO2 are effectively removed with the efficient transformation of their respective products. Based on the above analysis and relevant refs , , , – , and , this study proposes the following possible reaction pathways for the simultaneous removal of NO and SO2. During the oxidation stage, the majority of NO was efficiently oxidized to NO2, NO2 –, and NO3 – via reactions with H2O2, •OH, and 1O2. A minor fraction of SO2 competes with NO for active oxidative species, leading to its oxidation to SO4 2–. In the absorption stage, residuals of NO2 and SO2 were absorbed by Na2SO3 and NaOH, thereby generating NO2 – and SO3 2–. Finally, NO2 – and SO3 2– can be oxidized by the O2 to generate NO3 – and SO4 2–.
Oxidation process
| 10 |
| 11 |
| 12 |
| 13 |
| 14 |
| 15 |
| 16 |
Absorption process
| 17 |
| 18 |
| 19 |
| 20 |
4. Conclusions
An innovative LaCoO3‑δ/H2O2 Fenton-like process was systematically investigated for the simultaneous removal of NO and SO2 in the present study. Based on the experimental results and characterization analysis, the following conclusions were drawn:
-
(1)
Under the optimum operating conditions (molar ratio of H2O2 to NO, 3; LaCoO3‑δ dosage, 0.15 g; GHSV, 150,000 h–1; catalytic reaction temperature: 150 °C; and H2O2 pH, 5.3), 91.2% of NO and 100% of SO2 can be simultaneously removed, with energy consumption rates of 7.61 × 10–3 kWh/g (NO) and 1.74 × 10–3 kWh/g (SO2), showing a promising prospect for industrial application prospect. After undergoing ten cycles and 18 h of continuous operation, the denitrification efficiency of the system decreased by 0.7% and 0.9%, respectively, which demonstrates the exceptional stability of LaCoO3‑δ.
-
(2)
The surface Co3+/Co2+ redox pairs and abundant OV in the LaCoO3‑δ catalyst serve as the primary active sites for activating H2O2 to generate ROS (•OH, •O2 –, and 1O2). Among these, •OH and 1O2 are the predominant ROS responsible for the deep oxidation of NO.
-
(3)
NO2 (44.7%) and NO3 – (43.1%) were identified as the primary oxidation products, whereas NO2 – (37.7%), NO3 – (62.3%), and SO4 2– (100%) were the main absorption products. Furthermore, a good N/S balance suggests that both the LaCoO3‑δ/H2O2 advanced oxidation system and the Na2SO3/NaOH absorption system exhibit high efficiency and stability.
-
(4)
This study proposes a novel strategy for the effective utilization of H2O2 in Fenton-like advanced oxidation processes by the co-operation of 1O2 and •OH and offers sustainable implications for the development of simultaneous denitrification and desulfurization technologies.
Supplementary Material
Acknowledgments
The authors appreciate the financial support provided by the Science and Technology Project of Hebei Education Department (BJ2025008), the Natural Science Foundation of Hebei Province (No. E2024502082), and the Hebei Province Youth Talent Support Program (BJK2023089).
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.5c04716.
XRD analysis of dry gels prior to calcination; FTIR spectroscopy of LaCoO3‑δ and its dry gel precursor; adsorption/desorption isotherms of LaCoO3‑δ samples; XRD patterns of fresh and spent LaCoO3‑δ samples; pH measurement of precursor solution; average crystallite size of LaCoO3‑δ, La2O3, and Co3O4 from XRD; semiquantitative comparison results of EPR signals in the LaCoO3‑δ/H2O2 and Co3O4/H2O2 systems; calculation of energy consumption rates for simultaneous removal of NO and SO2; experimental conditions in this study; and comparative analysis of the present work and representative Fe-based Fenton-like systems (PDF)
Xingzhou Mao: Methodology, validation, formal analysis, and writingoriginal draft. Yujia Wang: Methodology, validation, and investigation. Zhipeng Ma: Methodology and investigation. Runlong Hao: Conceptualization, writingreview and editing, and supervision. Dong Fu: Conceptualization and supervision. Bo Yuan: Conceptualization, resources, writingreview and editing, and project administration. Yi Zhao: Conceptualization, resources, writingreview and editing, supervision, and project administration.
Natural Science Foundation of Hebei Province (No. E2024502082 and E2021502002) and Hebei Province Youth Talent Support Program (BJK2023089)
The authors declare no competing financial interest.
References
- Wang Y., Kong D., Gao X., Pan J., Liu Y.. Removal of SO2 and NO in flue gas using a vacuum ultraviolet light/Oxone/H2O2 oxidation system. Chem. Eng. J. 2025;390:134537. doi: 10.1016/j.fuel.2025.134537. [DOI] [Google Scholar]
- Yuan B., Mao X., Wang Z., Hao R., Zhao Y.. Radical-induced oxidation removal of multi-air-pollutant: A critical review. J. Hazard. Mater. 2020;383:121162. doi: 10.1016/j.jhazmat.2019.121162. [DOI] [PubMed] [Google Scholar]
- Liu X., Wang C., Zhu T., Lv Q., Che D.. Simultaneous removal of SO2 and NOx with •OH from the catalytic decomposition of H2O2 over Fe-Mo mixed oxides. J. Hazard. Mater. 2021;404:123936. doi: 10.1016/j.jhazmat.2020.123936. [DOI] [PubMed] [Google Scholar]
- Wu H., Li J., Gao M., Chen Y., Ren S., Yang J., Liu Q.. Deactivation mechanisms and strategies to mitigate deactivation of iron-based catalysts in NH3-SCR for NOx reduction: A comprehensive review. Sep. Purif. Technol. 2025;358:130268. doi: 10.1016/j.seppur.2024.130268. [DOI] [Google Scholar]
- Wang H., Yuan B., Hao R., Zhao Y., Wang X.. A critical review on the method of simultaneous removal of multi-air-pollutant in flue gas. Chem. Eng. J. 2019;378:122155. doi: 10.1016/j.cej.2019.122155. [DOI] [Google Scholar]
- Tang X., Wu P., Wang Y., Liu Y.. Recent advances in heavy metal poisoning mechanism and regeneration methods of selective catalytic reduction (SCR) denitration catalyst. Fuel. 2024;355:129429. doi: 10.1016/j.fuel.2023.129429. [DOI] [Google Scholar]
- Wang B., Song Z., Sun L.. A review: Comparison of multi-air-pollutant removal by advanced oxidation processes – Industrial implementation for catalytic oxidation processes. Chem. Eng. J. 2021;409:128136. doi: 10.1016/j.cej.2020.128136. [DOI] [Google Scholar]
- Fang P., Cen C., Wang X., Tang Z. j., Tang Z. x., Chen D. s.. Simultaneous removal of SO2, NO and Hg0 by wet scrubbing using urea + KMnO4 solution. Fuel Process. Technol. 2013;106:645–653. doi: 10.1016/j.fuproc.2012.09.060. [DOI] [Google Scholar]
- Huang X., Ding J., Zhong Q.. Catalytic decomposition of H2O2 over Fe-based catalysts for simultaneous removal of NOX and SO2 . Appl. Surf. Sci. 2015;326:66–72. doi: 10.1016/j.apsusc.2014.11.088. [DOI] [Google Scholar]
- Song Z., Wang B., Yang W., Chen T., Ma C., Sun L.. Simultaneous removal of NO and SO2 through heterogeneous catalytic oxidation-absorption process using magnetic Fe2.5M0.5O4 (M = Fe, Mn, Ti and Cu) catalysts with vaporized H2O2 . Chem. Eng. J. 2020;386:123883. doi: 10.1016/j.cej.2019.123883. [DOI] [Google Scholar]
- Jia S., Pu G., Gao J., Yuan C.. Oxidation-absorption process for simultaneous removal of NOx and SO2 over Fe/Al2O3@SiO2 using vaporized H2O2 . Chemosphere. 2022;291:133047. doi: 10.1016/j.chemosphere.2021.133047. [DOI] [PubMed] [Google Scholar]
- Choe Y., Kim J., Choi I., Kim S.. Metal oxides for Fenton reactions toward radical-assisted water treatment: A review. J. Ind. Eng. Chem. 2025;142:127–140. doi: 10.1016/j.jiec.2024.08.006. [DOI] [Google Scholar]
- Yang X., Tian P., Wang H., Xu J., Han Y.. Catalytic decomposition of H2O2 over a Au/carbon catalyst: A dual intermediate model for the generation of hydroxyl radicals. J. Catal. 2016;336:126–132. doi: 10.1016/j.jcat.2015.12.029. [DOI] [Google Scholar]
- Song Z., Wang B., Yang W., Chen T., Li W., Ma C., Sun L.. Research on NO and SO2 removal using TiO2-supported iron catalyst with vaporized H2O2 in a catalytic oxidation combined with absorption process. Environ. Sci. Pollut. Res. 2020;27:18329–18344. doi: 10.1007/s11356-020-08042-6. [DOI] [PubMed] [Google Scholar]
- Ding J., Zhong Q., Zhang S., Song F., Bu Y.. Simultaneous removal of NOX and SO2 from coal-fired flue gas by catalytic oxidation-removal process with H2O2 . Chem. Eng. J. 2014;243:176–182. doi: 10.1016/j.cej.2013.12.101. [DOI] [Google Scholar]
- Wu B., Xiong Y., Ru J., Feng H.. Enhancement of NO absorption in ammonium-based solution using heterogeneous Fenton reaction at low H2O2 consumption. Korean J. Chem. Eng. 2016;33:3407–3416. doi: 10.1007/s11814-016-0195-2. [DOI] [Google Scholar]
- Zhao Y., Yuan B., Hao R., Tao Z.. Low-temperature conversion of NO in flue gas by vaporized H2O2 and nanoscale zerovalent iron. Energy Fuels. 2017;31:7282–7289. doi: 10.1021/acs.energyfuels.7b00588. [DOI] [Google Scholar]
- Song Z., Wang B., Yu J., Ma C., Zhou C., Chen T., Yan Q., Wang K., Sun L.. Density functional study on the heterogeneous oxidation of NO over α-Fe2O3 catalyst by H2O2: Effect of oxygen vacancy. Appl. Surf. Sci. 2017;413:292–301. doi: 10.1016/j.apsusc.2017.04.011. [DOI] [Google Scholar]
- Ding J., Zhong Q., Zhang S., Cai W.. Size- and shape-controlled synthesis and catalytic performance of iron-aluminum mixed oxide nanoparticles for NOx and SO2 removal with hydrogen peroxide. J. Hazard. Mater. 2015;283:633–642. doi: 10.1016/j.jhazmat.2014.10.010. [DOI] [PubMed] [Google Scholar]
- Chen L., Li Y., Zhao Q., Wang Y., Liang Z., Lu Q.. Removal of NOX using hydrogen peroxide vapor over Fe/TiO2 catalysts and an absorption technique. Catal. 2017;7:386. doi: 10.3390/catal7120386. [DOI] [Google Scholar]
- Wu B., Xiong Y.. A novel low-temperature NO removal approach with •OH from catalytic decomposition of H2O2 over La1‑xCaxFeO3 oxides. J. Chem. Technol. Biotechnol. 2018;93:43–53. doi: 10.1002/jctb.5317. [DOI] [Google Scholar]
- Wu B., Xiong Y., Ge Y.. Simultaneous removal of SO2 and NO from flue gas with •OH from the catalytic decomposition of gas-phase H2O2 over solid-phase Fe2(SO4)3 . Chem. Eng. J. 2018;331:343–354. doi: 10.1016/j.cej.2017.08.097. [DOI] [Google Scholar]
- Liu X., Wang C., Zhu T., Lv Q., Li Y., Che D.. Simultaneous removal of NOx and SO2 from coal-fired flue gas based on the catalytic decomposition of H2O2 over Fe2(MoO4)3 . Chem. Eng. J. 2019;371:486–499. doi: 10.1016/j.cej.2019.04.028. [DOI] [Google Scholar]
- Yuan B., Zhao Y., Mao X., Zheng Z., Hao R.. Simultaneous removal of SO2, NO and Hg0 from flue gas using vaporized oxidant catalyzed by Fe/ZSM-5. Fuel. 2020;262:116567. doi: 10.1016/j.fuel.2019.116567. [DOI] [Google Scholar]
- Yang S., Xu D., Yan W., Xiong Y.. Effective NO and SO2 removal from fuel gas with H2O2 catalyzed by Fe3O4/Fe0/Fe3C encapsulated in multi-walled carbon nanotubes. J. Environ. Chem. Eng. 2021;9:105413. doi: 10.1016/j.jece.2021.105413. [DOI] [Google Scholar]
- Tan C., Xu Q., Sheng T., Cui X., Wu Z., Gao H., Li H.. Reactive oxygen species generation in FeOCl nanosheets activated peroxymonosulfate system: radicals and non-radical pathways. J. Hazard. Mater. 2020;398:123084. doi: 10.1016/j.jhazmat.2020.123084. [DOI] [PubMed] [Google Scholar]
- Wang H. Z., Guo W. Q., Si Q. S., Liu B., Zhao Q., Luo H., Ren N.. Non-covalent doping of carbon nitride with biochar: boosted peroxymonosulfate activation performance and unexpected singlet oxygen evolution mechanism. Chem. Eng. J. 2021;418:129504. doi: 10.1016/j.cej.2021.129504. [DOI] [Google Scholar]
- Sheng J., Lu A., Guo S., Shi Y., Wu H., Jiang H.. Superior bensulfuron-methyl degradation performance of sphercial magnetic Co3O4@C-500/peroxymonosulfate system by enhancing singlet oxygen generation: The effect of oxygen vacancies. Chem. Eng. J. 2023;471:143945. doi: 10.1016/j.cej.2023.143945. [DOI] [Google Scholar]
- Liang H., Zhu H., Zhang M., Hou S., Li Q., Yang J.. Oxygen vacancies promoted the generation of sulfate radicals and singlet oxygen by peroxymonosulfate activation with Co3O4 quantum dots/g-C3N4 nanosheets. Chem. Eng. Sci. 2024;284:119463. doi: 10.1016/j.ces.2023.119463. [DOI] [Google Scholar]
- Li Q., Zhao J., Shang H., Ma Z., Cao H., Zhou Y., Li G., Zhang D., Li H.. Singlet oxygen and mobile hydroxyl radicals co-operating on gas–solid catalytic reaction interfaces for deeply oxidizing NOx . Environ. Sci. Technol. 2022;56:5830–5839. doi: 10.1021/acs.est.2c00622. [DOI] [PubMed] [Google Scholar]
- Qian Z., Luo M., Feng X., Yang X., Guo Y., Yang L., Yuan B., Zhao Y., Hao R.. Unexpected role of singlet oxygen generated in urea-modulated LaFeO3 perovskite heterogeneous Fenton system in denitrification: Mechanism and theoretical calculations. Appl. Catal. B Environ. 2023;338:123042. doi: 10.1016/j.apcatb.2023.123042. [DOI] [Google Scholar]
- Ji H., Yan Q., Yin P., Mine S., Matsuoka M., Xing M.. Defects on CoS2–x: tuning redox reactions for sustainable degradation of organic pollutants. Angew. Chem., Int. Ed. 2021;60:2903–2908. doi: 10.1002/anie.202013015. [DOI] [PubMed] [Google Scholar]
- Yi Q., Ji J., Shen B., Dong C., Liu J., Zhang J., Xing M.. Singlet oxygen triggered by superoxide radicals in a molybdenum cocatalytic Fenton reaction with enhanced REDOX activity in the environment. Environ. Sci. Technol. 2019;53:9725–9733. doi: 10.1021/acs.est.9b01676. [DOI] [PubMed] [Google Scholar]
- Huang X., Wu Q., Tang J., Ning R., Yuan J., Fang S., Yu S., Hou L., You Q.. Unraveling the role of V modified CoO in a wide pH range Fenton-like process. J. Environ. Manage. 2025;375:124404. doi: 10.1016/j.jenvman.2025.124404. [DOI] [PubMed] [Google Scholar]
- Xie L., Liu X., Chang J., Zhang C., Li Y., Zhang H., Zhan S., Hu W.. Enhanced redox activity and oxygen vacancies of perovskite triggered by copper incorporation for the improvement of electro-Fenton activity. Chem. Eng. J. 2022;428:131352. doi: 10.1016/j.cej.2021.131352. [DOI] [Google Scholar]
- Zhang Y., Xiao R., Wang S., Zhu H., Song H., Chen G., Lin H., Zhang J., Xiong J.. Oxygen vacancy enhancing Fenton-like catalytic oxidation of norfloxacin over prussian blue modified CeO2: Performance and mechanism. J. Hazard. Mater. 2020;398:122863. doi: 10.1016/j.jhazmat.2020.122863. [DOI] [PubMed] [Google Scholar]
- Zhang C., Wang C., Zhan W., Guo Y., Guo Y., Lu G., Baylet A., Giroir-Fendler A.. Catalytic oxidation of vinyl chloride emission over LaMnO3 and LaB0.2Mn0.8O3 (B = Co, Ni, Fe) catalysts. Appl. Catal. B Environ. 2013;129:509–516. doi: 10.1016/j.apcatb.2012.09.056. [DOI] [Google Scholar]
- Zhang Q., Peng D., Su Z., Li A., Sun C., Wang L., Cui S., Yang S., Zheng X., Mo L., Zhang N., Gu F., Liu Y.. Selective oxidation of p-cymene over mesoporous LaCoO3 by introducing oxygen vacancies. Inorg. Chem. 2024;63:21499–21506. doi: 10.1021/acs.inorgchem.4c03521. [DOI] [PubMed] [Google Scholar]
- Bai D., He H., Wang Z., Dong H., Sun L., Yuan X.. Peracetic acid activation by oxygen vacancy-enriched LaCoO3 perovskite for expediting naproxen degradation: Structural properties and activation mechanism. Sep. Purif. Technol. 2025;365:132623. doi: 10.1016/j.seppur.2025.132623. [DOI] [Google Scholar]
- Wu T., Li X., Weng C., Ding F., Tan F., Duan R.. Highly efficient LaMO3 (M = Co, Fe) perovskites catalyzed Fenton’s reaction for degradation of direct blue 86. Environ. Res. 2023;227:115756. doi: 10.1016/j.envres.2023.115756. [DOI] [PubMed] [Google Scholar]
- Buxton G., Greenstock C., Helman W., Ross A.. Critical review of rate constants for reactions of hydrated electrons, hydrogen atoms and hydroxyl radicals (·OH/·O–) in aqueous solution. J. Phys. Chem. Ref. Data. 1988;17:513–886. doi: 10.1063/1.555805. [DOI] [Google Scholar]
- Appiani E., Ossola R., Latch D., Erickson P., McNeill K.. Aqueous singlet oxygen reaction kinetics of furfuryl alcohol: Effect of temperature, pH, and salt content. Environ. Sci.: Processes Impacts. 2017;19:507–516. doi: 10.1039/C6EM00646A. [DOI] [PubMed] [Google Scholar]
- Qi Z., Wu X., Li Q., Lu C., Carabineiro S., Zhao Z., Liu Y., Lv K.. Singlet oxygen in environmental catalysis: Mechanisms, applications and future directions. Coord. Chem. Rev. 2025;529:216439. doi: 10.1016/j.ccr.2025.216439. [DOI] [Google Scholar]
- Bielski B., Cabelli D., Arudi R., Ross A.. Reactivity of HO2·/O2·– with organic compounds in aqueous solution. J. Phys. Chem. Ref. Data. 1985;14:1041–1100. doi: 10.1063/1.555739. [DOI] [Google Scholar]
- Feng C., Gao Q., Xiong G., Chen Y., Pan Y., Fei Z., Li Y., Lu Y., Liu C., Liu Y.. Defect engineering technique for the fabrication of LaCoO3 perovskite catalyst via urea treatment for total oxidation of propane. Appl. Catal., B. 2022;304:121005. doi: 10.1016/j.apcatb.2021.121005. [DOI] [Google Scholar]
- Gao H., Song Z., Fan Y., Zhang J., Guo R., Zhang M., Chen X., Liu W., Huang Z., Zhang X.. Activation of Co–O bond in {110} facet exposed Co3O4 by La doping for the boost of toluene catalytic oxidation. Sep. Purif. Technol. 2022;431:128528. doi: 10.1016/j.jhazmat.2022.128528. [DOI] [Google Scholar]
- Malecka B., Lacz A., Drozdz E.. et al. Thermal decomposition of d-metal nitrates supported on alumina. J. Therm. Anal. Calorim. 2015;120:1053–1061. doi: 10.1007/s10973-014-4262-9. [DOI] [Google Scholar]
- Hao R., Yang S., Zhao Y., Zhang Y., Yuan B., Mao X.. Follow-up research of ultraviolet catalyzing vaporized H2O2 for simultaneous removal of SO2 and NO: Absorption of NO2 and NO by Na-based WFGD byproduct (Na2SO3) Fuel Process. Technol. 2017;160:64–69. doi: 10.1016/j.fuproc.2017.02.021. [DOI] [Google Scholar]
- Kasper J. M., Iii C. A. C., Cooper C. D.. Control of nitrogen oxide emissions by hydrogen peroxide-enhanced gas-phase oxidation of nitric oxide. J. Air Waste Manage. 1996;46:127–133. doi: 10.1080/10473289.1996.10467444. [DOI] [PubMed] [Google Scholar]
- Hao R., Zhao Y.. Macrokinetics of NO oxidation by vaporized H2O2 association with ultraviolet light. Energy Fuels. 2016;30:2365–2372. doi: 10.1021/acs.energyfuels.5b02748. [DOI] [Google Scholar]
- Seguel J., Leal E., Zarate X., Saavedra-Torres M., Schott E., Diaz de Leon J., Blanco E., Escalona N., Pecchi G., Sepulveda C.. Conversion of levulinic acid over Ag substituted LaCoO3 perovskite. Fuel. 2021;301:121071. doi: 10.1016/j.fuel.2021.121071. [DOI] [Google Scholar]
- Xie W., Xu G., Zhang Y., Yu Y., He H.. Mesoporous LaCoO3 perovskite oxide with high catalytic performance for NOx storage and reduction. J. Hazard. Mater. 2022;431:128528. doi: 10.1016/j.jhazmat.2022.128528. [DOI] [PubMed] [Google Scholar]
- Liu Y., Kong X., Guo X., Li Q., Ke J., Wang R., Li Q., Geng Z., Zeng J.. Enhanced N2 electroreduction over LaCoO3 by introducing oxygen vacancies. ACS Catal. 2020;10:1077–1085. doi: 10.1021/acscatal.9b03864. [DOI] [Google Scholar]
- Li C., Zhang N., Gao P.. Lessons learned: how to report XPS data incorrectly about lead-halide perovskites. Mater. Chem. Front. 2023;7:3797–3802. doi: 10.1039/D3QM00574G. [DOI] [Google Scholar]
- Hu J., Zeng X., Wang G., Qian B., Liu Y., Hu X., He B., Zhang L., Zhang X.. Modulating mesoporous Co3O4 hollow nanospheres with oxygen vacancies for highly efficient peroxymonosulfate activation. Chem. Eng. J. 2020;400:125869. doi: 10.1016/j.cej.2020.125869. [DOI] [Google Scholar]
- Wang Y., Wang R., Yu L., Wang Y., Zhang C., Zhang X.. Efficient reactivity of LaCu0.5Co0.5O3 perovskite intercalated montmorillonite and g-C3N4 nanocomposites in microwave-induced H2O2 catalytic degradation of bisphenol A. Chem. Eng. J. 2020;401:126057. doi: 10.1016/j.cej.2020.126057. [DOI] [Google Scholar]
- Zhan S., Zhang H., Mi X., Zhao Y., Hu C., Lyu L.. Efficient Fenton-like process for pollutant removal in electron-rich/poor reaction sites induced by surface oxygen vacancy over cobalt–zinc oxides. Environ. Sci. Technol. 2020;54:8333–8343. doi: 10.1021/acs.est.9b07245. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.







