Next Article in Journal
Controlled Synthesis of Nickel Phosphides in Hollow N, P Co-Doped Carbon: In Situ Transition to (Oxy)hydroxide Phases During Oxygen Evolution Reaction
Next Article in Special Issue
High-Pressure CO2 Photoreduction, Flame Spray Pyrolysis and Type-II Heterojunctions: A Promising Synergy
Previous Article in Journal
Synthesis of MnO2 with Different Crystal Phases via Adjusting pH for Ozone Decomposition Under Various Humidity Conditions and Monolithic Catalyst Development
Previous Article in Special Issue
Innovative Production of 3D-Printed Ceramic Monolithic Catalysts for Oxidation of VOCs by Using Fused Filament Fabrication
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Catalytic Reduction of SO2 with CO over LaCoO3 Perovskites Catalysts: Effect of Fe Doping and Pre-Sulfurization

by
Liang Yao
1,2,3,
Hao Wang
3,
Shuangde Li
1,* and
Yunfa Chen
1,2,*
1
State Key Laboratory of Mesoscience and Engineering, Institute of Process Engineering, Chinese Academy of Sciences, Beijing 100190, China
2
University of Chinese Academy of Sciences, No. 19A Yuquan Road, Beijing 100049, China
3
China ENFI Engineering Corporation, Beijing 100038, China
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(3), 291; https://doi.org/10.3390/catal15030291
Submission received: 11 February 2025 / Revised: 9 March 2025 / Accepted: 18 March 2025 / Published: 19 March 2025
(This article belongs to the Special Issue Advances in Catalysis for a Sustainable Future)

Abstract

:
SO2 emissions are a major source of air pollution, and the catalytic reduction of SO2 to elemental sulfur by CO represents a promising solution. This study investigates the effects of Fe doping and pre-sulfurization on the catalytic performance of LaCoO₃ perovskite catalysts. A series of Fe-doped LaCoO3 perovskites were synthesized via the sol–gel method and evaluated for the catalytic reduction of SO2 by CO. The results showed that LaCo0.8Fe0.2O3 exhibited the highest catalytic performance, achieving 84.0% SO2 conversion at 500 °C. The oxygen-free sulfurization (OFS) treatment compared with oxygen-assisted sulfurization (OAS) treatment significantly enhanced the activity, reaching a SO2 conversion of 95.9% from 80.0% at 450 °C with the lower byproduct generation. Characterization analyses demonstrated that the OFS treatment facilitated the formation of active sulfur species and oxygen vacancies on the catalyst surface while also enhancing the adsorption capacity of the catalyst for the reactant gases. These factors were identified as key contributors to the improved catalytic performance, driven by the combination of redox and carbonyl sulfide (COS) intermediate mechanism. The findings suggest that the OFS treatment is an effective strategy to improve the catalytic reduction of SO2 by CO, offering a more environmentally friendly solution for SO2 emission control through resource utilization.

1. Introduction

Due to the characteristics of China’s energy structure, sulfur dioxide (SO2) emissions from coal-fired flue gas represent the primary source of SO2 pollution [1,2]. Despite the widespread use of conventional desulfurization technologies, like absorption, adsorption, and catalysis, the reduction of SO2 emissions continues to face significant challenges, including secondary pollution and operational complexity [3,4]. Catalytic reduction of SO2 to sulfur in the reducing atmosphere (e.g., CO, CH4) of flue gas not only facilitates effective SO2 removal but also converts it into sulfur, a valuable chemical raw material for various industrial productions, like fertilizers, chemicals, and pharmaceuticals [5,6]. The technology, which promotes the resource utilization of SO2, supports both flue gas desulfurization and the recycling of valuable resources, making it of critical importance for sustainable development [7,8]. Furthermore, the investigation of the single-stage catalytic converter is a key step in advancing the technology. Catalytic materials derived from metal and rare-earth are extensively investigated in SO2 reduction. Ng reviewed dry-based and wet-based catalytic SO2 reduction technologies [8]. Lum summarized SO2 reduction for environmental sustainability and circular economy [9]. Sepehrian recently reviewed SO2 dry-based catalytic removal from flue gas leading to elemental sulfur production, encompassing various reductants employed, reduction mechanisms, catalyst types, methods for catalyst improvement, and diverse operational conditions [10]. Up to date, various reducing agents, such as carbon monoxide, hydrogen, methane, and carbon materials, have been investigated for SO2 reduction reactions [10,11,12,13]. CO, due to its low cost and widespread application in various industries, is a common gas used as a reducing agent in the catalytic conversion of SO2. Converting SO2 to sulfur with high efficiency, achieving a pure product, producing fewer byproducts, and requiring lower temperatures compared to other reductants are some of the main features of using CO reduction [10]. Catalytic SO2 reduction to sulfur via transition metal-based catalysts (like Cu, Ni, Ce loaded SiO2, γ-Al2O3, TiO2, and ZrO2) [6,14,15] and sulfided metal-based catalysts (like FeS, ZnS, Co9S8-N@NC) [7,16,17] appeared to be highly promising for high SO2 conversion while maintaining high sulfur selectivity. Ren reported a Fe-loaded catalyst derived from blast furnace slag (BFS) for the CO-SO2 reaction to produce sulfur due to the strong adsorption of BFS and the production pathway of H2S [18]. Cao investigated that the highest Ce3+ content, oxygen vacancies content, and Ir0 content over several Ir/CeO2 catalysts exhibited the strongest Ir-O-Ce interaction and caused better catalytic activity on the reduction of SO2 by CO [11]. Tian found the effect of temperature on the reaction path of pyrite (FeS2)-based catalyst catalyzed CO reduction of SO2 to sulfur due to the S-vacancy formation [7]. The activity selectivity of sulfur over the above-reported catalysts was mainly tested with the absence of O2, which may have shown better catalytic performance. However, conventional catalytic desulfurization for flue gas normally would turn to inactivation with oxidation by the existence of O2.
Owing to the diverse chemical properties and high electron mobility, metal oxide perovskite materials with an ABO3 structure have been widely studied in various catalytic reactions, like environmental pollution remediation [19,20,21,22,23,24], including the catalytic reduction of SO2 by CO [25,26]. Partial substitution of A- or B-site cations can effectively promote lattice distortion and enhance the mobility of oxygen species, thereby improving the catalytic activity toward reaction gas, like H2S oxidation and SO2 reduction [26,27]. Moreover, the unique structure of perovskites facilitates the formation of sulfide species during the SO2 reduction process, which generates a synergistic effect [28]. For instance, Hibbert observed that during the reduction of SO2 with La1−xSrxCoO3, the catalyst was fully converted into various metal sulfides when high catalytic activity was stabilized, indicating that sulfides serve as the active phase [28]. Additionally, catalyst sulfurization has also been shown to improve catalyst performance. For example, Ma found that the sulfides formed during the sulfurization process of LaCoO3 play a crucial role in promoting CO reduction of SO2 [29]. Wang investigated Cu-doping LaCoO3 in a pre-sulfurized process, given better SO2 reduction performance in the presence of oxygen due to the decreasing formation of carbonyl sulfide (COS) intermediate product [26]. A metal sulfide phase (MeS) is formed on the metal oxide catalysts, which were subjected to the sulfur-containing gases flow under certain temperature conditions. The performance of various pre-sulfided metal oxide catalysts such as Fe, Mo, Ni, and Co demonstrated that the formation of FeS2, MoS2, NiS2, and CoS2 metal sulfide active phases in the catalyst matrix would accelerate the catalytic process [10]. Furthermore, the sulfidation of the Fe catalyst can form Fe7S8, FeS2, and FeS mixed phases, which have higher access to the edge and corner unsaturated active sites on the surface with SO2 and reduction gas, together with the formation of the most important actives sites of S-vacancy [7,10].
Based on the extensive reported work, it is evident that perovskite-type catalysts show promising potential for application in SO2 catalytic reduction technology. However, the function of the doped LaCoO3 catalysts for catalytic activity and the pre-sulfurized process, together with the adjustment for catalytic performance and the mechanism in the presence of oxygen, are still undetermined. In the present work, a series of Fe-doped LaCoO3 perovskites were synthesized by the sol–gel method for the catalytic reduction of SO2 by CO under O2. Various parameters of the catalysts were then investigated, including the physical properties, surface chemical state, morphology, catalytic activity, and gas adsorption capacity to elucidate the effects of the doping and sulfurization process, together with the structure–activity relationship of the catalysts.

2. Results and Discussion

2.1. Physical Properties of XRD for Catalysts

Figure 1A presents the XRD patterns of the synthesized Fe-doped LaCoO3 and oxygen-assisted and oxygen-free sulfurization of LaCo0.8Fe0.2O3. The diffraction peaks corresponded to the orthorhombic structure of LaCoO3 (PDF#48-0123) and the cubic structure of LaFeO3 (PDF#75-0541) [27]. As the Fe content in LaCoO3 increased, the XRD peaks shifted slightly to lower angles, which could be attributed to the substitution of the larger Co3+ ions by the smaller Fe3+ ions, leading to an increase in the lattice constant [30]. Ko also reported that La0.7Ce0.1CoxNiyTi0.6O3 perovskites with tuned Co and Ni ratios reflect a structural relaxation of the lattice, which facilitates increased spacing between lattice planes as decreasing the larger ionic radii of Co content [31]. Notably, all diffraction peaks of LaCo1−xFexO3 were consistent with the perovskite phase, confirming the successful synthesis of stoichiometric LaCo1−xFexO3 [27]. Oxygen-assisted and oxygen-free sulfurization (OAS and OFS) experiments were conducted on LaCo0.8Fe0.2O3 to evaluate the impact of the sulfurization on the phase structure of catalysts. The results show a significant reduction in the intensity of the XRD diffraction peaks in Figure 1A. Although OAS-LaCo0.8Fe0.2O3 maintained relatively better phase stability compared to OFS-LaCo0.8Fe0.2O3, both processes indicate a decrease in the crystallinity of the catalysts. Furthermore, no new peaks corresponding to additional phases were detected in the XRD patterns of both sulfurized samples, suggesting that the sulfurization weakened the phase integrity without forming new phases. The observed loss of crystallinity in the XRD results for OAS and OFS-LaCo0.8Fe0.2O3 samples may be due to the effects of SO2 poisoning on the structure of the LaCo0.8Fe0.2O3 catalyst. The holding diffraction peaks of the perovskite structure were assumed to be due to the lesser SO2 sulfurization time. Zhu also reported that the perovskite structure of LaCoO3 was destroyed by SO2 and became amorphous due to a high temperature of 500 °C and long SO2 exposure [32].
The XRD patterns of LaCo1−xFexO3, OAS-LaCo0.8Fe0.2O3, and OFS-LaCo0.8Fe0.2O3 catalysts after catalytic testing under oxygen-assisted reaction (OAR) conditions are shown in Figure 1B. The decrease in intensity indicates a significant reduction in the crystallinity of the catalysts, while the main perovskite phase remained intact, accompanied by the emergence of weak peaks corresponding to new phases. Specifically, the peaks at 2θ = 25.6° and 28.5° were attributed to La2O2S (PDF 026-0825), which has proved the active sulfides formed in the reduction of SO2 [33]. The peaks at 2θ = 36.6° and 42.8° corresponded to CoO (PDF 01-1227), and the peaks at 2θ = 28.5° and 30° were assigned to La2O3 (PDF 74-2430). In addition, OFS-LaCo0.8Fe0.2O3 exhibited distinct characteristic peaks around 36.4° and 38.0°, which were not observed in other samples. These peaks can be attributed to FeS2 (PDF 03-0795), a high-activity species that has also been proven to promote SO2 treatment [7], indicating that OFS treatment effectively achieves structural change of the catalyst. Hibbert summarized the treatment of SO2 on perovskite oxides for the structural changes [34]. For example, the La0.5Sr0.5Co3 catalyst showed a mixture of perovskite, CoS2, SrS, and several unknown phases after SO2 adsorption flow reactor experiments [34]. Another example of higher concentrations of SO2 (10%) and CO (20%) reaction, the perovskite LaCoO3 decomposed, and the active catalyst was a mixture of La2O2S and CoS2 [34]. So, the presence of metal sulfide phases in the spent catalysts for OAR-OFS-LaCo0.8Fe0.2O3 may be due to the reaction of SO2 and the perovskite phase.

2.2. Physical Properties of BET for Catalysts

The Brunauer–Emmett–Teller (BET) surface area Barrett–Joyner–Halenda (BJH) pore size distribution of LaCo1−xFexO3 samples were investigated through N2 adsorption–desorption isotherms, as the results are shown in Figure 2. All samples exhibited a typical type IV adsorption isotherm with an H1 hysteresis loop in the relative pressure range of P/P0 = 0.6–1.0, indicating the presence of irregular mesopores resulting from the aggregation of nanocrystals. The BET surface area and pore structure are similar to the reported perovskite-based systems [35]. Table 1 summarizes the specific surface area, pore volume, and pore size of the samples. The increased Fe content may lead to a change in the crystal structure and particle growth during the synthesis process, resulting in a different particle size distribution varied from 23.23 to 19.05 nm. The specific surface area of all samples ranged from 12 m2/g to 16 m2/g, with pore sizes around 20 nm, which indicates that the textural properties of the catalysts have a negligible impact on the differences in their catalytic performance.

2.3. Morphological Characteristics

SEM images of the LaCo0.8Fe0.2O3 catalysts, the pre-sulfurization samples, and their catalytic reaction samples are presented in Figure 3. As shown in Figure 3A, the morphology of pristine LaCo0.8Fe0.2O3 consisted of large particles composed of tightly bound small particles. Similarly, the samples exhibit well-defined, small particles with an approximate diameter of 100 nm after sulfurization with OFS-LaCo0.8Fe0.2O3 and OAS-LaCo0.8Fe0.2O3, as depicted in Figure 3B,C, which could be expected to enhance the contact between the catalyst and the reactants. Heidinger reported the morphology evolution of Fe-LaCoO3 perovskite synthesized by reactive grinding, which also exhibited agglomerating particles with low-energy ball milling [23]. After undergoing SO2 catalytic testing under oxidative conditions, the sulfurized samples of OFS-LaCo0.8Fe0.2O3 and OAS-LaCo0.8Fe0.2O3 after SO2 treatment exhibited varying degrees of sintering, with the originally small particles turning to greater agglomeration (Figure 3D,E). EDS mapping was further employed to investigate the elemental distribution on the catalyst surface, with the results for OFS-LaCo0.8Fe0.2O3 shown in Figure 3F. The presence of Fe on the surface confirms the successful doping of Fe into the catalyst, while the S distribution indicates that sulfurization led to a significant accumulation of sulfur species on the surface. The atomic percentage data obtained from the EDS mapping for all samples are summarized in Table 2. The S atomic percentage on the surface of OFS-LaCo0.8Fe0.2O3 was higher than that of OAS-LaCo0.8Fe0.2O3. However, the OFS-LaCo0.8Fe0.2O3 catalyst exhibited a lower sulfur enrichment than OAS-LaCo0.8Fe0.2O3 after the oxygen-assisted reaction test, which suggests that the OFS treatment contributed to the formation of active sulfur species on the surface of LaCo0.8Fe0.2O3, thereby maybe enhancing its catalytic activity for SO2 conversion. The atomic percentage data on the surface of the catalysts were also detected by XPS and shown in Table 3. The phenomena above the general trend are similar to EDS but have slight differences. OFS-LaCo0.8Fe0.2O3 still owned a high S atomic percentage on the surface in comparison with OAS-LaCo0.8Fe0.2O3. OAR-OFS-LaCo0.8Fe0.2O3 catalyst showed a slightly higher sulfur enrichment than OAS-LaCo0.8Fe0.2O3, which was opposite to EDS. This may be due to the difference in depth of the sample being analyzed.

2.4. Catalytic Performance

The catalytic activities in the SO2 reduction over LaCo1−xFexO3 perovskites were first evaluated as a function of temperature under the oxygen-assisted reaction (OAR), and the results are presented in Figure 4A. As the reaction temperature increased, the SO2 conversion rate generally showed an upward trend for all catalysts except for LaFeO3, which exhibited deactivation, suggesting SO2 poisoning during the reaction. Among all the catalysts, LaCo0.8Fe0.2O3 demonstrated the highest activity, achieving an SO2 conversion rate of 84.0% at 500 °C, and showed lower COS production than that of Fe-doped catalysts, which is presented in Figure 5A. The temperatures up to 550 °C used in our experiments were indeed above the boiling point of sulfur (444.6 °C), which is essential to ensure the effective removal of sulfur from the catalyst surface.
Consequently, the catalyst was subjected to both OFS and OAS treatments, followed by tests under oxygen-free reaction (OFR) and oxygen-assisted reaction (OAR), as shown in Figure 4B. The results indicate that the OFS catalyst exhibited higher activity and lower COS production than those of OAS catalyst (Figure 5B), and the presence of oxygen during the reaction further enhanced the conversion rate, which reached 95.9% at 450 °C, indicating that the OFS treatment enhanced the performance of LaCo0.8Fe0.2O3 in SO2 catalytic reduction. The SO2 conversion for OFS-LaCo0.8Fe0.2O3 was even close to many reported catalysts mainly based on oxygen-free reactions. Xu reported that 15% Gd-Ce-Ox exhibited superior catalytic activity, achieving an impressive 97% conversion of SO2 toward sulfur by CO at 400 °C, but with the absence of O2 [1]. Li investigated that iron oxide supported on activated coke for SO2 to S by CO reaches 90.9% SO2 conversion at 400 °C [15]. Jin observed SO2 reduction for S production by CO over Ce-Al-Ox composite oxide catalyst reaching 98% beyond 425 °C [5]. At an inlet SO₂ concentration of 5%, the byproduct COS concentration was below 700 ppm, indicating a very low level of byproducts. For the OAS-LaCo0.8Fe0.2O3 catalyst at 500 °C, a relatively higher COS concentration was observed compared to other catalysts. This difference might be perceived due to the overall low concentration of COS. Subsequently, OFS-LaCo0.8Fe0.2O3 underwent a thermal stability test at 550 °C for 28 h, as well as a water resistance test with the introduction of theoretical 20% and 40% water during the reaction, both of which demonstrate the excellent stability of the catalyst, as shown in Figure 4C,D. The weight of the collected production solid sulfur was further measured. The results showed that the amount of sulfur collected corresponded well with the SO2 conversion rates, indicating a high degree of material balance. To gain deeper insights into the activation effect of the OFR process on LaCo0.8Fe0.2O3, further characterization and analyses were conducted.

2.5. TPD Analysis

TPD measurements were conducted on LaCo0.8Fe0.2O3, sulfurized LaCo0.8Fe0.2O3, LaCoO3, and LaFeO3 catalysts, with the results shown in Figure 6. It is generally accepted that the basic sites on the catalyst surface are crucial for SO2 adsorption [14]. The CO2-TPD method was used to investigate the change in surface basicity of the varied prepared perovskite catalysts. The CO2-TPD profiles (Figure 6A) revealed a prominent peak around 550 °C for OFS-LaCo0.8Fe0.2O3, suggesting the presence of strong basic sites that were absent in the other catalysts, thereby providing substantial support for SO2 adsorption. The CO2-TPD profiles were similar to the reported LaFe-based and LaCo1−xGaxO3 perovskites [36,37].
The finding was further corroborated by the SO2-TPD results, as shown in Figure 6C, where a clear SO2 desorption peak was observed around 550 °C on OFS-LaCo0.8Fe0.2O3, highlighting its excellent SO2 adsorption capability. In comparison to the LaCoO3 catalyst, Wu reported that the desorption peaks in the high-temperature region of the Ce-LaCoO3 catalyst were mainly attributed to the preferential adsorption of SO2 by Ce species as sacrificial sites [38]. The desorption peaks in the high-temperature section were mainly attributed to the metal sulfates on the surface of the catalysts, which were related to their strong ability to adsorb SO2 [38].
The CO-TPD result (Figure 6B) indicates that the sulfurization process significantly improved the CO adsorption capacity of the catalyst, which is essential for the reduction of SO2. Moreover, the higher desorption peak observed for OFS-LaCo0.8Fe0.2O3 suggests a stronger CO adsorption capacity, which contributes to its superior catalytic activity. Teng investigated that the total adsorption amount of CO on LaCoO3 nanowires is larger than that on the LaCoO3 particles, indicating that there may be more active centers on LaCoO3 nanowires than those on the latter, which would exhibit higher catalytic activity for CO oxidation, compared with the nanoparticles [39].

2.6. XPS Analysis

In order to investigate the surface states of the catalysts, XPS was employed to analyze Co 2p, O 1s, Fe 2p, and S 2p, with the corresponding spectra presented in Figure 7. In the Co 2p spectra (Figure 7A), peak deconvolution reveals two distinct peaks corresponding to Co3+ and Co2+. The peaks at 779.0–780.0 eV are attributed to octahedrally coordinated Co3+, while the peaks at 780.6–781.6 eV correspond to tetrahedrally coordinated Co2+ [40]. Additionally, the S1 and S2 peaks are satellite features of Co2+. Notably, the Co 2p peak of the OFS-LaCo0.8Fe0.2O3 shifts 0.9 eV toward lower binding energy compared to other catalysts, suggesting a strong interaction between Fe and Co in the catalyst. Electrons escaping from Fe ions are captured by Co3+, which is subsequently reduced to Co2+ [27]. The data obtained from peak deconvolution are summarized in Table 4. After sulfurization, the more pronounced Co2+/Co3+ observed in the OFS-LaCo0.8Fe0.2O3 suggests that the OFS treatment facilitates electron transfer between Co ions and Fe ions.
Figure 7B presents the Fe 2p spectra, where two broad peaks are observed in the binding energy ranges of 720–730 eV and 706–718 eV, corresponding to Fe 2p1/2 and Fe 2p3/2, respectively. After peak deconvolution, the peaks at 710.5–710.8 eV are assigned to Fe2+, while the peaks at 713.7–714.0 eV correspond to Fe3+, and the peaks at 718.0–718.8 eV are satellite peaks [41,42,43]. The shift of Fe species to higher binding energy in the OFS-LaCo0.8Fe0.2O3 further confirms the interaction between Co and Fe in the catalyst. The data in the table show that the Fe2+/Fe3+ ratio for the different samples follows the following trend: OFS-LaCo0.8Fe0.2O3 > OAS-LaCo0.8Fe0.2O3 > LaCo0.8Fe0.2O3, indicating that the OFS treatment enhanced the reduction of Fe3+. The resulting charge imbalance is expected to promote the formation of oxygen vacancies, as evidenced by the O 1s spectra.
O 1s spectra are shown in Figure 7C. The peaks at 528.9–529 eV are assigned to lattice oxygen (Olat) [44], the peaks at 531.3–531.7 eV to surface adsorbed oxygen (Oads) [45], and the peaks at 533.1–533.3 eV to hydroxyl groups (OOH) on the catalysts surface [46]. Kustov reported similar O 1s spectra over LaMO3 perovskite-like oxide materials (M: Fe, Co, Ni) [47]. Analysis of the peak areas after deconvolution reveals that the Oads/Ototal ratio is highest for the OFS-LaCo0.8Fe0.2O3, which reflects the result of charge compensation. This suggests that the OFS treatment facilitates the formation of oxygen vacancies on the catalyst surface, thereby enhancing the activation of both reactants and gaseous oxygen, ultimately improving catalytic activity.
Figure 7D presents the S 2p spectra for OFS-LaCo0.8Fe0.2O3, OAS-LaCo0.8Fe0.2O3, OAR-LaCo0.8Fe0.2O3, and the oxygen-assisted reaction (OAR) for the spent catalysts of the sulfurization of LaCo0.8Fe0.2O3. All the samples showed the binding energy of S 2p was about 168–170 eV, which can be mainly attributed to sulfate and sulfite, implying that sulfate and sulfite were formed on the surface due to both the pre-sulfurization and oxygen-assisted reaction [32]. Furthermore, according to the binding energy of the peak, it can be judged that the strong S 2p peaks of 162.4 eV existed only for OFS-LaCo0.8Fe0.2O3 and OAR-OFS-LaCo0.8Fe0.2O3 correspond to S2−, which may belong to FeS2 on the catalyst surface [7]. Tian investigated the effect of temperature on the reaction path of pyrite (FeS2)-based catalyst catalyzed CO reduction of SO2 to sulfur, in which they found that sulfate, sulfite, and S2− together with FeS2 by applying S 2p and Fe 2p to characterize the catalyst and reaction process [7]. The FeS2 detected by the XPS result perhaps was consistent with the XRD results for OAR-OFS-LaCo0.8Fe0.2O3 with the existence of the FeS2 phase.

2.7. Mechanism Explanation

XRD demonstrated that the crystallinity of OFS-LaCo0.8Fe0.2O3 was significantly reduced, as evidenced by the decreased intensity of the diffraction peaks. Moreover, after the oxygen-assisted reaction (OAR), the formation of the La2O2S phase was observed. The atomic percentage data obtained from SEM-EDX and XPS analyses indicated sulfur enrichment on the surface of OFS-LaCo0.8Fe0.2O3. The CO/CO₂/SO₂-TPD results further revealed that OFS-LaCo0.8Fe0.2O3 exhibited the strongest surface basicity, CO, and SO2 adsorption capacity among the catalysts. XPS also showed that OFS-LaCo0.8Fe0.2O3 had the highest Oads/Ototal ratio and surface active S2− species. These structural, physical, and chemical properties collectively suggested that OFS-LaCo0.8Fe0.2O3 possesses the optimal conditions for high SO2 conversion. Based on these characterization results, we further speculated the reaction mechanism responsible for the enhanced catalytic performance.
The SO2 catalytic reduction mechanism over transitional metal oxides by reduction gas, like CO, is normally related to the COS and redox intermediate mechanisms, which are ascribed to the active sites and oxygen vacancies. There is a need for a clear discussion of the reaction mechanism, particularly the identification of active sites and oxygen vacancies (Ov) during the catalytic process for the LaCoxFe1−xO3 perovskite system under Fe doping and sulfurization conditions. In the redox mechanism, the reaction pathway can be expressed as Equations (1)–(4), SO2 adsorption on the perovskite surface, which leads to the S formation through the SO2 direct reduction by CO [10]. So, the high Ov content may induce high SO2 conversion. The catalytic activities in the SO2 reduction show that LaCo0.8Fe0.2O3 owns higher SO2 conversion than those of LaCoxFe1−xO3 (Figure 4A). However, LaCo0.8Fe0.2O3 had less Oads/Ototal than that of LaCoO3 (0.47 vs. 0.59), so, besides the redox mechanism, there may be a related synergistic effect of Fe and LaCoO3. Fe pyrite (FeS2)-based catalysts were normally reported to promote the CO reduction of SO2 to S through the COS mechanism [7]. The phenomenon was identical to the observation from Figure 5A, which can find there is COS formation for LaCo0.8Fe0.2O3, but the absence of COS formation for LaCoO3 within 450–550 °C. Over LaCoO3, there may only be related to the redox mechanism, which almost zero COS production as some of the most important features at the relatively high reaction temperature with high capacity and mobility of Ov [10].
Oxygen-free sulfurization of LaCo0.8Fe0.2O3 perovskite owns the highest SO2 conversion than that of oxygen-assisted sulfurization and LaCoxFe1−xO3 (Figure 4A,B). One of the reasons for the phenomenon may be due to the OFS-LaCo0.8Fe0.2O3 owning the highest Oads/Ototal with 0.74 (Table 4). The other reason may be due to the promotion of the COS mechanism. In the COS mechanism, the reaction pathway can be expressed as Equations (5)–(7). COS can form through the reaction between CO and metal sulfide, then turn to S with SO2 [10]. OFS-LaCo0.8Fe0.2O3 exhibited the FeS2 phase detected from XRD (Figure 1B) and S 2p XPS (Figure 7D), which were unanimous with the optimal SO2 reduction performance. Moreover, the detection of La2O2S for OFS-LaCo0.8Fe0.2O3 and OAS-LaCo0.8Fe0.2O3 under oxygen-assisted reaction (OAR) catalytic testing indicated that the formation of La2O2S during the SO2 reduction, which was also reported to be the reactive sites based on the COS mechanism [29,33]. Based on the above analysis, LaCo0.8Fe0.2O3 by Fe doping exhibited increased SO2 conversion, maybe due to the main redox intermediate mechanism, together with the COS mechanism by the synergistic effect of Fe and LaCoO3. OFS-LaCo0.8Fe0.2O3 exhibited much higher SO2 conversion compared with OAS-LaCo0.8Fe0.2O3, perhaps due to both redox and COS intermediate mechanisms.
Cat-O + CO → Cat-□ + CO2
Cat-□ + SO2 → Cat-O + SOcat
Cat-□ + SOcat → Cat-O + Scat
Scat → Cat + 1/x Sx
M-Scat + CO → M + COS
SO2 + 2COS → 3S + 2CO2
S + M → M − Scat

3. Materials and Methods

3.1. Materials

La(NO3)3·6H2O (CAS: 100587-94-8), Co(NO3)2·6H2O (CAS: 10026-22-9), Fe(NO3)3·9H2O (CAS: 10421-48-4), citric acid, ammonium hydroxide, polyethyleneglycol, and quartz sands were purchased from Shanghai Macklin Biochemical Technology Co., Ltd. (Shanghai, China). All these chemicals with analytical purity were used as obtained without further purification.

3.2. Synthesis of LaCoxFe1−xO3 Catalysts

LaCoO3 was synthesized as follows: La(NO3)3·6H2O and Co(NO3)2·6H2O were dissolved in deionized water under magnetic stirring at a stoichiometric molar ratio of 1:1, resulting in a homogeneous solution. Subsequently, citric acid was added to the above solution in an amount equimolar to the total metal nitrates. Ammonia solution was then gradually added to adjust the pH of the mixture to a range of 6–7. After adding a little polyethyleneglycol, the resulting solution was stirred continuously at 75 °C until a stable gel was formed. The porous dry gel was obtained after the solution was dried overnight at 130 °C. The dry gel was subjected to self-combustion ignition and further calcined in a tubular furnace. Calcination was carried out at a heating rate of 2 °C·min−1, with the temperature maintained at 600 °C for 2 h.
LaCoxFe1−xO3 perovskites were synthesized as follows: Fe(NO3)3·9H2O was introduced in varying concentrations based on the molar ratio, while all other experimental conditions were kept identical to those used for preparing LaCoO3. The final samples were named as LaCoxFe1−xO3 (x = 0, 0.2, 0.5, 0.8).

3.3. Sulfurization of LaCo0.8Fe0.2O3 Catalyst

Catalysts were sulfurized as follows: 2 g of the catalysts (30–40 mesh) were placed in a tubular furnace and heated to 350 °C under a nitrogen flow of 100 mL/min. Subsequently, the nitrogen was replaced with a mixture of either 5% SO2/16% CO/3% O2/N2 or 5% SO2/10% CO/N2, and the temperature was maintained for 30 min. The catalysts were named OAS catalyst (oxygen-assisted sulfurization) and OFS catalyst (oxygen-free sulfurization) based on the presence or absence of oxygen in the sulfurization process.

3.4. Catalysts Characterization

The crystal structure of the catalysts was analyzed using a Rigaku SmartLab SE X-ray diffractometer (XRD, Tokyo, Japan), employing Cu-Kα radiation as the diffraction source (incident wavelength λ = 1.54 Å). The scanning range was typically set to 5–90°, with a scanning speed of 10°·min−1. The N2 adsorption–desorption isotherms were determined by a Micromeritics ASAP 2460 analyzer (Norcross, GA, USA), with BET surface area, BJH desorption cumulative volume of pores, and BJH desorption average pore diameter for analysis. The JSM-7800 (Prime, JEOL, Tokyo, Japan) scanning electron microscope (SEM) equipped with energy-dispersive X-ray spectroscopy (EDS) was used to investigate the morphology of samples. The temperature-programmed desorption (TPD) of CO2, CO, and SO2 (CO2-TPD, CO-TPD, and SO2-TPD) experiments were performed using a Micromeritics Autochem 2920II chemisorption analyzer. All the catalysts were pretreated in a He stream at 300 °C for 1 h before the measurement; subsequently, the adsorption of target gas was conducted on the catalysts. A thermal conductivity detector (TCD) detector was used to record the signal of gas desorbed through heating from 50 to 800 °C. The chemical states of the surface elements were characterized using X-ray photoelectron spectroscopy (XPS) with a Thermo Scientific K-Alpha instrument (Waltham, MA, USA). Al K-Alpha radiation (1486.6 eV) served as the excitation source, and the binding energy was calibrated to C 1s at 284.8 eV.

3.5. Catalytic Performance Tests

The catalyst evaluation experiment for CO reduction of SO2 was studied with 2 g (30–40 mesh) catalysts packed in the middle part of the quartz tube in a fixed-bed reactor system (Figure 8). The total gas flow rate was confirmed to be 100 mL/min adjusted with a mass flow controller (MFC), with a gas composition of 5% SO2, 16% CO/3% O2/N2 or 5% SO2 and 10% CO/N2 represented oxygen-assisted reaction (OAR) and oxygen-free reaction (OFR). The electric furnace was employed to regulate the temperature of the reactor. The reaction was run from 250 °C to 550 °C at an interval of 50 °C with a 15 min continuous test at each temperature. In addition, a cooling unit was incorporated into the sulfur collection process, and a filter was installed at the outlet of the condensation unit to prevent sulfur powder from being carried out with the reaction gas. The concentrations of SO2, COS, H2S, and CS2 in the outlet gas at each temperature interval were determined online by gas chromatography (Agilent 8860, Wilmington, DE, USA, columnHayeSep N) with TCD. The standard curve for SO₂ was obtained using the peak area from a two-point method with zero point and 5vol% SO2 (N2 balance gas) detected by TCD. The conversion rate of SO2 (XSO2) are calculated as follows:
X S O 2 = S O 2 i n S O 2 o u t S O 2 i n × 100 %
where [SO2]in is the concentration of SO2 in the inlet gas, while [SO2]out is the concentration of SO2 in the exhaust gas after the reaction.

4. Conclusions

This study provides significant insights into the catalytic reduction of SO2 to elemental sulfur by CO over Fe-doped LaCoO3 perovskite catalysts, together with pre-sulfurization treatment. The incorporation of Fe into the LaCoO3 structure enhanced the catalytic activity, with LaCo0.8Fe0.2O3 achieving the highest performance through oxygen-free sulfurization (OFS) under oxygen-assisted reaction. OFS-LaCo0.8Fe0.2O3 perovskite presented the highest SO2 conversion with 95.9% at 450 °C than that of LaCo0.8Fe0.2O3, reaching 84.0% SO2 conversion at 500 °C. XRD structure, CO/CO2/SO2-TPD, and XPS surface analysis revealed that OFS treatment promoted the formation of active sulfur species and oxygen vacancies, as well as the adsorption capacity of reactants, which were crucial for improving catalytic activity. Additionally, the catalysts exhibited excellent thermal stability and water resistance, making them suitable for practical applications in industrial SO2 reduction. These findings highlight the critical role of sulfurization treatments and surface properties in enhancing the catalytic performance of perovskite-based materials, offering insights for the development of more efficient and sustainable flue gas desulfurization technologies.

Author Contributions

Conceptualization, S.L. and Y.C.; methodology, L.Y., H.W. and S.L.; validation, L.Y., H.W. and S.L.; formal analysis, L.Y., H.W. and S.L.; investigation, L.Y., H.W. and S.L.; resources, L.Y. and Y.C.; visualization: S.L. and Y.C.; writing—original draft preparation, L.Y., H.W. and S.L.; writing—review and editing, S.L. and Y.C.; supervision, S.L. and Y.C.; project administration, S.L. and Y.C.; funding acquisition, Y.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China “grant number 22421003” and China ENFI Engineering Technology Co., Ltd. “grant number KFA2024-137”.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

Authors Liang Yao and Hao Wang were employed by the company China ENFI Engineering Corporation. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Xu, M.; Chen, L.; Cheng, X.; Meng, X.; Jin, Q.; Zhu, C.; Yang, J.; Xu, H. Reduction of SO2 into sulfur over Ce-based catalyst: Performance optimizations and reaction mechanisms. J. Environ. Chem. Eng. 2024, 12, 114064. [Google Scholar] [CrossRef]
  2. Wong, P.Y.; Zeng, Y.T.; Su, H.J.; Lung, S.C.C.; Chen, Y.C.; Chen, P.C.; Hsiao, T.C.; Adamkiewicz, G.; Wu, C.D. Effects of feature selection methods in estimating SO2 concentration variations using machine learning and stacking ensemble approach. Environ. Technol. Innov. 2025, 37, 103996. [Google Scholar] [CrossRef]
  3. Yang, L.; Zhong, W.; Sun, L.; Chen, X.; Shao, Y. Dynamic optimization oriented modeling and nonlinear model predictive control of the wet limestone FGD system. Chin. J. Chem. Eng. 2020, 28, 832–845. [Google Scholar] [CrossRef]
  4. Behzadi pour, G.; Kamel Oroumiyeh, M.; Fekri aval, L. Rapid SO2 gas removal using MgO/AC/CaCO3/Zeolite nanocomposite at room temperature. Case Stud. Chem. Environ. Eng. 2025, 11, 101042. [Google Scholar] [CrossRef]
  5. Jin, Q.; Meng, X.; Ji, W.; Wu, P.; Xu, M.; Zhang, Y.; Zhu, C.; Xu, H. SO2 reduction for sulfur production by CO over Ce-Al-Ox composite oxide catalyst. Catal. Commun. 2023, 174, 106587. [Google Scholar] [CrossRef]
  6. Mousavi, S.E.; Pahlavanzadeh, H.; Khalighi, R.; Khani, M.; Ebrahim, H.A.; Abbasizadeh, S.; Mozaffari, A. Reduction of SO2 to elemental sulfur in flue gas using Copper-Alumina catalysts. J. Nanotechnol. 2023, 2023, 3723612. [Google Scholar] [CrossRef]
  7. Tian, Y.; Zhou, X.; Liu, M.; Zhang, J.; Wang, W.; Song, Z.; Zhao, X. Effect of temperature on the reaction path of pyrite (FeS2)-based catalyst catalyzed CO reduction of SO2 to sulfur. Chemosphere 2023, 340, 139789. [Google Scholar] [CrossRef] [PubMed]
  8. Ng, K.H.; Lai, S.Y.; Jamaludin, N.F.M.; Mohamed, A.R. A review on dry-based and wet-based catalytic sulphur dioxide (SO2) reduction technologies. J. Hazard. Mater. 2022, 423, 127061. [Google Scholar] [CrossRef] [PubMed]
  9. Lum, M.M.X.; Ng, K.H.; Lai, S.Y.; Mohamed, A.R.; Alsultan, A.G.; Taufiq-Yap, Y.H.; Koh, M.K.; Mohamed, M.A.; Vo, D.-V.N.; Subramaniam, M.; et al. Sulfur dioxide catalytic reduction for environmental sustainability and circular economy: A review. Process Saf. Environ. Prot. 2023, 176, 580–604. [Google Scholar] [CrossRef]
  10. Sepehrian, M.; Anbia, M.; Hedayatzadeh, M.H.; Yazdi, F. SO2 dry-based catalytic removal from flue gas leading to elemental sulfur production: A comprehensive review. Process Saf. Environ. Prot. 2024, 182, 456–480. [Google Scholar] [CrossRef]
  11. Jinyan, C.; Yutao, L.; Tianrun, S.; Qiulin, Z.; Ping, N.; Jianjun, C.; Guocai, T.; Liangtao, Y.; Siyuan, X.; Rongbing, N. The positive role of Ir0 and OV on the reduction of SO2 by CO over Ir/CeO2. Appl. Surf. Sci. 2023, 620, 156826. [Google Scholar]
  12. Xia, X.; Zhao, X.; Zhou, P.; Feng, T.; Ma, C.; Song, Z. Reduction of SO2 to elemental sulfur with carbon materials through electrical and microwave heating methods. Chem. Eng. Process. 2020, 150, 107877. [Google Scholar]
  13. Feng, T.; Huo, M.; Zhao, X.; Wang, T.; Xia, X.; Ma, C. Reduction of SO2 to elemental sulfur with H2 and mixed H2/CO gas in an activated carbon bed. Chem. Eng. Res. Des. 2017, 121, 191–199. [Google Scholar] [CrossRef]
  14. Lu, Y.; Jin, Q.; Ji, W.; Zhu, C.; Xu, M.; Zhu, Y.; Xu, H. Resource utilization of high concentration SO2 for sulfur production over La–Ce-Ox@ZrO2 composite oxide catalyst. J. Rare Earths 2023, 41, 1945–1952. [Google Scholar]
  15. Li, S.; Feng, T.; Kong, Q.; Li, J.; Liu, P.; Ni, P.; Wang, C. Investigation of iron oxide supported on activated coke for catalytic reduction of sulfur dioxide by carbon monoxide. J. Anal. Appl. Pyrolysis 2024, 179, 106488. [Google Scholar]
  16. He, L.; Hu, J.; Yang, Y.; Zheng, Z.; Deng, Y.; Xin, Y.; Zou, G.; Hou, H.; Ji, X. Hierarchical cocoon-like Co9S8-N@NC for ambient reduction of sulphur dioxide to sulphur via proton-coupled electron transfer. Appl. Catal. B Environ. 2025, 364, 124860. [Google Scholar]
  17. AlQahtani, M.S.; Knecht, S.D.; Wang, X.; Bilén, S.G.; Song, C. One-step low-temperature reduction of sulfur dioxide to elemental sulfur by plasma-enhanced catalysis. ACS Catal. 2020, 10, 5272–5277. [Google Scholar]
  18. Ren, W.; Zhou, P.; Tian, Y.; Wang, W.; Dong, Y.; Wang, T.; Zhang, L.; Ma, C.; Zhao, X. Catalytic performance and reaction mechanism of an iron-loaded catalyst derived from blast furnace slag for the CO-SO2 reaction to produce sulfur. Appl. Catal. A Gen. 2020, 606, 117810. [Google Scholar]
  19. Teixeira, G.F.; Silva, E., Jr.; Vilela, R.; Zaghete, M.A.; Colmati, F. Perovskite structure associated with precious metals: Influence on feterogenous catalytic process. Catalysts 2019, 9, 721. [Google Scholar]
  20. Jain, A.; Tamhankar, S.; Jaiswal, Y. Role of La-based perovskite catalysts in environmental pollution remediation. Rev. Chem. Eng. 2024, 40, 193–228. [Google Scholar]
  21. Chen, K.; Xu, L.; Li, Y.; Xiong, J.; Han, D.; Ma, Y.; Zhang, P.; Guo, H.; Wei, Y. Cerium doping effect in 3DOM perovskite-type La2−xCexCoNiO6 catalysts for boosting soot oxidation. Catalysts 2024, 14, 18. [Google Scholar]
  22. Fotovat, F.; Beyzaei, M.; Ebrahimi, H.; Mohebolkhames, E. Synthesis, characterization, and attrition resistance of kaolin and boehmite alumina-reinforced La0.7Sr0.3FeO3 perovskite catalysts for chemical looping partial oxidation of methane. Catalysts 2024, 14, 670. [Google Scholar] [CrossRef]
  23. Heidinger, B.; Royer, S.; Giraudon, J.-M.; Simon, P.; Bion, N.; Alamdari, H.; Lamonier, J.-F. Properties evolution of LaCoO3 perovskite synthesized by reactive grinding-application to the toluene oxidation reaction. J. Environ. Chem. Eng. 2024, 12, 112107. [Google Scholar]
  24. Moschos, M.; Evdou, A.; Zaspalis, V. (La1−xCax)MnO3−δ (x = 0, 0.2, 0.3, 0.4) perovskites as redox catalysts in chemical looping hydrogen production process: The relation between defect chemistry and redox performance. Catalysts 2024, 14, 431. [Google Scholar] [CrossRef]
  25. Ji, W.; Jin, Q.; Xu, M.; Chen, Y.; Yang, B.; Li, X.; Shen, Y.; Wang, Y.; Xu, H. Resource utilization of high-concentration SO2 for sulfur production over La-Ce-Ox composite oxide catalyst. Environ. Sci. Pollut. Res. 2023, 30, 21756–21768. [Google Scholar]
  26. Wang, G.-j.; Qin, Y.-n.; Ma, Z.; Qi, X.-z.; Ding, T. Study on the catalytic reduction mechanism of SO2 by CO over doped copper perovskite catalyst in presence of oxygen. React. Kinet. Catal. Lett. 2006, 89, 229–236. [Google Scholar]
  27. Zheng, X.; Li, B.; Shen, L.; Cao, Y.; Zhan, Y.; Zheng, S.; Wang, S.; Jiang, L. Oxygen vacancies engineering of Fe doped LaCoO3 perovskite catalysts for efficient H2S selective oxidation. Appl. Catal. B Environ. 2023, 329, 122526. [Google Scholar]
  28. Hibbert, D.B.; Campbell, R.H. Flue gas desulphurisation: Catalytic removal of sulphur dioxide by carbon monoxide on sulphided La1−xSrxCoO3: II. Reaction of sulphur dioxide and carbon monoxide in a flow system. Appl. Catal. 1988, 41, 289–299. [Google Scholar]
  29. Ma, J.; Fang, M.; Lau, N. On the synergism between La2O2S and CoS2 in the reduction of SO2 to elemental sulfur by CO. J. Catal. 1996, 158, 251–259. [Google Scholar]
  30. Wang, T.; Zhang, C.; Wang, J.; Li, H.; Duan, Y.; Liu, Z.; Lee, J.Y.; Hu, X.; Xi, S.; Du, Y.; et al. The interplay between the suprafacial and intrafacial mechanisms for complete methane oxidation on substituted LaCoO3 perovskite oxides. J. Catal. 2020, 390, 1–11. [Google Scholar]
  31. Ko, Y.; Kim, H.; Kim, S.; Lee, C.; Lee, S.S.; Roh, H.-S.; Shin, J.; Jeon, Y. CO management for hydrogen processes through a catalytic oxidation mechanism on dual-doped perovskites with tuned Co and Ni ratios. Catalysts 2025, 15, 45. [Google Scholar] [CrossRef]
  32. Zhu, Y.; Tan, R.; Feng, J.; Ji, S.; Cao, L. The reaction and poisoning mechanism of SO2 and perovskite LaCoO3 film model catalysts. Appl Catal. A Gen. 2001, 209, 71–77. [Google Scholar]
  33. Lau, N.; Fang, M.; Chan, C. Reduction of SO2 by CO and COS over La2O2S—A mechanistic study. J. Mol. Catal. A Chem. 2003, 203, 221–229. [Google Scholar] [CrossRef]
  34. Hibbert, D.B. Reduction of sulfur dioxide on perovskite oxides. Catal. Rev. 1992, 34, 391–408. [Google Scholar]
  35. Osti, A.; Rizzato, L.; Cavazzani, J.; Meneghello, A.; Glisenti, A. Perovskite oxide catalysts for enhanced CO2 reduction: Embroidering surface decoration with Ni and Cu nanoparticles. Catalysts 2024, 14, 313. [Google Scholar] [CrossRef]
  36. Zheng, J.-n.; An, K.; Wang, J.-m.; Li, J.; Liu, Y. Direct synthesis of ethanol via CO2 hydrogenation over the Co/La-Ga-O composite oxide catalyst. J. Fuel Chem. Technol. 2019, 47, 697–708. [Google Scholar]
  37. Ma, L.H.; Gao, X.H.; Zhang, J.L.; Ma, J.J.; Hu, X.D.; Guo, Q.J. Effects of metal doping on the catalytic performance of LaFe-based perovskites for CO2 hydrogenation to light olefins. J. Fuel Chem. Technol. 2023, 51, 101–110. [Google Scholar]
  38. Wu, S.; Zhan, s.; Dong, F.; Song, X.; Han, W.; Han, W.; Zhang, H.; Dong, X.; Tang, Z. Engineering surface exposed LaCoO3 perovskite nanotubular catalyst for catalytic combustion of toluene through acid etching. J. Mater. Chem. A 2025, 10, 1039. [Google Scholar]
  39. Teng, F.; Liang, S.; Gaugeu, B.; Zong, R.; Yao, W.; Zhu, Y. Carbon nanotubes-templated assembly of LaCoO3 nanowires at low temperatures and its excellent catalytic properties for CO oxidation. Catal. Commun. 2007, 8, 1748–1754. [Google Scholar]
  40. Zhang, X.; Cheng, X.; Chen, F.; Xu, L.; Wang, Y.; Qian, J.; Wu, Z.; Zhang, Q. Facile loading carbon dots on Co3O4 as an enhanced oxygen reduction reaction catalyst. Chem. Phys. Lett. 2020, 740, 137058. [Google Scholar]
  41. Cao, Z.; Li, J.; Zhao, Y.; Mei, Q.; Wang, Q.; Cheng, H. Meta-kaolinite/LaFeCoO3 microsphere catalyst for photocatalytic persulfate activation: Enhanced removal of tetracycline hydrochloride. Chem. Eng. J. 2023, 466, 143076. [Google Scholar]
  42. He, N.; Yu, Z.; Yang, G.; Tan, Q.; Wang, J.; Chen, Y. Designing with A-site cation defects in LaFeO3: Removal of tetracycline hydrochloride in complex environments using photo-Fenton synergy. Chem. Eng. J. 2024, 484, 149613. [Google Scholar]
  43. Kostyukhin, E.M.; Kustov, A.L.; Evdokimenko, N.V.; Bazlov, A.I.; Kustov, L.M. Hydrothermal microwave-assisted synthesis of LaFeO3 catalystfor N2O decomposition. J. Am. Ceram. Soc. 2021, 104, 492–503. [Google Scholar]
  44. Feng, X.; Tian, M.; He, C.; Li, L.; Shi, J.; Yu, Y.; Cheng, J. Yolk-shell-like mesoporous CoCrOx with superior activity and chlorine resistance in dichloromethane destruction. Appl. Catal. B Environ. 2020, 264, 118493. [Google Scholar]
  45. Lv, Y.; Li, Y.; Ta, N.; Shen, W. Co3O4 nanosheets: Synthesis and catalytic application for CO oxidation at room temperature. Sci. China Chem. 2014, 57, 873–880. [Google Scholar]
  46. Wen, M.; Dong, F.; Tang, Z.; Zhang, J. Engineering order mesoporous CeCoOx catalyst via in-situ confined encapsulation strategy for VOCs catalytic combustion. Mol. Catal. 2022, 519, 112149. [Google Scholar] [CrossRef]
  47. Zemlianskii, P.; Morozov, D.; Kapustin, G.; Davshan, N.; Kalmykov, K.; Chernyshev, V.; Kustov, A.; Kustov, L. Correlations between synthetic conditions and catalytic activity of LaMO3 perovskite-like oxide materials (M: Fe, Co, Ni): The key role of glycine. ChemPhysMater 2025, 2, 59. [Google Scholar]
Figure 1. XRD patterns of (A) as-synthesized catalysts and sulfurized catalysts and (B) spent catalysts.
Figure 1. XRD patterns of (A) as-synthesized catalysts and sulfurized catalysts and (B) spent catalysts.
Catalysts 15 00291 g001
Figure 2. (A) N2 adsorption–desorption curve and (B) pore diameter distribution of LaCo1−xFexO3.
Figure 2. (A) N2 adsorption–desorption curve and (B) pore diameter distribution of LaCo1−xFexO3.
Catalysts 15 00291 g002
Figure 3. SEM of (A) LaCo0.8Fe0.2O3, (B) OFS-LaCo0.8Fe0.2O3, (C) OAS-LaCo0.8Fe0.2O3, (D) OAR-OFS-LaCo0.8Fe0.2O3, (E) OAR-OAS-LaCo0.8Fe0.2O3, (F) EDS mapping of OFS-LaCo0.8Fe0.2O3.
Figure 3. SEM of (A) LaCo0.8Fe0.2O3, (B) OFS-LaCo0.8Fe0.2O3, (C) OAS-LaCo0.8Fe0.2O3, (D) OAR-OFS-LaCo0.8Fe0.2O3, (E) OAR-OAS-LaCo0.8Fe0.2O3, (F) EDS mapping of OFS-LaCo0.8Fe0.2O3.
Catalysts 15 00291 g003
Figure 4. SO2 conversion with temperature over (A) LaCo1−xFexO3 catalysts under the oxygen-assisted reaction (OAR) and (B) oxygen-free sulfurized OFS-LaCo0.8Fe0.2O3 catalysts and oxygen-assisted sulfurized OAS-LaCo0.8Fe0.2O3 catalysts under the oxygen-assisted reaction (OAR) or oxygen-free reaction (OFR); (C) catalytic thermal stability test for SO2 conversion with reaction time under OAR at 550 °C over OFS-LaCo0.8Fe0.2O3 catalysts, and (D) catalytic stability test for SO2 conversion with water tolerance under OAR at theoretical water vapor of 20% and 40% at 550 °C over OFS-LaCo0.8Fe0.2O3 catalysts.
Figure 4. SO2 conversion with temperature over (A) LaCo1−xFexO3 catalysts under the oxygen-assisted reaction (OAR) and (B) oxygen-free sulfurized OFS-LaCo0.8Fe0.2O3 catalysts and oxygen-assisted sulfurized OAS-LaCo0.8Fe0.2O3 catalysts under the oxygen-assisted reaction (OAR) or oxygen-free reaction (OFR); (C) catalytic thermal stability test for SO2 conversion with reaction time under OAR at 550 °C over OFS-LaCo0.8Fe0.2O3 catalysts, and (D) catalytic stability test for SO2 conversion with water tolerance under OAR at theoretical water vapor of 20% and 40% at 550 °C over OFS-LaCo0.8Fe0.2O3 catalysts.
Catalysts 15 00291 g004aCatalysts 15 00291 g004b
Figure 5. COS production over (A) LaCo1−xFexO3 catalysts under the oxygen-assisted reaction and (B) sulfurized LaCo0.8Fe0.2O3 catalysts under the oxygen-assisted reaction or oxygen-free reaction.
Figure 5. COS production over (A) LaCo1−xFexO3 catalysts under the oxygen-assisted reaction and (B) sulfurized LaCo0.8Fe0.2O3 catalysts under the oxygen-assisted reaction or oxygen-free reaction.
Catalysts 15 00291 g005
Figure 6. (A) CO2-TPD profiles, (B) CO-TPD profiles, (C) SO2-TPD profiles of catalysts.
Figure 6. (A) CO2-TPD profiles, (B) CO-TPD profiles, (C) SO2-TPD profiles of catalysts.
Catalysts 15 00291 g006
Figure 7. XPS spectra of catalysts: (A) Co 2p; (B) Fe 2p; (C) O 1s; (D) S 2p.
Figure 7. XPS spectra of catalysts: (A) Co 2p; (B) Fe 2p; (C) O 1s; (D) S 2p.
Catalysts 15 00291 g007
Figure 8. Schematic diagram of the fixed-bed catalytic reaction.
Figure 8. Schematic diagram of the fixed-bed catalytic reaction.
Catalysts 15 00291 g008
Table 1. Textural data of the samples.
Table 1. Textural data of the samples.
SamplesSBET (m2/g)Pore Volume (cm3/g)Pore Size (nm)
LaCoO312.860.1023.23
LaCo0.8Fe0.2O314.670.1021.21
LaCo0.5Fe0.5O316.040.1019.05
LaCo0.2Fe0.8O314.820.1020.44
LaFeO314.480.1021.88
Table 2. The atomic percentage data on the surface of the catalysts detected by SEM-EDX.
Table 2. The atomic percentage data on the surface of the catalysts detected by SEM-EDX.
SampleO (%)S (%)Fe (%)Co (%)La (%)
LaCo0.8Fe0.2O370.3104.1813.5012.01
OFS-LaCo0.8Fe0.2O375.453.853.487.919.31
OAS-LaCo0.8Fe0.2O371.411.123.9611.2212.29
OAR-OFS-LaCo0.8Fe0.2O373.7212.451.365.437.06
OAR-OAS-LaCo0.8Fe0.2O339.4131.732.6810.5915.59
Table 3. The atomic percentage data on the surface of the catalysts detected by XPS.
Table 3. The atomic percentage data on the surface of the catalysts detected by XPS.
SampleO (%)S (%)Fe (%)Co (%)La (%)
LaCo0.8Fe0.2O353.7003.746.1412.29
OFS-LaCo0.8Fe0.2O344.8113.202.885.216.76
OAS-LaCo0.8Fe0.2O351.325.773.134.239.22
OAR-OFS-LaCo0.8Fe0.2O350.0315.901.932.786.31
OAR-OAS-LaCo0.8Fe0.2O350.5713.142.144.476.39
Table 4. XPS results of LaCoO3, LaCo0.8Fe0.2O3, OAS-LaCo0.8Fe0.2O3, and OFS-LaCo0.8Fe0.2O3.
Table 4. XPS results of LaCoO3, LaCo0.8Fe0.2O3, OAS-LaCo0.8Fe0.2O3, and OFS-LaCo0.8Fe0.2O3.
SampleCo2+/Co3+Fe2+/Fe3+Oads/OtotalBinding Energy (eV)
Co2+(2p3/2)Co3+(2p3/2)Fe2+(2p3/2)Fe3+(2p3/2)
LaCoO30.80-0.59781.6780.0--
LaCo0.8Fe0.2O30.610.740.47781.6780.0710.5713.7
OAS-LaCo0.8Fe0.2O30.820.840.59781.4779.8710.6713.8
OFS-LaCo0.8Fe0.2O31.291.250.74780.6779.0710.8714.0
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yao, L.; Wang, H.; Li, S.; Chen, Y. Catalytic Reduction of SO2 with CO over LaCoO3 Perovskites Catalysts: Effect of Fe Doping and Pre-Sulfurization. Catalysts 2025, 15, 291. https://doi.org/10.3390/catal15030291

AMA Style

Yao L, Wang H, Li S, Chen Y. Catalytic Reduction of SO2 with CO over LaCoO3 Perovskites Catalysts: Effect of Fe Doping and Pre-Sulfurization. Catalysts. 2025; 15(3):291. https://doi.org/10.3390/catal15030291

Chicago/Turabian Style

Yao, Liang, Hao Wang, Shuangde Li, and Yunfa Chen. 2025. "Catalytic Reduction of SO2 with CO over LaCoO3 Perovskites Catalysts: Effect of Fe Doping and Pre-Sulfurization" Catalysts 15, no. 3: 291. https://doi.org/10.3390/catal15030291

APA Style

Yao, L., Wang, H., Li, S., & Chen, Y. (2025). Catalytic Reduction of SO2 with CO over LaCoO3 Perovskites Catalysts: Effect of Fe Doping and Pre-Sulfurization. Catalysts, 15(3), 291. https://doi.org/10.3390/catal15030291

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop