Next Article in Journal
(1E)-1,2-Diaryldiazene Derivatives Containing a Donor–π-Acceptor-Type Tolane Skeleton as Smectic Liquid–Crystalline Dyes
Previous Article in Journal
Performance of Stainless-Steel Bipolar Plates (SS-BPPs) in Polymer Electrolyte Membrane Water Electrolyser (PEMWE): A Comprehensive Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

NOx Storage and Reduction (NSR) Performance of Sr-Doped LaCoO3 Perovskite Prepared by Glycine-Assisted Solution Combustion

1
School of Water Conservancy and Environment, University of Jinan, Jinan 250022, China
2
Zaozhuang Ecological Environment Monitoring Center of Shandong Province, Zaozhuang 277800, China
3
School of Materials Science and Engineering, University of Jinan, Jinan 250022, China
*
Author to whom correspondence should be addressed.
Compounds 2024, 4(2), 268-287; https://doi.org/10.3390/compounds4020014
Submission received: 15 January 2024 / Revised: 26 March 2024 / Accepted: 1 April 2024 / Published: 8 April 2024

Abstract

:
Here, we successfully synthesized Sr-doped perovskite-type oxides of La1−xSrxCo1−λO3−δ, “LSX” (x = 0, 0.1, 0.3, 0.5, 0.7), using the glycine-assisted solution combustion method. The effect of strontium doping on the catalyst structure, NO to NO2 conversion, NOx adsorption and storage, and NOx reduction performance were investigated. The physicochemical properties of the catalysts were studied by XRD, SEM-EDS, N2 adsorption–desorption, FTIR, H2-TPR, O2-TPD, and XPS techniques. The NSR performance of LaCoO3 perovskite was improved after Sr doping. Specifically, the perovskite with 50% of Sr doping (LS5 sample) exhibited excellent NOx storage capacity within a wide temperature range (200–400 °C), and excellent stability after hydrothermal and sulfur poisoning. It also displayed the highest NOx adsorption–storage capacity (NAC: 1889 μmol/g; NSC: 1048 μmol/g) at 300 °C. This superior performance of the LS5 catalyst can be attributed to its superior reducibility, better NO oxidation capacity, increased surface Co2+ concentration, and, in particular, its generation of more oxygen vacancies. FTIR results further revealed that the LSX catalysts primarily store NOx through the “nitrate route”. During the lean–rich cycle tests, we observed an average NOx conversion rate of over 50% in the temperature range of 200–300 °C, with a maximum conversion rate of 61% achieved at 250 °C.

1. Introduction

Nitrogen oxides (NOx) are considered to be significant air pollutants generated in diverse combustion processes. Researchers have explored and recognized NOx storage and reduction (NSR) [1] and selective catalytic reduction (SCR) [2] as effective methods for removing NOx from exhaust emissions. NSR catalysts typically consist of active components such as noble metals (Pt, Pd, Rh) [3,4], storage components such as alkali or alkaline earth metals, and supports such as Al2O3 or CeO2 [5,6]. A representative catalyst commonly employed in NOx storage and reduction (NSR) systems is Pt-BaO/Al2O3. The operation of this catalyst involves alternating cycles of lean and rich conditions. Specifically, in the lean phase, the platinum (Pt) active sites facilitate the oxidation of NO to NO2. Following this conversion, the resulting NO2 species are subsequently stored as nitrates and nitrites on the barium oxide (BaO) sites [7]. In the rich phase, the oxidation of NOx occurs, followed by its reduction to N2 through the catalytic action of a reducing agent [8,9]. However, the use of Pt-based catalysts is limited due to their high cost, low activity at low temperatures [10], susceptibility to sintering at high temperatures, and poor resistance to sulfur [11]. Therefore, there is an urgent need to develop NSR catalysts with excellent low-temperature activity.
Perovskite (ABO3) catalysts have been widely employed in environmental catalysis due to their low cost, superior thermal stability, and catalytic activity [12]. Alkali or alkaline earth metals have the potential to act as substitutes for the A-site, while transition metals can replace the B-site in perovskite catalysts. Incorporation of alkali or alkaline earth metal dopants plays a vital role in controlling the performance of perovskite catalysts by facilitating the creation of more oxygen vacancies and stabilizing the perovskite structure. The study conducted by Ueda et al. [13] revealed that the incorporation of Ba2+ ions amounting to 30% into La3+-based perovskite (LaFe0.97Pd0.03O3) remarkably tripled the capacity for NOx storage. Similarly, Panunzi et al. [14] synthesized lanthanum strontium ferrite (LSFPt), and investigated how the oxygen vacancy content and the catalyst surface activity were enhanced in combined Sr substitution at the A-site and Pt doping at the B-site of lanthanum ferrites, leading to a substantial improvement in the activity and stability of methane oxidation. Concomitantly, Ji et al. [15] proposed that Sr dopants played a crucial role in enhancing the performance of catalysts during low-temperature reduction. Moreover, substitution of La with Sr in perovskite compounds like LaCoO3 and LaMnO3 exhibited enhanced oxidation activity, effectively converting NO to NO2 [16,17]. Thus, Sr doping at the A site in perovskite catalysts yields improved NO oxidation performance and thermal stability.
NSR for environmental catalysis is known to have various drawbacks associated with thermal aging and sulfur poisoning [18,19]. The catalytic materials are typically exposed to high temperature, resulting in sintering of active sites/promoters/catalytic support materials and loss of specific surface area (SSA) and functionality [20]. Furthermore, since acidic NO2(g) and SO2(g) adsorbates compete for similar adsorption sites on the catalyst surface, SOx species gradually accumulate over the NOx-storage components and form more stable sulfides, diminishing the NSC of the catalyst [21]. Therefore, the ABO3 surface chemistry and composition of catalytic materials need to be tuned at the nanoscale to improve their NSC, thermal stability, and resistance to sulfur neutrality.
In this study, we aimed to enhance the NOx storage and reduction performance of LaCoO3 catalysts by Sr doping at the A-site. The solution combustion method was employed to prepare the oxide materials. This method is attractive and simple [22], as it produces pure, uniform, and crystalline products with the desired composition and structure. In this study, we employed a rational design strategy and optimized the synthesis process, leading to the successful synthesis of perovskite-type catalysts. Specifically, La1−xSrxCo1−λO3−δ (x = 0, 0.1, 0.3, 0.5, 0.7) catalysts were synthesized via the glycine-assisted solution combustion method. Our research focused on elucidating the influence of Sr doping on the NSR performance. Additionally, Infrared Spectroscopy (FTIR) was employed to analyze the NOx storage route on the catalyst surface.

2. Experimental

2.1. Catalyst Preparation

The perovskite-type catalysts La1−xSrxCoO3 (x = 0, 0.1, 0.3, 0.5, 0.7) were synthesized using the glycine-assisted solution combustion method, denoted as LSX (X = 0, 1, 3, 5, 7). The theoretical value of (La + Sr)/Co is set to 1:1. A precise amount of La(NO3)3·6H2O, Co(NO3)2·4H2O, and Sr(NO3)2·4H2O was dissolved in deionized water. Following the addition of glycine (G/N = 1.6) to the mixture of nitrates, the resulting solution was stirred at ambient temperature for a duration of 2 h. Subsequently, the solution underwent evaporation under constant stirring at 80 °C until a viscous gel with a distinctive purple hue was obtained. The gel was combusted in an oven at 200 °C. After the reaction was completed, the fluffy flocculent product was ground to a powder and calcined in a muffle furnace at 700 °C for 4 h.

2.2. Catalysts Characterization

X-ray diffraction (XRD) patterns of the samples were obtained using the German D8 advance. The radiation source used was Cu Kα. A scanning range of 10 to 90° was employed, with a scanning speed of 6° per minute.
The specific surface area, pore volume, and size of the samples were measured at −196 °C through utilization of an ASAP 2020 automated specific surface area and porosity analyzer. In preparation for the assessment, all samples were subjected to a 6 h vacuum degassing process at 300 °C to eliminate any adsorbed species. The surface areas of the samples were determined by employing the Brumauer–Emmett–Teller (BET) technique.
The surface morphology and microstructure of the specimens were examined utilizing a Quanta FEG 250 scanning electron microscope (Carl Zeiss AG, Oberkochen, Germany). The applied working voltage was 10 kV, and a thin layer of gold was deposited on the specimens for 60 s prior to the scanning process.
The infrared spectra (FTIR) of the specimens were recorded with a Bruker tensor 27 infrared spectrometer (Bruker, Saarbrucken, Germany) located in Germany. The spectra were acquired by conducting 16 scans at a resolution of 4 cm−1. Prior to the analysis, the specimens were mixed with dry KBr in a ratio of 1:100 and subsequently compressed into tablets using a tablet press under a pressure of 10 MPa.
Using a quartz tube reactor equipped with a TCD (thermal conductivity detector) (Lunan Ruihong Chemical, Zaozhuang, China), we conducted temperature-programmed reduction with H2 (H2-TPR) measurements. The catalysts, weighing 0.05 g, were progressively heated from ambient temperature to 900 °C in the presence of a 5 vol.% H2/N2 mixture. The heating rate was set at 10 °C per minute, while maintaining a total flow rate of 30 mL/min.
O2 temperature-programmed desorption (O2-TPD) tests were conducted to observe the dynamic process of oxygen adsorption and desorption on perovskite. A quartz tube reactor with a capacity of approximately 150 mg of catalyst sample was utilized. The sample was subjected to a purification procedure involving a 30 min purge using a mixture of 10 vol% O2 in N2 flow (50 mL/min) at a temperature of 500 °C. Following the purification step, the sample was cooled to 50 °C and underwent an additional 30 min purge with N2 flow (50 mL/min) to eliminate any residual adsorbed oxygen. Subsequently, a gradual heating process was conducted from 50 to 900 °C at a heating rate of 10 °C/min under a N2 flow (30 mL/min), while simultaneously monitoring the released oxygen employing a TCD. Integration of the resulting TPD curves facilitated the determination of the quantities of oxygen species present.
By employing the Thermo Scientific X-ray photoelectron spectrometer (Bruker, Saarbrucken, Germany) equipped with an Al Kα X-ray radiation source, we were able to precisely determine the chemical composition and oxidation states of surface elements in catalysts. To calibrate the binding energy measurements, we utilized the C1s peak (BE = 284.6 eV) as a reference standard.

2.3. Catalytic Activity Measurements

2.3.1. NO Oxidation Experiments

The investigation into the oxidation process of NO was conducted using the temperature-programmed oxidation (NO-TPO) technique. Prior to the experimentation, the catalysts underwent a pretreatment process involving exposure to N2 at a temperature of 500 °C for a duration of one hour, followed by natural cooling to room temperature. Subsequently, the gas mixture was converted to contain 1000 ppm NO/N2 and 5% O2/N2, and left to stabilize for a certain period. Finally, once the NOx concentration at the outlet remained constant, the temperature was increased to 700 °C at a rate of 5 °C per minute. The concentration of NOx was measured using a chemiluminescence NOx analyzer.

2.3.2. NOx Adsorption–Desorption Measurements

The catalysts (50 mg, 40–80 mesh) were examined for their NOx adsorption and storage capacity (NAC/NSC) using a custom-built catalyst sample evaluation system equipped with a chemiluminescence NOx analyzer (Thermo 42i–HL) (Thermo Fisher Scientific, Waltham, MA, USA). Initially, the feed gas was composed of 1000 ppm of NO and 5% O2 for NOx adsorption, while N2 was utilized as the balance gas. After adsorbing NOx for 1 h, the system was purged by introducing N2 into the reaction chamber. Subsequently, the temperature was ramped up to 700 °C at a heating rate of 10 °C/min. The total flow rate was maintained at 100 mL/min. The LS5 catalyst after hydrothermal aging and sulfur poisoning was re-tested under the same conditions as those used for the fresh catalyst. The NAC/NSC was determined by calculating the NOx adsorption–desorption curves using Equations (1) and (2).
NAC = ( 0 t ( NO X , in NO X , out ) × V ) / ( 22.4 × m c )
NSC = ( 0 t ( NO X , out ) × V ) / ( 22.4 × R × m c )
where NOx,in is the NOx concentration at the inlet, NOx,out is the NOx concentration at the outlet, t is time, V is the flow rate, mc is the catalyst quantity, and R is the heating rate.

2.3.3. NOx Storage and Reduction Measurements

To assess the efficiency of catalysts in reducing NOx emissions, a comprehensive array of NOx storage and reduction experiments was executed, spanning diverse temperature ranges. The experimental design entailed 15 cycles, wherein each lean phase persisted for 2 min, incorporating a gas mixture comprising 500 ppm of NO/N2 and 7.5 vol.% of O2/N2. Conversely, the rich phase spanned 1 min, during which a gas composition of 500 ppm of NO/N2 and 1200 ppm of C3H8/He was employed. The overall volumetric flow rate was consistently maintained at 50 mL/min (GHSV: 60,000 h−1). The chemiluminescence NOx analyzer was used to measure the NOx concentration at the outlet, and subsequently, the NOx removal efficiency of the catalysts was determined using Equation (3).
NO X   conversion   ( % ) = 0 t ( NO X . in NO X . out ) d t 0 t ( NO X . in ) d t · 100 %

3. Results and Discussion

3.1. XRD Analysis

The XRD patterns of LSX perovskite catalysts are illustrated in Figure 1a. Pure perovskite diffraction patterns were obtained with Sr doping amounts lower than 30%, indicating the absence of impurities. However, after Sr doping reached 50%, trace amounts of impurities in the form of SrCO3 and SrCoOx were detected on the LS5 catalyst. In addition, the impurity phase SrCoOx was distinctly observed on the LS7 sample. Therefore, it can be concluded that the LS5 catalyst allows for higher Sr accommodation in the lattice compared to the LS7 catalyst. The cobalt in perovskite serves as the active oxidation site [23], while the hybrid SrCoOx species may occupy the LS7 catalyst surface and cover the active component, thereby influencing the catalyst activity.
An enlarged version of the peak from the crystal plane (110) of LSX perovskite catalysts is presented in Figure 1b. The peak of the crystal plane (110) shifts to a lower 2θ angle as the Sr doping content increases. This shift is caused by lattice expansion resulting from the isomorphic substitution of Sr2+ cations with larger ionic radii replacing La3+ ions. Furthermore, the LSX catalysts exhibit a rhomboid bimodal structure in the absence of, or with low, strontium doping (X ≤ 1) [24]. As the Sr doping content increases, this typical rhomboid bimodal structure weakens and gradually transforms into a cubic symmetric singlet structure [25].

3.2. SEM and EDS Mapping Analysis

The scanning electron microscope (SEM) images of LS0 and LS5 catalysts are presented in Figure 2a,b. Both catalysts exhibit similar loose and porous spongy morphological structures, both before and after Sr doping. The formation of abundant pores can be attributed to the release of CO2, N2, and H2O during the combustion of the precursor materials [26,27]. This porous structure enhances the dispersion, absorption, and desorption of gas reactants. However, there are differences in grain size and dispersion between the two catalysts, which can be attributed to the variation in Sr doping. The LS5 catalyst has smaller particle diameters and appears to be in a looser state. A smaller grain size provides more catalytic active sites, leading to higher catalytic activity.
The energy-dispersive X-ray spectroscopy (EDS) mapping of LS0 and LS5 catalysts is shown in Figure 2c,e, revealing the uniform distribution of La, Sr, Co, and O atoms across their respective surfaces. Furthermore, the distribution diagram depicted in Figure 2d,f facilitates the determination of the (La + Sr)/Co ratio on the catalyst surfaces, which aligns with the theoretical outcomes as illustrated in Table 1.

3.3. N2 Adsorption–Desorption Analysis

The N2 adsorption–desorption curves of the samples are depicted in Figure 3a. All catalysts exhibit IUPAC type II isotherms, and a H3 type hysteresis loop is observed in the range of p/p0 = 0.8–1.0, indicating the presence of macroporous materials [28]. The pore size distribution curve in Figure 3b combined with SEM images reveals that all catalysts possess a multistage pore structure consisting of both mesopores and macropores.
In Table 1, the LSX catalysts are characterized by their BET surface areas (SBET), pore volume (Vp), and average pore size (Dp). Among them, the LS0 sample possesses the smallest surface area of 5.58 m2/g and a corresponding pore volume of 0.05 cm3/g. Through Sr doping, the LSX catalysts exhibit an increasing trend in both surface area and pore volume with the escalating Sr doping amount, while the pore size demonstrates the opposite relationship. The LS5 sample demonstrates the largest surface area of 17.71 m2/g, coupled with a pore volume of 0.11 cm3/g. However, as the Sr doping content further rises (X = 7), LS7 experiences a slight reduction in both surface area and pore volume. This may be due to impurities, such as SrCoOx and SrCO3, which block the catalyst’s pores and affect its textural parameters. Previous studies [29,30] have proposed that optimizing the pore structure by reducing its size can significantly improve gas transport properties and boost catalytic activity. Furthermore, Table 1 reveals that Sr doping significantly increases the mesoporous rate of the catalysts, leading to a larger surface area in the corresponding samples. This larger surface area exposes more active sites, facilitating contact between the catalyst and gas, and thus improving reaction efficiency [31].

3.4. FTIR and H2-TPR Analysis

The FTIR spectra of the LSX catalysts are presented in Figure 4a. Among the catalysts with varying amounts of Sr doping, it is observed that the bending vibration of Co-O bonds in the octahedral structure of perovskite BO6 is assigned to approximately 580 cm−1 [32]. This suggests the successful preparation of the perovskite structure. Additionally, the peaks at 858 and 1451 cm−1 are attributed to surface carbonate [33,34]. When considering the XRD results, it is further confirmed that SrCoO3 exists in the samples. Comparing the catalysts before and after Sr doping, it is evident that the characteristic peak of perovskite is significantly weakened upon the introduction of Sr, indicating that Sr has a certain impact on the integrity of the perovskite crystal.
The reducibility of the LSX samples was assessed via H2-TPR, as depicted in Figure 4b. The peaks observed in the temperature range of 100 °C to 800 °C can all be attributed solely to the reduction of Co3+. This finding is consistent with the notion that La3+ and Sr2+ are impervious to reduction under the specific experimental conditions employed [24]. In all catalysts, the reduction peaks witnessed in the 300–430 °C range correspond to the reduction of Co3+, whereas those observed within the 440–620 °C range align with the reduction of Co2+ [35,36]. Specifically, between 300 and 430 °C, there are two reduction peaks for Co3+ with x > 1. The peak at the lower temperature is related to surface/highly reactive Co3+, which is in accordance with the stepwise reduction behavior of fine Co3O4 particles [37]. Notably, the reduction temperature of Co species on the doped catalysts exhibits a noticeable decrease compared to that of the undoped catalyst when the Sr doping level is maintained below 50%. This noteworthy trend alludes to the beneficial influence of Sr doping, which serves to enhance the reduction performance of the catalysts and facilitate the reduction of surface oxygen [24]. However, when X ≥ 5, the intensity of the reduction peak in the range of 500–540 °C decreased significantly. Combined with the XRD results, it is speculated that the reason for this occurrence may be the inhibitory effect of the presence of SrCoOx and SrCO3 impurities on the reduction of Co2+. Furthermore, exceeding a certain threshold of Sr doping leads to a slight elevation in the reduction temperature of Co species, signifying that excessive Sr doping does not favor the reduction process of the catalysts.
The H2 consumption of all samples is listed in Table 2. It can be concluded that Sr doping increases the H2 consumption of the catalysts. Noticeably, the LS3 catalyst possesses the highest H2 consumption (8.1 mmol/g), followed by the LS5 catalyst (7.5 mmol/g). The result shows that LS3 and LS5 catalysts have better reducibility.

3.5. O2-TPD

O2 temperature-programmed desorption tests were conducted on the LSX catalysts and the results are presented in Figure 5. The integral areas of the α and β peaks were determined from the O2-TPD spectra, as shown in Table 2. Except for LS0, all LSX catalysts exhibited a α-oxygen desorption peak below 300 °C and a β-oxygen desorption peak above 600 °C. The α-peak in the graph represents the desorption of surface-adsorbed oxygen (O2−) species, whereas the β-peak corresponds to the liberation of lattice oxygen (O2−) ions [38]. The graph displays an upsurge in the intensity of the α-peak during Sr doping, indicating that Sr doping facilitates the desorption of surface oxygen species. Typically, an increase in the concentration of oxygen species associated with the α-peak enhances catalytic activity [12]. Among the catalysts, LS5 demonstrated the lowest α-peak temperature of 118 °C and the largest peak area, indicating a higher concentration of surface oxygen species and stronger oxidation properties. Furthermore, the intensity of the β-oxygen desorption peak increased with Sr doping, with the maximum observed for the LS7 catalyst.

3.6. XPS

The surface properties of the synthesized LSX catalysts were confirmed using X-ray photoelectron spectroscopy (XPS), as illustrated in Figure 6. The Sr 3d XPS peak exhibited two distinct sets of doublets, specifically Sr 3d5/2 and 3d3/2. The peaks observed at binding energies of 131.7 and 134.1 eV can be assigned to Srlatt, while the peaks detected at 133.1 and 135.2 eV can be attributed to Srsurf [11,39]. The ratios of Srsurf/Srlatt are summarized in Table 1. The proportion of Srsurf increases with an increase in Sr doping. It is likely that Srsurf species are present in the form of SrCO3, as suggested by XRD and FTIR results. An increase in the amount of Srsurf signifies an increase in the content of SrCO3. This is significant because SrCO3 is an important storage phase for NOx adsorption, suggesting that the LS5 catalyst may possess a strong capacity for NOx adsorption and storage.
The Co 2p XPS peak was fitted with two sets of Co 2p3/2 and 2p1/2 doublets. The peaks at 779.7 and 794.9 eV were assigned to Co3+, while the peaks at 781 and 796.5 eV were attributed to Co2+. The peaks at 790 and 804 eV were considered satellite peaks [40,41]. The bands of the fitted O1s spectra at 528.8–529.1 eV, 530.3–531.5 eV, and 533.0–533.8 eV could be assigned to lattice oxygen (OI), surface reactive oxygen (OII), and chemisorbed water (OIII), respectively [42]. Table 3 presents the (La + Sr)/Co, Co2+/Co3+, and OII/(OI + OII + OIII) ratios. The (La + Sr)/Co ratio increases with the decrease in the Sr doping amount, indicating that Sr doping results in the partial bulk Co ion segregation on the LSX surface. The high value of the (La + Sr)/Co ratio can be attributed to various reasons, one of which is the significant variation in this molar ratio from one site to another. This variation is caused by the swift application of high power densities, which leads to a highly heterogeneous structure [43]. As a result, there are abrupt changes in the (La + Sr)/Co values, both chemically and morphologically [44]. Therefore, this change is denoted by “1 − λ” in La1−xSrxCo1−λO3−δ. The quantities of both Co2+ cations and OII species also increase with the Sr doping content. Notably, the LS3 catalyst exhibits the highest number of reactive oxygen species on its surface. This suggests that the content of reactive oxygen does not completely determine the activity of the catalysts. It has been shown that the oxygen vacancy is formed on the site of oxygen, coupled with cobalt [45]. Therefore, variations in OII fractions may be due to differences in oxygen vacancies or defects on the sample surface, or differences in oxygen-deficient regions in the cobalt-containing structures [46]. Changes in the concentration of oxygen vacancies or defects in the sample may favor the chemisorption process. The LS5 catalyst contains a higher amount of surface Co2+, which leads to the formation of oxygen vacancies and improves the NO oxidation capacity of the catalyst, thereby enhancing the NOx storage performance.

3.7. NO Oxidation

The NOx storage capacity of non-noble metal catalysts is subject to the influence of their NO oxidation capability [47]. The temperature-programmed oxidation of NO was conducted on LSX catalysts and the corresponding results are illustrated in Figure 7. The presence of the catalysts facilitates the conversion of NO to NO2 through oxidation reactions. Remarkably, even at temperatures as low as 100–150 °C, a considerable fraction of NO, amounting to approximately 15%, can be converted into NO2. The LSX catalysts reach their peak NO conversions within the temperature range of 300–365 °C. As the temperature continues to rise, the decline in NO conversion can be attributed to the thermodynamic equilibrium associated with NO oxidation. Notably, LaCoO3 exhibits remarkable catalytic activity towards NO oxidation, resulting in a peak conversion rate of 49.5% at 365 °C. The achievement of a higher conversion rate at a lower peak temperature indicates an enhanced capability for NO oxidation. Based on the data presented in Figure 7, it can be deduced that Sr doping enhances the NO oxidation capability of the catalysts, leading to a conversion rate surpassing 50%. Among the catalysts examined, LS3 demonstrates the optimal NO oxidation ability, achieving a maximum conversion rate of 76.8% at 305 °C. This observation can be attributed to an elevated presence of surface oxygen species and Co2+ in LS3. The generation of a substantial amount of NO2 during NOx adsorption proves advantageous for adsorption via the “nitrate” pathway [48].

3.8. NOx Adsorption/Storage Capacity

To assess the impact of Sr doping on the efficiency of NOx storage, experiments were conducted to measure the adsorption and desorption of NOx at a temperature of 300 °C. The results are presented in Figure 8. The adsorption curve (Figure 8a) clearly demonstrates that at the initial stage of adsorption, NOx is almost entirely adsorbed. As adsorption progresses, the concentration of NOx at the outlet gradually increases and stabilizes, suggesting that NOx adsorption reaches saturation. Throughout this process, NO2 is detected at the outlet, indicating the oxidation of NO to NO2. Previous studies have indicated that NO2 can be readily stored in the form of nitrate, and the conversion of NO to NO2 is the rate-determining step for NOx storage in Pt/BaO/Al2O3 catalysts [49]. Figure 8b illustrates the desorption behavior of the LSX catalysts. All catalysts exhibit a one-step desorption process. With increasing Sr doping content, the desorption peak of the catalysts shifts to higher temperatures, indicating that the addition of more strontium to perovskite leads to greater stability of the nitrate species.
Table 4 provides the data for the NOx adsorption capacity (NAC) and NOx storage capacity (NSC) of the LSX catalysts, as determined from their NOx adsorption–desorption profiles. It is important to note that the NAC and NSC show significant increases after Sr doping, suggesting that Sr enhances the storage of NOx. Additionally, both the NAC and NSC of the catalysts increase with higher levels of Sr doping (X ≤ 5). Among the catalysts, LS5 exhibits the highest adsorption and storage capacity, with an NAC of 1889 µmol/g and an NSC of 1084 µmol/g. This can be attributed to its larger surface area, strong reducibility, excellent NO oxidation activity, and the charge imbalance caused by Sr doping at the A-site, which promotes the generation of more active Co2+ cations and surface reactive oxygen species [24]. However, when Sr doping continues to increase, the NAC and NSC of the catalysts decrease. This is likely due to an excessive amount of Sr doping, which results in the formation of large amounts of SrCO3 and SrCoOx. These heterogeneous compounds do not disperse well on the surface, and the presence of a significant quantity of SrCoOx reduces the availability of effective storage sites and active sites, ultimately weakening the storage capacity of the catalysts.
Wang et al. [50] developed LaCo0.90Pt0.10O3 catalysts based on chalcogenides. These catalysts achieved a maximum NSC of 22.6 μmol/g at 300 °C. Additionally, Wen et al. [51] studied the impact of adding LaCoO3 chalcogenide to 1 wt% Pt/LaCoO3/K/Al2O3 on NSR performance. Their results showed that a maximum NSC of 44.8 μmol/g could be achieved at 400 °C. Furthermore, Xie et al. [52] tested the catalytic performance of LaCoO3-Meso with a high surface area in the NSR reaction. LaCoO3-Meso demonstrated a maximum NSC of 124 μmol/g at 300 °C. Although this is only a rough comparison of the data, it still demonstrates the superiority of the LS5 catalyst in terms of NOx adsorption/storage capacity.
Based on the above studies, it was observed that the LS5 catalyst demonstrates remarkable performance in NOx storage. Therefore, an investigation into its NOx storage capabilities at varying adsorption temperatures (200–400 °C) was conducted. The findings are presented in Figure 9. Initially, the LS5 catalyst completely traps NO, resulting in minimal NOx levels. However, as the reaction time increases, the concentration of NOx gradually rises until it reaches a certain threshold. Based on the NOx concentration at the outlet, the conversion rate of NO can be calculated and is shown in Table 5: 300 °C (63%) > 250 °C = 350 °C = 400 °C (42%) > 200 °C (30%). Analysis of the NOx desorption curve reveals that catalyst desorption follows a one-step process at different temperatures. Moreover, the desorption peak of the catalysts shifts to higher temperatures with increasing temperature. This suggests that elevating the temperature enhances the stability of nitrate.
The NAC and NSC of the LS5 catalysts were presented in Table 5. The catalyst’s NAC and NSC reach their maximum values at 300 °C. However, as the adsorption temperature increases, the capacity for NOx adsorption/storage decreases due to the exothermic nature of the process [48].
To investigate the impact of hydrothermal aging and sulfur on the NOx storage performance of LS0 and LS5 catalysts, experiments were performed to analyze NOx adsorption and desorption at 300 °C. Table 6 and Table 7 summarize the NOx adsorption capacity and RNO2 of LS0 and LS5 catalysts. The results reveal that the presence of water reduces the catalyst’s acidity after hydrothermal aging, leading to decreased efficiency in NO to NO2 conversion and a subsequent loss in NOx storage performance. However, perovskite promotes Sr dispersion and stability during hydrothermal aging, mitigating negative effects on the catalyst. The presence of impurities SrCoOx and SrCO3 also affected the NOx storage performance of the catalysts. It is hypothesized that hydrothermal aging increases the likelihood of SrCoOx and SrCO3 production, which has a reverse effect on catalysis. Moreover, the smaller decrease in NAC of Sr-doped catalysts suggests that Sr doping enhances the hydrothermal resistance of the catalysts. Table 7 depicts the effect of sulfur on the LS0 and LS5 catalysts. The NAC decrease rate of the LS5 sample is lower than that of the LS0 sample. This may be because Sr completely enters the perovskite lattice in the LS0 catalyst. After the introduction of SO2, SO2 directly poisons perovskite. However, in the LS5 catalyst, Sr exists in the form of perovskite and SrCO3, and after the introduction of SO2, SO2 easily reacts with SrCO3. This weakens the toxicity of the perovskite structure, resulting in a small decrease in the NAC of the catalysts. This indicates that Sr doping enhances the sulfur-resistant performance of the LS5 catalyst. Additionally, Sr doping increases the specific surface area, which supports better dispersion of the storage sites. Therefore, the small amount of sulphate species generated by the reaction of SO2 with SrCO3 is not sufficient to completely inactivate the NOx storage sites of the catalyst. This is another reason for the superior NOx storage performance of the sulfided LS5 chalcogenide catalyst.
Overall, the LS5 catalyst exhibits a wide operating temperature range (200–400 °C), with higher NAC (750–1899 μmol/g) and NSC (80–1084 μmol/g), which indicates its excellent ability to capture NOx at low to medium temperatures. In addition, the LS5 catalyst also exhibits excellent stability after hydrothermal and sulfur poisoning. Consequently, LS5 can be considered a promising material for NOx storage.

3.9. IR Study of NOx Storage

In order to determine the storage species of the samples, FTIR analysis was conducted on the catalysts after NOx storage. The results, depicted in Figure 10, reveal that the predominant adsorbed species on the LSX surface are primarily bulk nitrate (736 cm−1, 814 cm−1), nitrate ions (1360 cm−1, 1384 cm−1) [53], and bidentate nitrate (1502 cm−1) (1502 cm−1) [54,55]. Additionally, the characteristic peak at 580 cm−1 corresponds to the perovskite structure, while peaks at 667, 858, and 1451 cm−1 indicate the presence of carbonate species. Lastly, the peak at 1633 cm−1 is attributed to H2O. Overall, the NOx storage on LS0 and LS5 catalysts follows the “nitrate” route, aligning with the findings of the NOx desorption stage (Figure 8 and Figure 9).
As illustrated in Figure 10a, the LS0 catalyst exhibits nitrate ions and bidentate nitrate species as the stored NOx, with the storage phase consisting solely of the perovskite phase. On the other hand, the LS5 catalyst displays stored NOx in the form of nitrate ions and bulk nitrate species. Notably, the peak of carbonate diminishes while, the peak of bulk nitrate species emerges after storage, suggesting a possible conversion from carbonate to bulk nitrate. Moreover, the storage phase consists of perovskite and strontium carbonate following Sr doping [56]. Consequently, it can be inferred that the LS5 catalyst demonstrates favorable NAC and NSC, with the presence of SrCO3 contributing to this effect.
Based on Figure 10b, LS5 adsorbs NOx at different storage temperatures and stores it as bulk nitrates and nitrate ions, indicating that the storage phase encompasses both perovskite and strontium carbonate. The intensity of the nitrate species peaks increases as the temperature rises from 200 °C to 300 °C. However, beyond 350 °C, the intensity diminishes, suggesting a decline in the storage phase’s effectiveness at this temperature. Notably, the catalyst exhibits the strongest infrared peak of nitrate at 300 °C, aligning with the LS5’s maximum adsorption–desorption capacity at this temperature.

3.10. NOx Storage and Reduction Performance

Figure 11 displays the results of lean–rich cycling tests performed on the LS5 sample at different temperatures. During the lean periods, NO undergoes oxidation to NO2 and is subsequently stored as nitrates. In the rich periods, the stored nitrates decompose and release NOx [31]. The released NOx is reduced by C3H8. At different reaction temperatures, there is minimal release of NOx in the first cycle, indicating complete reduction of NOx by C3H8. However, as the number of cycles increases, the concentration of NOx gradually rises while the conversion of NOx decreases. This can be attributed to the stable storage of NOx, which occupies the storage sites, and the partial deactivation of active sites on the catalysts, resulting in lower reduction efficiency.
Figure 12 displays the average NOx conversion of the LS5 catalyst at different temperatures. As shown, the catalyst achieves an average NOx removal rate of over 52% within the temperature range of 200–300 °C, with maximum NOx removal of 61% observed at 250 °C. Subsequently, as the reaction temperature increases, the average NOx conversion decreases. These findings align with the adsorption and desorption behavior of NOx. It is likely that the positive correlation between temperature and nitrate stability hampers the decomposition of nitrate and catalyst regeneration, thereby reducing the average NOx removal rate. In conclusion, when utilizing C3H8 as the reducing gas, the LS5 catalyst exhibits a superior average NOx conversion rate at low temperatures.

4. Conclusions

In summary, the perovskite La1−xSrxCo1−λO3−δ nanomaterials were successfully synthesized using the solution combustion method. The incorporation of Sr strongly influences the crystalline phase, textural properties, desorption of surface oxygen species, and redox performance of the perovskite catalysts. As a result, it promotes the NO oxidation, NOx adsorption and storage, and NOx reduction. In addition, the incorporation of Sr enables the catalysts to exhibit superior resistance to hydrothermal aging and sulfur. The adsorption of NOx over LSX catalysts at 300 °C occurs predominantly via the “nitrate route”. Substituting 50% of La3+ with Sr2+ doubled the NOx adsorption capacity of perovskite LaCoO3 at 300 °C, which could be attributed to the large surface area, strong reducibility, more surface oxygen species, and elevated surface Co2+ concentration. During the lean–rich cycling experiments, the LS5 catalyst achieved an average NOx conversion rate of over 50% in the temperature range of 200–300 °C, with a maximum NOx conversion of 61% recorded at 250 °C. Thus, Sr-doped LaCoO3 perovskite catalysts prepared using the solution combustion method demonstrate outstanding NOx storage performance at low–medium temperatures, making them a potential material for De-NOx applications.

Author Contributions

Conceptualization: Z.W. and L.G.; Methodology: Z.W., W.L. and L.W.; Formal analysis and investigation: X.L., X.W., T.Z., J.L. and Y.Z.; Writing—original draft: X.L.; Writing—review and editing: Z.W.; Funding acquisition: Z.W.; Resources: Z.W.; Supervision: Z.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Natural Science Foundation of Shandong Province (Nos. ZR2023MB100, ZR2021MB063, ZR2020MB120), National Natural Science Foundation of China (No. 21777055), and Innovation ability improvement project of technology-based small and medium-sized enterprises in Shandong Province (Nos. 2022TSGC2043, 2021TSGC1358).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Epling, W.S.; Parks, J.E.; Campbell, G.C.; Yezerets, A.; Currier, N.W.; Campbell, L.E. Further evidence of multiple NOx sorption sites on NOx storage/reduction catalysts. Catal. Today 2004, 96, 21–30. [Google Scholar] [CrossRef]
  2. Pereda-Ayo, B.; De La Torre, U.; Illán-Gómez, M.J.; Bueno-López, A.; González-Velasco, J.R. Role of the different copper species on the activity of Cu/zeolite catalysts for SCR of NOx with NH3. Appl. Catal. B Environ. 2014, 147, 420–428. [Google Scholar] [CrossRef]
  3. Salasc, S.; Skoglundh, M.; Fridell, E. A comparison between Pt and Pd in NOx storage catalysts. Appl. Catal. B Environ. 2002, 36, 145–160. [Google Scholar] [CrossRef]
  4. Castoldi, L.; Matarrese, R.; Morandi, S.; Righini, L.; Lietti, L. New insights on the adsorption, thermal decomposition and reduction of NOx over Pt- and Ba-based catalysts. Appl. Catal. B Environ. 2018, 224, 249–263. [Google Scholar] [CrossRef]
  5. Ecker, S.I.; Dornseiffer, J.; Werner, J.; Schlenz, H.; Sohn, Y.J.; Sauerwein, F.S.; Baumann, S.; Bouwmeester, H.J.M.; Guillon, O.; Weirich, T.E.; et al. Novel low-temperature lean NOx storage materials based on La0.5Sr0.5Fe1−xMxO3−δ/Al2O3 infiltration composites (M = Ti, Zr, Nb). Appl. Catal. B Environ. 2021, 286, 119919. [Google Scholar] [CrossRef]
  6. You, R.; Zhang, Y.; Liu, D.; Meng, M.; Jiang, Z.; Zhang, S.; Huang, Y. A series of ceria supported lean-burn NOx trap catalysts LaCoO3/K2CO3/CeO2 using perovskite as active component. Chem. Eng. J. 2015, 260, 357–367. [Google Scholar] [CrossRef]
  7. Onrubia-Calvo, J.A.; Pereda-Ayo, B.; De-La-Torre, U.; Gonzalez-Velasco, J.R. Strontium doping and impregnation onto alumina improve the NOx storage and reduction capacity of LaCoO3 perovskites. Catal. Today 2019, 333, 208–218. [Google Scholar] [CrossRef]
  8. Pereda-Ayo, B.; De La Torre, U.; Pilar Gonzalez-Marcos, M.; Gonzalez-Velasco, J.R. Influence of ceria loading on the NOx storage and reduction performance of model Pt-Ba/Al2O3 NSR catalyst. Catal. Today 2015, 241, 133–142. [Google Scholar] [CrossRef]
  9. Mráček, D.; Kočí, P.; Choi, J.-S.; Partridge, W.P. New operation strategy for driving the selectivity of NOx reduction to N2, NH3 or N2O during lean/rich cycling of a lean NOx trap catalyst. Appl. Catal. B Environ. 2016, 182, 109–114. [Google Scholar] [CrossRef]
  10. Fernandes, D.M.; Scofield, C.F.; Neto, A.A.; Cardoso, M.J.B.; Zotin, F.M.Z. Thermal deactivation of Pt/Rh commercial automotive catalysts. Chem. Eng. J. 2010, 160, 85–92. [Google Scholar] [CrossRef]
  11. Peng, Y.; Si, W.; Luo, J.; Su, W.; Chang, H.; Li, J.; Hao, J.; Crittenden, J. Surface Tuning of La0.5Sr0.5CoO3 Perovskite Catalysts by Acetic Acid for NOx Storage and Reduction. Environ. Sci. Technol. 2016, 50, 6442–6448. [Google Scholar] [CrossRef]
  12. Uran, L.; Gallego, J.; Li, W.Y.; Santamaria, A. Effect of catalyst preparation for the simultaneous removal of soot and NOX. Appl. Catal. A Gen. 2019, 569, 157–169. [Google Scholar] [CrossRef]
  13. Ueda, A.; Yamada, Y.; Katsuki, M.; Kiyobayashi, T.; Xu, Q.; Kuriyama, N. Perovskite catalyst (La, Ba)(Fe, Nb, Pd)O3 applicable to NOx storage and reduction system. Catal. Commun. 2009, 11, 34–37. [Google Scholar] [CrossRef]
  14. Panunzi, A.P.; Duranti, L.; Luisetto, I.; Lisi, N.; Marelli, M.; Di Bartolomeo, E. Triggering electrode multi-catalytic activity for reversible symmetric solid oxide cells by Pt-doping lanthanum strontium ferrite. Chem. Eng. J. 2023, 471, 144448. [Google Scholar] [CrossRef]
  15. Ji, K.; Dai, H.; Deng, J.; Jiang, H.; Zhang, L.; Zhang, H.; Cao, Y. Catalytic removal of toluene over three-dimensionally ordered macroporous Eu1−xSrxFeO3. Chem. Eng. J. 2013, 214, 262–271. [Google Scholar] [CrossRef]
  16. Onrubia-Calvo, J.A.; Pereda-Ayo, B.; De-La-Torre, U.; González-Velasco, J.R. Key factors in Sr-doped LaBO3 (B = Co or Mn) perovskites for NO oxidation in efficient diesel exhaust purification. Appl. Catal. B Environ. 2017, 213, 198–210. [Google Scholar] [CrossRef]
  17. Kim, C.H.; Qi, G.; Dahlberg, K.; Li, W. Strontium-doped perovskites rival platinum catalysts for treating NOx in simulated diesel exhaust. Science 2010, 327, 1624–1627. [Google Scholar] [CrossRef]
  18. Say, Z.; Mihai, O.; Kurt, M.; Olsson, L.; Ozensoy, E. Trade-off between NOx storage capacity and sulfur tolerance on Al2O3/ZrO2/TiO2–based DeNOx catalysts. Catal. Today 2019, 320, 152–164. [Google Scholar] [CrossRef]
  19. Say, Z.; Vovk, E.I.; Bukhtiyarov, V.I.; Ozensoy, E. Enhanced Sulfur Tolerance of Ceria-Promoted NOx Storage Reduction (NSR) Catalysts: Sulfur Uptake, Thermal Regeneration and Reduction with H2(g). Top. Catal. 2013, 56, 950–957. [Google Scholar] [CrossRef]
  20. Theis, J.R. An assessment of Pt and Pd model catalysts for low temperature NOx adsorption. Catal. Today 2016, 267, 93–109. [Google Scholar] [CrossRef]
  21. Elbouazzaoui, S.; Corbos, E.C.; Courtois, X.; Marecot, P.; Duprez, D. A study of the deactivation by sulfur and regeneration of a model NSR Pt/Ba/Al2O3 catalyst. Appl. Catal. B Environ. 2005, 61, 236–243. [Google Scholar] [CrossRef]
  22. Varma, A.; Mukasyan, A.S.; Rogachev, A.S.; Manukyan, K.V. Solution Combustion Synthesis of Nanoscale Materials. Chem. Rev. 2016, 116, 14493–14586. [Google Scholar] [CrossRef]
  23. Park, J.-H.; Cho, H.J.; Park, S.J.; Nam, I.-S.; Yeo, G.K.; Kil, J.K.; Youn, Y.K. Role of cobalt on γ-Al2O3 based NOx storage catalyst. Top. Catal. 2007, 42, 61–64. [Google Scholar] [CrossRef]
  24. Ao, M.; Pham, G.H.; Sage, V.; Pareek, V. Structure and activity of strontium substituted LaCoO3 perovskite catalysts for syngas conversion. J. Mol. Catal. A Chem. 2016, 416, 96–104. [Google Scholar] [CrossRef]
  25. Hueso, J.L.; Holgado, J.P.; Pereñíguez, R.; Mun, S.; Salmeron, M.; Caballero, A. Chemical and electronic characterization of cobalt in a lanthanum perovskite. Effects of strontium substitution. J. Solid State Chem. 2010, 183, 27–32. [Google Scholar] [CrossRef]
  26. Villoria, J.A.; Alvarez-Galvan, M.C.; Al-Zahrani, S.M.; Palmisano, P.; Specchia, S.; Specchia, V.; Fierro, J.L.G.; Navarro, R.M. Oxidative reforming of diesel fuel over LaCoO3 perovskite derived catalysts: Influence of perovskite synthesis method on catalyst properties and performance. Appl. Catal. B Environ. 2011, 105, 276–288. [Google Scholar] [CrossRef]
  27. Specchia, S.; Galletti, C.; Specchia, V. Solution Combustion Synthesis as intriguing technique to quickly produce performing catalysts for specific applications. In Studies in Surface Science and Catalysis; Gaigneaux, E.M., Devillers, M., Hermans, S., Jacobs, P.A., Martens, J.A., Ruiz, P., Eds.; Elsevier: Amsterdam, The Netherlands, 2010; pp. 59–67. [Google Scholar]
  28. Feng, N.; Meng, J.; Wu, Y.; Chen, C.; Wang, L.; Gao, L.; Wan, H.; Guan, G. KNO3 supported on three-dimensionally ordered macroporous La0.8Ce0.2Mn1−xFexO3 for soot removal. Catal. Sci. Technol. 2016, 6, 2930–2941. [Google Scholar] [CrossRef]
  29. Groen, J.C.; Peffer, L.A.A.; Pérez-Ramírez, J. Pore size determination in modified micro- and mesoporous materials. Pitfalls and limitations in gas adsorption data analysis. Microporous Mesoporous Mater. 2003, 60, 1–17. [Google Scholar] [CrossRef]
  30. Wang, Y.; Zhu, Y.; Liu, S.; Zhang, R. Pore characterization and its impact on methane adsorption capacity for organic-rich marine shales. Fuel 2016, 181, 227–237. [Google Scholar] [CrossRef]
  31. Cui, C.; Ma, J.; Wang, Z.; Liu, W.; Liu, W.; Wang, L. High Performance of Mn-Doped MgAlOx Mixed Oxides for Low Temperature NOx Storage and Release. Catalysts 2019, 9, 677. [Google Scholar] [CrossRef]
  32. Liang, H.; Mou, Y.; Zhang, H.; Li, S.; Yao, C.; Yu, X. Sulfur resistance and soot combustion for La0.8K0.2Co1−yMnyO3 catalyst. Catal. Today 2017, 281, 477–481. [Google Scholar] [CrossRef]
  33. Montanari, T.; Castoldi, L.; Lietti, L.; Busca, G. Basic catalysis and catalysis assisted by basicity: FT-IR and TPD characterization of potassium-doped alumina. Appl. Catal. A Gen. 2011, 400, 61–69. [Google Scholar] [CrossRef]
  34. Li, Q.; Wang, X.; Xin, Y.; Zhang, Z.; Zhang, Y.; Hao, C.; Meng, M.; Zheng, L.; Zheng, L. A unified intermediate and mechanism for soot combustion on potassium-supported oxides. Sci. Rep. 2014, 4, 4725. [Google Scholar] [CrossRef] [PubMed]
  35. Morales, M.; Segarra, M. Steam reforming and oxidative steam reforming of ethanol over La0.6Sr0.4CoO3-delta perovskite as catalyst precursor for hydrogen production. Appl. Catal. A Gen. 2015, 502, 305–311. [Google Scholar] [CrossRef]
  36. Li, Z.; Meng, M.; Zha, Y.; Dai, F.; Hu, T.; Xie, Y.; Zhang, J. Highly efficient multifunctional dually-substituted perovskite catalysts La1−xKxCo1−yCuyO3−δ used for soot combustion, NOx storage and simultaneous NOx-soot removal. Appl. Catal. B Environ. 2012, 121–122, 65–74. [Google Scholar] [CrossRef]
  37. Stangeland, K.; Kalai, D.Y.; Ding, Y.; Yu, Z.X. Mesoporous manganese-cobalt oxide spinel catalysts for CO2 hydrogenation to methanol. J. CO2 Util. 2019, 32, 146–154. [Google Scholar] [CrossRef]
  38. Wang, X.; Zuo, J.; Luo, Y.; Jiang, L. New route to CeO2/LaCoO3 with high oxygen mobility for total benzene oxidation. Appl. Surf. Sci. 2017, 396, 95–101. [Google Scholar] [CrossRef]
  39. Crumlin, E.J.; Mutoro, E.; Liu, Z.; Grass, M.E.; Biegalski, M.D.; Lee, Y.-L.; Morgan, D.; Christen, H.M.; Bluhm, H.; Shao-Horn, Y. Surface strontium enrichment on highly active perovskites for oxygen electrocatalysis in solid oxide fuel cells. Energy Environ. Sci. 2012, 5, 6081–6088. [Google Scholar] [CrossRef]
  40. Bai, B.; Arandiyan, H.; Li, J. Comparison of the performance for oxidation of formaldehyde on nano-Co3O4, 2D-Co3O4, and 3D-Co3O4 catalysts. Appl. Catal. B Environ. 2013, 142, 677–683. [Google Scholar] [CrossRef]
  41. Zhang, B.; Yu, C.; Li, Z. Enhancing the Electrochemical Properties of LaCoO(3)by Sr-Doping, rGO-Compounding with Rational Design for Energy Storage Device. Nanoscale Res. Lett. 2020, 15, 184. [Google Scholar] [CrossRef]
  42. Arandiyan, H.; Dai, H.; Deng, J.; Liu, Y.; Bai, B.; Wang, Y.; Li, X.; Xie, S.; Li, J. Three-dimensionally ordered macroporous La0.6Sr0.4MnO3 with high surface areas: Active catalysts for the combustion of methane. J. Catal. 2013, 307, 327–339. [Google Scholar] [CrossRef]
  43. Akay, G. Plasma Generating-Chemical Looping Catalyst Synthesis by Microwave Plasma Shock for Nitrogen Fixation from Air and Hydrogen Production from Water for Agriculture and Energy Technologies in Global Warming Prevention. Catalysts 2020, 10, 152. [Google Scholar] [CrossRef]
  44. Akay, G. Hydrogen, Ammonia and Symbiotic/Smart Fertilizer Production Using Renewable Feedstock and CO2 Utilization through Catalytic Processes and Nonthermal Plasma with Novel Catalysts and In Situ Reactive Separation: A Roadmap for Sustainable and Innovation-Based Technology. Catalysts 2023, 13, 1287. [Google Scholar] [CrossRef]
  45. Zakaryan, H.A.; Aroutiounian, V.M. Investigation of cobalt doped tin dioxide structure and defects: Density functional theory and empirical force fields. J. Contemp. Phys. Armen. Acad. Sci. 2017, 52, 227–233. [Google Scholar] [CrossRef]
  46. Catto, A.C.; da Silva, L.F.; Bernardi, M.I.B.; Bernardini, S.; Aguir, K.; Longo, E.; Mastelaro, V.R. Local Structure and Surface Properties of CoxZn1−xO Thin Films for Ozone Gas Sensing. ACS Appl. Mater. Interfaces 2016, 8, 26066–26072. [Google Scholar] [CrossRef] [PubMed]
  47. Zhang, Y.; Yu, Y.; Shan, W.; Lian, Z.; He, H. Effect of support preparation with different concentration precipitant on the NOx storage performance of Pt/BaO/CeO2 catalysts. Catal. Today 2020, 339, 135–147. [Google Scholar] [CrossRef]
  48. Liu, W.; Cao, H.; Wang, Z.; Cui, C.; Gan, L.; Liu, W.; Wang, L. A novel ceria hollow nanosphere catalyst for low temperature NOx storage. J. Rare Earths 2022, 40, 626–635. [Google Scholar] [CrossRef]
  49. Fridell, E.; Persson, H.; Westerberg, B.; Olsson, L.; Skoglundh, M. The mechanism for NOx storage. Catal. Lett. 2000, 66, 71–74. [Google Scholar] [CrossRef]
  50. Wang, X.Y.; Qi, X.X.; Chen, Z.L.; Jiang, L.L.; Wang, R.H.; Wei, K.M. Studies on SO2 Tolerance and Regeneration over Perovskite-Type LaCo1−xPtxO3 in NOx Storage and Reduction. J. Phys. Chem. C 2014, 118, 13743–13751. [Google Scholar] [CrossRef]
  51. Wen, W.; Wang, X.Y.; Jin, S.; Wang, R.H. LaCoO3 perovskite in Pt/LaCoO3/K/Al2O3 for the improvement of NOx storage and reduction performances. RSC Adv. 2016, 6, 74046–74052. [Google Scholar] [CrossRef]
  52. 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. [Google Scholar] [CrossRef]
  53. Xian, H.; Zhang, X.; Li, X.; Li, L.; Zou, H.; Meng, M.; Li, Q.; Tan, Y.; Tsubaki, N. BaFeO3−x Perovskite: An Efficient NOx Absorber with a High Sulfur Tolerance. J. Phys. Chem. C 2010, 114, 11844–11852. [Google Scholar] [CrossRef]
  54. You, R.; Zhang, Y.; Liu, D.; Meng, M.; Zheng, L.; Zhang, J.; Hu, T. YCeZrO Ternary Oxide Solid Solution Supported Nonplatinic Lean-Burn NOx Trap Catalysts Using LaCoO3 Perovskite as Active Phase. J. Phys. Chem. C 2014, 118, 25403–25420. [Google Scholar] [CrossRef]
  55. Hadjiivanov, K. Identification of neutral and charged NxOy surface species by IR spectroscopy. Catal Rev. Sci. Eng. 2000, 42, 144–171. [Google Scholar] [CrossRef]
  56. Li, X.-G.; Dong, Y.-H.; Xian, H.; Hernandez, W.Y.; Meng, M.; Zou, H.-H.; Ma, A.-J.; Zhang, T.-Y.; Jiang, Z.; Tsubaki, N.; et al. De-NOx in alternative lean/rich atmospheres on La1−xSrxCoO3 perovskites. Energy Environ. Sci. 2011, 4, 3351–3354. [Google Scholar] [CrossRef]
Figure 1. (a) XRD and (b) the peak of crystal plane (110) of LSX catalysts.
Figure 1. (a) XRD and (b) the peak of crystal plane (110) of LSX catalysts.
Compounds 04 00014 g001
Figure 2. SEM and EDS mapping images of (a,c,d) LS0 and (b,e,f) LS5 catalysts.
Figure 2. SEM and EDS mapping images of (a,c,d) LS0 and (b,e,f) LS5 catalysts.
Compounds 04 00014 g002aCompounds 04 00014 g002b
Figure 3. (a) N2 adsorption–desorption curves and (b) pore size distributions of LSX catalysts.
Figure 3. (a) N2 adsorption–desorption curves and (b) pore size distributions of LSX catalysts.
Compounds 04 00014 g003
Figure 4. (a) FTIR spectra and (b) H2-TPR profiles of the LSX catalysts.
Figure 4. (a) FTIR spectra and (b) H2-TPR profiles of the LSX catalysts.
Compounds 04 00014 g004
Figure 5. O2-TPD spectra of the LSX catalysts.
Figure 5. O2-TPD spectra of the LSX catalysts.
Compounds 04 00014 g005
Figure 6. Sr 3d, Co 2p, O 1s XPS spectra of the LSX catalysts.
Figure 6. Sr 3d, Co 2p, O 1s XPS spectra of the LSX catalysts.
Compounds 04 00014 g006
Figure 7. Oxidation curves of NO on the LSX catalysts.
Figure 7. Oxidation curves of NO on the LSX catalysts.
Compounds 04 00014 g007
Figure 8. (a) NOx absorption curves and (b) desorption curves (at 300 °C) of the LSX catalysts.
Figure 8. (a) NOx absorption curves and (b) desorption curves (at 300 °C) of the LSX catalysts.
Compounds 04 00014 g008
Figure 9. NOx absorption–desorption curves of the LS5 catalyst at different temperatures (200–400 °C).
Figure 9. NOx absorption–desorption curves of the LS5 catalyst at different temperatures (200–400 °C).
Compounds 04 00014 g009
Figure 10. (a) IR spectra of the LSX catalysts before and after NOx storage at 300 °C; (b) IR spectra of LS5 at different NOx storage temperatures.
Figure 10. (a) IR spectra of the LSX catalysts before and after NOx storage at 300 °C; (b) IR spectra of LS5 at different NOx storage temperatures.
Compounds 04 00014 g010
Figure 11. NOx concentration curves of the LS5 catalyst during the NSR process at different temperatures (200–350 °C).
Figure 11. NOx concentration curves of the LS5 catalyst during the NSR process at different temperatures (200–350 °C).
Compounds 04 00014 g011
Figure 12. Average NOx conversion of the LS5 catalyst at different temperatures during lean–rich cycling experiments.
Figure 12. Average NOx conversion of the LS5 catalyst at different temperatures during lean–rich cycling experiments.
Compounds 04 00014 g012
Table 1. Textural parameters for SBET, Vp, and Dp of the LSX catalysts.
Table 1. Textural parameters for SBET, Vp, and Dp of the LSX catalysts.
CatalystSBET
(m2/g)
Vp
(cm3/g)
Dp
(nm)
Mesopore Fraction(La + Sr)/Co
LS05.580.05410.361.20
LS112.380.08320.66-
LS313.440.13290.76-
LS517.710.11220.751.09
LS715.960.09230.70-
(La + Sr)/Co: EDS results; Mesopore Fraction: Percentage of mesopore to total pore; Dp: Average Pore diameter.
Table 2. H2-TPR and O2-TPD results.
Table 2. H2-TPR and O2-TPD results.
CatalystH2 Consumption
(mmol/g)
Peak Area
α-Peakβ-Peak
LS05.7100-
LS16.217739
LS38.1194369
LS57.5219578
LS76.9172668
Table 3. XPS results.
Table 3. XPS results.
CatalystOII/(OI + OII + OIII)
(%)
Co2+/Co3+(La + Sr)/CoSrsurf/Srlatt
LS057.61.132.33-
LS368.61.411.281.23
LS565.81.431.051.6
OI: lattice oxygen; OII: surface-active oxygen; OIII: Chemical adsorption of oxygen in water; Srlatt: Sr lattice components; Srsurf: Sr surface components.
Table 4. NAC and NSC of the LSX catalysts at 300 °C.
Table 4. NAC and NSC of the LSX catalysts at 300 °C.
CatalystsNAC
(μmol/g)
NSC
(μmol/g)
LS083123
LS186548
LS31485635
LS518891084
LS71556810
Table 5. NAC, NSC, and RNO2 at different adsorption temperatures on the LS5 catalyst.
Table 5. NAC, NSC, and RNO2 at different adsorption temperatures on the LS5 catalyst.
Adsorption Temperature
(°C)
NAC
(μmol/g)
NSC
(μmol/g)
RNO2
(%)
20093760930
250129798442
3001889108463
3501633100742
4007508042
RNO2: NO to NO2 conversion after reaching a steady state during the adsorption process.
Table 6. Comparison of NAC and RNO2 after hydrothermal aging of LS0 and LS5 catalysts.
Table 6. Comparison of NAC and RNO2 after hydrothermal aging of LS0 and LS5 catalysts.
CatalystNAC (μmol/g)RNO2The Decline Rate of NAC
FreshHydrothermal AgingFreshHydrothermal Aging
LS0831461523344.5
LS518891262633833.2
Table 7. Comparison of NAC and RNO2 before and after sulfurization of LS0 and LS5 catalysts.
Table 7. Comparison of NAC and RNO2 before and after sulfurization of LS0 and LS5 catalysts.
CatalystNAC (μmol/g)RNO2The Decline Rate of NAC
FreshSulfur PoisoningFreshSulfur Poisoning
LS0831554521933.3
LS518891434632824.2
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

Luan, X.; Wang, X.; Zhang, T.; Gan, L.; Liu, J.; Zhai, Y.; Liu, W.; Wang, L.; Wang, Z. NOx Storage and Reduction (NSR) Performance of Sr-Doped LaCoO3 Perovskite Prepared by Glycine-Assisted Solution Combustion. Compounds 2024, 4, 268-287. https://doi.org/10.3390/compounds4020014

AMA Style

Luan X, Wang X, Zhang T, Gan L, Liu J, Zhai Y, Liu W, Wang L, Wang Z. NOx Storage and Reduction (NSR) Performance of Sr-Doped LaCoO3 Perovskite Prepared by Glycine-Assisted Solution Combustion. Compounds. 2024; 4(2):268-287. https://doi.org/10.3390/compounds4020014

Chicago/Turabian Style

Luan, Xinru, Xudong Wang, Tianfei Zhang, Liangran Gan, Jianxun Liu, Yujia Zhai, Wei Liu, Liguo Wang, and Zhongpeng Wang. 2024. "NOx Storage and Reduction (NSR) Performance of Sr-Doped LaCoO3 Perovskite Prepared by Glycine-Assisted Solution Combustion" Compounds 4, no. 2: 268-287. https://doi.org/10.3390/compounds4020014

Article Metrics

Back to TopTop