Next Article in Journal
One-Step Biodiesel Production from Waste Cooking Oil Using CaO Promoted Activated Carbon Catalyst from Prunus persica Seeds
Next Article in Special Issue
Impact of Hydrothermally Prepared Support on the Catalytic Properties of CuCe Oxide for Preferential CO Oxidation Reaction
Previous Article in Journal
Low-Temperature Selective NO Reduction by CO over Copper-Manganese Oxide Spinels
Previous Article in Special Issue
Advanced Synthesis and Characterization of Vanadia/Titania Catalysts through a Molecular Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances in Designing Efficient La-Based Perovskites for the NOx Storage and Reduction Process

1
Research Institute of Petroleum Processing, Sinopec, Beijing 100083, China
2
Collaborative Innovation Center of Chemical Science and Engineering (Tianjin), Tianjin Key Laboratory of Applied Catalysis Science and Technology, School of Chemical Engineering and Technology, Tianjin University, Tianjin 300072, China
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(6), 593; https://doi.org/10.3390/catal12060593
Submission received: 17 April 2022 / Revised: 20 May 2022 / Accepted: 23 May 2022 / Published: 30 May 2022

Abstract

:
To overcome the inherent challenge of NOx reduction in the net oxidizing environment of diesel engine exhaust, the NOx storage and reduction (NSR) concept was proposed in 1995, soon developed and commercialized as a promising DeNOx technique over the past two decades. Years of practice suggest that it is a tailor-made technique for light-duty diesel vehicles, with the advantage of being space saving, cost effective, and efficient in NOx abatement; however, the over-reliance of NSR catalysts on high loadings of Pt has always been the bottleneck for its wide application. There remains fervent interest in searching for efficient, economical, and durable alternatives. To date, La-based perovskites are the most explored promising candidate, showing prominent structural and thermal stability and redox property. The perovskite-type oxide structure enables the coupling of redox and storage centers with homogeneous distribution, which maximizes the contact area for NOx spillover and contributes to efficient NOx storage and reduction. Moreover, the wide range of possible cationic substitutions in perovskite generates great flexibility, yielding various formulations with interesting features desirable for the NSR process. Herein, this review provides an overview of the features and performances of La-based perovskite in NO oxidation, NOx storage, and NOx reduction, and in this way comprehensively evaluates its potential to substitute Pt and further improve the DeNOx efficiency of the current NSR catalyst. The fundamental structure–property relationships are summarized and highlighted to instruct rational catalyst design. The critical research needs and essential aspects in catalyst design, including poisoner resistance and catalyst sustainability, are finally addressed to inspire the future development of perovskite material for practical application.

Graphical Abstract

1. Introduction

The diesel engines operated under a high air-to-fuel (A/F) ratio in the range of 20 to 65 achieve higher fuel economy and better driving performance than the stoichiometric gasoline engines [1,2,3], but the higher combustion temperature and pressure in the diesel cylinder lead to increased NOx emissions, and the net oxidizing environment of the exhaust brings huge challenges for NOx reduction [2]. The conventional three-way catalysis (TWC) system developed for the removal of NOx in gasoline engine exhaust is not as efficient as the A/F ratio shifts towards the lean-burn region. Thereafter, three alternative DeNOx techniques were developed, that is, the direct decomposition of NO, selective catalytic reduction (SCR), and NOx storage and reduction (NSR) or lean NOx trap (LNT). NO decomposition in the absence of reductant was proposed as an ideal DeNOx concept with thermodynamic feasibility, but was demonstrated to be kinetically sluggish, showing insufficient NOx conversions that do not satisfy the requirements of real automotive emission control systems [4,5]. However, the other two techniques have been fully developed and commercialized over the last several decades.
An SCR technique with ammonia from urea decomposition as the reductant was proposed in 1979, which was demonstrated to be efficient and selective for NOx reduction to N2 in net oxidizing atmosphere, especially in the temperature range of 250–450 °C [6,7,8,9,10]. However, the complex hardware system (including heated tanks and tubes, injector nozzle, and mixing device, to ensure storage and complete decomposition of urea, homogeneous flow distribution, and accurate dosing of NH3) means that NH3-SCR is not a suitable NOx abatement technology for vehicles with space constraints and cost considerations [11]. The intrinsic risk of ammonium salt [(NH4)2SO4] deposition and NH3 slip also drove the exploration of novel NOx removal techniques. Several research organizations then focused on the selective reduction of NOx with hydrocarbons as the reductant in the 1990s, but this technique posed a couple of problems, including inadequate NOx conversion and a narrow temperature window [12,13]. In this context, the NSR technique was proposed by Toyota in 1995 and is characteristic of incorporating basic components in DeNOx catalyst formulation for NOx storage, adopting alternating lean-burn/fuel-rich operations, and not requiring external reductants [14]. In the lean-burn periods of a typical NSR process, NO is oxidized to NO2 and diffuses to the neighboring basic sites to form nitrates or nitrites species [15,16]. Then, the fuel or fuel-rich pulse is injected into the catalytic converter chamber to reduce the stored NOx to N2, thereby regenerating the occupied NOx storage sites and preparing them for the next lean-burn period [15]. Adopting this alternating operation is a smart innovation to avoid the inherent challenge of NOx reduction in the net oxidizing atmosphere.
An NSR system is ~50% space saving compared to NH3-SCR [11], which exhibits qualified NOx removal efficiency as well. NSR can also be applied to lean-burn gasoline engine vehicles or coupled with SCR, hence removing the urea injection system of SCR and making it more cost effective [17,18]. However, the application of NSR also suggests technical and cost concerns: (1) basic sites are irreversibly occupied by SO, so that long-term usage leads to continuous deactivation of the catalyst; and (2) high loadings of noble metals (mainly Pt) are demanded in commercial NSR catalysts to ensure sufficient NO oxidation and NOx reduction efficiency in frequently switching operation mode [11,16]. To date, the former constraint has been greatly alleviated with the execution of increasingly stringent legislation to reduce automotive SOx emissions, realized by the deep desulfurization of diesel by hydrotreating in refineries [19,20], while the challenges of high cost and low thermal stability of Pt have become the bottleneck for further application of NSR, and the attempts to find economical and durable alternatives have never stopped [11,16].
Perovskite-type oxide ABO3 has been proven to be a promising candidate, with the advantage of high thermal stability, prominent redox property, and low cost [21,22,23,24,25,26]. In a cubic perovskite-type oxide ABO3 with space group type Pm-3m, the A-site cation is 12-fold coordinated by oxygen anions, while the B-site cations are octahedrally coordinated (6-fold) [23]. A-site cations are large and generally basic, serving as NOx storage sites; and B-site cations are small and chemically active, functioning as the redox center for NO oxidation and NOx reduction [27,28,29,30]. The perovskite-type oxide structure of perovskite allows the integration of redox and storage components with homogeneous distribution and strong mutual interaction that substantially reduces the distance of NOx diffusion, which is desirable for achieving a high lean-burn NOx trapping rate and preventing fuel-rich NOx slip, as compared to other NSR catalysts that are typically composed of multiple active metals/oxides and basic components with limited interface [31].
The perovskite possesses excellent compositional versability, as it is able to accommodate nearly 90% of metals in the periodic table and form thermally stable oxides and able to endure partial cationic substitution, enabling flexible tuning of physio-chemical properties of the catalyst according to the reaction features [23,32,33,34]. Notably, any cation alternation or substitution can result in lattice expansion/shrinkage or a change in the extent of lattice distortion. For a specific formula, the selection of cations can refer to the tolerance factor criterion, that is, 0.8 < (rA + rO)/[√2(rB + rO)] < 1, which is reliable in most cases for predicting the structural stability of perovskite-type oxides [35]. Most of the research before 2010 investigated the perovskite simply as an NOx storage material, as the substitute for BaO in a commercial NSR formulation. Alkaline-earth-metal-based perovskites, such as BaCeO3, BaSnO3, BaFeO3, CaTiO3, and SrTiO3, were predominantly reported in this stage [36,37,38,39,40,41,42]. The main strategy was to use the perovskite-type oxide structure to modify sulfur- and sinter-resistant properties of the basic metals, whereas most of these formulations exhibited relatively inert redox properties and still required the addition of noble metals to facilitate NO oxidation and NOx reduction. Over the last decade, La has gradually been accepted as an ideal option for A-site cation after extensive formulation screening. It was found to provide excellent structural thermal stability [3,16]. Most importantly, the proper ion radius of La3+ favors the accommodation of the transitional metals with prominent redox properties, such as Mn and Co. With the development of these perovskite formulations, noble metals have appeared less indispensable in NSR formulation.
In this review, we summarize the recent advances of the perovskites (mainly La-based) applied in the NSR field. Apart from highlighting the promising catalyst formulations and activity results, the review provides insights into the structure–property relationship that would instruct the rational design of perovskite-based NSR catalysts. Moreover, we disassemble the complex NSR reaction process into three general steps, that is, NO oxidation, NOx storage/trapping, and NOx desorption and reduction, and separately review the performance of perovskite catalysts in each step to provide deep and comprehensive perspectives on the advantages and potential problems of La-based perovskite employed in NSR reaction. The critical research needs are finally addressed for further development of this type of material.

2. NO Oxidation over La-Based Perovskites

NO oxidation to NO2 as the first step of NOx storage substantially impacts the overall storage efficiency in lean-burn periods [43,44], especially at low temperatures, when kinetic limitations dominate NO oxidation. This necessitates the incorporation of active oxidizing components, such as noble metals and perovskites, in NSR catalyst formulation. Over the Pt-based catalysts, NO oxidation was reported to proceed following the Langmuir–Hinshelwood (L-H) and/or Eley–Rideal (E-R) mechanism [45,46,47]. Both mechanisms describe O2 adsorption and activation on noble metal clusters as the rate-limiting step of NO oxidation, but there is a divergence of opinion on whether the adsorbed NO on acid sites or gaseous NO takes part in the reaction. Weiss et al. [48,49] and Mulla et al. [50] pointed out that NO oxidation activity cannot be simply interpreted to be in inverse proportion to the cluster size of noble metal. Instead, a slight increase in cluster size decreased the binding energy of oxygen on coordinatively unsaturated Pt atoms, and thus increased the concentrations of oxygen vacancies and enhanced NO oxidation activity. However, long-term usage inevitably leads to the severe growth of Pt clusters and a dramatic activity decline. This motivates the later development of more thermally stable active centers for NO oxidation, such as perovskite.
Compared to noble metals, fewer mechanism studies have carried out on perovskites. Constantinou et al. [51] reported the reaction orders for NO2, NO, and O2 over LaMnO3 to be around −1, 1, and 1, respectively, in accordance with the data collected over Pd- and Pt-based catalysts [49,52], which might indicate that NO oxidation over LaMnO3 perovskite and noble metal catalysts share the same kinetic step. The negative reaction order of NO2 implies a product inhibition effect. NO2 strongly absorbs on NO oxidation sites with a high sticking coefficient [50,53,54,55,56], thereby inhibiting NO adsorption and oxidation. Considering the compositional versability of perovskite-based catalysts, the NO oxidation mechanism might vary as a function of the formulations. Besides L-H and E-R, the Mars−van Krevelen (MvK) mechanism might also function with lattice oxygen participating in the reaction.
As a general trend, perovskites possess comparable or even higher NO oxidation activity and much better durability than noble metal catalysts. He et al. [57] investigated the total substitution of Pt by LaCoO3 perovskite in an NSR formulation and obtained an enhanced NO-to-NO2 conversion of 38% over K2CO3/LaCoO3/ZrTiO4, compared to 18% over K2CO3/Pt/ZrTiO4. The promising NO oxidation activity of the perovskite-based formulation was attributed to the low electron density in the d-orbit of Co3+, which was beneficial for accepting a lone pair on the N atom in NO. Afterwards, Chen et al. [58] also reported that the high oxidation state of B-site cations would be favorable for NO oxidation and identified the lattice oxygen bonded to Mn4+ as the active species. In this study, a series of nonstoichiometric LaxMnO3 (0.9 < x < 1.11) catalysts were prepared. As the ratio of La/Mn decreased, the transformation of Mn3+ to Mn4+ was triggered to compensate for the charge imbalance caused by nonstoichiometry. As a concomitant of that, the NO-to-NO2 conversion was promoted. As revealed in these studies, the high oxidation states of B-site cations, such as Co3+ and Mn4+, can be stabilized by a perovskite-type structure, which is considered a major reason for the superior oxidation activity of perovskite compared to single-transition metal oxides [59].
However, the B-site oxidation state does not always correlate well with activity trends. As a typical example, Li and coworkers [60] studied NO oxidation over Sr-doped LaCoO3 and LaMnO3 perovskites. The transformation of Mn3+ to Mn4+ induced by low-valence Sr2+ doping did not lead to a notable activity enhancement in this case. In contrast, the NO conversion over the LaCoO3 increased from 69% to 86% at 300 °C after Sr doping. This was associated with the formation of oxygen vacancies in La0.9Sr0.1CoO3, on which the weakly adsorbed oxygen species were highly active for NO oxidation. These observations were confirmed in a recent study by Onrubia et al. [61], which looked deeply into the effects of Sr doping on NO oxidation activity (Figure 1). They revealed that the amount of oxygen vacancies in the La1−xSrxMnO3 did not further increase as Sr substitution reached 10%, and the charge imbalance caused by further Sr doping was compensated by Mn4+ formation, which, however, did not promote the NO oxidation activity. On the other hand, the generation of oxygen vacancy appeared to be the only mechanism for charge neutralization of La1−xSrxCoO3. This leads to their conclusion that the amount of oxygen vacancies is the principal factor governing the NO oxidation activity of perovskites. In addition, Sr doping causes notable increases in the specific surface areas (SSA) of LaMnO3 and LaCoO3 perovskites.
Sr doping also alleviates the [CoO6] octahedral distortion, and in some cases leads to the transition of the crystal structure of La-based perovskites from rhombohedral to cubic [62]. The Co3+ electronic configuration correspondingly switched from high/intermediate spin state to low spin state with empty eg filling. This favors the adsorption of electrophilic absorbates, and hence enhances NO oxidation.
Ag+ doping induced a more substantial charge imbalance than Sr2+ doping, which is beneficial to the formation of oxygen vacancies. With 10% of Ag doped into the A site of the LaMnO3 perovskite, the activation energy for NO oxidation declined from 66.7 ± 2.1 to 47.9 ± 1.8 kJ mol−1 [63]. However, the maximum substitution amount of Ag in LaMnO3 was only 20% at the A site.
Compared to A-site substitution, the alternation of B-site redox cations has a more direct impact on NO oxidation activity. Chen et al. [64] first investigated the activity of different B-site cations and provided the NO oxidation activity order of LaCoO3 > LaMnO3 > LaFeO3, in accordance with the reducibility of the B-site cation. The results suggest that the adsorbed NO may react with lattice oxygen species. However, further examination reveals the inadequacies of such a simple correlation. For example, LaNiO3 shows a higher reduction temperature than LaCoO3 in H2-TPR, but a lower NO-to-NO2 conversion [65]. In effect, when lattice oxygen participates in NO oxidation, this reaction is based on the cycling of reversible redox couples, such as Co3+/Co2+ and Mn4+/Mn3+, corresponding to the cyclic lattice oxygen storage and reaction process. Reducibility is apparently not enough to describe the cycling property, which is also determined by other factors, including the ability of the perovskite to host oxygen vacancies, filling rate of gaseous O2 to oxygen vacancies, and lattice oxygen mobility [61]. Furthermore, in some reported perovskite systems, the adsorbed oxygen species appear more active for NO oxidation than lattice oxygen. For example, Zhong et al. [65] prepared a series of Co-doped LaNiO3 perovskites by a co-precipitation approach. As a general trend, Co doping increased the NO and O2 adsorption amount, and the adsorption strength of O2. The highest NO oxidation activity was observed over the LaNi0.7Co0.3O3 catalyst, and this was ascribed to moderate adsorption strength and appropriate amount of adsorbed O2.
Besides transition metals, the perovskite structure allows the substitution of noble metals into the B site as well, but this is always limited to a trace amount due to cost considerations and ionic size disparity [66,67]. Pt doping generally promotes NO oxidation of perovskite since Pt itself is a highly active oxidation center. For example, LaCo0.92Pt0.08O3 shows a higher NO-to-NO2 conversion of 66.7% at 300 °C than the Pt-free counterpart of 42.0% [66], but the high Pt loading (~6.0 wt.%) makes the results less promising. Pd was demonstrated in our study to be less active than Pt in NO oxidation [67]. Only a slight increase in NO oxidation rate could be observed for Pd-doped La0.7Sr0.3CoO3 perovskite compared to the Pd-free counterpart in the kinetic regime, and the Pd-supported perovskite even showed a lower NO oxidation activity. The kinetics show that the incorporation of Pd in either way does not alter the NO oxidation pathway. Perovskite continues to be the dominant oxidation center.

3. NOx Storage

Upon NO being oxidized to NO2 and diffused to neighboring basic sites, NOx storage takes place on the catalyst surface, which is a unique behavior of NSR catalysts. Considering the essential impact of NOx storage efficiency in lean-burn periods on overall DeNOx activity, great efforts have been devoted to developing functional basic material that satisfies NOx storage capacity (NSC) in a wide temperature window. Most of the studies before the year 2010 focused on single oxides with varying basicity, morphology, and interaction with supports, especially on the Pt/BaO/Al2O3 model catalyst [11,16]. Perovskite accommodates basic metals in a complex oxide lattice with synergetic interaction with B-site cations, thus exhibiting new features for NOx storage, distinct from the conventional basic components, which will be highlighted in this section with Pt/BaO/Al2O3 as the benchmark.
Table 1 summarizes the NSC of La-based perovskite catalysts reported in the recent literature. Note that NSC cannot be simply described as an inherent attribute of the basic component, which is comprehensively affected by multiple factors [51]. As a general trend, NSC varies with the temperature in a volcanic shape, restricted by insufficient NO-to-NO2 conversions and sluggish NOx diffusion at low temperatures (<250 °C), and by the loss of thermal stability of nitrates in the high-temperature region (>450 °C) [68]. With prominent NO oxidation ability, and a negligible NOx diffusion barrier determined by the nature of the perovskite-type oxide structure, La-based perovskites generally exhibit higher NSC at low temperatures than the Pt/BaO/Al2O3 model catalyst, as observed over LaMn0.9Fe0.1O3 and alumina-supported La0.5Sr0.5Fe1−xTixO3-δ (Figure 2) [59,69], showing the potential to reduce the NOx slip during cold start of the diesel engine, a long-lasting headache for automotive NOx removal. Along this line, it should be mentioned that a prominent NO oxidation and NOx storage performance was recently observed over Sr-based perovskites, such as Mn, Fe, Co-doped SrTiO3 and SrFeO3−δ, and Ruddlesden−Popper (RP)-type layered Sr3Fe2O7−δ perovskite, in the low-temperature range [70,71,72], which was ascribed to the high mobility and reactivity of lattice oxygen. Among them, the Mn-doped SrTiO3 exhibits a unique NO adsorption behavior at 323 K, that is, NO is adsorbed on the perovskite surface using lattice oxygens without the occurrence of NO oxidation, and this enables the catalyst to show significant NOx storage capacity in a kinetically restricted temperature region for NO oxidation. The excellent NOx storage performance at low temperatures makes them not only promising NSR catalysts, but also potential passive NOx adsorber (PNA) materials.
At medium temperatures (250–450 °C), the NSC of Pt/BaO/Al2O3 substantially increases as a consequence of improved NO-to-NO2 conversions (Figure 2). In comparison, the NSC of La-based perovskites appears less promising, especially for the LaCoO3 and LaMnO3 without cationic substitution, the NSCs of which were reported to be lower than 100 μmol g−1 at 300 °C [73,74], as compared to ~140 μmol g−1 for Pt/BaO/Al2O3 with 20% of Ba loadings, reported in [59]. Say et al. [75] investigated the NOx adsorption behavior of these two perovskites using in situ infrared spectroscopy. The NOx adsorption yielded multiple types of N-bearing species, including monodentate nitrite and nitrate, bidentate nitrate, bridging nitrate, and a small amount of bridging/chelating nitrites, but the free nitrate ions were not detected, suggesting that only surface La sites in perovskite are functional for NOx storage. The inaccessibility of bulk storage sites can be ascribed to the high structural stability of La-based perovskite, which hinders the bulk phase transformation during NOx adsorption and desorption [40]. This constitutes the major difference between NOx trapping on perovskites and on single basic oxides, that is, the latter allows the inward NOx diffusion with phase transformation to nitrate [11]. The surface adsorption nature, to a large extent, accounts for the inadequate NSC of LaCoO3 and LaMnO3 perovskites at medium temperatures, even with the La-enriched surface [75], especially considering the poor texture property resulting from crystal growth of oxides at high calcination temperature. The surface areas of most proposed perovskites lie in the range of 5–20 m2 g−1.
Aiming to improve NSC, additional alkali or alkaline earth metals are always incorporated in the perovskite formulation. For La-based perovskites, Sr is a common choice, since the doped Sr not only provides a stronger NOx storage site on the surface, but also, as stated earlier, enhances NO oxidation activity and increases the surface area of the perovskite [61]. Our study [76] first investigated NOx storage on Sr-doped LaCoO3 perovskite through FT-IR and demonstrated three possible pathways (Figure 3): (1) NO2 combines with surface lattice oxygen and forms monodentate nitrate; (2) NO2 is directly trapped onto the oxygen vacancies and forms the “nitrate on perovskite” species; (3) NO2 is trapped by segregated SrCO3 and forms free nitrate ions. The presence of segregated SrCO3 greatly increases NSC, and the nitrates formed on SrCO3 show higher thermal stability than those on perovskite sites. After NOx storage, a surface enrichment of Sr was observed, indicating possible exsolution of Sr2+ from the perovskite lattice and migration to the catalyst surface owing to the interaction with NOx.
Dong et al. [77] substituted 30% of La with Sr and prepared La0.7Sr0.3MnO3 via a sol-gel procedure under optimized synthesis conditions. The calefactive velocity of 2 °C min−1, calcination temperature of 700 °C, and pH of 8 were identified to favor the formation of pure perovskite phase and reduce particle aggregation, yielding the best NO oxidation and NOx storage performance. Nonetheless, the best NSC of La0.7Sr0.3MnO3 was only 170.4 μmol g−1 at 350 °C (Table 1). As a comparison, the La0.7Sr0.3CoO3 catalysts prepared via the same procedure present NSC in the range of 700.0–967.3 μmol g−1 [67,76]. In effect, the type of B-site cations affects the existing state of Sr. Compared to Mn-based perovskite, Co-based perovskite is less prone to accommodate Sr into perovskite lattice. The segregated SrCO3 phase can always be detected in XRD patterns of La1−xSrxCoO3 [67,76].
Even though NSC is a significant descriptor for evaluating or predicting NOx storage efficiency in lean-burn periods of NSR reaction, it only reflects the quantity of storage sites, rather than the quality. Indeed, the NSR reaction does not even demand huge NSC, but enough “reversible NOx storage sites” that can rapidly trap NOx and be effectively regenerated [67]. In this sense, compared to those sites embedded inside bulk, the surface sites are more desirable, since they skip the sluggish NOx diffusion in the radial direction [31]. Along this line, efforts were devoted to exposing more surface storage sites of perovskite, and two strategies have proven effective.
(1) Preparing porous perovskite material. The construction of the porous structure of perovskite requires a hard or soft template. For example, Ye et al. [73] reported a nano-casting method for preparing mesoporous perovskite, with Santa Barbara Amorphous (SBA) as the hard template that was finally removed by NaOH solution. The as-prepared LaCoO3 shows a high SSA of 75.0 m2 g−1 and NSC of 252 µmol g−1, as compared to 93 µmol g−1 for the catalyst synthesized by the conventional sol-gel method. In a recent study, Xie et al. [78] also used the nano-casting method to synthesize mesoporous LaCoO3 with KIT-6 as the hard template. The as-prepared LaCoO3-meso exhibits an even higher SSA of 246 m2 g−1 and two times higher NSC compared to the counterpart prepared via the sol-gel method. More oxygen defects were produced on the surface of LaCoO3-meso with the confinement of the porous structure of the KIT-6 template, and this is revealed to improve oxygen mobility and favor fast NOx storage and desorption. The majority of NOx adsorbed on oxygen defects desorbed at low temperatures (<240 °C), indicating the weak basicity of such storage sites. Polymethyl methacrylate (PMMA) microspheres are the soft template generally used for synthesizing three-dimensionally ordered macroporous (3DOM) perovskite, and they can be completely removed by calcination. The as-prepared 3DOM perovskites are widely used for the simultaneous removal of soot and NOx, since the macropores (>50 nm) can accommodate soot particles (<25 nm). However, for the application of porous perovskite materials, the mechanical strength and cost issues should be considered [23]
(2) Dispersing perovskite on oxide supports. This strategy has a simpler preparation procedure and is more commonly used. The properties of the support, to a large extent, affect the NOx storage performance of perovskite. Supporting LaCoO3 on ZrTiO4 was found to be effective in improving NSC and reducing perovskite sintering [57]. Afterwards, You et al. [79,80] loaded 10% LaCoO3 perovskite on CeO2 and ternary (Y,Ce,Zr)O2 oxide supports. Despite the low loading, the catalyst achieved promising NO-to-NO2 conversions and NSC, which was attributed to the high dispersion of perovskite and abundant surface-adsorbed oxygen species on the Ce-based support. More recently, Onrubia-Calvo et al. [81,82,83] conducted a series of studies comprehensively investigating the La1−xSrxCoO3 perovskite supported on Al2O3, a cheap sinter-resistant support with a large surface area. At low loadings, Co preferably combined with Al2O3 and induced a solid phase transformation to CoAl2O4; high loadings benefited the formation of perovskite phase, but the less active bulk perovskite phase was also generated. Moreover, 30% perovskite loading on alumina was identified as an efficient formulation to balance dispersion and amount of active component (Figure 4), achieving a high NOx storage capacity of 306 μmol (g perovskite)−1 at 400 °C, compared to the bulky counterpart of 115 μmol (g perovskite)−1 [83]. The particle size of perovskite decreased from 50–150 nm to 20–50 nm in diameter after being supported on alumina. In a later study by Ecker et al. [59], alumina was again examined as support for perovskite. Fe was adopted as the B-site cation, partially doped by Nb, Ti, or Zr, to enhance sulfur resistance and phase stability. These catalysts, particularly La0.5Sr0.5Fe1−xTixO3-δ, exhibited promising hydrothermal stability, and the NSC decrease was less than 20% after aging at 750 °C. To date, the most promising results were reported by an early study conducted by Ding et al. [84], obtained by loading La0.7Sr0.3CoO3 on the mesoporous SiO2. The small crystalline size of perovskite confined in the mesoporous SiO2 increased the surface area. The as-prepared catalyst has an SSA of 237.6 m2 g−1, more than ten times higher than the perovskite prepared by the conventional sol-gel process (18.7 m2 g−1). The mesoporous SiO2-supported perovskite achieved an NSC of 5269.2 μmol g−1 in 3 h. Since NO oxidation activity was not substantially improved with the increase in SSA, in this case, the enhanced NSC was mainly related to the high dispersion of perovskite particles. Additionally, owing to the interactions between perovskite and mesoporous SiO2, the catalyst exhibited higher stability in the reducing atmosphere than perovskite prepared by the conventional sol-gel method.
At high temperatures (>450 °C), the storage sites on La-based perovskite surface are basically not functional, due to the relatively weak basicity, which is unable to stabilize nitrates [59]. This further necessitates the incorporation of strong basic components into perovskite-based catalyst formulation [81]. Besides the aforementioned Sr-containing perovskites, various formulations were proposed along this line, such as 5% K2CO3 impregnated on mesoporous LaCoO3 [73], Ba/Al2O3 ball-milled with LaMnO3 [85], and LaxBa1−xCoO3 prepared via sol-gel procedure [86,87,88]. The results suggest that the high-temperature NSR performance was enhanced in the presence of these strong basic components. In fact, the Ba-based perovskite exhibits better high-temperature NOx storage performance than the La-based counterparts, such as the BaTi0.8Cu0.2O3 proposed in Albaladejo-Fuente’s study [89]. However, the Ba-based perovskite shows lower chemical and structural stability, as indicated by the phase segregation and transformation during the NOx storage reaction [89]. Ba2+ in the A site is also less resistant to CO2 in the atmosphere than La3+ due to the strong basicity, and hence it is more prone to exsolution from perovskite matrix to form a surface BaCO3 layer [90]. These issues require special consideration in future studies, and the potential competitive or synergetic effect between different storage sites in NOx adsorption or desorption at various temperatures needs to be specifically investigated.
Table 1. Review of NOx storage capacity of the La-based perovskites.
Table 1. Review of NOx storage capacity of the La-based perovskites.
Catalyst Formulation (Preparation Method)Reaction ConditionsNOx Storage, μmol g−1
(Temperature, °C)
Ref.
K2O/LaCoO3/ZrTiO4 (impregnation)400 ppm NO/5% O2/N2; GHSV = 60,000 mL g−1 h−1325.0 (350)[57]
Pt/La0.5Sr0.5Fe0.5Ti0.5O3/Al2O3 (impregnation)500 ppm NO/500 ppm CO/200 ppm C3H6/9% H2O/9% CO2/6.5% O2; GHSV = 80,000 h−1117.0 (250)
150.0 (350)
50.0 (450)
[59]
La0.7Sr0.3CoO3 (sol-gel)400 ppm NO/ 5% O2/ balanced N2; GHSV = 120,000 mL g−1 h−1967.3 (300)[67]
LaMn0.9Fe0.1O3 (combustion)500 ppm NO/8% O2/N2; GHSV = 30,000 h−1371.6 (100)
392.3 (300)
306.3 (400)
[69]
LaCoO3 (sol-gel)400 ppm NO/5% O2/N2; GHSV = 80,000 mL g−1 h−193.0 (350)[73]
Mesoporous LaCoO3 (nano-casting)400 ppm NO/5% O2/N2; GHSV = 80,000 mL g−1 h−1981.0 (350)[73]
La0.7Sr0.3CoO3 (sol-gel)800 ppm NO/5% O2/ N2; GHSV = 80,000 mL g−1 h−1700.0 (300)[76]
La0.7Sr0.3MnO3 (sol-gel)800 ppm NO/5% O2/ N2; GHSV = 80,000 mL g−1 h−1170.4 (350)[77]
K2O/LaCoO3/(Y,Ce,Zr)O2 (impregnation)400 ppm NO/5% O2/N2; GHSV = 45,000 h−1639.0 (350)[79]
LaCoO3/3% K2O/CeO2 (impregnation)400 ppm NO/5% O2/N2; GHSV = 45,000 h−1512.4 (350)[80]
30% Pd/La0.7Sr0.3CoO3/Al2O3 (impregnation)500 ppm NO/6% O2/Ar; GHSV = 123,500 h−197.3 (350)[83]
La0.7Sr0.3CoO3/mesoporous SiO2 (deposition)500 ppm NO/ 5% O2/ N2; GHSV = 80,000 mL g−1 h−15269.2 (300)[84]
La0.7Sr0.3CoO3 (sol-gel)500 ppm NO/5% O2/ N2; GHSV = 80,000 mL g−1 h−12405.0 (300)[84]

4. NOx Desorption and Reduction

The engines are periodically shifted to fuel-rich conditions by introducing reductant pulses to regenerate the saturated storage sites, and NOx is desorbed and reduced on the redox center. Since the desorption and the reduction occur in series in a rather short time, these two steps are basically inseparable and always examined in combination over the catalysts. Four factors mainly determine the efficiency of fuel-rich regeneration: (1) reaction temperature; (2) intrinsic reducing ability of the reductant (i.e., H2 > CO > propene > propane); (3) stability of nitrate species, determined by the basicity of storage site; (4) redox property of the catalyst. In an early study [75], perovskite materials were found to be capable of activating N-O bonds at ca. 277 °C (even in the absence of reductants), with the formation of N2 and N2O as the NOx decomposition products. This observation has long been considered to be an important indicator of La-based perovskites being promising catalysts for fuel-rich reactions. However, more detailed examinations revealed the deficiency of perovskite in NOx reduction.
Table 2 summarizes the reaction conditions and DeNOx activity of La-based perovskite catalysts in the recent literature. Our previous study [76] reported 71.4% DeNOx activity and nearly 100% N2 selectivity over the La0.7Sr0.3CoO3 perovskite under alternating 3 min lean-burn/1 min fuel-rich conditions. However, only the NOx species stored on perovskite sites were reduced in fuel-rich periods, while the saturated SrCO3 site could hardly be regenerated, especially at low temperatures. Considering that such irreversible storage sites do not contribute to DeNOx activity and may even cover the active sites on the surface, Peng et al. [91] used the diluted HNO3 to remove the bulk SrCO3 on La0.5Sr0.5CoO3 perovskite, and observed a DeNOx activity enhancement from 11.9 to 42.4% at 250 °C. These results indicate the mismatch between the strong basicity of SrCO3 and relatively weak NOx reduction ability of La-based perovskites. A-site cationic deficiency can also inhibit the formation of SrCO3, which serves as an extra driving force for Sr doping into the perovskite structure, as observed with the La0.5Sr0.3MnO3 catalyst in our study [31]. However, the DeNOx activity of 43.2% over this catalyst at 350 °C was still insufficient for practical application, and the absence of strong basic sites inevitably leads to activity loss in the high-temperature region. It should be mentioned that the cationic deficiency affects the stability of the ABO3 perovskite structure, and it can result in the formation of a complex layered intergrown structure and impurity phases [92], which should be carefully examined.
Shi [69] demonstrated that even with the excellent NO oxidation ability, the overall DeNOx activity of LaMn0.9Fe0.1O3 was still less than half that of conventional Pt/BaO/Al2O3 catalyst in the whole temperature range (30–400 °C), due to the weak NOx reduction ability. In this case, a non-thermal H2-plasma was employed to assist in activation of reductants in the fuel-rich periods (Figure 5a), and then an excellent DeNOx activity of around 90% was achieved over perovskite with temperatures as low as 30 °C. Interestingly, the plasma did not significantly improve the DeNOx activity of the conventional Pt/BaO/Al2O3 catalyst, over which the rate-limiting step was NO oxidation and NOx storage (Figure 5b,c).
Even though the plasma technique provides great reinforcement for La-based perovskites, its application in a real aftertreatment system is restricted by cost constraints and product selectivity considerations. The noble metal is still required in the perovskite-based formulation to facilitate NOx reduction. Wen et al. [93] synthesized a series of Pt/LaCoO3/K2O/Al2O3 with different Pt loadings (0, 0.3%, and 1.0%). On one hand, the addition of Pt enhanced the NOx reduction ability of LaCoO3/K2O/Al2O3. On the other hand, the presence of LaCoO3 increased the NO oxidation ability and NSC of Pt/K2O/Al2O3, thereby reducing the Pt loading from 1.0% to 0.3% for achieving the equivalent DeNOx activity in a wide temperature range (Figure 6). The combination of perovskite and a trace amount of noble metal seems to be a promising and economical formulation for efficient NOx removal.
Compared to Pt, Pd is more commonly used together with perovskite, due to the lower activation temperature for hydrocarbons and better thermal stability [94]. The inadequate NO oxidation ability of Pd will not be problematic when perovskite is present in a catalyst formulation. Nishihata et al. [95,96] proposed a famous intelligent Pd-doped perovskite catalyst (LaFe0.57Co0.38Pd0.05O3) in 2002 characterized by the ‘‘self-regeneration” of Pd, which showed the behavior of repeatedly moving into and out of the perovskite lattice in response to lean/rich atmospheric alternations, and in this way the particle growth could be suppressed. Along this line, Ueda et al. [97] added a basic component and prepared an La0.7Ba0.3Fe0.776Nb0.194Pd0.03O3 catalyst, which showed a DeNOx activity of 47% at 250 °C. Afterwards, we reported [98] a La0.7Sr0.3Co0.97Pd0.03O3 perovskite formulation prepared via the sol-gel method with comparable activity and selectivity to Pt-based catalyst. A combination of characterizations, including XRD, XPS, and EAXFS, were used to identify the dissolution of Pd into the perovskite lattice.
With the aim of achieving higher DeNOx activity, in a more recent study, we investigated the optimal combination methods of perovskite and Pd by comparing the DeNOx performance of doped Pd cation and PdO impregnated on perovskite surface [67]. As a general trend, both Pd species significantly promote the DeNOx activity of La0.7Sr0.3CoO3 perovskite by accelerating NOx desorption and reduction. The impregnated PdO on the surface more effectively accelerated these steps due to the higher accessibility, and the efficient regeneration in turn increased NOx storage efficiency in the lean-burn periods. The effects of the spatial locations of Pd species were confirmed in a later study by Onrubia et al. [84], which also demonstrated the poorer accessibility of Pd cations accommodated in perovskite structure. In this study, alumina was used as the support to disperse Pd and perovskite. The formulated 1.5% Pd/30% La0.7Sr0.3CoO3/Al2O3 catalyst shows a prominent DeNOx activity of 86.2% at 350 °C.
Besides the effect of spatial location, there is also debate concerning the active chemical state of Pd on perovskite, which is more complicated, because the chemical state may change dynamically during lean-burn/fuel-rich alternations. By properly tuning metal–support interactions, an interesting in situ activation behavior was observed over the Pd/La0.7Sr0.3MnO3 catalyst in our study (Figure 7a) [99]. The DeNOx activity substantially increased from 56.1% to 90.1%, with N2O production simultaneously suppressed. This behavior resulted from the transformation of Pd2+ to Pd0 in the reaction atmosphere. The generated Pd0 showed a stronger ability to activate C3H6, and thus achieved eight-fold higher TOF for NOx reduction than Pd2+ (Figure 7b). Notably, the excessive strength of interactions suppressed the in situ formation of Pd0, thus lowering the De-NOx activity even with high dispersion of Pd. These studies provide important information for designing efficient Pd-perovskite NSR catalysts.
Table 2. Review of the reaction conditions and DeNOx activity of the La-based perovskites.
Table 2. Review of the reaction conditions and DeNOx activity of the La-based perovskites.
Catalyst (Preparation Method)Reaction ConditionsDeNOx Activity, % (Temperature, °C)Ref.
La0.5Sr0.3MnO3 (sol-gel)Lean (50 s): 400 ppm NO/5% O2/ N2; rich: phase (10 s): 1000 ppm C3H6/N2; GHSV = 120,000 mL g−1 h−143.2 (350)[31]
LaCo0.92Pt0.08O3 (sol-gel)Lean (120 s): 280 ppm NO/8% O2/8% CO2/N2; rich (30 s): 280 ppm NO/ 8% CO2/3.5% H2/ N2, GHSV = 72,000 h−1 (SO2 treatment: 100 ppm SO2/8% O2/ N2 for 45 min; regeneration: 3.5 vol% H2 at 500 °C for 12 h)90.9 (350)
25.5 (350, SO2 treatment)
75.5 (350, regeneration)
[66]
La0.7Sr0.3CoO3 (sol-gel)Lean (50 s): 400 ppm NO/5% O2/ N2; rich (10 s): 1000 ppm C3H6/N2; GHSV = 120,000 mL g−1 h−151.6 (300)[67]
La0.7Sr0.3Co0.97Pd0.03O3 (sol-gel)Lean (50 s): 400 ppm NO/ 5% O2/N2; rich (10 s): 1000 ppm C3H6/N2; GHSV = 120,000 mL g−1 h−174.4 (300)[67]
1.4% Pd/La0.7Sr0.3CoO3 (impregnation)Lean (50 s): 400 ppm NO/5% O2/N2; rich (10 s): 1000 ppm C3H6/ N2; GHSV = 120,000 mL g−1 h−190.4 (300)[67]
LaMn0.9Fe0.1O3 (sol-gel)Lean (10 min): 500 ppm NO/8% O2/N2; rich (2 min): 1% H2/ N2; GHSV = 10,000 h−1, plasma assisted>80% (30–300)[69]
LaCoO3 (sol-gel)Lean (3 min): 500 ppm NO/ 1000 ppm C3H6/6.7% O2/N2; rich (1 min): 500 ppm NO/ 1000 ppm C3H6/ N2; GHSV = 80,000 mL g−1 h−171.4 (300)[76]
1.5% Pd/30% La0.7Sr0.3CoO3/Al2O3 (impregnation)Lean (150 s): 500 ppm NO/6% O2/Ar; rich (20 s): 512 ppm NO/ 3% H2/ Ar; GHSV = 123,500 mL g−1 h−186.2% (350)[82]
La0.7Sr0.3CoO3 (sol-gel, acid wash)Lean (3 min): 500 ppm NO/5% O2/ N2; rich (1 min): 500 ppm NO/1000 ppm C3H6/ N2; GHSV = 120,000 mL g−1 h−155.0% (300)[91]
0.3% Pt/LaCoO3/K2O/Al2O3 (impregnation)Lean (120 s): 500 ppm NO/ 8% O2/ Ar2, rich (120 s): 500 ppm NO/3.5%H2/Ar; GHSV = 72,000 mL g−1 h−180.0 (300)[93]
La0.7Ba0.3Fe0.776Nb0.194Pd0.03O3 (glycine-nitrate method)Lean (54 s): 512 ppm NO/200 ppm C3H6/10% O2/Ar; rich (6 s): 512 ppm NO/200 ppm C3H6/4% CO/Ar; GHSV = 60,000 mL g−1 h−147.0 (250)[97]
La0.7Sr0.3Co0.97Pd0.03O3 (sol-gel)Lean (2 min): 0 or 100 ppm SO2/ 500 ppm NO/ 6.7% O2/ N2; rich (1 min); 0 or 100 ppm SO2/ 500 ppm NO/ 1000 ppm C3H6/ N2; GHSV = 32,000 h−1100 (325)
99.2 (325, with SO2)
[98]
2.1%Pd/La0.7Sr0.3MnO3 (sol-gel)Lean (50 s): 400 ppm NO/5% O2/ N2; rich: phase (10 s): 1000 ppm C3H6/ N2; GHSV = 120,000 mL g−1 h−190.1 (350)[99]
2.1%Pd/La0.7Sr0.3MnO3 (impregnation)Lean (50 s): 400 ppm NO/5% O2/N2; rich: phase (10 s): 1000 ppm C3H6/N2; GHSV = 120,000 mL g−1 h−172.0 (350)[99]

5. Resistance to Poisoners

An inherent defect of NSR catalyst is the susceptibility to sulfur poisoning, due to the close binding affinity between basic components and SO2 that easily leads to the formation of thermally stable sulfate, which causes the severe loss of NSC and continuous deactivation of the catalyst [11]. The sulfur tolerance is a major concern for the application of an NSR catalyst. Our recent studies illustrated that La-based perovskites had stronger sulfur tolerance than conventional Pd/BaO/Al2O3 and Pt/BaO/Al2O3 catalysts, which was tentatively assigned to the relatively weak basicity of perovskite storage sites and the stability of perovskite structure [31,99]. The incorporation of Fe into the B site of perovskite could further enhance the sulfur tolerance of La-based perovskite [100]. After the presulfation, the NSC of the La0.7Sr0.3CoO3 dropped by 43.6%, while that of the La0.7Sr0.3Co0.8Fe0.2O3 only dropped by 6.4%, since SO2 preferentially bonded with Fe to form Fe2(SO4)3, which circumvented the poisoning of the Sr site.
The regeneration of sulfated catalyst would be enhanced in the presence of noble metals. For example, Pd0 species was revealed to be effective in reducing sulfur poisoning through promoting hydrogen spillover to regenerate adjacent sulfated storage sites before they developed into the irreducible bulk phase [98]. The DeNOx activity of La0.7Sr0.3Co0.97Pd0.03O3 declined from 100% to 70.2% after the presulfating treatment but was fully restored by H2 reduction at a mild temperature (325 °C). In contrast, the DeNOx activity of Pt/BaO/Al2O3 catalyst declined from 99.2 to 54.9% and could not be recovered by H2 reduction. The study also reveals that the sulfur species are less likely to deposit on the catalyst in the co-feeding condition, as compared to long-term sulfur pretreatment. However, once the bulk sulfate species form on the basic sites, they can hardly be regenerated in H2 under 450 °C. The influence of sulfur poisoning was also comparatively studied with Pt/BaO/Al2O3 and LaCo0.92Pt0.08O3 [66]. Both catalysts suffered a drastic loss of NSC after a long-term presulfation (perovskite: 55.5 to 6.0 μmol g−1, Pt/BaO/Al2O3: 53.9 to 2.9 μmol g−1), as well as the decline of NO-to-NO2 conversion. The recovery of NSC and NO oxidation activity of the catalysts by 3.5% H2 reduction at 350 °C was time dependent. A long regeneration time benefited the removal of sulfate species from the catalysts but led to the damage to perovskite structure. The perovskite catalyst exhibited better regenerability (Figure 8). This was assigned to the formation of reducible cobalt sulfate on perovskite, which is less stable than bulk BaSO4 formed on Pt/BaO/Al2O3.
H2O is another poisoner for NSR catalysts and is known to impact NOx storage performance via competitively adsorbing on basic sites, forming hydroxide [11]. Our study [99] revealed that compared with the conventional Pt/BaO/Al2O3 catalyst, the La-based perovskite showed slightly higher H2O resistance, which was tentatively attributed to the higher NO-to-NO2 conversion [99], since NO2 could be more easily stored on basic sites than NO. Different from SO2, H2O poisoning is reversible and less severe; the perovskite catalyst’s activity can be gradually restored after shutting off the H2O feed [66].
Besides SO2 and H2O, soot in the exhaust can also affect NSR performance. Toyota proposed a diesel particulate NOx reduction (DPNR) filter system, which combines the functions of diesel particulate filter (DPF) and NSR and is operated under NSR cyclic conditions. Previous studies of Pt-based model NSR catalysts reveal that soot deposited on the surface of NSR catalyst reduces the NSC but promotes NOx desorption and reduction [101]. However, there is a lack of research on perovskite-based catalysts for DPNR systems. Although some authors focus on the simultaneous removal of soot and NOx with perovskite-based catalysts, they study the oxidation of soot and simultaneous reduction of NOx [102,103]. This operation mode is very different from NSR.

6. Conclusions and Perspective

Diesel engines achieve lower fuel consumption and CO2 emissions; however, the net oxidizing environment of exhaust is problematic for NOx reduction. The conventional three-way catalysis designed for gasoline vehicles is not functional under lean-burn conditions, which drives the development of NSR and SCR techniques that basically cover different market niches. The NSR technique does not demand a complex hardware system to realize NOx removal, so it is ideal for light-duty diesel vehicles with space constraints and cost considerations. However, NSR catalysts demand high loadings of Pt to achieve efficient NOx removal, which adds to the cost and reduces the thermal stability of the catalysts. The La-based perovskites, considered an efficient, durable, and economical alternative to Pt, have been extensively researched over the past decade. Through comprehensively reviewing these reported formulations and activities, a global evaluation of La-based perovskite materials applied in the NSR process is provided.
Perovskite-type structure stabilizes the uncommonly high oxidation state of B-site cations, and thus exhibits significant NO oxidation activity, especially LaMnO3 and LaCoO3, which show even higher oxidation activity than Pt-based catalysts. This advantage is further strengthened by the partial substitution of A-site La by low-valence cations, such as Sr2+, Ag+, or Ba2+, since this triggers the formation of oxygen vacancies to neutralize charge imbalance. The prominent NO oxidation activity and eliminated NOx diffusion barrier endow La-based perovskites with superior low-temperature NOx storage capacity compared to conventional Pt/BaO/Al2O3, showing the potential to reduce the NOx slip during cold start; however, at medium and high temperatures, La-based perovskites are less efficient for NOx storage, which is ascribed to the surface adsorption nature, poor texture property, and relatively weak basicity of the La site. This motivates the adoption of various synthetic strategies to increase the exposed surface area of perovskite, and to incorporate stronger basic components into perovskite-based catalyst formulations. The major problem of La-based perovskites is the inadequate NOx reduction performance in fuel-rich periods. This shows that the total substitution of noble metal by perovskite is not applicable. Nonetheless, a lower amount of noble metal is required in the presence of perovskite to reach the same DeNOx performance, since perovskite covers the shortage in NO oxidation and low-temperature NSC of noble metal catalysts. The simultaneous use of perovskite and a trace amount of noble metal is proven to be a feasible strategy for achieving high DeNOx activity in an economical way; therefore, how to increase the accessibility of noble metal and keep it in an active metallic state in a predominantly oxidizing atmosphere will be critical. Furthermore, the storage sites on La-based perovskite show strong sulfur tolerance, especially when Fe and Pd are incorporated. This increases the applicability to real automotive aftertreatment systems.
However, in-depth studies are still urgently needed for the future development of perovskite-based catalysts for NSR applications. The critical research areas and aspects that should be considered in future research mainly focus on the following aspects:
(1)
In-depth understanding of the NSR reaction mechanism in perovskites. To date, most mechanism studies concerning the NSR reaction were conducted with a Pt-based model catalyst, while for other types of catalysts, including perovskites, only vague pathways and plausible lines have been proposed. Considering the complexity of the reaction process, more in-depth experimental and theoretical studies should be conducted to unravel the detailed reaction routes, active sites, and intermediates to support the rational design of perovskite formulations with predictable properties.
(2)
Comprehensive comparison of NSR performance of La-based perovskites with others. The majority of perovskite-based materials reported for NSR applications over the last decade have been La-based. The obtained data reveals that La3+ provides prominent structural integrity with few impurities and little phase segregation, as well as thermal and chemical stability resistant to phase transformation during NOx storage and reduction. Furthermore, the extensively studied LaCoO3 and LaMnO3 catalytic systems exhibit a good redox property, which is desirable for the NSR reaction. However, the weak basicity of A-site La3+ brings challenges for high-temperature NOx storage, and the usage of La and Co increases environmental burdens from the perspective of sustainability. Some interesting data have been recently reported for Sr- and Ba-based perovskite formulations. On one hand, the stronger basicity of A-site alkaline earth metal provides better high-temperature NOx storage performance, and the +2 oxidation state of the A-site cation increases the B-site oxidation state and/or generates oxygen vacancies desirable for NO oxidation and NOx storage. On the other hand, these strongly basic cations are generally less resistant to CO2, NOx, and SO2 in the reaction atmosphere, and hence are prone to exsolution from the perovskite matrix, leading to the risk of collapse of the perovskite structure, but this tendency may vary depending on specific formulations. At this stage, a comprehensive study of the effects of different A-site cations on the activity and stability of perovskite catalysts in NSR reactions will be rather helpful.
(3)
Synergetic/competitive relation between different storage sites. As stated earlier, a practical NSR catalyst should be capable of efficiently trapping and releasing NOx in a wide range of temperatures. This indicates that multiple storage components with different basicity should be used in combination; however, almost no comprehensive research has been devoted to this line. Current research suggests that La-based perovskite is an excellent low-temperature NOx storage component, but whether the combination of perovskite and stronger basic sites efficient for high-temperature NSR reaction can work as they are intended remains unclear.
(4)
Redox stability and resistance to H2O, CO2, and SO2 poisoning. Most perovskite formulations are designed and tested under ideal laboratory conditions. The durability of perovskite catalysts in reducing atmosphere or when exposed to the common poisoners in automotive exhaust should be carefully considered and further clarified.
(5)
Sustainability has become an essential concern in catalyst design with the guidance of legislation and increasing environmental awareness of the community. However, this is entirely neglected in the currently reviewed publications. The environmental burden of using Pt on a per kilogram basis is calculated to be 3–4 orders of magnitude larger than base metals, in the aspects of global warming potential, cumulative energy demand, terrestrial acidification, freshwater eutrophication, and human toxicity [104], which necessitates the development of environmentally friendly base metal alternatives. Along this line, in perovskite formulation design, the metals at the lower end of the scale of environmental impacts should be preferably considered, such as manganese, iron, and titanium at B-site. These metals also result in decent NSR performance, while the usage of nickel and cobalt should be more prudent. For A-site cations, Sr and Ba are better choices than La.

Author Contributions

Conceptualization, D.Z.; formal analysis, D.Z., J.L. and Q.J.; writing—original draft preparation, D.Z.; writing—review and editing, X.L.; supervision, H.S.; project administration, H.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Sinopec Research Program (119043 and 121011).

Acknowledgments

We acknowledge the inspiring discussion with Chunyu Xin and Zhongnan Gao in Tianjin University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bai, Z.F.; Zhang, Z.S.; Chen, B.B.; Zhao, Q.; Crocker, M.; Shi, C. Non-thermal plasma enhanced NSR performance over Pt/M/Ba/Al2O3 (M = Mn, Co, Cu) catalysts. Chem. Eng. J. 2017, 314, 688–699. [Google Scholar] [CrossRef] [Green Version]
  2. Parks, J.E. Less costly catalysts for controlling engine emissions. Science 2010, 327, 1584–1585. [Google Scholar] [CrossRef] [PubMed]
  3. Jabłońska, M.; Palkovits, R. Perovskite-based catalysts for the control of nitrogen oxide emissions from diesel engines. Catal. Sci. Technol. 2019, 9, 2057–2077. [Google Scholar] [CrossRef]
  4. Zhang, Y.; Cao, G.; Yang, X. Advances in De-NOx methods and catalysts for direct catalytic decomposition of NO: A review. Energy Fuels 2021, 35, 6443–6464. [Google Scholar] [CrossRef]
  5. Ishihara, T.; Ando, M.; Sada, K.; Takiishi, K.; Yamada, K.; Nishiguchi, H.; Takita, Y. Direct decomposition of NO into N2 and O2 over La(Ba)Mn(In)O3 perovskite oxide. J. Catal. 2003, 220, 104–114. [Google Scholar] [CrossRef]
  6. Zhao, H.W.; Han, L.; Wang, Y.J.; Zheng, J.D. Insight into platinum poisoning effect on Cu-SSZ-13 in selective catalytic reduction of NOx with NH3. Catalysts 2021, 11, 796. [Google Scholar] [CrossRef]
  7. Chen, C.; Cao, Y.; Liu, S.; Chen, J.; Jia, W. Review on the latest developments in modified vanadium-titanium-based SCR catalysts. Chin. J. Catal. 2018, 39, 1347–1365. [Google Scholar] [CrossRef]
  8. Gao, F. Fe-exchanged small-pore zeolites as ammonia selective catalytic reduction (NH3-SCR) catalysts. Catalysts 2020, 10, 1324. [Google Scholar] [CrossRef]
  9. Sun, H.; Park, S. Recent advances in MNOx/CeO2-based ternary composites for selective catalytic reduction of NOx by NH3: A review. Catalysts 2021, 11, 1519. [Google Scholar] [CrossRef]
  10. Zhao, X.; Mao, L.; Dong, G. Mn-Ce-V-WOx/TiO2 SCR catalysts: Catalytic activity, stability and interaction among catalytic oxides. Catalysts 2018, 8, 76. [Google Scholar] [CrossRef] [Green Version]
  11. Roy, S.; Baiker, A. NOx storage-reduction catalysis: From mechanism and materials properties to storage-reduction performance. Chem. Rev. 2009, 109, 4054–4091. [Google Scholar] [CrossRef]
  12. Mrad, R.; Aissat, A.; Cousin, R.; Courcot, D.; Siffert, S. Catalysts for NOx selective catalytic reduction by hydrocarbons (HC-SCR). Appl. Catal. A 2015, 504, 542–548. [Google Scholar] [CrossRef]
  13. Liu, Z.; Ihl Woo, S. Recent advances in catalytic DeNOx science and technology. Catal. Rev. 2006, 48, 43–89. [Google Scholar] [CrossRef]
  14. Takahashi, N.; Shinjoh, H.; Iijima, T.; Suzuki, T.; Yamazaki, K.; Yokota, K.; Suzuki, H.; Miyoshi, N.; Matsumoto, S.; Tanizawa, T.; et al. The new concept 3-way catalyst for automotive lean-burn engine: NOx storage and reduction catalyst. Catal. Today 1996, 27, 63–69. [Google Scholar] [CrossRef]
  15. Yin, M.X.; Liu, D.S.; Zhao, D.Y.; Ding, T.; Tian, Y.; Li, X.G. Effect of copper doping on lean NOx Trap Performance of Pt/Ba/CuxMg1-xAl2O4 catalysts at high temperatures. Chem. J. Chin. Univ. 2019, 40, 2170–2177. [Google Scholar]
  16. Epling, W.S.; Campbell, L.E.; Yezerets, A.; Currier, N.W.; Parks, J.E. Overview of the fundamental reactions and degradation mechanisms of NOx storage/reduction catalysts. Catal. Rev. 2004, 46, 163–245. [Google Scholar] [CrossRef]
  17. Park, C.; Lee, S.; Yi, U. Effects of engine operating conditions on particle emissions of lean-burn gasoline direct-injection engine. Energy 2016, 115, 1148–1155. [Google Scholar] [CrossRef]
  18. Praveena, V.; Martin, M.L.J. A review on various after treatment techniques to reduce NOx emissions in a CI engine. J. Energy Inst. 2018, 91, 704–720. [Google Scholar] [CrossRef]
  19. Song, C. An overview of new approaches to deep desulfurization for ultra-clean gasoline, diesel fuel and jet fuel. Catal. Today 2003, 86, 211–263. [Google Scholar] [CrossRef]
  20. Pérot, G. Hydrotreating catalysts containing zeolites and related materials-mechanistic aspects related to deep desulfurization. Catal. Today 2003, 86, 111–128. [Google Scholar] [CrossRef]
  21. Yang, X.; Gao, Q.; Zhao, Z.; Guo, Y.; Guo, Y.; Wang, L.; Wang, Y.; Zhan, W. Surface tuning of noble metal doped perovskite oxide by synergistic effect of thermal treatment and acid etching: A new path to high-performance catalysts for methane combustion. Appl. Catal. B 2018, 239, 373–382. [Google Scholar] [CrossRef]
  22. Xu, J.; Liu, J.; Zhao, Z.; Zheng, J.; Zhang, G.; Duan, A.; Jiang, G. Three-dimensionally ordered macroporous LaCoxFe1−xO3 perovskite-type complex oxide catalysts for diesel soot combustion. Catal. Today 2010, 153, 136–142. [Google Scholar] [CrossRef]
  23. Hwang, J.; Rao, R.R.; Giordano, L.; Katayama, Y.; Yu, Y.; Shao-Horn, Y. Perovskites in catalysis and electrocatalysis. Science 2017, 358, 751–756. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Peng, Y.; Si, W.; Li, J.; Crittenden, J.; Hao, J. Experimental and DFT studies on Sr-doped LaMnO3 catalysts for NOx storage and reduction. Catal. Sci. Technol. 2015, 5, 2478–2485. [Google Scholar] [CrossRef]
  25. Li, F.; Tang, J.; Ke, Q.; Guo, Y.; Ha, M.N.; Wan, C.; Lei, Z.; Gu, J.; Li, Q.; Nguyen, V.N.; et al. Investigation into enhanced catalytic performance for epoxidation of styrene over LaSrCoxFe2–xO6 double perovskites: The role of singlet oxygen species promoted by the photothermal effect. ACS Catal. 2021, 11, 11855–11866. [Google Scholar] [CrossRef]
  26. Si, W.; Wang, Y.; Peng, Y.; Li, J. Selective dissolution of A-site cations in ABO3 perovskites: A new path to high-performance catalysts. Angew. Chem. 2015, 127, 8065–8068. [Google Scholar] [CrossRef]
  27. Gao, Z.N.; Guo, L.H.; Zhao, D.Y.; Li, X.G. Effect of A site-deficiency on the structure and catalytic oxidation activity of the La-Sr-Co-O perovskite. Chem. J. Chin. Univ. 2021, 42, 2869–2877. [Google Scholar]
  28. Hernández-Giménez, A.; Castelló, D.; Bueno-López, A. Diesel soot combustion catalysts: Review of active phases. Chem. Pap. 2014, 68, 1154–1168. [Google Scholar] [CrossRef] [Green Version]
  29. Zhu, J.; Thomas, A. Perovskite-type mixed oxides as catalytic material for NO removal. Appl. Catal. B 2009, 92, 225–233. [Google Scholar] [CrossRef]
  30. Royer, S.; Duprez, D.; Can, F.; Courtois, X.; Batiot-Dupeyrat, C.; Laassiri, S.; Alamdari, H. Perovskites as substitutes of noble metals for heterogeneous catalysis: Dream or reality. Chem. Rev. 2014, 114, 10292–10368. [Google Scholar] [CrossRef]
  31. Zhao, D.Y.; Yang, Y.X.; Gao, Z.N.; Tian, Y.; Zhang, J.; Jiang, Z.; Li, X.G. A-site defects in perovskite-based catalyst promoting NOx storage and reduction for lean-burn exhausts. J. Rare Earth 2021, 39, 959–968. [Google Scholar] [CrossRef]
  32. Švarcová, S.; Wiik, K.; Tolchard, J.; Bouwmeester, H.J.; Grande, T. Structural instability of cubic perovskite BaxSr1−xCo1−yFeyO3−δ. Solid State Ion. 2008, 178, 1787–1791. [Google Scholar] [CrossRef]
  33. Zhu, J.; Li, H.; Zhong, L.; Xiao, P.; Xu, X.; Yang, X.; Zhao, Z.; Li, J. Perovskite oxides: Preparation, characterizations, and applications in heterogeneous catalysis. ACS Catal. 2014, 4, 2917–2940. [Google Scholar] [CrossRef]
  34. Labhasetwar, N.; Saravanan, G.; Megarajan, S.K.; Manwar, N.; Khobragade, R.; Doggali, P.; Grasset, F. Perovskite-type catalytic materials for environmental applications. Sci. Technol. Adv. Mater. 2015, 16, 036002. [Google Scholar] [CrossRef]
  35. Li, C.; Soh, K.C.K.; Wu, P. Formability of ABO3 perovskites. J. Alloy Compd. 2004, 372, 40–48. [Google Scholar] [CrossRef]
  36. 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]
  37. Xian, H.; Zhang, X.W.; Li, X.G.; Zou, H.H.; Meng, M.; Zou, Z.Q.; Guo, L.H.; Tsubaki, N. Effect of the calcination conditions on the NOx storage behavior of the perovskite BaFeO3−x catalysts. Catal. Today 2010, 158, 215–219. [Google Scholar] [CrossRef]
  38. Xian, H.; Li, F.L.; Li, X.G.; Zhang, X.W.; Meng, M.; Zhang, T.Y.; Tsubaki, N. Influence of preparation conditions to structure property, NOx and SO2 sorption behavior of the BaFeO3−x perovskite catalyst. Fuel Process. Technol. 2011, 92, 1718–1724. [Google Scholar] [CrossRef]
  39. Ge, C.; Li, L.; Xian, H.; Yan, H.; Meng, M.; Li, X. Effects of Ti-doping on the NOx storage and the sulfur resistance of the BaFe1−xTixO3−y perovskite-type catalysts for lean-burn exhausts. Fuel Process. Technol. 2014, 120, 1–7. [Google Scholar] [CrossRef]
  40. Hodjati, S.; Vaezzadeh, K.; Petit, C.; Pitchon, V.; Kiennemann, A. Absorption/desorption of NOx process on perovskites: Performances to remove NOx from a lean exhaust gas. Appl. Catal. B 2000, 26, 5–16. [Google Scholar] [CrossRef]
  41. Hodjati, S.; Petit, C.; Pitchon, V.; Kiennemann, A. Absorption/desorption of NOx process on perovskites: Impact of SO2 on the storage capacity of BaSnO3 and strategy to develop thioresistance. Appl. Catal. B 2001, 30, 247–257. [Google Scholar] [CrossRef]
  42. Li, X.; Chen, J.; Lin, P.; Meng, M.; Fu, Y.; Tu, J.; Li, Q. A study of the NOx storage catalyst of Ba-Fe-O complex oxide. Catal. Commun. 2004, 5, 25–28. [Google Scholar] [CrossRef]
  43. Olsson, L.; Persson, H.; Fridell, E.; Skoglundh, M.; Andersson, B. A kinetic study of NO oxidation and NOx storage on Pt/Al2O3 and Pt/BaO/Al2O3. J. Phys. Chem. B 2001, 105, 6895–6906. [Google Scholar] [CrossRef]
  44. Bhatia, D.; McCabe, R.W.; Harold, M.P.; Balakotaiah, V. Experimental and kinetic study of NO oxidation on model Pt catalysts. J. Catal. 2009, 266, 106–119. [Google Scholar] [CrossRef]
  45. Hauff, K.; Tuttlies, U.; Eigenberger, G.; Nieken, U. Platinum oxide formation and reduction during NO oxidation on a diesel oxidation catalyst–experimental results. Appl. Catal. B 2012, 123, 107–116. [Google Scholar] [CrossRef]
  46. Li, L.D.; Shen, Q.; Cheng, J.; Hao, Z.P. Catalytic oxidation of NO over TiO2 supported platinum clusters. II: Mechanism study by in situ FTIR spectra. Catal. Today 2010, 158, 361–369. [Google Scholar] [CrossRef]
  47. Hong, Z.; Wang, Z.; Li, X.B. Catalytic oxidation of nitric oxide (NO) over different catalysts: An overview. Catal. Sci. Technol. 2017, 7, 3440–3452. [Google Scholar] [CrossRef]
  48. Weiss, B.M.; Iglesia, E. NO oxidation catalysis on Pt clusters: Elementary steps, structural requirements, and synergistic effects of NO2 adsorption sites. J. Phys. Chem. C 2009, 113, 13331–13340. [Google Scholar] [CrossRef] [Green Version]
  49. Weiss, B.M.; Iglesia, E. Mechanism and site requirements for NO oxidation on Pd catalysts. J. Catal. 2010, 272, 74–81. [Google Scholar] [CrossRef]
  50. Mulla, S.S.; Chen, N.; Cumaranatunge, L.; Blau, G.E.; Zemlyanov, D.Y.; Delgass, W.N.; Epling, W.S.; Ribeiro, F.H. Reaction of NO and O2 to NO2 on Pt: Kinetics and catalyst deactivation. J. Catal. 2006, 241, 389–399. [Google Scholar] [CrossRef]
  51. Constantinou, C.; Li, W.; Qi, G.; Epling, W.S. NOx storage and reduction over a perovskite-based lean NOx trap catalyst. Appl. Catal. B 2013, 134, 66–74. [Google Scholar] [CrossRef]
  52. Mulla, S.S.; Chen, N.; Delgass, W.N.; Epling, W.S.; Ribeiro, F.H. NO2 inhibits the catalytic reaction of NO and O2 over Pt. Catal. Lett. 2005, 100, 267–270. [Google Scholar] [CrossRef]
  53. Sivachandiran, L.; Thévenet, F.; Gravejat, P.; Rousseau, A. Investigation of NO and NO2 adsorption mechanisms on TiO2 at room temperature. Appl. Catal. B 2013, 142, 196–204. [Google Scholar] [CrossRef]
  54. 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]
  55. Parker, D.H.; Koel, B.E. Chemisorption of high coverages of atomic oxygen on the Pt (111), Pd (111), and Au (111) surfaces. J. Vac. Sci. Technol. A 1990, 8, 2585–2590. [Google Scholar] [CrossRef]
  56. Segner, J.; Vielhaber, W.; Ertl, G. Interaction of NO2 with a Pt (111) surface. Isr. J. Chem. 1982, 22, 375–379. [Google Scholar] [CrossRef]
  57. He, X.; Meng, M.; He, J.; Zou, Z.; Li, X.; Li, Z.; Jiang, Z. A potential substitution of noble metal Pt by perovskite LaCoO3 in ZrTiO4 supported lean-burn NOx trap catalysts. Catal. Commun. 2010, 12, 165–168. [Google Scholar] [CrossRef]
  58. Chen, J.; Shen, M.; Wang, X.; Qi, G.; Wang, J.; Li, W. The influence of nonstoichiometry on LaMnO3 perovskite for catalytic NO oxidation. Appl. Catal. B 2013, 134, 251–257. [Google Scholar] [CrossRef]
  59. 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 2021, 286, 119919. [Google Scholar] [CrossRef]
  60. 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]
  61. Onrubia, 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 2017, 213, 198–210. [Google Scholar] [CrossRef]
  62. Guo, L.H.; Bo, L.; Li, Y.; Jiang, Z.; Tian, Y.; Li, X.G. Sr doping effect on the structure property and NO oxidation performance of dual-site doped perovskite La(Sr)Co(Fe)O3. Solid State Sci. 2021, 113, 106519. [Google Scholar] [CrossRef]
  63. Lim, E.; Kim, Y.J.; Kim, J.H.; HoKim, J.; Ryu, T.; Lee, S.; Cho, B.K.; Nam, I.; Choung, J.W.; Yoo, S. NO oxidation activity of Ag-doped perovskite catalysts. J. Catal. 2014, 319, 182–193. [Google Scholar]
  64. Chen, J.; Shen, M.; Wang, X.; Wang, J.; Su, Y.; Zhao, Z. Catalytic performance of NO oxidation over LaMeO3 (Me = Mn, Fe, Co) perovskite prepared by the sol-gel method. Catal. Commun. 2013, 37, 105–108. [Google Scholar] [CrossRef]
  65. Zhong, S.; Sun, Y.; Xin, H.; Yang, C.; Chen, L.; Li, X. NO oxidation over Ni-Co perovskite catalysts. Chem. Eng. J. 2015, 275, 351–356. [Google Scholar] [CrossRef]
  66. Wang, X.; Qi, X.; Chen, Z.; Jiang, L.; Wang, R.; Wei, K. 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]
  67. Zhao, D.; Gao, Z.; Xian, H.; Xing, L.; Yang, Y.; Tian, Y.; Ding, T.; Jiang, Z.; Zhang, J.; Zheng, L.R.; et al. Addition of Pd on La0.7Sr0.3CoO3 perovskite to enhance catalytic removal of NOx. Ind. Eng. Chem. Res. 2018, 57, 521–531. [Google Scholar] [CrossRef]
  68. Epling, W.S.; Yezerets, A.; Currier, N.W. The effects of regeneration conditions on NOx and NH3 release from NOx storage/reduction catalysts. Appl. Catal. B 2007, 74, 117–129. [Google Scholar] [CrossRef]
  69. Shi, C.; Zhang, Z.; Crocker, M.; Xu, L.; Wang, C.Y.; Au, C.; Zhu, A.M. Non-thermal plasma-assisted NOx storage and reduction on a LaMn0.9Fe0.1O3 perovskite catalyst. Catal. Today 2013, 211, 96–103. [Google Scholar] [CrossRef]
  70. Yoshiyama, Y.; Hosokawa, S.; Tamai, K.; Kajino, T.; Yoto, H.; Asakura, H.; Teramura, K.; Tanaka, T. NOx storage performance at low temperature over platinum group metal-free SrTiO3-based material. ACS Appl. Mater. Interfaces 2021, 13, 29482–29490. [Google Scholar] [CrossRef]
  71. Tamai, K.; Hosokawa, S.; Kato, K.; Asakura, H.; Teramura, K.; Tanaka, T. Low-temperature NO oxidation using lattice oxygen in Fe-site substituted SrFeO3−δ. Phys. Chem. Chem. Phys. 2020, 22, 24181–24190. [Google Scholar] [CrossRef]
  72. Takamatsu, A.; Tamai, K.; Hosokawa, S.; Tanaka, T.; Ehara, M.; Fukuda, R. Oxidation and storage mechanisms for nitrogen oxides on variously terminated (001) surfaces of SrFeO3−δ and Sr3Fe2O7−δ perovskites. ACS Appl. Mater. Interfaces 2021, 13, 7216–7226. [Google Scholar] [CrossRef]
  73. Ye, J.; Yu, Y.; Meng, M.; Jiang, Z.; Ding, T.; Zhang, S.; Huang, Y. Highly efficient NOx purification in alternating lean/rich atmospheres over non-platinic mesoporous perovskite-based catalyst K/LaCoO3. Catal. Sci. Technol. 2013, 3, 1915–1918. [Google Scholar] [CrossRef]
  74. Qi, G.; Li, W. NOx adsorption and reduction over LaMnO3 based lean NOx trap catalysts. Catal. Lett. 2014, 144, 639–647. [Google Scholar] [CrossRef]
  75. Say, Z.; Dogac, M.; Vovk, E.I.; Kalay, Y.E.; Kim, C.H.; Li, W.; Ozensoy, E. Palladium doped perovskite-based NO oxidation catalysts: The role of Pd and B-sites for NOx adsorption behavior via in-situ spectroscopy. Appl. Catal. B 2014, 154, 51–61. [Google Scholar] [CrossRef]
  76. Li, X.G.; Dong, Y.H.; Xian, H.; Hernández, 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]
  77. Dong, Y.H.; Xian, H.; Lv, J.L.; Liu, C.; Guo, L.; Meng, M.; Tan, Y.S.; Tsubaki, N.; Li, X.G. Influence of synthesis conditions on NO oxidation and NOx storage performances of La0.7Sr0.3MnO3 perovskite-type catalyst in lean-burn atmospheres. Mater. Chem. Phys. 2014, 143, 578–586. [Google Scholar] [CrossRef]
  78. Xie, W.; Xu, G.Y.; Zhang, Y.; Yu, Y.B.; He, H. Mesoporous LaCoO3 perovskite oxide with high catalytic performance for NOx storage and reduction. J. Hazard. Mater. 2022, 431, 128528. [Google Scholar] [CrossRef]
  79. 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]
  80. You, R.; Zhang, Y.; Liu, F.; 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]
  81. Onrubia-Calvo, J.A.; Pereda-Ayo, B.; Gonzalez-Velasco, J.R. Perovskite-based catalysts as efficient, durable, and economical NOx storage and reduction systems. Catalysts 2020, 10, 208. [Google Scholar] [CrossRef] [Green Version]
  82. Onrubia-Calvo, J.A.; Pereda-Ayo, B.; Bermejo-Lopez, A.; Caravaca, A.; Vernoux, P.; González-Velasco, J.R. Pd-doped or Pd impregnated 30% La0.7Sr0.3CoO3/Al2O3 catalysts for NOx storage and reduction. Appl. Catal. B 2019, 259, 118052. [Google Scholar] [CrossRef]
  83. Onrubia-Calvo, J.A.; Pereda-Ayo, B.; De-La-Torre, U.; González-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]
  84. Ding, Q.; Xian, H.; Tan, Y.; Tsubaki, N.; Li, X. Mesoporous SiO2-confined La0.7Sr0.3CoO3 perovskite nanoparticles: An efficient NOx adsorber for lean-burn exhausts. Catal. Sci. Technol. 2013, 3, 1493–1496. [Google Scholar] [CrossRef]
  85. Qi, G.; Li, W. Pt-free, LaMnO3 based lean NOx trap catalysts. Catal. Today 2012, 184, 72–77. [Google Scholar] [CrossRef]
  86. Onrubia-Calvo, J.A.; Pereda-Ayo, B.; Caravaca, A.; De-La-Torre, U.; Vernoux, P.; González-Velasco, J.R. Tailoring perovskite surface composition to design efficient lean NOx trap Pd–La1-xAxCoO3/Al2O3-type catalysts (with A = Sr or Ba). Appl. Catal. B 2020, 266, 118628. [Google Scholar] [CrossRef]
  87. Onrubia-Calvo, J.A.; Pereda-Ayo, B.; Cabrejas, I.; De-La-Torre, U.; González-Velasco, J.R. Ba-doped vs. Sr-doped LaCoO3 perovskites as base catalyst in diesel exhaust purification. Mol. Catal. 2020, 488, 110913. [Google Scholar] [CrossRef]
  88. Onrubia-Calvo, J.A.; Pereda-Ayo, B.; Urrutxua, M.; De La Torre, U.; González-Velasco, J.R. Boosting NOx removal by perovskite-based catalyst in NSR–SCR diesel aftertreatment systems. Ind. Eng. Chem. Res. 2021, 60, 6525–6537. [Google Scholar] [CrossRef]
  89. Albaladejo-Fuentes, V.; Sánchez-Adsuar, M.S.; Anderson, J.A.; Illán-Gómez, M.J. NOx storage on BaTi0.8Cu0.2O3 perovskite catalysts: Addressing a feasible mechanism. Nanomaterials 2021, 11, 2133. [Google Scholar] [CrossRef]
  90. Yi, J.; Schroeder, M.; Weirich, T.; Mayer, J. Behavior of Ba (Co, Fe, Nb)O3-δ perovskite in CO2-containing atmospheres: Degradation mechanism and materials design. Chem. Mater. 2010, 22, 6246–6253. [Google Scholar] [CrossRef]
  91. Peng, Y.; Si, W.; Luo, J.; Su, W.K.; Chang, H.Z.; Li, J.H.; Hao, J.M.; 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]
  92. Zhu, Y.; Zhou, W.; Yu, J.; Chen, Y.; Liu, M.; Shao, Z. Enhancing electrocatalytic activity of perovskite oxides by tuning cation deficiency for oxygen reduction and evolution reactions. Chem. Mater. 2016, 28, 1691–1697. [Google Scholar] [CrossRef]
  93. Wen, W.; Wang, X.; Jin, S.; Wang, R. 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]
  94. Ciuparu, D.; Bensalem, A.; Pfefferle, L. Pd–Ce interactions and adsorption properties of palladium: CO and NO TPD studies over Pd–Ce/Al2O3 catalysts. Appl. Catal. B 2000, 26, 241–255. [Google Scholar] [CrossRef]
  95. Nishihata, Y.; Mizuki, J.; Akao, T.; Tanaka, H.; Uenishi, M.; Kimura, M.; Okamoto, T.; Hamada, N. Self-regeneration of a Pd-perovskite catalyst for automotive emissions control. Nature 2002, 418, 164–167. [Google Scholar] [CrossRef]
  96. Nishihata, Y.; Mizuki, J.; Tanaka, H.; Uenishi, M.; Kimura, M. Self-regeneration of palladium-perovskite catalysts in modern automobiles. J. Phys. Chem. Solids 2005, 66, 274–282. [Google Scholar] [CrossRef]
  97. 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]
  98. Li, X.G.; Chen, C.; Liu, C.; Xian, H.; Guo, L.; Lv, J.; Jiang, Z.; Vernoux, P. Pd-doped perovskite: An effective catalyst for removal of NOx from lean-burn exhausts with high sulfur resistance. ACS Catal. 2013, 3, 1071–1075. [Google Scholar] [CrossRef]
  99. Zhao, D.; Yang, Y.; Gao, Z.; Yin, M.; Tian, Y.; Zhang, J.; Jiang, Z.; Yu, X.B.; Li, X. Promoting NOx reduction via in situ activation of perovskite supported Pd catalysts under alternating lean-burn/fuel-rich operating atmospheres. Chin. J. Catal. 2021, 42, 795–807. [Google Scholar] [CrossRef]
  100. Ma, A.J.; Wang, S.Z.; Liu, C.; Xian, H.; Ding, Q.; Guo, L.; Meng, M.; Tan, Y.S.; Tsubaki, N.; Zhang, J.; et al. Effects of Fe dopants and residual carbonates on the catalytic activities of the perovskite-type La0.7Sr0.3Co1-xFexO3 NOx storage catalyst. Appl. Catal. B 2014, 146, 24–34. [Google Scholar] [CrossRef]
  101. Artioli, N.; Matarrese, R.; Castoldi, L.; Lietti, L.; Forzatti, P. Effect of soot on the storage-reduction performances of PtBa/Al2O3 LNT catalyst. Catal. Today 2011, 169, 36–44. [Google Scholar] [CrossRef]
  102. Castoldi, L. An overview on the catalytic materials proposed for the simultaneous removal of NOx and soot. Materials 2020, 13, 3551. [Google Scholar] [CrossRef] [PubMed]
  103. Yang, L.; Hu, J.; Tian, G.; Zhu, J.; Song, Q.; Wang, H.; Zhang, C. Efficient catalysts of K and Ce Co-doped LaMnO3 for NOx–soot simultaneous removal and reaction kinetics. ACS Omega 2021, 6, 19836–19845. [Google Scholar] [CrossRef] [PubMed]
  104. Nuss, P.; Eckelman, M.J. Life cycle assessment of metals: A scientific synthesis. PLoS ONE 2014, 9, e101298. [Google Scholar] [CrossRef]
Figure 1. NO-to-NO2 conversions over (a) La1−xSrxCoO3 and (b) La1−xSrxMnO3 perovskites with x ranging from 0 to 0.5, with model Pt-based catalyst as the benchmark. Adapted with permission from [61]. Copyright 2017, Elsevier.
Figure 1. NO-to-NO2 conversions over (a) La1−xSrxCoO3 and (b) La1−xSrxMnO3 perovskites with x ranging from 0 to 0.5, with model Pt-based catalyst as the benchmark. Adapted with permission from [61]. Copyright 2017, Elsevier.
Catalysts 12 00593 g001
Figure 2. Comparison of the temperature-dependent NOx storage capacities of infiltration composite 15 wt.% La0.5Sr0.5Fe1−xTixO3-δ loading (◆) and a conventional Pt/BaO/Al2O3 (20% BaO) catalyst (★). Reprinted with permission from [59]. Copyright 2021, Elsevier.
Figure 2. Comparison of the temperature-dependent NOx storage capacities of infiltration composite 15 wt.% La0.5Sr0.5Fe1−xTixO3-δ loading (◆) and a conventional Pt/BaO/Al2O3 (20% BaO) catalyst (★). Reprinted with permission from [59]. Copyright 2021, Elsevier.
Catalysts 12 00593 g002
Figure 3. FT-IR spectra (left) of La0.7Sr0.3CoO3: fresh (a); after NOx storage at 200 °C (b), 250 °C (c), 300 °C (d), 350 °C (e), and 400 °C (f); the sample (c) reduced by 5% H2 at 300 °C for 10 min (g). Possible NOx storage routes on La1−xSrxCoO3 (right). Adapted from [76].
Figure 3. FT-IR spectra (left) of La0.7Sr0.3CoO3: fresh (a); after NOx storage at 200 °C (b), 250 °C (c), 300 °C (d), 350 °C (e), and 400 °C (f); the sample (c) reduced by 5% H2 at 300 °C for 10 min (g). Possible NOx storage routes on La1−xSrxCoO3 (right). Adapted from [76].
Catalysts 12 00593 g003
Figure 4. NO oxidation rate and NOx storage capacity as functions of the loading of La0.7Sr0.3CoO3 perovskite on alumina support. Reprinted with permission from [83]. Copyright 2019, Elsevier.
Figure 4. NO oxidation rate and NOx storage capacity as functions of the loading of La0.7Sr0.3CoO3 perovskite on alumina support. Reprinted with permission from [83]. Copyright 2019, Elsevier.
Catalysts 12 00593 g004
Figure 5. (a) Illustrative profile of NSR reaction with the assistance of H2-plasma; NOx conversion during lean/rich cycling at different temperatures over LaMn0.9Fe0.1O3 and Pt/BaO/Al2O3: (b) without H2-plasma, (c) with H2-plasma. Reprinted with permission from [69]. Copyright 2013, Elsevier.
Figure 5. (a) Illustrative profile of NSR reaction with the assistance of H2-plasma; NOx conversion during lean/rich cycling at different temperatures over LaMn0.9Fe0.1O3 and Pt/BaO/Al2O3: (b) without H2-plasma, (c) with H2-plasma. Reprinted with permission from [69]. Copyright 2013, Elsevier.
Catalysts 12 00593 g005
Figure 6. NOx conversion of 1.0 wt.% Pt/K2O/Al2O3 and x wt% Pt/LaCoO3/K2O/Al2O3. Reprinted with permission from [93]. Copyright 2011, Royal Society of Chemistry.
Figure 6. NOx conversion of 1.0 wt.% Pt/K2O/Al2O3 and x wt% Pt/LaCoO3/K2O/Al2O3. Reprinted with permission from [93]. Copyright 2011, Royal Society of Chemistry.
Catalysts 12 00593 g006
Figure 7. (a) De-NOx activity as a functions of reaction time at 350 °C of La0.7Sr0.3MnO3 (▲), Pd/La0.7Sr0.3MnO3-sol-gel (■), Pd/La0.7Sr0.3MnO3-impregnation (●), and pretreated Pd/La0.7Sr0.3MnO3--ol-gel (▼); and (b) illustrative profile of in situ activation of Pd/La0.7Sr0.3MnO3-sol-gel catalyst. Adapted with permission from [99]. Copyright 2021, Elsevier.
Figure 7. (a) De-NOx activity as a functions of reaction time at 350 °C of La0.7Sr0.3MnO3 (▲), Pd/La0.7Sr0.3MnO3-sol-gel (■), Pd/La0.7Sr0.3MnO3-impregnation (●), and pretreated Pd/La0.7Sr0.3MnO3--ol-gel (▼); and (b) illustrative profile of in situ activation of Pd/La0.7Sr0.3MnO3-sol-gel catalyst. Adapted with permission from [99]. Copyright 2021, Elsevier.
Catalysts 12 00593 g007
Figure 8. Outlet NOx concentration at 350 °C under lean conditions for fresh, sulfated, and regenerated (R) catalysts: (a) LaCo0.92Pt0.08O3 and (b) Pt/BaO/Al2O3. Reprinted with permission from [66]. Copyright 2014, American Chemical Society.
Figure 8. Outlet NOx concentration at 350 °C under lean conditions for fresh, sulfated, and regenerated (R) catalysts: (a) LaCo0.92Pt0.08O3 and (b) Pt/BaO/Al2O3. Reprinted with permission from [66]. Copyright 2014, American Chemical Society.
Catalysts 12 00593 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhao, D.; Song, H.; Liu, J.; Jiang, Q.; Li, X. Advances in Designing Efficient La-Based Perovskites for the NOx Storage and Reduction Process. Catalysts 2022, 12, 593. https://doi.org/10.3390/catal12060593

AMA Style

Zhao D, Song H, Liu J, Jiang Q, Li X. Advances in Designing Efficient La-Based Perovskites for the NOx Storage and Reduction Process. Catalysts. 2022; 12(6):593. https://doi.org/10.3390/catal12060593

Chicago/Turabian Style

Zhao, Dongyue, Haitao Song, Jun Liu, Qiuqiao Jiang, and Xingang Li. 2022. "Advances in Designing Efficient La-Based Perovskites for the NOx Storage and Reduction Process" Catalysts 12, no. 6: 593. https://doi.org/10.3390/catal12060593

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