Next Article in Journal
Photocatalytic Degradation of Sulfolane Using a LED-Based Photocatalytic Treatment System
Previous Article in Journal
Electrocatalysis for Oxygen Reduction Reaction on EDTAFeNa and Melamine co-Derived Self-Supported Fe-N-C Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review on the Catalytic Decomposition of NO by Perovskite-Type Oxides

Marine Engineering College, Dalian Maritime University, Dalian 116000, China
*
Author to whom correspondence should be addressed.
Submission received: 12 April 2021 / Revised: 6 May 2021 / Accepted: 10 May 2021 / Published: 12 May 2021

Abstract

:
Direct catalytic decomposition of NO has the advantages of being a simple process, producing no secondary pollution, and being good for the economy, which has attracted extensive research in recent years. Perovskite-type mixed oxides, with an ABO3 or A2BO4 structure, are promising materials as catalysts for NO decomposition due to their low cost, high thermal stability, and, of course, their good catalytic performances. In this review, the influence factors, such as A-site substitution, B-site substitution and reaction conditions on the catalytic performance of catalysts have been expounded. The reaction mechanisms of direct NO decomposition are also discussed. Finally, major conclusions are drawn and some research challenges are highlighted.

1. Introduction

NOx is one of the main pollutants that can seriously endanger human health and the ecological environment. It is the main pollutant in the formation of acid rain: NOx and HC react chemically under the action of sunlight to generate photochemical smog. NOx can destroy the ozone layer causing global climate change, therefore, the harm caused by NOx cannot be ignored [1,2].
Direct catalytic decomposition of NO has the advantages of being a simple process, producing no secondary pollution, and being good for the economy, which has attracted extensive research in recent years [3,4].
2NO(g) = N2(g) + O2(g)
Δ H f ( 298 K ) 0 = 90.2   kJ · mol 1
As can be seen from the above equation, the reaction is thermodynamically favorable. This decomposition reaction has highly negative Gibbs free energy (∆rGm = −86 kJ/mol), and the tendency of NO decomposition to generate N2 and O2 is large. Although it is thermodynamically feasible, the reaction needs to overcome the high activation energy of the reaction (~335 kJ/mol−1) [5,6]. Accordingly, NO can decompose smoothly under a catalyst and certain temperature conditions, and the development of greatly active and stable catalysts have always been a popular research direction and challenging task. The process is free of secondary pollution and does not require a reducing agent, which largely reduces the operating costs.
Most of the studies of NO direct catalytic decomposition catalysts focus on six types of catalysts, including noble metals, metal oxides, perovskite-type mixed oxides, ion-exchange ZSM-5 type molecular sieves, multi-compounds, and hydrotalcite type materials [6]. Although some noble metals and metal oxides have high activity, there is catalyst poisoning during NO decomposition. Besides, metal oxide catalysts are also very sensitive to SO2. When SO2 is present it is easy to generate sulfate, which will poison and deactivate the catalyst. Noble metals are expensive and resources are scarce, whereas perovskite-type mixed oxides have the advantages of good thermal stability and stable oxygen vacancies on the surface. Therefore, the removal of NO by perovskite-type composite oxides has become one of the research hotspots of catalytic decomposition denitrification technology. The research progress and the current state of direct catalytic decomposition of NO by perovskite-type oxides will be reviewed, and the future developmental direction will be discussed.

2. Mechanism of NO Decomposition on Perovskite-Type Oxides

The perovskite-type oxide structure is ABO3, which is a cubic crystal with A-site or B-site cations as nodes, and is consistent with the perovskite (CaTiO3) structure. In general, the metal ions at position A are mostly rare earth metal, alkaline earth metal, and lanthanide metal, which are located at the center of the structure and coordinate with the surrounding 12 oxygen atoms. The metal ions at position B are mostly transition metal ions, which are located in the body center of the octahedral structure and coordinate with the surrounding six oxygen atoms [7]. Among them, metal ion B determines the catalytic performance of the catalyst, and the role of metal ion A is mainly to stabilize the crystal structure. The ideal crystal structure of ABO3 perovskite is shown in Figure 1.
There are defects in perovskite oxide crystals like most crystals. According to the electric neutral principle, ABO3 can be composed of three kinds: A+1B+5O3, A+2B+4O3, A+3B+3O3. In the perovskite structure, the larger radius of the A ion can be partially or completely lost, so A-site ion defects are common. The B-site ions have a relatively large charge and small ionic radius. From a thermodynamic point of view, the B-site is unfavorable. There are several ways to generate ionic defects:
(1)
Substitution of high valence ions (rare earth elements) with low valence ions A′ (generally alkaline earth metals). According to the principle of electroneutrality, high valence B-site ions and oxygen vacancies can be generated;
(2)
Substitution of low valence ion A with high valence ion A′ can generate low B valence ion and A vacancy;
(3)
Substitution of high valence B ions with low valence B′ ions.
The changes and catalytic reactions of various valence states and oxygen defects of A and B above can generally occur simultaneously.
Zhu et al. [8] believed that oxygen vacancy played a crucial role in the decomposition of NO, and the active center of NO decomposition was composed of oxygen vacancy and transition metal with low oxidation state. A large number of oxygen vacancies generally exist in the perovskite-type oxide structure, which provides space and convenience for the adsorption and decomposition of NO. The adsorbed NO obtains an electron from the transition metal with a lower oxidation state. Meanwhile, it is activated by the electron to participate in the reaction.
For the decomposition mechanism of NO on perovskite-type catalysts, Shin et al. [9] thought that, at first, two molecules of NO attached on the Fe3+ with some oxygen vacancies, and then the NO molecules coordinated with the iron ions. At the same time, the iron atoms changed their valence state from Fe3+ to Fe4+, giving one electron per one coordinating NO molecule. The N–O bond in the coordinating nitrosyl ion (NO) is so weak that cleavage of the N–O bond may occur, yielding a free molecule of N2. Subsequently, the thermal energy supplied to the reaction system at a certain temperature will cause the release of the oxygen atoms from the Fe4+. Then, an oxygen molecule will be also isolated from the reaction system, and again the Fe3+, with accompanying oxygen vacancies, are regenerated, and the reaction mechanism is shown in Figure 2. This scheme is thought to be credible for the thermal decomposition of NO over an oxygen-deficient perovskite SrFeO3−x because the unique lattice of SrFeO3−x itself easily stabilizes an unusual valence state of +4 for the irons, and it also tends to accommodate appreciable oxygen defects with increasing temperature.
Teraoka et al. [10] proposed the reaction mechanism of direct decomposition of NO on perovskite oxides. A pair of NO molecules was first adsorbed onto the adjacent oxygen vacancies on the surface of the catalyst, and then the two adjacent N–O bonds were dissociated to form a free N2 molecule, and finally released into the gas phase, as shown in Figure 3. Since the adsorption of the second NO molecule at the paired position is a rate-determined step, the important characteristics of this reaction are NO adsorption capacity and oxygen ion vacancy. Besides, the cation activity of the B-site greatly affects the catalytic activity, because B-site ions are adjacent to oxygen anions, and O2 desorption plays an important role in the regeneration of cation vacancies.
Zhu et al. [11] explored the decomposition mechanism of NO on perovskite-like oxides by studying the O2-TPD experiment and NO decomposition activity. The NO was first absorbed and dissociated into N2 and atomic oxygen, which then reacted with another NO to form the adsorbed NO2 and, after that, the adsorbed NO2 desorbed and dissociated into NO and O2. In this process, NO2 acted as an intermediate of O2 formation. The results showed that the decomposition of NO on perovskite-like oxides was carried out in a cycle manner, in which the cycle (generation and decomposition) of NO2 played an important role in the reaction and was closely related to the activity.
(a) NO2 is mainly generated on the surface of the catalyst:
NO(g) + O(a) = NO2(a)
(b) Oxygen desorption step:
O(a) + O(a) = O2(g)
NO2 dissociation reaction (2NO2 = 2NO + O2) is an important step in the process of NO decomposition as the catalytic activity of NO decomposition on perovskite-like oxides only occurs at high temperatures. The cyclic decomposition mechanism of NO on perovskite-like oxides with NO2 as intermediate species is shown in Figure 4.
A large number of oxygen vacancies or oxygen excess in perovskite-like oxides favor the decomposition of NO; NdSrCu1−xCoxO4−δ has a higher density of oxygen vacancies and a stronger Cu3+/Cu2+ redox capacity, which is more likely to activate NO and is responsible for the enhancement of catalytic performance [12,13,14].
Perovskite composite oxides have good thermal stability and oxygen defects, so they have been considered promising catalysts for catalytic NO decomposition [15]. On the one hand, oxygen vacancies provide vacancies for NO adsorption, while on the other, electrons generated in the structure can activate NO. The existence of oxygen vacancies and redox properties of B-site ions are two main factors affecting NO adsorption and activation, which play an important role in the reaction process. In recent years, perovskite catalysts have become a research hotspot in the field of catalytic decomposition of NO due to their good direct decomposition performance [16,17].
It is generally believed that the catalytic activity mainly depends on the B-site ions, while the A-site ions mainly play a role in stabilizing the crystal structure. However, the substitution of A-site atoms by other valence atoms can change the valence of B-site ions and the content of oxygen vacancies in the structure. For example, the content of Fe2+ and Fe3+ ions and oxygen vacancy (δ) in La1−xSrxFeO3+δ can be changed by the substitution of Sr2+ with La3+. Due to the unique structural characteristics of perovskite, the original crystal structure of perovskite will not be destroyed by doping different valence metal ions at A- or B-site within the allowable range of the tolerance factor. Therefore, the decomposition activity of NO can be improved by doping different metal ions [18].

3. Effect of A-Site Substitution

The A-site ions of perovskite are larger radius ions. The ions on the A-site will indirectly promote catalytic activity. When the A-site ions are substituted, there are more defects produced and they will affect the electronic state of the B-site ions [19]. Meanwhile, A-site doping can change the oxidation state of B-site ions, and B-site ions are easily reduced to highly disperse metal centers [20,21,22]. A- and B-sites play different roles in the reaction process. B-site mainly affects the temperature range of the reaction, and A-site mainly affects the selectivity of the reaction. Higher lattice oxygen mobility and lower active oxygen adsorption are also the key factors to determining catalyst activity. Divalent alkaline earth ions are often used to partially replace the A-site ions to increase the alkalinity and generate anion oxide vacancies through charge compensation.

3.1. Effect of Dopants

Zhu et al. [23] found that if external A-site cations could not change the oxidation state of B-site cations, the effect of A-site substitution on catalytic performance was indirect, i.e., by changing the oxidation state of B-site cations or generating oxygen vacancies. This indicates that although Sr2+ had no activity on the decomposition of NO, it could affect the reaction rate by changing the average oxidation of B-site cations.
Yasuda et al. [24] studied the effect of Sr substitution on NO decomposition by La2−xSrxCuO4 and demonstrated that the substitution influence was caused by changing the average oxidation state of Cu. The catalytic activity is closely related to the average oxidation state of Cu, and its value is closely related to the amount of Sr2+ substitution. He et al. [25] prepared La1−xSrxFeO3 perovskite composite oxides. After La was partially replaced by Sr, more oxygen vacancies were generated and the catalytic performance was improved. In recent years, barium has been considered as a promising A-site ion for perovskite catalysts because it not only has a low oxidation state (+2), it also has the potential to capture NOx, thus ensuring the supply of NOx in the reaction.
An A-site cation does not necessarily have no catalytic activity, but as long as the A-site cation has a variable oxidation state, it can indeed participate in the reaction process. Notably, Ce enters the structural framework of La1−xCexSrNiO4 up to 30% (x = 0.3), which is much higher than the reported ABO3 structure of perovskite-type oxides (La1−xA′xBO3 is usually below 10%) [26,27,28]. Belessi et al. [29] found that the substitution of Ce in La0.5SrxCeyFeOz was very difficult. No perovskite oxide containing Ce was detected in this system.
Zhu et al. [30] added Sr in La2CuO4 and the NO decomposition conversion rate was low (34.3%, t = 850 °C). For a La2−xThxCuO4 (0 ≤ x ≤ 0.4) catalyst, Zhu et al. [22] showed that Cu+ was easily oxidized in the redox cycle of Cu+↔Cu2+, so that the active adsorption rate of Cu+ at 850 °C was the highest, reaching about 42%. Chen et al. [31] prepared a La1.6Ba0.4NiO4-x%BaO (x = 0, 5, 10, 15, 20, 25, 30) catalytic by heating a mixture of Ba(NO3)2 and La1.6Ba0.4NiO4, and evaluated their catalytic performance. The results showed that the activity of the La1.6Ba0.4NiO4-x%BaO catalysts increased with the increase of the amount of BaO.
Iwakuni et al. [32] showed that La1.6Ba0.4NiO4-20% BaO had the highest activity, and the yield of nitrogen increased from 51.7% to 98.8% at 600–850 °C. The results showed that the addition of BaO increased the number of oxygen vacancies and the mobility of lattice oxygen. Besides, BaO played an important role in the transport and storage of NO and facilitated the reconstruction of active centers. Zhang et al. [33] prepared a series of perovskite-type La1−xCexFeO3 (x = 0–0.5) nanocomposites by sol-gel method and studied the effect of Ce doping ratio on NO conversion rate. When the doping amount is 0.3, the conversion of NO can reach 80% at low temperatures.
Li et al. [34] prepared a series of perovskite La1−xCexMnO3 (x = 0–0.2) nanocomposites by sol-gel method and studied the effect of a Ce4+ doping ratio on NO conversion. The results showed that the catalytic activity of active perovskite species was improved by adding the appropriate amount of Ce4+ to the La3+ site.
Dong et al. [35] prepared a series of La0.7Sr0.3MnO perovskite catalysts by the sol-gel method under different synthesis conditions. The results showed that the initial pH value of the precursor solution had a great influence on the shape and size of the catalyst, which directly related to the NOx storage capacity of the catalyst. Also, the oxidation capacity of the NO was determined by the amount of excess oxygen in the perovskite.
Gao et al. [36] further found that after a reduction in H2 (H2/He = 1/9, flow rate: 20 mL·min−1) at 300 °C, the catalytic performance of La1.867Th0.100CuO4 was effectively improved, and 91% NO conversion (converted to N2) was obtained at 650 °C.
In addition, sometimes the substitution of A-site cations changes the crystalline phase from an ordered structure to a disordered structure, thus affecting the catalytic performance. Shin et al. [25] reported that Ca2Fe2O5+δ with ordered oxygen vacancies had no activity to decompose NO even at 900 °C, while Sr was substituted for Ca, i.e., Sr2Fe2O5+δ with disordered oxygen vacancies. Besides, a Ce3+/4+ partial substitution of site ions facilitated the regeneration of the active site due to the redox properties of the cerium ion, which can release oxygen adsorbed on the oxide anion defects. This redox effect is more obvious in K2NiF4 perovskite oxides deposited alternately between the ABO3 layer and AO layer, but not in ABO3 type oxides. Because the cavity space of the K2NiF4 lattice is larger than that of the ABO3 lattice, larger Ce ions can replace some A-site ions. Besides, Sr2+ can compensate charge at the B-site.

3.2. Effect of Coexisting Gases

In the absence of coexisting gases, the N2 yield of the NO decomposition of a perovskite catalyst was over 70%. However, the reaction was strongly inhibited by the presence of O2 and CO2. Ba0.8La0.2Mn0.8Mg0.2O3, the most active catalyst among the ABO3-type oxides, gave about 80% N2 yield for NO decomposition at 850 °C and in the absence of coexisting gases. In contrast, under the presence of 5% CO2, the concentration was reduced to less than 20% due to the dissociative adsorption of CO2 on the catalyst surface and the strong adsorption of the resulting CO on the active site [37]. CO2 poisoning has an enormous negative impact on the catalytic performance of NO decomposition. In addition, when the catalyst was heated at 700 °C, the adsorbed CO molecules were desorbed from the surface, without desorption at a lower temperature of 600 °C.
For La0.8Sr0.2CoO3, CO2 inhibited NO decomposition through carbonate formation, which is different from the poisoning mechanism observed in Ba0.8La0.2Mn0.8Mg0.2O3, although they are similar perovskite oxides [38]. A La0.8Sr0.2CoO3 catalyst needs to be heated above 750 °C to decompose carbonate. Therefore, to effectively decompose NO in the presence of CO2, a new and advanced catalyst is needed, which can inhibit the reaction with CO2 and limit the adsorption of CO2.

4. Effect of B-Site Substitution

The substitution effect of B-site cations is much more complex than that of A-site cations. Because B-site cations can not only change their oxidation state, but also an important part of the active site.

4.1. Effect of Dopants

LaMO3 lanthanide is a common B-site substituted perovskite system. It is generally believed that La3+ mainly fixes the crystal structure of the catalyst, and the main catalytic effect is controlled by B-site ions. Yokoi and Uchida [39] studied the reactivity of different B-site cations to NO decomposition directly in LaMO3 (M = Cr, Mn, Fe, Co, and Ni).
Figure 5 compares the NO decomposition catalytic activity of some LaMO3-based perovskite [30,31,32,33,34,35,36,37,38,39,40,41,42,43].
For BaMnO3-based perovskite oxides, Iwakuni et al. [37] studied the effect of dopants on NO direct decomposition activity. Goto et al. [44] investigated the effect of B-site substitution on the NO decomposition activity of SrFe0.7M0.3O3 activity in the order of Mg > Sn > Ni > Ce > Zr. The NO decomposition activity was Mg > Sn > Ni > Ce > Zr. Notably, Kazuya [45] found that Ba3Y3.4Sc0.6O9 with a perovskite structure composed of alkaline earth metals and rare earth ions exhibited high NO decomposition activity in the temperature range of 600–850 °C.
The application of lanthanide-free perovskite catalysts in NO decomposition reactions has been studied by many researchers. Traditional three-way catalysts (TWCs) in reducing NOx, CO, and HC are very effective, considering the need for the development of high-efficiency catalyst. Glisenti et al. [46] studied the effect of rare-earth element (typically lanthanum) perovskite oxides on the activity of the catalyst: Co and Cu were doped active cations and the results showed that Co appeared to be active in oxidation, while Cu was necessary for reduction. Strontium titanates (Sr1−xKxTiO3, x = 0.1–0.5), partially replaced with potassium, were prepared by the citric acid sol-gel method [47]. The promoting effect of potassium on the formation of oxygen vacancy and surface alkalinity was discussed, in which molecular oxygen or (NO, NO2) was absorbed to form basic surface oxygen species soot oxidation.
Carlotto et al. [48] studied the NO reduction in a CO-NO atmosphere at the Co and Cu-doped SrTiO3 surface by density function theory. They found that the step to determine the rate is the formation of oxygen vacancy, which is beneficial to the doping of Cu- and Co. Due to its higher ability to stabilize oxygen vacancy, the catalytic performance of the Cu-doped surface was improved at the same time. While the solubility of Cu was low in SrTiO3, Cu-doping is predicted to be particularly effective.

4.2. Effect of Coexisting Gases

To improve the oxygen resistance of the catalyst, Zhu et al. [49] added an appropriate amount of Ce in LaSrNiO4. Among these La1−xCexSrNiO4, La0.7Ce0.3SrNiO4 showed high direct decomposition activity at 900 °C even in the presence of 6.0% O2. At 800 °C, the yield of nitrogen reached about 75%. In the presence of 1.0% and 4.0% oxygen, the yield of nitrogen reached 50.5% and 39.8%, respectively [50]. The results of Zhu et al. [51] showed that the NO decomposition activity of laser crystal could be improved by replacing manganese with nickel. LaSrMn1−xNixO4+δ (x = 0.8) had the highest nitrogen yield at about 85% at 850 °C and about 75% even in the presence of 2.5% oxygen. XPS characterization showed that the stronger the catalyst bonded to oxygen, the lower the catalytic decomposition activity. Ma et al. [52] studied nickel-containing A2BO4 oxide catalysts. Through research, it was found that these catalysts have excellent catalytic performance at low temperatures, which is very suitable for simultaneous removal of NOx and diesel soot.
Gao et al. [53] synthesized a special structure of La2CuO4 nanofibers using single-walled carbon nanotubes (CNT) as a template. At temperatures above 300 °C, 100% NO conversion could be obtained without other by-products. Wang et al. [54] combined a BaBi0.05Co0.8Nb0.15O3−δ (BBCN) hollow fiber membrane with a nickel silicate hollow sphere catalyst to fabricate a novel bifunctional catalytic peroxide hollow fiber membrane reactor. Using this new catalytic membrane reactor, the oxygen in the NO decomposition reaction system was instantaneously removed, and NO was completely converted to N2 at low temperatures (675 °C). Dhal et al. [55] applied the solution of combustion synthesis technology to the production of LaFeO3 and improved its NO purification activity by increasing the specific surface area of LaFeO3 with high exothermic and self-sustaining reactions.
Using these efficient composite materials, combined with the existing knowledge of structure, the catalytic activity of perovskite oxide was analyzed. It is hoped that an efficient perovskite oxide catalyst can be successfully found to remove NO and meet the requirements of a large-scale application.
As can be seen from Table 1, the BaMnO3-based perovskite shows low catalytic activity. Meanwhile, the catalytic activity increased after loading Pt, indicating that the adsorption strength of surface oxygen was weakened by Pt. The adsorption capacity and efficiency of oxygen vacancy could be improved by doping Ba in perovskite catalyst. The difference of the activity between Ce-doped perovskite in the presence of 1% O2 and the absence of oxygen was very slight, which indicates that the Ce-doped perovskite sample had a strong resistance to oxidation. The direct decomposition capacity of perovskite oxide doped with Sr was greatly affected by Fe doping. Fe was replaced by a low valence cation in SrFeO3, and a small amount of Pt was loaded at the same time, which was very effective for a decomposition reaction of NO.
In conclusion, there has been much research done on the doping of a perovskite oxide catalyst with La. The high activity of a perovskite oxide catalyst can be obtained by doping La and the catalytic performance of the catalyst can be improved by doping Ba, Ni, Ce, etc.

5. Conclusions

The regeneration of surface oxygen defects in the perovskite catalyst system is an important factor affecting the reaction; the oxygen defects in peroxide-based catalysts are more stable and easily regenerated at high temperatures. However, due to its stability, the initial activity of NO catalytic decomposition on perovskite-type oxide catalysts is also relatively high. On the other hand, because the formation of perovskite requires a high temperature and long calcination time, while the specific surface area of perovskite oxide is small, it can limit the catalytic ability of the catalyst in catalytic reactions. In general, catalysts require a relatively large specific surface area to effectively adsorb and activate the reactants. Presently, work in this area has focused on changing the composition, structure, and specific surface area of the system to obtain catalysts with high activity for NO decomposition at low temperatures. Moreover, the SO2 resistance of perovskite-type oxides catalysts needs to be further improved.
Perovskite oxide is a promising NO removal catalyst. Its high activity and low cost make it competitive. Therefore, the search for an efficient perovskite-type catalyst is a subject of great interest. In this paper, the structure, catalytic mechanism and research status of perovskite oxides were reviewed, and the developmental direction of perovskite oxides was discussed. The possibility of modifying the catalyst and the method of screening the active catalyst was also indicated.

Author Contributions

Conceptualization, Q.S.; data collection, S.D.; writing, Q.S. and S.D.; figures, S.D.; tables, S.D.; supervision, Q.S.; literature search, S.L.; review and editing, S.L., G.Y. and X.P.; funding acquisition, Q.S., S.L. and G.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by THE NATIONAL NATURAL SCIENCE FOUNDATION OF CHINA OF FUNDER, grant number 52001045, CHINA POSTDOCTORAL SCIENCE FOUNDATION OF FUNDER, grant number 2019M651097 and 2019M651094, NATURAL SCIENCE FOUNDATION OF LIAONING PROVINCE OF FUNDER, grant number 2019-ZD-0154 and 2020-HYLH-38, and DA-LIAN CITY INNOVATIVE SUPPORT PROGRAM FOR HIGH-LEVEL TALENTS OF FUNDER, grant number 2019RQ036. The APC was funded by THE FUNDAMENTAL RESEARCH FUNDS FOR THE CENTRAL UNIVERSITIES OF CHINA OF FUNDER, grant number 3132019327.

Acknowledgments

This work was supported by the National Natural Science Foundation of China.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fino, D.; Russo, N.; Saracco, G.; Specchia, V. Catalytic removal of NOx and diesel soot over nanostructured spinel-type oxides. J. Catal. 2006, 242, 38–47. [Google Scholar] [CrossRef]
  2. Reichert, D.; Finke, T.; Atanassova, N.; Bockhorn, H.; Kureti, S. Global kinetic modelling of the reaction of soot with O2 and NOx on Fe2O3 catalyst. Appl. Catal. B Environ. 2008, 84, 803–812. [Google Scholar] [CrossRef]
  3. 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]
  4. Xie, S.; Rosynek, M.P.; Lunsford, J.H. NOx storage on barium-containing three-way catalyst in the presence of CO2. J. Catal. 2001, 72, 59–64. [Google Scholar]
  5. Wise, H.; Frech, M.F. Kinetics of Decomposition of Nitric Oxide at Elevated Temperatures. I. Rate Measurements in a Quartz Vessel. J. Chem. Phys. 1952, 20, 22. [Google Scholar] [CrossRef]
  6. Sun, Q.; Wang, Z.; Wang, D.; Hong, Z.; Zhou, M.; Li, X. A review on the catalytic decomposition of NO to N2 and O2: Catalysts and processes. Catal. Sci. Technol. 2018, 8, 4563–4575. [Google Scholar] [CrossRef]
  7. Zhu, J.; Thomas, A. Perovskite-type mixed oxides as catalytic material for NO removal. Appl. Catal. B Environ. 2009, 92, 225–233. [Google Scholar] [CrossRef]
  8. Carlotto, S.; Natile, M.M.; Glisenti, A.; Vittadini, A. Catalytic Mechanisms of NO Reduction in a CO–NO Atmosphere at Co- and Cu-Doped SrTiO3(100) Surfaces. J. Phys. Chem. C 2017, 122, 449–454. [Google Scholar] [CrossRef]
  9. Shin, S.; Arakawa, H.; Hatakeyama, Y.; Ogawa, K.; Shimomura, K. Absorption of NO in the lattice of an oxygen-deficient perovskite SrFeO3−x and the infrared spectroscopic study of the system NO-SrFeO3−x. Mater. Res. Bull. 1979, 14, 633–639. [Google Scholar] [CrossRef]
  10. Teraoka, Y.; Harada, T.; Kagawa, S. Reaction mechanism of direct decomposition of nitric oxide over Co- and Mn-based perovskite-type oxides. J. Chem. Soc. Faraday Trans. 1998, 94, 1887–1891. [Google Scholar] [CrossRef]
  11. Zhu, J.; Xiao, D.; Li, J.; Xie, X.; Yang, X.; Wu, Y. Recycle—New possible mechanism of NO decomposition over perovskite(-like) oxides. J. Mol. Catal. A Chem. 2005, 233, 29–34. [Google Scholar] [CrossRef]
  12. Zhu, J.J.; Yang, X.G.; Xu, X.L. Active site structure of NO decomposition on perovskite(-like) oxides: An investigation from experiment and density functional theory. J. Phys. Chem. C 2007, 111, 1487–1490. [Google Scholar] [CrossRef]
  13. Xu, W.; Yu, Y.; Zhang, C.; He, H. Selective catalytic reduction of NO by NH3 over a Ce/TiO2 catalyst. Catal. Commun. 2008, 9, 1453–1457. [Google Scholar] [CrossRef]
  14. Wang, Z.; Jiang, Z.; Shangguan, W. Simultaneous catalytic removal of NOx and soot particulate over Co-Al mixed oxide catalysts derived from hydrotalcites. Catal. Commun. 2007, 8, 1659–1664. [Google Scholar] [CrossRef]
  15. Xu, P.; Tu, R.; Zhang, S.; Yang, M.; Li, Q.; Goto, T.; Zhang, L. Catalytic Decomposition of Nitric Oxide by LaCoO3 Nano-particles Prepared by Rotary CVD. J. Wuhan Univ. Technol. Sci. Ed. 2018, 33, 368–374. [Google Scholar] [CrossRef]
  16. Imanaka, N.; Masui, T. Advances in direct NOx decomposition catalysts. Appl. Catal. A Gen. 2012, 431, 1–8. [Google Scholar] [CrossRef]
  17. Royer, S.; Duprez, D.; Can, F.; Courtois, X.; Batiot-Dupeyrat, C.; Laassiri, S.; Alamdari, H. ChemInform Abstract: Perovskites as Substitutes of Noble Metals for Heterogeneous Catalysis: Dream or Reality. Am. Chem. Soc. 2014, 114, 10292–10368. [Google Scholar] [CrossRef]
  18. Hu, J.; Ma, J.; Wang, L.; Huang, H. Synthesis and photocatalytic properties of LaMnO3–graphene nanocomposites. J. Alloys Compd. 2014, 583, 539–545. [Google Scholar] [CrossRef]
  19. Yang, P.; Li, N.; Teng, J.J.; Wu, J.; Ma, H. Influence of A-site cation in La0.7M0.3Ni0.7Fe0.3O3 (M = Pr, Y, Sr, Zr, Ce) perovskite-type oxides on the catalytic performance. Chem. Res. Appl. 2018, 30, 1979–1985. [Google Scholar]
  20. Biendicho, J.J.; Shafeie, S.; Frenck, L.; Gavrilova, D.; Böhme, S.; Bettanini, A.M.; Svedlindh, P.; Hull, S.; Zhao, Z.; Istomia, Z.Y.; et al. Synthesis and characterization of perovskite-type SrxY1-xFeO3-δ, (0.63 ≤ x < 1.0) and Sr0.75Y0.25Fe1-yMyO3-δ, (M = Cr, Mn, Ni), (y = 0.2, 0.33, 0.5). J. Solid State Chem. 2013, 200, 30–38. [Google Scholar]
  21. Kabra, S. Perovskite LaxM1−xNi0.8Fe0.2O3catalyst for steam reforming of toluene: Crucial role of alkaline earth metal at low steam condition. Appl. Catal. B Environ. 2014, 148, 231–242. [Google Scholar]
  22. Su, H.-Y.; Sun, K. DFT study of the stability of oxygen vacancy in cubic ABO3 perovskites. J. Mater. Sci. 2014, 50, 1701–1709. [Google Scholar] [CrossRef]
  23. Zhu, J.J.; Yang, X.G.; Xu, X.L. Effect of strontium substitution on the activity of La2−xSrxNiO4 (x = 0.0–1.2) in NO decomposition. Sci. China Ser. B Chem. 2007, 50, 41–46. [Google Scholar] [CrossRef]
  24. Yasuda, H.; Mizuno, N.; Misono, M. Role of valency of copper in the direct decomposition of nitrogen monoxide over well characterized La2xA′xCu1-yB′yO4. Chem. Soc. Chem. Commun. 1990, 16, 1094–1096. [Google Scholar] [CrossRef]
  25. He, F.; Li, X.; Zhao, K.; Huang, Z.; Wei, G.; Li, H. The use of La1−xSrxFeO3 perovskite-type oxides as oxygen carriers in chemical-looping reforming of methane. Fuel 2013, 108, 465–473. [Google Scholar] [CrossRef]
  26. Kirchnerova, J.; Alifanti, M.; Delmon, B. Evidence of phase cooperation in the LaCoO3–CeO2–Co3O4 catalytic system in relation to activity in methane combustion. Appl. Catal. A Gen. 2002, 231, 65–80. [Google Scholar] [CrossRef]
  27. Belessi, V.C.; Costa, C.N.; Baka, T.V. Catalytic behavior of La-Sr-Ce-Fe-O mixed oxidic/perovskitic systems for the NO+CO and NO+CH4+O2 (lean-NOx) reactions. Catal. Today 2000, 59, 347–363. [Google Scholar] [CrossRef]
  28. Zhao, B.; Wang, R.; Yang, X.X. Simultaneous catalytic removal of NOx and diesel soot particulates over La1-xCexNiO3 perovskite oxide catalysts. Catal. Commun. 2009, 10, 1029–1033. [Google Scholar] [CrossRef]
  29. Shin, S.; Hatakeyama, Y.; Ogawa, K.; Shimomura, K. Catalytic decomposition of NO over brownmillerite-like compounds, Ca2Fe2O5 and Sr2Fe2O5. Mater. Res. Bull. 1979, 14, 133–136. [Google Scholar] [CrossRef]
  30. Zhu, J.; Zhao, Z.; Xiao, D.; Li, J.; Yang, X.; Wu, Y. CO Oxidation, NO Decomposition, and NO + CO Reduction over Perovskite-like Oxides La2CuO4 and La2-xSrxCuO4: An MS−TPD Study. Ind. Eng. Chem. Res. 2005, 44, 4227–4233. [Google Scholar] [CrossRef]
  31. Chen, L.; Niu, X.; Li, Z.; Dong, Y.; Wang, N.; Yuan, F.; Zhu, Y. The effects of BaO on the catalytic activity of La1.6Ba0.4NiO4 in direct decomposition of NO. J. Mol. Catal. A Chem. 2016, 423, 277–284. [Google Scholar] [CrossRef]
  32. Iwakuni, H.; Shinmyou, Y.; Yano, H.; Matsumoto, H.; Ishihara, T. Direct decomposition of NO into N2 and O2 on BaMnO3-based perovskite oxides. Appl. Catal. B Environ. 2007, 74, 299–306. [Google Scholar] [CrossRef]
  33. Zhang, H.; Li, X.; Hui, Y.; Yu, L.; Xia, Q.; Luo, S.; Yao, C. Development of La1−xCexFeO3/attapulgite nanocomposites for photocatalytic reduction of NO at low temperature. J. Mater. Sci. Mater. Electron. 2017, 28, 9371–9377. [Google Scholar] [CrossRef]
  34. Li, X.; Yin, Y.; Yao, C.; Zuo, S.; Lu, X.; Luo, S.; Ni, C. La1−Ce MnO3/attapulgite nanocomposites as catalysts for NO reduction with NH3 at low temperature. Particuology 2016, 26, 66–72. [Google Scholar] [CrossRef]
  35. 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]
  36. Gao, L.; Au, C. Studies on the redox behaviour of La1.867Th0.100CuO4 and its catalytic performance for NO decomposition. Catal. Lett. 2000, 65, 91–98. [Google Scholar] [CrossRef]
  37. Iwakuni, H.; Shinmyou, Y.; Yano, H.; Goto, K.; Matsumoto, H.; Ishihara, T. Effects of Added CO2and H2 on the Direct Decomposition of NO over BaMnO3-Based Perovskite Oxide. Bull. Chem. Soc. Jpn. 2008, 81, 1175–1182. [Google Scholar] [CrossRef]
  38. Teraoka, Y.; Torigoshi, K.; Kagawa, S. Inhibition of NO Decomposition Activity of Perovskite-type Oxides by Coexisting Carbon Dioxide. Bull. Chem. Soc. Jpn. 2001, 74, 1161–1162. [Google Scholar] [CrossRef]
  39. Yokoi, Y.; Uchida, H. Catalytic activity of perovskite-type oxide catalysts for direct decomposition of NO: Correlation between cluster model calculations and temperature-programmed desorption experiments. Catal. Today 1998, 42, 167–174. [Google Scholar] [CrossRef]
  40. Zhao, Z.; Yang, X.; Wu, Y. Comparative study of Nickel-based perovskite-like mixed oxide catalysts for direct decomposition of NO. Appl. Catal. B Environ. 1996, 8, 281–297. [Google Scholar] [CrossRef]
  41. Zhu, Y.; Wangm, D.; Yuan, F.; Zhang, G.; Fu, H. Direct NO decomposition over La2xBaxNiO4 catalysts containing BaCO3 phase. Appl. Catal. B Environ. 2008, 82, 255–263. [Google Scholar] [CrossRef]
  42. Zhu, J.J.; Zhao, Z.; Xiap, D.; Li, J.; Yang, X.; Wu, Y. Effect of valence of copper in La2-xThxCuO4 on NO decomposition reaction. Catal. Commun. 2006, 7, 29–32. [Google Scholar] [CrossRef]
  43. Iwakuni, H.; Shinmyou, Y.; Matsumoto, H.; Ishihara, T. Direct Decomposition of NO into N2 and O2 on SrFe0.7Mg0.3O3 Perovskite Oxide. Bull. Chem. Soc. Jpn. 2007, 80, 2039–2046. [Google Scholar] [CrossRef]
  44. Goto, K.; Matsumoto, H.; Ishihara, T. Direct decomposition of NO on Ba/Ba-Y-O catalyst. Top. Catal. 2009, 52, 1776–1780. [Google Scholar] [CrossRef]
  45. Xu, W.; Cai, J.; Zhou, J.; Ou, Y.; Long, W.; You, Z.; Luo, Y. Highly Effective Direct Decomposition of Nitric Oxide by Microwave Catalysis over BaMeO3 (Me = Mn, Co, Fe) Mixed Oxides at Low Temperature under Excess Oxygen. ChemCatChem 2015, 8, 417–425. [Google Scholar] [CrossRef]
  46. Glisenti, A.; Natile, M.M.; Carlotto, S.; Vittadini, A. Co- and Cu-Doped Titanates: Toward a New Generation of Catalytic Converters. Catal. Lett. 2014, 144, 1466–1471. [Google Scholar] [CrossRef]
  47. Ura, B.; Trawczyński, J.; Zawadzki, M.; Gomez, M.I.; López, A.B.; Suárez, F.L. Sr1-xKxTiO3 catalysts for diesel soot combustion. Catal. Today 2011, 176, 169–172. [Google Scholar] [CrossRef]
  48. Zhu, J.J.; Xiao, D.H.; Li, J. Mechanism of NO decomposition on perovskite (-like) catalysts. Chin. Sci. Bull. 2005, 50, 707–710. [Google Scholar] [CrossRef]
  49. Zhu, J.J.; Xiao, D.H.; Li, J.; Yang, X.; Wu, Y. Effect of Ce on NO direct decomposition in the absence/presence of O2 over La1-xCexSrNiO4 (0 ≤ x ≤ 0.3). J. Mol. Catal. A Chem. 2005, 234, 99–105. [Google Scholar] [CrossRef]
  50. Zhu, J.; Xiao, D.; Li, J.; Yang, X.; Wei, K. Effect of Ce and MgO on NO decomposition over La1−x–Cex–Sr–Ni–O/MgO. Catal. Commun. 2006, 7, 432–435. [Google Scholar] [CrossRef]
  51. Zhu, J.J.; Xiao, D.H. Perovskite-like mixed oxides (LaSrMn1-xNixO4+δ, 0≤x≤1) as catalyst for catalytic NO decomposition: TPD and TPR Studies. Catal. Lett. 2009, 129, 240–246. [Google Scholar] [CrossRef]
  52. Ma, Z.; Gao, X.; Yuan, X.; Zhang, L.; Zhu, Y.; Li, Z. Simultaneous catalytic removal of NOx and diesel soot particulates over La2−xAxNi1−yByO4 perovskite-type oxides. Catal. Commun. 2011, 12, 817–821. [Google Scholar] [CrossRef]
  53. Gao, L.; Chua, H.T.; Kawi, S. The direct decomposition of NO over the La2CuO4 nanofiber catalyst. J. Solid State Chem. 2008, 181, 2804–2807. [Google Scholar] [CrossRef]
  54. Wang, Z.; Li, Z.; Cui, Y.; Chen, T.; Hu, J.; Kawi, S. Highly Efficient NO Decomposition via Dual-Functional Catalytic Perovskite Hollow Fiber Membrane Reactor Coupled with Partial Oxidation of Methane at Medium-Low Temperature. Environ. Sci. Technol. 2019, 53, 9937–9946. [Google Scholar] [CrossRef] [PubMed]
  55. Dhal, G.C.; Dey, S.; Mohan, D. Solution combustion syntheses of perovskite-type catalysts for diesel engine exhaust gas purification. Mater. Today Proc. 2017, 4, 10489–10493. [Google Scholar] [CrossRef]
Figure 1. A schematic of the ideal ABO3 perovskite unit cell.
Figure 1. A schematic of the ideal ABO3 perovskite unit cell.
Catalysts 11 00622 g001
Figure 2. A possible mechanism of NO decomposition on the perovskite-type catalyst. Adapted with permission from [9], 2021, Elsevier.
Figure 2. A possible mechanism of NO decomposition on the perovskite-type catalyst. Adapted with permission from [9], 2021, Elsevier.
Catalysts 11 00622 g002
Figure 3. A possible reaction mechanism for the direct decomposition reaction of NO on perovskite-type oxides. Adapted with permission from [10], 2021, Royal Society of Chemistry.
Figure 3. A possible reaction mechanism for the direct decomposition reaction of NO on perovskite-type oxides. Adapted with permission from [10], 2021, Royal Society of Chemistry.
Catalysts 11 00622 g003
Figure 4. The mechanism of cyclic decomposition of NO on perovskite-like oxides with NO2 as an intermediate species.
Figure 4. The mechanism of cyclic decomposition of NO on perovskite-like oxides with NO2 as an intermediate species.
Catalysts 11 00622 g004
Figure 5. NO decomposition catalytic activity of some LaMO3-based perovskite.
Figure 5. NO decomposition catalytic activity of some LaMO3-based perovskite.
Catalysts 11 00622 g005
Table 1. A summary of the reaction conditions and maximum nitrogen yield of major perovskite catalysts.
Table 1. A summary of the reaction conditions and maximum nitrogen yield of major perovskite catalysts.
CatalystsN2 Yield (%)Reaction ConditionsRefs.
La1.4Sr0.6NiO4801.0% NO, 1.34 g s cm−3, 850 ℃[7]
La1.2Sr0.8NiO4201.0% NO, 1.34 g s cm−3, 700 °C[7]
La1.6Ba0.4NiO4-20%BaO98.81.0% NO, 1.2 g s cm−3, 850 °C[31]
Ba0.9La0.1Mn0.8La0.2O374.11.0% NO, 3.0 g s cm−3, 850 °C[32]
Ba0.9Pr0.1Mn0.8Pr0.2O369.61.0% NO, 3.0 g s cm−3, 850 °C[32]
Ba0.8La0.2Mn0.8Mg0.2O3751.0% NO, 3.0 g s cm−3, 850 °C[32]
BaMn0.8Mg0.2O366.81.0% NO, 3.0 g s cm−3, 850 °C[32]
BaMn0.8Zr0.2O351.61.0% NO, 3.0 g s cm−3, 850 °C[32]
BaMnO316.01.0% NO, 3.0 g s cm−3, 850 °C[32]
SrMn0.8Mg0.2O337.81.0% NO, 3.0 g s cm−3, 850 °C[32]
SrMn0.8Fe0.2O339.71.0% NO, 3.0 g s cm−3, 850 °C[32]
La1.2Ba0.8NiO4904.0% NO, 1.2 g s cm−3, 850 °C[41]
La1.2Ba0.8NiO4764.0% NO, 0.75 g s cm−3, 850 °C[41]
La1.6Ba0.4NiO4/BaCO3984.0% NO, 1.2 g s cm−3, 850 °C[41]
La0.87Sr0.13Mn0.2Ni0.8O3401.0% NO, 1.2 g s cm−3, 650 °C[41]
La0.8Ce0.5NiO4701.0% NO, 1.2 g s cm−3, 800 °C[41]
La1.867Th0.1CuO4.005651.0% NO, 1.33g s cm−3, 500 °C[42]
La1.6Th0.4CuO4421.0% NO, 1.33 g s cm−3, 850 °C[42]
SrFe0.7Mg0.3O347.31.0% NO, 3.0 g s cm−3, 850 °C[43]
Pt/SrFe0.7Mg0.3O3561.0% NO, 3.0 g s cm−3, 850 °C[43]
SrFe0.7Sn0.3O342.71.0% NO, 3.0 g s cm−3, 850 °C[43]
SrFe0.7Ni0.3O341.61.0% NO, 3.0 g s cm−3, 850 °C[43]
La0.5Ce0.5SrNiO/MgO751.0% NO, 2.4 g s cm−3, 800 °C[52]
LaSrMn0.2Ni0.8O4+δ851.0% NO, 1.2 g s cm−3, 850 °C[53]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shen, Q.; Dong, S.; Li, S.; Yang, G.; Pan, X. A Review on the Catalytic Decomposition of NO by Perovskite-Type Oxides. Catalysts 2021, 11, 622. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11050622

AMA Style

Shen Q, Dong S, Li S, Yang G, Pan X. A Review on the Catalytic Decomposition of NO by Perovskite-Type Oxides. Catalysts. 2021; 11(5):622. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11050622

Chicago/Turabian Style

Shen, Qiuwan, Shuangshuang Dong, Shian Li, Guogang Yang, and Xinxiang Pan. 2021. "A Review on the Catalytic Decomposition of NO by Perovskite-Type Oxides" Catalysts 11, no. 5: 622. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11050622

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