Next Article in Journal
Achieving Net Zero Carbon Performance in a French Apartment Building?
Previous Article in Journal
An Ejector and Flashbox-Integrated Approach to Flue Gas Waste Heat Recovery: A Novel Systematic Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Study of SCR Mechanism on LaMn1−xFexO3 Catalyst Surface Based DFT

1
School of Environmental and Municipal Engineering, Qingdao University of Technology, Qingdao 266520, China
2
Ganjiang Innovation Academy, Chinese Academy of Sciences, No. 1, Science Academy Road, Ganzhou 341000, China
3
School of Energy and Environment, Southeast University, Nanjing 210096, China
*
Author to whom correspondence should be addressed.
Energies 2023, 16(22), 7609; https://doi.org/10.3390/en16227609
Submission received: 18 October 2023 / Revised: 7 November 2023 / Accepted: 10 November 2023 / Published: 16 November 2023
(This article belongs to the Section I3: Energy Chemistry)

Abstract

:
Perovskite SCR catalysts have become a hot research topic in the field of de-NOx catalyst development. This article selects LaMnO3 with high performance as the research object, modifies the catalyst by doping some iron elements instead of manganese elements, and applies density functional theory to study its reaction mechanism, providing theoretical reference for further research on perovskite. Research has found that several main reactants such as NH3, NO, and O2 can form stable adsorption at the active site, with NO more inclined to adsorb at the nitrogen atom end at the active site. The oxidation of O2 molecules after adsorption is greater than that of the active site. The adsorption capacity of the Mn active site of the catalyst before modification on the above molecules is weaker than that of the Fe active site introduced after modification. Under both anaerobic and aerobic conditions in the SCR reaction process, NH3 molecules are first adsorbed at the active site, and then influenced by lattice oxygen under anaerobic conditions. Under aerobic conditions, they are gradually dehydrogenated and produce NH2 and NH radicals. These two radicals react with NO molecules to form intermediate products in the form of NH2NO and NHNO molecules. Due to the instability of the intermediate products, they ultimately decompose into N2 and H2O molecules. The introduction of Fe active sites can increase the generation of NH2 and NH radicals during the reaction process and simplify the reaction process between NH2 radicals and NO molecules, which will be conducive to the completion of the reaction.

1. Introduction

NOx is the main atmospheric pollutant and one of the common major polluting gases in industry today. SCR is currently one of the most widely used technologies in NOx control, with a de-NOx efficiency of over 90% [1]. SCR (selective catalytic reduction) technology uses NH3 generated from urea, ammonia water, or liquid ammonia as a reducing agent, which adsorbs onto the catalyst surface and selectively reacts with NOx to generate nitrogen, water, and other products, thereby achieving the goal of removing NOx [2].
The core of SCR technology lies in catalysts which have excellent performance and are characterized by good catalytic activity, high selectivity, high mechanical strength, and stable operation [3]. Therefore, selecting appropriate catalysts has become the focus of research by many scholars. Platinum group elements have been widely used in the treatment of automotive exhaust [4]. Vanadium–titanium-based catalysts have also matured in recent decades of application. However, the reaction temperature of vanadium –titanium-based catalysts is relatively high, ranging from 300 to 400 °C, with an optimal activity window temperature of 340 to 380 °C [5]. However, due to the high sulfur content of the fuel used, there are also by-products such as sulfur oxides (SOx) in the tail gas [6]. To prevent the poisoning and deactivation of denitrification catalysts, the de-NOx device is usually placed behind the electrostatic precipitator and desulfurization device, which leads directly to a significant decrease in the temperature of the tail gas processed by the denitrification device. In order to avoid fuel loss caused by the demand for exhaust gas reheating, it is urgent to develop reasonable, effective, low-temperature, and high-performance denitrification catalysts to meet industrial needs [7].
“Perovskite oxides” is a general term for oxides with an ABO3 structural form, with an ideal cubic crystal system and a microscopic symmetry group belonging to Pm3m. Perovskite oxides have many advantages, such as strong redox ability, thermal stability, and variety diversity, which have attracted the attention and research of many scholars. LaMnO3 perovskite oxides exhibit robust stability and excellent selective catalytic reduction (SCR) activity, making them promising candidates for low-temperature catalysts [8]. Scholars have shown a keen interest in their potential [9]. Zhang’s investigation of LaMnO3 catalysts revealed impressive SCR performance, achieving a notable 78% NO conversion at 250 °C, with increasing conversion as the temperature ranged from 100 °C to 300 °C [10]. Guo et al. also examined LaMnO3 catalysts, demonstrating remarkable results with over 90% NOx conversion at 135 °C and within the 135–260 °C temperature range [11]. Wang et al. extended these findings by utilizing a LaMnO3-supported iron ore catalyst, achieving an exceptional NOx conversion efficiency of up to 98% at 180 °C [12]. These findings collectively underscore the catalyst’s exceptional SCR capabilities. Nevertheless, the precise mechanism governing NOx conversion over LaMnO3 catalysts remains unclear.
However, the catalytic activity and N2 selectivity of pure ABO3 perovskite catalysts are not ideal, which limits their application. It is generally believed that there are two types of SCR mechanisms; one is the E–R mechanism, where NH3 adsorbs onto the catalyst surface and reacts with weakly adsorbed or non-adsorbed NO to generate N2 and H2O. The other mechanism is the L–H mechanism, where NH3 and NO simultaneously adsorb onto the catalyst surface and react to generate N2 and H2O. Zhang et al. [13]. prepared perovskite materials doped with Sr and Ce at the A site and Fe and V at the B site based on LaMnO3 perovskite. They found that when doped with Fe at the B site, the denitrification efficiency at the medium- to high-temperature end was significantly improved, reaching 80% at 200 °C, maintaining over 90% efficiency at 225–350 °C, and reaching 100% at 250–300 °C. At the same time, the N2 selectivity was also slightly improved. Other literature shows that LaMnxFe1−xO3 perovskite catalysts can be used as an effective SCR denitrification catalyst for development and application [14]. However, there is currently no research on the denitrification reaction mechanism of this catalyst. Although many researchers have studied the mechanisms of Mn-based and Fe-based catalysts, there are currently no reports on the co-acting catalysts of Mn and Fe, and there is no relevant research on the synergistic mechanism between Mn and Fe. Therefore, this article applies density functional theory (DFT) to study the mechanism of LaMnxFe1−xO3 type perovskite catalysts.
DFT is an effective approach in the material design and development stage and also has a wide range of applications in de-NOx field. Yao et al. [15] used density functional theory to study the reaction mechanism of NO reduction by NH3 on a V2O5 (001) surface and proposed a new reaction mechanism via calculation. The mechanism consists of four steps: the first step is to convert surface L-acid to B-acid, the second step is to generate double B-acid, the third step is to generate a vacancy, O, on the V atom, and the fourth step is to introduce O2 to regenerate the catalyst. Yang et al. [16] used density functional theory to study the surface SCR reaction process of CeO2/TiO2 catalysts. The calculations showed that NH3 was first adsorbed onto the surface L-acid site in molecular form, then activated and dehydrogenated to generate NH2, and then reacted with NO to generate N2 and H2O. This reaction follows the E–R mechanism. Via this calculation, the energy barrier and reaction heat of each elementary reaction step in the entire SCR reaction can be obtained, and its possible paths can be analyzed. Song et al. [17] used density functional theory to investigate the behavior of the acatalytic reduction of NOx by H2O2 on the surface of an α-Fe2O3 catalyst, and the surface reaction path was determined based on this. The results showed that both NOx and H2O2 would react on the surface of an α-Fe2O3 catalyst undergoing adsorption, and H2O2 undergoes dissociation under the action of the catalyst, thereby reacting with the adsorbed NO to generate N2 and H2O. At the same time, some scholars have also studied the surface catalytic reaction of perovskite-type catalysts using DFT. Zhang et al. conducted DFT calculations to determine atomic charges for BiMnO3 and LaMnO3, highlighting the relationship between smaller O charges, higher electronegativity, increased acidity, and enhanced NH3 adsorption and activation, thus boosting low-temperature NH3-SCR activity [13]. Yan et al. examined the CO-SCR reaction process on LaMnO3 catalyst surfaces, defining key steps, activation energy barriers, and reaction paths [18]. Ren et al. used the DFT method to study the SCR reaction process on the surface of LaMnO3 and obtained the surface SCR reaction pathway, determining the E–R mechanism of the surface reaction [19]. Araujo et al. performed functional correction and conducted higher-level calculations on small metal clusters to improve the adsorption energy and potential barrier of periodic band structures and compared them with experimental results. This functional can be widely used for surface science evaluation and testing [20]. However, DFT research on NH3-SCR surface reactions with LaMnO3 catalysts remains unexplored.
In summary, the perovskite-structured LaMnxFe1−xO3 catalyst exhibits strong potential as a de-NOx catalyst. However, its mechanism requires further investigation. Employing the mature and reliable DFT method can uncover surface reaction mechanisms, providing valuable guidance for the development and application of low-temperature de-NOx catalysts. This paper leverages DFT to study surface reaction mechanisms, capturing parameters such as molecular adsorption, elementary reaction steps, and activation energies, offering insights into adsorption behavior and reaction processes while identifying the most probable reaction pathway on the catalyst surface.

2. Model and Method

Figure 1 shows the structure of LaMn1−xFexO3 catalyst, which is established based on LaMnO3. LaMnO3 perovskite belongs to the category of rhombic structures; the space group is Pnma; and the cell parameters are as follows: a = 5.724, b = 7.690, and c = 5.534. Figure 1 shows the LaMnO3 crystal model after doping Fe elements, in which two perspectives are selected for representation. The crystal model structure consists of twelve oxygen atoms, four lanthanum atoms, three manganese atoms, and one iron atom, in which the doping amount of iron element is 0.25. The XRD characterization results of iron-doped LaMnO3 crystals show that the XRD spectra of iron-doped LaMnO3 crystals have not changed, indicating that the main composition of iron-doped LaMnO3 crystals is still LaMnO3 [21]. Therefore, when the model in this paper was built, based on LaMnO3 crystal, some iron elements were doped on the manganese site so as to modify the LaMnO3 crystal.
In this paper, surface 001 in the LaMnO3 crystal was selected as the reaction active surface for study, and a periodic plate model was established on this basis. At the same time, the 2 × 1 surface of the LaMnO3 crystal was selected as the size of reactive active surface. On the surface of the reactive active surface, this paper uses the model to add a 30 Å vacuum layer to prevent interaction between adjacent layers. In addition, in the whole calculation process, in the whole calculation process, the bottom atoms are fixed and other atoms were set to be able to relax [19].
During the computational process, calculations were conducted using the DMol3 package, employing a spin-polarized DFT model [22]. The GGA method was utilized for the electron exchange correlation function, along with the application of the effective core potentials (ECP) method for Fe and Mn atom core electrons. To ensure accurate calculations, double-numeric quality plus polarization (DNP) functions were applied to all atoms, with a selected real space cutoff radius of 5.0 Å [23]. Brillouin zone integration was performed using 5 × 5 × 1 Monkhorst–Pack grids for the adsorption and reaction system [24].
To assess the adsorption properties of the catalysts, the adsorption energy was thoroughly evaluated, which was defined as follows, according to [25]:
E a d s = E s l a b + m o l e c u l e E s l a b E m o l e c u l e
where E a d s means the adsorption energy, E s l a b + m o l e c u l e means the energy of slab and molecule after adsorption, and E s l a b and E m o l e c u l e are the energy of the catalyst slab before adsorption and that of NH3 and NO molecule before adsorption, respectively.
The NH3-SCR process on the Ce-doped γ-Fe2O3 catalyst displayed distinct stages, each with unique intermediate products (IM), transition states (TS), and final states (FS). In the computational procedure, the LST–QST method was applied to locate the transition states (TS), and a careful vibration frequency analysis was conducted to verify the presence of only a single virtual frequency [26].
The activation energy barrier (Ea) is defined as the following equation [27,28]:
Ea = ETSEIM
where ETS is intermediate energy and EIM represent transition state energy.

3. Results and Discussion

3.1. Fe Doping on LaMnO3 Crystal Surface

The active reaction surface of the modified LaMnO3 crystal model is shown in Figure 2. The crystal model shown in the figure is divided into upper and lower layers, in which iron is doped in the upper layer and the lower layer is a fixed bottom layer. Table 1 shows the parameter changes of the crystal at the site before and after doping. It can be seen that doping iron will lead to lattice deformation, which will affect the performance of the catalyst in the SCR reaction process. At the same time, the Mulliken charge of Mn before doping is 0.658, while after Fe doping, the Mulliken charge is 0.701. The reason for the increase of the charge is that the electronegativity of iron is greater than that of manganese, and electrons are attracted by Fe ions. Therefore, doping with Fe will bring more valence electrons and increase the free electrons on the catalyst surface, thus changing the redox ability of the catalyst surface. Chen [29] and others found that the NO molecules adsorbed onto the surface of the perovskite may react with the perovskite lattice oxygen, but this process has a weak impact. Yan [18] and others pointed out that the active sites in the LaMnO3 catalyst are mainly Mn sites. Considering the influence of doping Fe on the catalyst, Mn and Fe ion sites in the upper layer of Figure 2b were selected as the active sites for this study. The influence of some molecules adsorbed by lattice oxygen is not considered.

3.2. Adsorption of Reactant Molecules on Modified LaMnO3 (001) Surface during NH3-SCR Process

There are L–H mechanisms and E–R mechanisms in the NH3-SCR de-NOx process involving catalysts. “L–H mechanism” refers to the heterogeneous catalytic mechanism in which two reactant molecules first adsorb onto the catalyst surface and then desorb after reaction to produce products, while “E–R mechanism” refers to the mechanism in which one reactant molecule adsorbs onto another to produce products. Many scholars have no consensus on this, but both reaction mechanisms clearly indicate that NH3 will be adsorbed onto the catalyst surface first. As the reactant NH3 molecule, its adsorption capacity on the catalyst surface has a great impact on NOx removal [30], so this paper will first study the adsorption capacity of NH3 molecule. Figure 3 shows the schematic diagram of NH3 molecules adsorbed onto Mn and Fe active sites, while Table 2 shows the bond angle, bond length, released adsorption energy, and Mulliken charge of adsorbed molecules after NH3 molecules adsorbed onto Mn and Fe active sites.
It can be seen in this situation that NH3 can be absorbed onto the Mn and Fe sites with the N atom end, and a new chemical bond is formed. The NH3 bond angle adsorbed onto Mn active sites is greater than that of NH3 molecules adsorbed onto Fe active sites. This is because the number of electrons lost by the H atoms of NH3 molecules adsorbed onto Mn active sites is greater than that adsorbed onto Fe, which leads to the weaker repulsion of H atoms to hydrogen atoms, leads to a larger bond angle. The N-Mn bond length of the NH3 molecule adsorbed onto the Mn active site is longer than that adsorbed onto the Fe active site because the radius of the Mn ion is larger than that of the Fe ion. The adsorption energy released after adsorption and the energy released by NH3 adsorbed onto the Mn active site is less than that released by NH3 molecule adsorbed onto Fe active site, indicating that the NH3 molecule adsorbed onto the Fe active site is conducive to reaction. In addition, the Mulliken charges are all positive, indicating that electrons migrate from the NH3 molecule to the catalyst surface, which is due to the existence of lone pair electrons in the NH3 molecule, and the existence of a 3d empty orbital on the catalyst surface is sufficient to accept excess electrons. Moreover, the charge amount of the NH3 molecule adsorbed by the Mn active site is less than that of the Fe active site, because the electronegativity of Fe is greater than that of Mn. Therefore, in both adsorption processes, a large amount of energy is released, and a large number of electrons are transferred from NH3 molecules to the catalyst. Since the electronegativity of Fe ions is greater than that of Mn ions, there is more electron migration at the active site of Fe, the interaction between them is stronger, and more energy is released.
As one of the main reactants in SCR de-NOx reaction, and under normal combustion temperature, the NOx contained in the flue gas generated after the combustion of coal and other fuels is mainly NO and NO2, of which NO accounts for more than 90% of the total NOx, while NO2 only accounts for 5–10%, so it is necessary to study its adsorption capacity on the catalyst surface [31].
It has been shown that NO will approach the active surface sites in the form of N atoms and O atoms, respectively, to form a stable adsorption configuration. The corresponding parameters are list in Table 3.
There are two kinds of adsorption of NO reactants in Figure 4 on Mn and Fe active sites. It can be seen from Table 3 that when NO is adsorbed onto the catalyst surface as a N atom end, its bond length is shorter than that in the form of O atom end proximity, and the adsorption energy in the case of N proximity is far greater than that in the case of O proximity. Therefore, the adsorption of NO molecules should mostly occur on Mn and Fe ions on the catalyst surface as a N atom end. During the adsorption process, no matter whether the NO molecule on Mn active site is adsorbed by a nitrogen atom or an oxygen atom, the energy released after adsorption is less than the energy released by the corresponding end of the NO molecule adsorbed onto the Fe active site, which also indicates that the NO molecule is more easily adsorbed onto the Fe active site. This is likely due to the fact that the electronegativity of Fe ions is greater than that of Mn ions, and their oxidation is stronger, which is also true of NH3 molecules. This also shows that the doping of Fe is conducive to promoting the adsorption of NO. When the nitrogen atom is close, the bond length is shorter and more electrons are transferred, whereas when the oxygen atom is close, the bond length is longer and fewer electrons are transferred.
In the process of the NH3-SCR de-NOx reaction, O2 plays an important role as a reactant, which is helpful to the smooth progress of the reaction. Figure 5 shows the schematic diagram of O2 molecules adsorbed onto Mn and Fe active sites, and Table 4 shows the corresponding adsorption parameters. Table 4 shows that the N-Mn bond length of O2 adsorbed onto Mn active site is longer than that on Fe active site. During the adsorption process, the energy released by O2 adsorption on the Mn active site is less than that adsorbed onto Fe active site, indicating that O2 is more easily adsorbed onto Fe. Since the oxidizability of oxygen atom in O2 is stronger than that of Mn and Fe active sites, O2 molecules obtain electrons during adsorption. Since the electronegativity of Fe ions is greater than that of Mn ions, the charge amount of O2 molecule adsorbed by Mn active site is greater than that of O2 molecules adsorbed by an Fe active site.

3.3. Study on Surface Reaction Process of Modified LaMnO3 (001) Surface

A large number of results show that NH3 will first be adsorbed onto the catalyst surface, regardless of the E–R mechanism or L–H mechanism, and will decompose to produce NH2, NH, and other intermediates under the action of lattice oxygen. Hence, it is necessary to study the process of NH3 decomposition, which is shown in Figure 6.
Figure 6 shows the process of dehydrogenation of NH3 molecules into NH2, NH, and N after adsorption at the Mn and Fe active sites in the absence of lattice oxygen. The NH3 molecule is adsorbed onto the Mn active site. At this time, the N-Mn bond length is 2.075 Å, and the distances between the three hydrogen atoms of the NH3 molecule and the nearest lattice oxygen are 2.738 Å, 2.747 Å, and 2.822 Å, respectively. Then, under the action of lattice oxygen, the hydrogen atom nearest to the lattice oxygen desorbs and combines with the lattice oxygen to form hydroxyl. The energy barrier and reaction heat of this process are 87.034 kJ/mol and 71.445 kJ/mol, respectively, indicating that this process is an endothermic process. After the formation of NH2 radicals on the catalyst surface, the dehydrogenation process will continue under the action of the lattice oxygen. The energy barrier and reaction heat of the process are 148.695 kJ/mol and 72.741 kJ/mol, respectively, indicating that the process is also an endothermic process. After the formation of NH radicals on the catalyst surface, the dehydrogenation process will continue under the action of the lattice oxygen. The energy barrier and reaction heat of the process are 86.322 kJ/mol and 5.556 kJ/mol, respectively. While the NH3 molecule removes three hydrogen atoms in turn, the energy barrier of NH2 radical removing hydrogen atom is the highest, indicating that the occurrence of this process is difficult, which leads to less NH radical generation, and fewer NH radical sources also lead to fewer nitrogen ions finally formed. The same process will occur at the Fe site. The NH3 molecule is adsorbed onto the Fe active site, which is an exothermic process. At this time, the N-Fe bond length is 1.999 Å, and the distances between the three hydrogen atoms of the NH3 molecule and the nearest lattice oxygen are 2.620 Å, 2.659 Å, and 2.848 Å, respectively. It can be seen that the average distance between hydrogen atom and lattice oxygen is shorter than that adsorbed onto the Mn ion at this time. This is because the electronegativity of Fe ions is stronger than that of Mn ions. As a result, after the NH3 is adsorbed onto the active site, the N atom adsorbed onto the Fe ion obtains more electrons, which weakens the repulsion of hydrogen atoms in NH3 molecules, thus reducing the distance between hydrogen atom and lattice oxygen. Subsequently, under the action of the lattice oxygen, the hydrogen atom nearest to the lattice oxygen desorbs and combines with lattice oxygen to form hydroxyl. The energy barrier and reaction heat of this process are 98.583 kJ/mol and 53.631 kJ/mol, respectively. NH2 radicals continue to dehydrogenate, and the energy barrier and reaction heat of this process are 136.380 kJ/mol and 80.549 kJ/mol, respectively. However, the NH radical continues the dehydrogenation process, and the energy barrier and reaction heat of the process are 119.181 kJ/mol and 60.320 kJ/mol, respectively. In comparison with the above analysis, the dehydrogenation process of NH3 molecule at Mn and Fe active sites is basically the same, and the energy barrier of NH2 radical removal of hydrogen atom is the highest. Therefore, NH2 should be the main dehydrogenation product, which is the same as that of other catalysts [32].
Figure 7 shows the decomposition process of NH2NO. On the catalyst surface, after the formation of NH2, it will react with gaseous NO to generate NH2NO. In this step, the reaction potential barrier can be ignored, and the reaction is very rapid, and then NH2NO will decompose. At Mn active site, NH2NO molecule removes a hydrogen atom to the lattice oxygen to form the NHNO molecular structure, which is a linear structure. The energy barrier and reaction heat of the process are 28.128 kJ/mol and 24.510 kJ/mol, respectively, indicating that the process is endothermic. The generated NHNO molecule is unstable, which will further undergo dehydrogenation and finally decompose into N2 and hydroxyl. The energy barrier and reaction heat of this process are 204.640 kJ/mol and −206.557 kJ/mol, respectively, indicating that this process is an exothermic process. Since the product N2 has been generated, the reaction is complete, and the generated hydroxyl group will combine with the free hydrogen atom to generate the product H2O. At the Fe active site, the NH2NO molecule will also break an N-H bond and adsorb the desorbed hydrogen atom onto the oxygen ion in the NH2NO molecule to form an O-H bond. The energy barrier and reaction heat of the process are 155.849 kJ/mol and −39.559 kJ/mol, respectively, indicating that the process is exothermic. The NHNO generated is unstable, so dehydrogenation will occur further and eventually produce N2 and H2O. The energy barrier and reaction heat of the process are 58.345 kJ/mol and −201.643 kJ/mol respectively, which also indicates that the process is exothermic. Compared with the above analysis, the reaction process of NH2 free radicals at Mn and Fe active sites is the same, and there is a higher energy barrier step in both reaction processes, but the reaction will eventually release a lot of energy. Combined with the change in reaction energy, the occurrence of the reaction will gradually be easy to realize as the reaction proceeds.
Yang et al. pointed out that during the process of SCR, the surface oxygen on the catalyst surface has a high mobility, and its oxidation is stronger than that of lattice oxygen. Therefore, the decomposition process of O2 molecules on the catalyst surface was studied in this paper [16]. The corresponding process is shown in Figure 8.
Figure 8 shows the decomposition process of O2 after adsorption on Mn and Fe active sites. The O2 molecule is adsorbed onto the active site, and then under the influence of the active site, O2 decomposes and adsorbs an oxygen atom on the active site respectively. At the Mn site, the energy barrier and reaction heat of this process are 220.707 kJ/mol and −2.854 kJ/mol, respectively, and at the Fe site, they are 254.764 kJ/mol and 12.224 kJ/mol, respectively. By comparing the above data, it can be seen that the energy barrier of the decomposition process of O2 molecule on the active site is relatively high, among which the energy barrier of the decomposition process of O2 molecule adsorbed onto the Fe active site is higher, which indicates that the breaking of O=O bond requires more energy, and also indicates that the decomposition process of O2 molecule on Mn and Fe active site is difficult to occur. This result is similar to the results on the surface of iron oxide [33]. The decomposition process of O2 molecule adsorbed onto Mn active site is exothermic, while the decomposition process of O2 molecule adsorbed onto Fe active site is endothermic process, indicating that O2 molecule adsorbed onto Mn active site is easier to decompose.
Figure 9 shows the dehydrogenation process of NH3 molecules at Mn and Fe active sites. During this process, NH3 adsorbed onto Mn and Fe active sites undergoes dehydrogenation by oxygen ions generated by the decomposition of O2 molecules, producing NH2 radicals and hydroxyl groups. The hydroxyl groups combine with free hydrogen atoms on the catalyst surface to form the product H2O. The energy barriers of this process are 50.761 kJ/mol and 63.411 kJ/mol, respectively, and the reaction heats are 25.879 kJ/mol and 46.599 kJ/mol, respectively, indicating that this process is endothermic. Comparing the above process with the process of removing one hydrogen atom from NH3 molecules under anaerobic conditions, it can be found that under the influence of surface oxygen, the energy barrier for NH3 molecule dehydrogenation is lower, and the reaction heat absorption is also lower. This indicates that NH3 molecules are more inclined to react with surface oxygen and undergo dehydrogenation when surface oxygen is present. This is because surface oxygen has more coordination numbers to accommodate more hydrogen atoms. The NH2 free radicals generated by the reaction will react with NO molecules in the air to generate NH2NO molecules, which will then decompose and ultimately generate N2 and H2O products, as shown in Figure 7. In addition, the generated NH2 radicals will continue to react with surface oxygen and undergo dehydrogenation.
Figure 10 depicts the dehydrogenation process of NH2 radicals at Mn and Fe active sites. During this process, NH3 adsorbed onto Mn and Fe active sites undergoes dehydrogenation by oxygen ions generated by the decomposition of O2 molecules, resulting in the formation of NH2 radicals and hydroxyl groups, which combine with free hydrogen atoms on the catalyst surface to form the product H2O. The energy barriers of this process at Mn and Fe sites are 374.292 kJ/mol and 101.089 kJ/mol, respectively, and the reaction heats are 27.39 kJ/mol and 50.537 kJ/mol, respectively, indicating that this process is endothermic.
By comparing the above processes, it can be seen that the energy barrier for the dehydrogenation process of NH2 radicals adsorbed onto Fe active sites is significantly lower than that of NH2 radicals adsorbed onto Mn active sites. This indicates that increasing Fe active sites by doping iron is beneficial for the dehydrogenation process of NH2 radicals adsorbed onto active sites under aerobic conditions. The generated NH radicals will continue to react with surface oxygen and undergo a dehydrogenation process to generate nitrogen ions. However, based on the previous analysis, it can be seen that the generation of NH radicals and oxygen ions is relatively small. Therefore, this article does not consider the reaction between NH radicals and surface oxygen and the dehydrogenation process.
The generated NH radicals will react with NO molecules in the air to generate NHNO molecules, as shown in Figure 11, which shows the formation process of NHNO on the catalyst surface.
Figure 11 shows that NH radicals adsorbed onto Mn and Fe active sites combine with NO molecules in the air to form NHNO molecules. In this structure, N-Mn and N-Fe bonds still exist, while oxygen ions in NHNO molecules do not form stable chemical bonds with the active sites. The energy barriers of this process are 28.537 kJ/mol and 30.629 kJ/mol, respectively, and the reaction heats are −150.812 kJ/mol and −132.462 kJ/mol, respectively, indicating that this process is an exothermic process.
By comparing the above processes, it can be found that the reaction process between NH radicals adsorbed onto Mn and Fe active sites and NO molecules in the air is different, and the reaction energy barrier and reaction heat also differ significantly. The reaction energy barrier between NH radicals adsorbed onto Mn active sites and NO molecules in the air is lower than that between NH radicals adsorbed onto Fe active sites and NO molecules in the air. The reaction heat is that the energy released by the reaction between NH radicals adsorbed onto Mn active sites and NO molecules in the air is higher than that released by the reaction between NH radicals adsorbed onto Fe active sites and NO molecules in the air. The above data indicates that NO molecules are more inclined to react with NH radicals adsorbed onto Mn active sites. However, since both reaction processes are exothermic, the occurrence of the reaction will gradually become easier to achieve as the reaction progresses.
From the analysis of the adsorption of NO2 molecules on Mn and Fe active sites mentioned earlier, it can be concluded that the adsorption of NO2 molecules near the nitrogen atom end on the active site or the adsorption of NO2 molecules on the Fe active site is beneficial for the reaction, and when both conditions are met simultaneously, the reaction is most favorable. Therefore, this article studied the generation process of NO2 at Mn and Fe active sites using this approach, as shown in Figure 12.
Figure 12 shows the formation process of NO2 molecules at Mn and Fe active sites, with energy barriers of 199.664 kJ/mol and 205.752 kJ/mol, respectively, and reaction heats of −18.586 kJ/mol and −8.590 kJ/mol, respectively, indicating that this process is exothermic.
By comparing the above processes, it can be found that the energy barrier of NO2 generation at both Mn and Fe active sites is relatively high. The energy barrier of NO2 generation at Mn active sites is slightly higher than that of NO2 generation at Fe active sites. At the same time, the reaction heat released by NO2 generation at Mn active sites is greater than that released by NO2 generation at Fe active sites. The results also indicate that NO2 molecules are more easily generated at Mn active sites. However, since both reaction processes are exothermic, the occurrence of the reaction will gradually become easier to achieve as the reaction progresses. There is literature indicating that when NO2 is present, nitric acid will be generated to further generate NH4NO3. Under the action of surface acidic sites, NH4NO3 will be reduced by NO to NH4NO2, which is extremely unstable and can decompose into N2 and H2O at 100 °C, which is a rapid SCR reaction [34,35,36].

3.4. Analysis of NH3-SCR De-NOx Reaction Mechanism

The NH3-SCR denitrification reaction process on the surface of modified LaMnO3 (001) catalysts can be divided into three processes: adsorption, reaction, and desorption. The adsorption of NH3 molecules on the catalyst surface can be divided into adsorption at Mn active sites and adsorption at Fe active sites, with NH3 molecules having the highest adsorption energy at Fe active sites. The adsorption of NO molecules on the catalyst surface can be divided into two types: adsorption on the Mn active site, adsorption on the Fe active site, or adsorption on the active site near the nitrogen atom end of NO molecules; and adsorption on the active site near the oxygen atom end of NO molecules. Among them, the adsorption energy of NO molecules on the Fe active site near the nitrogen atom end is the highest.
The reaction process is shown in Figure 13, listing the relationships between the various elemental reactions. Under anaerobic conditions, the adsorbed NH3 molecules will undergo dehydrogenation with lattice oxygen on the catalyst surface to generate NH2 radicals. Subsequently, NH2 radicals will continue to dehydrogenate with lattice oxygen to form NH radicals. Figure 13a shows that the energy barrier of this step is 148.695 kJ/mol, while Figure 13b shows that the energy barrier of this step is 136.380 kJ/mol, indicating that the reaction energy barrier of this step is the highest under anaerobic conditions. Therefore, the generation of NH radicals is relatively low, leading to a very low generation of nitrogen ions. Under aerobic conditions, NH3 molecules will also react and dehydrogenate with oxygen ions adsorbed onto the catalyst surface to generate NH2 radicals. NH2 radicals will also continue to dehydrogenate with surface oxygen to form NH radicals. The NH2 radicals generated under anaerobic and aerobic conditions will react with NO in the air and generate NH2NO molecules. Due to the instability of NH2NO molecules, they will subsequently decompose and ultimately produce N2 and H2O products. The generated NH radicals will also react with NO molecules in the air and generate NHNO molecules. Similarly, due to the instability of NHNO molecules, they will also undergo decomposition and ultimately produce N2 and H2O products. After N2 and H2O molecules detach from the catalyst surface, the SCR reaction is completed. Comparing the generation of NH2 and NH radicals under anaerobic and aerobic conditions, it can be found that the reaction proceeds more easily under aerobic conditions. By comparing the reactions between Mn active sites and Fe active sites, it can be seen that the reactions involving NH3 molecules, NH2 radicals, and NH radicals located at Fe active sites are more easily carried out. In addition, under aerobic conditions, the participation of O2 will cause the oxidation of NO molecules into NO2 molecules, and the participation of NO2 molecules will undergo a “rapid SCR reaction”, accelerating the progress of the SCR reaction.

4. Conclusions

On the active site, NH3, NO, and O2 molecules can all adsorb, with NH3 molecules as one of the main reactants releasing the most energy and losing the most charge after adsorption. NO molecules undergo chemical adsorption at the nitrogen atom end, while adsorption at the oxygen atom end is only physical adsorption. O2 molecules can also adsorb well onto the surface of the catalyst.
In the SCR reaction process, under anaerobic conditions, NH3 molecules adsorbed onto the active site are gradually dehydrogenated and generate NH2, NH, and N. The generated NH2 radicals will react with NO molecules to form the intermediate product NH2NO, but due to the instability of the intermediate product, it will further react and ultimately decompose into N2 and H2O molecules. Under aerobic conditions in the presence of O2, NH2 radicals will also undergo dehydrogenation via surface oxygen and generate NH radicals. NH radicals also react with NO molecules to generate the intermediate product NHNO, which is unstable and ultimately decomposes into N2 and H2O molecules. In the SCR reaction at Mn active sites, the NH2 radical dehydrogenation process has the highest energy barrier under aerobic conditions, which is the rate control step of Mn active sites. In the SCR reaction at Fe active sites, the NH2 radical dehydrogenation process in the absence of oxygen has the highest energy barrier, indicating that this process is the rate control step of Fe active sites. Comparing the SCR reaction process between Mn and Fe active sites, it can be found that the introduction of Fe active sites doped with some iron elements can increase the generation of NH2 and NH radicals and simplify the reaction process between NH2 radicals and NO molecules, which will be conducive to the completion of the reaction.

Author Contributions

Conceptualization, D.R.; Methodology, D.R. and T.M.; Formal analysis, K.L. (Kangshuai Lin), Z.Z. and K.G.; Data curation, K.L. (Kangshuai Lin) and T.M.; Writing—original draft, D.R.; Funding acquisition, S.L. (ZR2023ME2116) and K.L. (Kaijie Liu, 52304429, 20212BAB213032) All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Natural Science Foundation of Shandong Province, China, (ZR2021QE295), Natural Science Foundation of Shandong Province, China, (ZR2023ME2116), National Natural Science Foundation of China (52304429), and Jiangxi Provincial Natural Science Foundation (20212BAB213032).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liu, Z.; Ihl Woo, S. Recent Advances in Catalytic DeNOX Science and Technology. Catal. Rev. 2006, 48, 43–89. [Google Scholar] [CrossRef]
  2. Liu, K.; Li, J.; Yu, Q.; Han, X.; Bian, M.; Zhang, Y.; Yi, T. Optimization and comprehensive mechanism of environment-friendly bimetal oxides catalysts for efficient removal of NO in ultra-low temperature flue gas. Sep. Purif. Technol. 2023, 311, 123324. [Google Scholar] [CrossRef]
  3. Liu, K.; Yu, Q.; Wu, T.; Wang, B.; Duan, W.; Qin, Q. Manganese-based catalyst for NO removal at low temperatures: Thermodynamics analysis and experimental validation. Pet. Sci. Technol. 2018, 36, 1663–1671. [Google Scholar] [CrossRef]
  4. Omrani, M.; Goriaux, M.; Liu, Y.; Martinet, S.; Jean-Soro, L.; Ruban, V. Platinum group elements study in automobile catalysts and exhaust gas samples. Environ. Pollut. 2020, 257, 113477. [Google Scholar] [CrossRef] [PubMed]
  5. Forzatti, P. Environmental catalysis for stationary applications. Catal. Today 2000, 62, 51–65. [Google Scholar] [CrossRef]
  6. Li, Z.; Zhong, Z.; Yang, Q.; Ben, H.; Seufitelli, G.; Resende, F. Parametric study of catalytic hydropyrolysis of rice husk over a hierarchical micro-mesoporous composite catalyst for production of light alkanes, alkenes, and liquid aromatic hydrocarbons. Fuel 2022, 310, 122457. [Google Scholar] [CrossRef]
  7. Yu, S.; Yang, Q.; Li, Z.; Liu, T.; Tao, L.; Yao, E.; Zhang, Y. Mechanism analysis of gas products from catalytic pyrolysis of tire rubber based on reaction thermodynamics and kinetics. Fuel Process. Technol. 2022, 227, 107134. [Google Scholar] [CrossRef]
  8. Tarjomannejad, A.; Farzi, A.; Gómez, M.; Niaei, A.; Salari, D.; Albaladejo-Fuentes, V. Catalytic Reduction of NO by CO over LaMnFeO and LaAMnFeO (A = Sr, Cs, Ba, Ce) Perovskite Catalysts. Catal. Lett. 2016, 146, 2330–2340. [Google Scholar] [CrossRef]
  9. Li, X.; Yin, Y.; Yao, C.; Zuo, S.; Lu, X.; Luo, S.; Ni, C. La1−xCexMnO3 attapulgite nanocomposites as catalysts for NO reduction with NH3 at low temperature. Particuology 2016, 26, 66–72. [Google Scholar] [CrossRef]
  10. Zhang, R.; Luo, N.; Yang, W.; Liu, N.; Chen, B. Low-temperature selective catalytic reduction of NO with NH3 using perovskite-type oxides as the novel catalysts. J. Mol. Catal. A Chem. 2013, 371, 86–93. [Google Scholar] [CrossRef]
  11. Guo, J.; Shi, X.; Fan, A.; Li, J.; Chu, Y.; Yuan, S. Study on the Preparation and Denitration Performance of Ce Modified La–Mn Perovskite Catalyst. Adv. Eng. Sci. 2021, 4, 233–239. [Google Scholar]
  12. Wang, R.; Gui, K.; Liang, H. Effect of Ce-doped on performance of supported perovskite catalyst LaMnO3/hematite for SCR of NO by NH3. Chem. Ind. Eng. Process 2016, 35, 192–199. [Google Scholar]
  13. Zhang, Y.; Wang, D.; Wang, J.; Chen, Q.; Zhang, Z.; Pan, X.; Miao, Z.; Zhang, B.; Wu, Z.; Yang, X. BiMnO3 perovskite catalyst for selective catalytic reduction of NO with NH3 at low temperature. Chin. J. Catal. 2012, 33, 1448–1454. [Google Scholar] [CrossRef]
  14. Magalhães, F.; Moura, F.C.C.; Ardisson, J.D.; Lago, R.M. LaMn1-xFexO3 and LaMn0.1-xFe0.90MoxO3 perovskites: Synthesis, characterization and catalytic activity in H2O2 reactions. Mater. Res. 2008, 11, 307–312. [Google Scholar] [CrossRef]
  15. Yao, H.; Chen, Y.; Zhao, Z.; Wei, Y.; Liu, Z.; Zhai, D.; Liu, B.; Xu, C. Periodic DFT study on mechanism of selective catalytic reduction of NO via NH3 and O2 over the V2O5 (001) surface: Competitive sites and pathways. J. Catal. 2013, 305, 67–75. [Google Scholar] [CrossRef]
  16. Yang, Y.; Liu, J.; Wang, Z.; Liu, F. A skeletal reaction scheme for selective catalytic reduction of NOx with NH3 over CeO2/TiO2 catalyst. Fuel Process. Technol. 2018, 174, 17–25. [Google Scholar] [CrossRef]
  17. Song, Z.; Wang, B.; Yu, J.; Ma, C.; Zhou, C.; Chen, T.; Yan, Q.; Wang, K.; Sun, L. Density functional study on the heterogeneous oxidation of NO over α-Fe2 O3 catalyst by H2O2: Effect of oxygen vacancy. Appl. Surf. Sci. 2017, 413, 292–301. [Google Scholar] [CrossRef]
  18. Yan, X.; Liu, J.; Yang, Y.; Wang, Z.; Zheng, Y. A catalytic reaction scheme for NO reduction by CO over Mn-terminated LaMnO3 perovskite: A DFT study. Fuel Process. Technol. 2021, 216, 106798. [Google Scholar] [CrossRef]
  19. Ren, D.; Wu, K.; Luo, S.; Li, Y.; Gui, K.; Zuo, Z.; Guo, X. Study of the NH3-SCR Mechanism on LaMnO3 Surfaces Based on the DFT Method. Energies 2022, 15, 9099. [Google Scholar] [CrossRef]
  20. Araujo, R.B.; Rodrigues, G.L.; Dos Santos, E.C.; Pettersson, L.G. Adsorption energies on transition metal surfaces: Towards an accurate and balanced description. Nat. Commun. 2022, 13, 6853. [Google Scholar] [CrossRef]
  21. Zhang, R.; Yang, W.; Luo, N.; Li, P.; Lei, Z.; Chen, B. Low-temperature NH3-SCR of NO by lanthanum manganite perovskites: Effect of A-/B-site substitution and TiO2/CeO2 support. Appl. Catal. B Environ. 2014, 146, 94–104. [Google Scholar] [CrossRef]
  22. Delley, B. From molecules to solids with the DMol3 approach. J. Chem. Phys. 2000, 113, 7756–7764. [Google Scholar] [CrossRef]
  23. Yang, Y.; Liu, J.; Zhang, B.; Liu, F. Mechanistic studies of mercury adsorption and oxidation by oxygen over spinel-type MnFe2O4. J. Hazard. Mater. 2017, 321, 154–161. [Google Scholar] [CrossRef] [PubMed]
  24. Bergner, A.; Dolg, M.; Küchle, W.; Stoll, H.; Preuß, H. Ab initio energy-adjusted pseudopotentials for elements of groups 13–17. Mol. Phys. 1993, 80, 1431–1441. [Google Scholar] [CrossRef]
  25. Ren, D.; Gui, K. Study of the adsorption of NH3 and NOx on the nano-γFe2O3 (001) surface with density functional theory. Appl. Surf. Sci. 2019, 487, 171–179. [Google Scholar] [CrossRef]
  26. Halgren, T.A.; Lipscomb, W.N. The synchronous-transit method for determining reaction pathways and locating molecular transition states. Chem. Phys. Lett. 1977, 49, 225–232. [Google Scholar] [CrossRef]
  27. Yang, Y.; Liu, J.; Liu, F.; Wang, Z.; Ding, J. Comprehensive Hg/Br reaction chemistry over Fe2O3 surface during coal combustion. Combust. Flame 2018, 196, 210–222. [Google Scholar] [CrossRef]
  28. Liu, Y.; Liu, J.; Lin, Y.S.; Chang, M. Effects of water vapor and trace gas impurities in flue gas on CO2/N2 separation using ZIF-68. J. Phys. Chem. C 2014, 118, 6744–6751. [Google Scholar] [CrossRef]
  29. 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]
  30. Ren, D.; Gui, K.; Gu, S. Quantum chemistry study of SCR-NH3 nitric oxide reduction on Ce-doped γFe2O3 catalyst surface. Mol. Catal. 2021, 502, 111373. [Google Scholar] [CrossRef]
  31. Yao, G.H.; Wang, F.; Wang, X.B.; Gui, K.T. Magnetic field effects on selective catalytic reduction of NO by NH3 over Fe2O3 catalyst in a magnetically fluidized bed. Energy. 2010, 35, 2295–2300. [Google Scholar] [CrossRef]
  32. Anstrom, M.; Topsøe, N.Y.; Dumesic, J.A. Density functional theory studies of mechanistic aspects of the SCR reaction on vanadium oxide catalysts. J. Catal. 2003, 213, 115–125. [Google Scholar] [CrossRef]
  33. Ren, D.; Gui, K.; Gu, S.; Wei, Y. Study of the nitric oxide reduction of SCR-NH3 on nano-γFe2O3 catalyst surface with quantum chemistry. Appl. Surf. Sci. 2020, 509, 144659. [Google Scholar] [CrossRef]
  34. Ciardelli, C.; Nova, I.; Tronconi, E.; Chatterjee, D.; Bandl-Konrad, B. A “Nitrate Route” for the low temperature “Fast SCR” reaction over a V2O5–WO3/TiO2 commercial catalyst. Chem. Commun. 2004, 23, 2718–2719. [Google Scholar] [CrossRef] [PubMed]
  35. Savara, A.; Li, M.J.; Sachtler, W.M.; Weitz, E. Catalytic reduction of NH4NO3 by NO: Effects of solid acids and implications for low temperature DeNOx processes. Appl. Catal. B Environ. 2008, 81, 251–257. [Google Scholar] [CrossRef]
  36. Wang, D.; Zhang, L.; Kamasamudram, K.; Epling, W.S. In Situ-DRIFTS Study of Selective Catalytic Reduction of NOx by NH3 over Cu-Exchanged SAPO-34. ACS Catal. 2013, 3, 871–881. [Google Scholar] [CrossRef]
Figure 1. A unit of LaMn1−xFexO3.
Figure 1. A unit of LaMn1−xFexO3.
Energies 16 07609 g001
Figure 2. (a) LaMnO3 (001) 2 × 1 surface structure. (b) Modified LaMnO3 (001) 2 × 1 surface structure.
Figure 2. (a) LaMnO3 (001) 2 × 1 surface structure. (b) Modified LaMnO3 (001) 2 × 1 surface structure.
Energies 16 07609 g002
Figure 3. (a) NH3 adsorbed onto Mn ions. (b) NH3 adsorbed onto Fe ions.
Figure 3. (a) NH3 adsorbed onto Mn ions. (b) NH3 adsorbed onto Fe ions.
Energies 16 07609 g003
Figure 4. The NO adsorption configuration on catalyst surface. (a) Adsorption on Mn with N, (b) Adsorption on Mn with O, (c) Adsorption on Fe with N, (d) Adsorption on Fe with O.
Figure 4. The NO adsorption configuration on catalyst surface. (a) Adsorption on Mn with N, (b) Adsorption on Mn with O, (c) Adsorption on Fe with N, (d) Adsorption on Fe with O.
Energies 16 07609 g004
Figure 5. (a) O2 adsorbed onto Mn ions. (b) O2 adsorbed onto Fe ions.
Figure 5. (a) O2 adsorbed onto Mn ions. (b) O2 adsorbed onto Fe ions.
Energies 16 07609 g005
Figure 6. Catalytic decomposition of NH3 at surfactant sites.
Figure 6. Catalytic decomposition of NH3 at surfactant sites.
Energies 16 07609 g006
Figure 7. The decomposition of NH2NO.
Figure 7. The decomposition of NH2NO.
Energies 16 07609 g007
Figure 8. The decomposition of O2.
Figure 8. The decomposition of O2.
Energies 16 07609 g008
Figure 9. Dehydrogenation of NH3 in the presence of O.
Figure 9. Dehydrogenation of NH3 in the presence of O.
Energies 16 07609 g009
Figure 10. Dehydrogenation of NH2 in the presence of O.
Figure 10. Dehydrogenation of NH2 in the presence of O.
Energies 16 07609 g010
Figure 11. Generation of NHNO at Active Sites.
Figure 11. Generation of NHNO at Active Sites.
Energies 16 07609 g011
Figure 12. Generation of NO2 at Active Sites.
Figure 12. Generation of NO2 at Active Sites.
Energies 16 07609 g012
Figure 13. (a) Diagram of SCR Reaction at Mn Active Sites (b) Diagram of SCR Reaction at Fe Active Sites.
Figure 13. (a) Diagram of SCR Reaction at Mn Active Sites (b) Diagram of SCR Reaction at Fe Active Sites.
Energies 16 07609 g013
Table 1. Crystal parameters at sites before and after doping.
Table 1. Crystal parameters at sites before and after doping.
(a)O-Mn Bond Length (Å)(b)O-Fe Bond Length (Å)
O11.927O61.917
O21.929O71.922
O31.974O81.977
O31.991O91.975
O52.001O102.013
Table 2. Adsorption Parameters of NH3 Molecule at Mn and Fe Active Sites.
Table 2. Adsorption Parameters of NH3 Molecule at Mn and Fe Active Sites.
Active SitesNH3 Bond Angle
(°)
N-Me Bond Length
(Å)
Adsorption Energy
(kJ/mol)
Mulliken Charge
(a.u)
Mn110.0332.075−125.7070.349
Fe109.8641.999−152.7090.367
Table 3. Adsorption Parameters of NO Molecule at Mn and Fe Active Sites.
Table 3. Adsorption Parameters of NO Molecule at Mn and Fe Active Sites.
Absorption SitesN/O-Me Bond Length
(Å)
Adsorption Energy (kJ/mol)Mulliken Charge
(a.u)
A1.737−67.7930.254
B2.452−18.0970.063
C1.765−82.4140.244
D2.132−28.9470.086
Table 4. Adsorption Parameters of O2 Molecule at Mn and Fe Active Sites.
Table 4. Adsorption Parameters of O2 Molecule at Mn and Fe Active Sites.
Absorption SitesO-Me Bond Length
(Å)
Adsorption Energy
(kJ/mol)
Mulliken Charge
(a.u)
Mn1.988−57.211−0.105
Fe1.916−72.050−0.073
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ren, D.; Lin, K.; Mao, T.; Luo, S.; Liu, K.; Zuo, Z.; Gui, K. The Study of SCR Mechanism on LaMn1−xFexO3 Catalyst Surface Based DFT. Energies 2023, 16, 7609. https://doi.org/10.3390/en16227609

AMA Style

Ren D, Lin K, Mao T, Luo S, Liu K, Zuo Z, Gui K. The Study of SCR Mechanism on LaMn1−xFexO3 Catalyst Surface Based DFT. Energies. 2023; 16(22):7609. https://doi.org/10.3390/en16227609

Chicago/Turabian Style

Ren, Dongdong, Kangshuai Lin, Taipeng Mao, Siyi Luo, Kaijie Liu, Zongliang Zuo, and Keting Gui. 2023. "The Study of SCR Mechanism on LaMn1−xFexO3 Catalyst Surface Based DFT" Energies 16, no. 22: 7609. https://doi.org/10.3390/en16227609

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