Next Article in Journal
Iron-Based Alumina-Supported Catalysts for Clean Hydrogen Production from Ammonia
Previous Article in Journal
Metal-Free Carbon Catalysis for Flue Gas Pollutants Purification: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Selective Catalytic Reduction of NO by NH3 and SOx Poisoning Mechanisms on Mn3O4 Catalysts: A Density Functional Investigation

1
Shandong Key Laboratory of Intelligent Energy Materials, School of Materials Science and Engineering, China University of Petroleum (East China), Qingdao 266580, China
2
State Key Laboratory of Petroleum Molecular & Process Engineering, SINOPEC Research Institute of Petroleum Processing Co., Ltd., Beijing 100083, China
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(3), 241; https://doi.org/10.3390/catal15030241
Submission received: 23 January 2025 / Revised: 25 February 2025 / Accepted: 28 February 2025 / Published: 3 March 2025

Abstract

:
Mn-based oxides are promising catalysts for the selective catalytic reduction (SCR) of NOx by NH3 at low temperatures. However, fundamental NH3-SCR mechanisms and resistance mechanisms against SOx remain controversial. This study employed density functional theory (DFT) calculations to explore the intrinsic mechanisms of NH3-SCR and SOx poisoning on Mn3O4(001). Both NH3 and NO adsorb atop the surface Mn site (the Lewis acid site). In contrast to the traditional Langmuir–Hinshelwood (L-H) mechanism in which gaseous NO is first oxidized to form adsorbed nitrites or nitrates and then react with adsorbed NHx species to produce H2O and N2, a new potential L-H pathway is proposed that involves gaseous NO first adsorbing and then reacting with NH* to generate the key intermediate NHNO*, followed by the formation of H2O and N2. This L-H pathway is more efficient as it bypasses the NO oxidation step and is more selective for N2 formation by avoiding N2O production. In addition, the L-H mechanism is more favorable than the Eley–Rideal (E-R) mechanism because of the lower free energy profile. SO2 exhibits limited poisoning effects, whereas SO3 strongly poisons the Mn3O4(001) surface by occupying adsorption sites, hindering intermediate formation and producing ammonium bisulfate.

Graphical Abstract

1. Introduction

The selective catalytic reduction (SCR) of nitrogen oxides (NOx) by ammonia (NH3) is a pivotal technology for mitigating NOx emissions, which are notorious for their contribution to acid rain, photochemical smog, and respiratory health issues [1,2,3,4,5,6,7,8,9]. SCR catalysts are employed to facilitate the conversion of nitrogen oxides (NOx) into nitrogen (N2) and water (H2O), typically through a redox reaction with ammonia (NH3) as the reductant: 4NH3 + 4NO + O2 → 4N2 + 6H2O. The efficiency and durability of this process are heavily dependent on the performance of the catalyst used. Traditionally, V2O5-WO3 (MoO3)/TiO2 has been the predominant commercial catalyst for NH3-SCR due to its high activity and stability at temperatures of 300–400 °C [7,8,10,11,12]. However, the placement of these catalysts upstream in flue gas treatment processes can expose them to dust, sulfur dioxide (SO2), and other impurities, which can significantly diminish their denitration efficiency and lead to catalyst poisoning or deactivation over time. In response to these challenges, there is a growing trend toward integrating SCR catalysts in low-dust arrangements, where the inlet flue gas temperature is approximately 200 °C [7,8,13,14,15,16]. This shift highlights the importance of developing catalysts that exhibit robust catalytic activity at lower temperatures, and Mn-based oxide catalysts have shown promise in this regard due to their enhanced performance in the low-temperature range [2,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38].
Mn3O4 has garnered considerable attention due to its tetragonally distorted spinel structure and complex canted spin configuration, making it an intriguing subject for both experimental and theoretical studies [27,28,39,40,41,42,43]. Kapteijn et al. [43] reported that the SCR activity of Mn oxides decreased in the order of MnO2 > Mn5O8> Mn2O3 > Mn3O4. Despite MnO2 exhibiting the highest activity per unit surface area, it is more susceptible to sulfur poisoning due to its high surface active oxygen content, which reacts readily with SO2. On the other hand, Mn3O4, though displaying lower inherent NO oxidizing ability due to the reduced Mn valence state, offers promise as a low-temperature SCR deNOx catalyst with enhanced SO2 resistance. Previous studies highlighted by Kang et al. [44] and Xie et al. [45] have emphasized the presence of Mn3O4 as a significant component in active catalysts, suggesting its potential when modified appropriately. Mn3O4 offers a promising balance for enhancing both catalytic activity and durability. While not as extensively studied as other manganese oxides, an in-depth examination of the surface chemistry in SCR de-NOx processes using Mn3O4 could greatly advance our molecular-level understanding of Mn-based catalysts, particularly at low temperatures.
The NH3-SCR mechanism at low temperatures over MnOx-based catalysts has been extensively studied both experimentally and theoretically [7,8,46,47]. The widely accepted reaction pathway involves adsorbed NHx species reacting either with adsorbed nitrites/nitrates via the Langmuir–Hinshelwood (L-H) mechanism [30,44,48] or directly with gaseous NO and NO2 through the Eley–Rideal (E-R) mechanism [43,49,50], forming NHx−NOx intermediates that decompose into N2 and H2O. In the L-H mechanism, reactions occur between two adsorbed species. For instance, Wu et al. [51] demonstrated that the SCR de-NOx reaction on FeOx-MnOx/TiO2 catalysts follows the L-H mechanism, where stable bidentate nitrates react with adsorbed NH3 to form less stable monodentate nitrates and NH4+, which subsequently react with NO to produce N2 and H2O. Similar findings have been reported in other studies [52,53]. Kapteijn et al. [43] proposed a model for pure MnOx, where adsorbed NH3 is progressively dehydrogenated by surface oxygen. Qi et al. [30] observed that gaseous NH3 adsorbed on MnOx-CeO2 forms NH3, NH2, and OH species, with NO oxidized to NO2, which then reacts with adsorbed NH3 to form nitrates and nitrites that decompose into N2 and H2O. Conversely, Marbán et al. [49,50] suggested that SCR reactions over carbon-supported Mn3O4 catalysts proceed via the E-R mechanism, where surface-activated NH3 reacts primarily with gaseous NO2 and, to a lesser extent, NO. Xiong et al. [42] also found that the E-R mechanism dominates in low-temperature SCR reactions over Mn3O4 and (Cu1.0Mn2.0)1–δO4 spinels. Using DFT calculations, Yang et al. [41] investigated the SCR process on Mn3O4 (110) surfaces, revealing that NH3 adsorbs on Lewis acid sites and dehydrogenates into NH2* with the assistance of lattice oxygen. NH2* then reacts with gaseous NO to form NH2NO, which tends to convert into N2O rather than N2, with residual H atoms interacting with O2 to form H2O. Despite extensive research and the potential of Mn-based oxide catalysts, the fundamental NH3-SCR mechanism at low temperatures remains controversial. Additionally, N2 selectivity is a critical factor in de-NOx performance, as side reactions can produce N2O, a potent greenhouse gas. N2O formation can occur through several pathways: (1) 3NO → N2O + NO2 [54], (2) 2NH3 + 2O2 → N2O + 3H2O [55], (3) 4NO + 4NH3 + 3O2 → 4N2O + 6H2O [43,56], and (4) 4NO2 + 4NH3 + O2 → 4N2O + 6H2O [57,58]. Reaction intermediates significantly influence product distribution, making it essential to understand the pathways leading to nonselective NH3 oxidation. These findings underscore the need for a reassessment of the NH3-SCR mechanism on Mn3O4 surfaces to optimize N2 selectivity and minimize N2O formation.
In practical applications, residual SOx (<100 ppm) persists in flue gas post-desulfurization, adversely affecting catalyst performance [7,8,59]. SOx induces the formation of ammonium sulfate, which deposits on catalyst surfaces, deactivating them by blocking active sites. Metal oxide catalysts, particularly Mn-based ones, are vulnerable to deactivation due to ammonium sulfate and metal sulfates, which either block or destroy active sites [27,52,60,61,62,63,64,65]. The decomposition of ammonium sulfate occurs between 300 and 400 °C; thus, at lower temperatures, its accumulation significantly impairs catalyst denitration activity. For Mn-based oxide catalysts, SOx leads to permanent deactivation through sulfate deposition, limiting their practical application. The deactivation process by SO2 involves three stages: adsorption of SO2, its oxidation to SO3, and the subsequent deposition of ammonium sulfate or sulfation of active components. Wei et al. [66] highlighted that Mn-Ce/TiO2 catalysts exhibit resistance to sulfur poisoning in low-temperature SCR, attributing deactivation to SO2 outcompeting NH3 for active sites, as evidenced by DRIFT and DFT analyses. Despite these insights, understanding the resistance mechanism against SOx on Mn-based oxide catalysts remains challenging for low-temperature applications. Further research is essential to elucidate how SOx suppresses de-NOx activity at low temperatures, thereby enhancing the poison-resistant performance of Mn-based oxide catalysts.
Experimental studies on SCR have been limited by technical constraints, leaving critical questions about the reaction pathways and intermediates unresolved [7,8]. Therefore, an in-depth theoretical investigation employing DFT calculations is crucial for unveiling the fundamental aspects of the SCR process. In this work, DFT calculations of NH3-SCR mechanism were carried out on the Mn3O4(001) surface. The adsorption configurations and adsorption sites of NH3 and NO were determined. Through the calculated free energy changes in the reaction process, NH3 dehydrogenation, the formation of important intermediates (NH2NO and NHNO) and byproduct (N2O) via N−N bonding, and the desorption of N2 and H2O were analyzed in detail. In addition, the adsorption of SO2 and SO3 on pure and NHx pre-adsorbed Mn3O4(001) were discussed to explore the possible mechanisms of SOx poisoning.

2. Results and Discussion

2.1. SCR Mechanism

2.1.1. Adsorption Properties of NH3 and NO

Figure 1 shows the most stable adsorption configurations of NH3 and NO on the Mn3O4(001) surface, and Table 1 lists the corresponding geometric parameters and adsorption energies. NH3 could stably stay through the N atom at the top-Mn site with the N−Mn axis perpendicular to the surface. The N−Mn bond length is 2.088 Å, and the adsorption energy (ΔEads) is calculated to be −0.971 eV. Similar adsorption configurations were also obtained for NH3 on other Mn-based oxide surfaces [41,67,68,69]. Yang et al. [41] proposed that the most stable adsorption site of NH3 on Mn3O4(110) is the Mn site with the ΔEads of −1.00 eV. Chen et al. [40] also identified a top-Mn adsorbed NH3 on Mn3O4(101), and the corresponding ΔEads is −0.94 eV. NO adsorption involves three geometric types:
(1) NO binds to the Mn atom via the O atom with the N−O axis vertical to the surface. The O−Mn bond length is 2.115 Å and the adsorption strength is weakly mirrored by the ΔEads of −0.375 eV. (2) When NO begins to adsorb in a horizontal orientation, a bridge-type adsorption configuration forms accordingly. In this configuration, the N and O atoms bond with two Mn atoms on the surface, and the corresponding bond lengths are 1.758 Å and 2.241 Å, respectively. The adsorption strength between NO and the surface increases accordingly, reflected by the lower ΔEads of −1.431eV. (3) NO can also adsorb vertically on the surface via the N−Mn bonding. The N−Mn distance of 1.671 Å is substantially shorter than above N−Mn and O−Mn bond lengths, indicating a stronger N−Mn interaction. This accounts for the lowest ΔEads value of −1.864 eV for NO binding at the top-Mn (N−Mn) site. On Mn3O4(101), for comparison, Chen et al. [40,70] confirmed a vertical adsorption configuration of NO via the N−Mn bond with the ΔEads of −2.37 eV. Zhang et al. [71] built a Mn3O4/TiO2 catalyst model to study the adsorption properties of NOx. They also identified the stable NO adsorption on the Mn site with the N−Mn bond length of 1.863 Å and the ΔEads of −2.66 eV.

2.1.2. Reaction Pathways

The SCR process can generally be divided into four stages: molecular adsorption and NH3 dehydrogenation, N−N bonding, the formation of N2 and H2O, and surface regeneration [72]. The calculated free energy profiles for NH3-SCR on Mn3O4(001) are displayed in Figure 2 with the corresponding intermediate structures presented in Figure 3. Two main reaction pathways are discussed in detail. One is the “E-R” mechanism, in which NH3 adsorbs first, followed by dehydrogenation, and then the adsorbed NH2 reacts with gas-phase NO. The other is the “L-H” mechanism, which starts with co-adsorbed NO and NH3.
  • The E-R Mechanism
The oxidative dehydrogenation of NH3 to active NH2 species was decisive for the E-R mechanism because NH2 species could directly react with gaseous NO to generate important NH2NO intermediate that subsequently decomposed to N2 and H2O. The NH3 activation to V-NH2 amide species over vanadyl centers has been demonstrated over bulk V2O5 [73], V2O5−TiO2 [74], and V2O5−WO3−TiO2 [75] catalysts. Adsorbed NH3* dissociates into NH2* and H* species (Figure 3a) with an increase in free energy of 0.585 eV. Further step-wise dehydrogenations of NH2* to produce NH* (Figure 3h) and N* (Figure 3i) are thermodynamically unfavorable due to the high endothermicity with the respective ΔG values of 0.620 and 1.609 eV. Following the classic E-R pathway [7,8], adsorbed NH2* can react with a gas-phase NO molecule, forming a stable N–N bond with a bond length of 1.459Å. This step is endothermic, reflected by the ΔG value of 0.333 eV. The newly formed intermediate, NONH2* (Figure 3b), still binds to the surface Mn site through the N atom of the NH2 group. Then, NONH2* converts to NH2NO* (Figure 3c), in which the adsorbed configuration switches to the NO group bonding with the Mn site rather than NH2. For NH2NO*, the N−Mn bond length between the N atom of NO and Mn decreases to 2.152 Å compared with that of 2.181 Å between NH2 and Mn for NONH2*, indicating that this geometric transition from NONH2* to NH2NO* increases the interaction between adsorbed intermediate and surface, reflected by the exothermicity of this step (ΔG = −0.305 eV). The NH2NO* intermediate further undergoes dehydrogenation by transferring one H atom to an adjacent surface O site, forming cis-NHNO* species (Figure 3o) and a second surface H* species. The free energy is reduced by 0.228 eV in this step. To facilitate the removal of the third H atom, cis-NHNO* converts to trans-NHNO* (Figure 3d) with the N−H axis pointing to the surface (ΔG = 0.238 eV). Accompanied by the cleavage of the N−H bond, a new N−Mn bond is formed, generating a π-bonding configuration of ONN* (Figure 3e) at the Mn site. This third dehydrogenation step is endothermic with a ΔG of 0.543 eV. Afterward, an adsorbed H2OL molecule (“L” represents a surface lattice O atom) can be produced by the combination of two H atoms and one surface lattice OL atom (Figure 3f). This step requires the breaking of one O−H bond, which accounts for the high endothermicity with the ΔG of 0.959 eV. Subsequent H2O desorption leaves one oxygen vacancy (OV) on the surface (Figure 3g), followed by the refilling of the OV by the O atom of the ONN* and the N2 desorption. The formation of gas-phase H2O and N2 releases significant amounts of energy, as indicated by their respective ΔG values of −1.016 and −3.769 eV. These energy releases are primarily due to the refilling process of OV by the O atom from the ONN*, which stabilizes the system and contributes to the overall exergonic nature of the reactions. After the reactions in the aforementioned three stages are completed, an adsorbed H atom remains atop one surface lattice O atom of Mn3O4(001) surface, forming a Brønsted acid site denoted as H* (Figure 3t). During the final stage of catalyst surface regeneration, the two residual H atoms combine with a lattice O atom from the surface to form H2O, which then desorbs from the surface. The OV created on the surface as a result of this process is subsequently replenished by 1/2 O2.
For the H* Brønsted acid site formed at the end of the third stage (formation of H2O and N2), we further investigated if this H* site could act as an active center for the NH3- SCR reaction. Generally, there are two kinds of acid sites on the surface of metal oxide catalysts, the Lewis acid site and Brønsted acid site. The adsorption of NH3 can be divided into two categories according to the type of acid site: one is the NH3 adsorbed on Brønsted acid sites, labeled as NH4+; the other one is the NH3 adsorbed on Lewis acid sites, labeled as NH3*, which has been described above. Herein, the adsorption of NH3 at the H* site to form ammonium species (NH4*) was evaluated (Figure S1 in the Supplementary Materials). The results showed that the interaction between NH3 and H* is characterized by a ΔEads value of only −0.559 eV, with no indication of N−H bond formation (dN−H = 1.706 Å). This suggests that the binding of NH3 at the H* site is weak, resulting in an unstable intermediate. Consequently, it is inferred that the Lewis acid sites (surface Mn sites), rather than the H* Brønsted acid sites, are the principal active centers for NH3 adsorption. Moreover, the H* Brønsted acid site does not favorably interact with NO either. Therefore, the H* Brønsted acid site likely plays a passive role in the NH3-SCR mechanism, acting more as a spectator than an active participant. In the final stage of the catalyst surface regeneration, the NH3 and NO species adsorbed at the Mn-based Lewis acid sites proceed through the aforementioned three reaction stages twice, leading to the production of two surface H*. Then, adsorbed H2O is formed through H transfer, and H2O desorption leads to the formation of OV. Finally, the OV is replenished by 1/2 O2, which regenerates the catalyst surface to the initial state and completes the NH3-SCR catalytic cycle. The above results underscore the importance of the Lewis acid sites in facilitating the overall SCR process while highlighting the limited involvement of the Brønsted acid sites in the key reaction steps.
It is important to note that after the formation of the ONN* intermediate, besides the original E-R mechanism where H transfer occurs and results in the formation of H2OL, there is also a possibility for ONN* to directly desorb, producing gaseous N2O (Figure 3j). In contrast to the highly endothermic pathway leading to H2O formation, the N2O desorption process is significantly exothermic, mirrored by the ΔG of −1.125 eV. Therefore, from a thermodynamic perspective, the production of N2O is more favorable than the E-R pathway that produces H2O followed by N2. This is in good agreement with the DFT results on Mn3O4(110) by Yang et al. [41], in which the adsorbed NH2NO*, generated by the reaction of adsorbed NH2* and gaseous NO, tends to convert into N2O rather than N2. Our results show that on the Mn3O4 surface, the selectivity toward the byproduct N2O is higher than that toward N2, suggesting that the classic E-R mechanism for the NH3-SCR of NO might be flawed and cannot fully account for the product distribution of the SCR process on Mn3O4 surface.
  • The L-H Mechanism
The L-H pathway begins with the co-adsorption of NH3 and NO (Figure 3k) on the Mn3O4(001) surface. NO and NH3 adsorb on two neighboring Mn sites. Adsorbed NH3 undergoes dehydrogenation, releasing one H atom (Figure 3l). This step is energetically favorable in the L-H mechanism (ΔG = 0.350 eV) compared to the E-R mechanism (ΔG = 0.585 eV), as indicated by the lower energy required for dehydrogenation. The presence of adsorbed NO alters the electronic environment of the surface oxygen atoms, stabilizing the adsorption of hydrogen and creating conditions that favor the further dehydrogenation of NH2* to NH* (Figure 3m). Notice that the free energy changes in the N−N coupling between adsorbed NH2* and NO* to form NONH2* (1.203 eV) and NH2NO* (0.898 eV) are higher than the energy required to dissociate the second hydrogen from NH2* to form NH* + H* (0.637 eV), indicating that the reaction steps of NH2*+ NO* → NONH2*/NH2NO* are thermodynamically unfavorable. The transition from NH2* to NH* is still endothermic with the ΔG of 0.636 eV. Once formed, NH* interacts with NO* to produce the intermediate NONH* (Figure 3n). The intermediate NONH* is then transformed into cis-NHNO* (Figure 3o), which serves as a common intermediate where the E-R and L-H pathways converge. The next stage is the formation of H2O and N2. Adsorbed cis-NHNO* is initially isomerized to iso-NHNO* (Figure 3p), in which the O atom of the NO group binds to the Mn site instead of the N atom. This isomerization step is slightly exothermic with a ΔG of −0.127 eV. Then, one adsorbed H atom approaches a surface OL−H group, forming an adsorbed H2OL (Figure 3q). Similar to the case in the E-R path, this H transfer process to generate H2O is difficult, mirrored by the high ΔG value of 0.719 eV. Afterward, the O atom of iso-NHNO* occupies the position of OL, producing adsorbed HNN*, while the original H2OL is lifted up over the surface and becomes adsorbed H2O*, binding to an adjacent Mn site (Figure 3r). This step is slightly exothermic with a ΔG of −0.251 eV. The subsequent desorption of H2O is also slightly exothermic (ΔG = −0.052 eV), but it does not involve the formation of an OV (Figure 3s), which is different from the case in the E-R path (Figure 3g). After H2O desorption, the N−N group within the HNN* intermediate subsequently desorbs from the surface in the form of N2. Concurrently, the N−H bond breaks and the H atom is captured by a surface O atom, leading to the formation of a highly stable O−H bond (Figure 3t). The formation of this bond is substantially exothermic, resulting in a significant change in free energy (ΔG = −3.333 eV) for this N2 desorption step. For the catalyst surface regeneration, the final stage of the L-H path merges with that of the E-R path.
In summary, the L−H mechanism is more favorable than the E−R mechanism because of the lower free energy profile. The key intermediates of the N−N bonding are the NH2HO* in the E−R path and NHNO* in the L−H path, respectively, and the N−N bonding is easier for the L−H path thermodynamically. The potential-determining step (PDS) is the H transfer step for both the E−R and L−H paths, in which the cleavage of one O−H bond occurs accompanied by the formation of adsorbed H2OL. In the traditional L-H mechanism, gaseous NO is first oxidized to form adsorbed nitrites or nitrates, which then react with adsorbed NHx species to ultimately produce H2O and N2. In contrast, a potential L-H pathway identified in this work involves gaseous NO first adsorbing and then reacting with NH* to form the intermediate NHNO*, which decomposes into H2O and N2. This new L-H pathway is more efficient as it bypasses the NO oxidation step and is more selective for N2 formation by avoiding N2O production.

2.2. SOx Poisoning Mechanisms

2.2.1. Adsorption Properties of SO2 and SO3

SO2: Three distinct adsorption configurations of SO2 on the Mn3O4(001) surface were investigated (Figure 4). In the first configuration, the S atom of SO2 is directly bonded to a surface Mn atom, forming an S−Mn bond with a length of 2.763 Å. This interaction results in a relatively weak ΔEads of −0.284 eV, indicating that this structure is less stable. The second configuration involves both O atoms of SO2 being attracted by adjacent surface Mn atoms, leading to the formation of two O−Mn bonds with lengths of 1.955 Å and 1.949 Å, respectively. The strong interaction between the Mn and O atoms shortens these bond lengths compared to the first configuration, resulting in a more stable geometry with the ΔEads of −0.833 eV. The third and most stable configuration features SO2 interacting with the Mn-O-Mn structure of the surface, forming one S−O bond and two O−Mn bonds. The corresponding bond lengths are 1.662 Å and 1.983/2.021 Å, respectively. This configuration exhibits the most negative ΔEads of −1.451 eV, signifying its stability and preference over the other two configurations. For comparison, Liu et al. [72] reported that the most stable adsorption of SO2 on Ti-doped CeO2 catalysts involves a direct S−Ti bond with a ΔEads of −3.010 eV, suggesting that Mn3O4 has a much weaker affinity for SO2 than Ti-doped CeO2 and thereby possesses certain anti-sulfurization properties. Xiong et al. [42] determined that adsorbed SO2 forms two bonds with the Mn3O4(101) surface: one S−O bond with a surface lattice oxygen atom and one O−Mn bond with a surface Mn atom. Compared with the configuration on Mn3O4(001), where both oxygen atoms of SO2 form bonds with surface Mn atoms, the SO2 on Mn3O4(101) forms one fewer O−Mn bond but exhibits a stronger binding strength, as evidenced by its more negative ΔEads value of −1.79 eV. This indicates that Mn3O4 is facet-sensitive, and different crystal facets may exhibit different anti-poisoning performances. Compared with NO, which has an adsorption energy of −1.864 eV, SO2 exhibits a lower adsorption strength on the Mn3O4(001) surface with an ΔEads value of −1.451 eV, indicating a small impact of SO2 on NO adsorption.
SO3: SO3 exhibits two stable adsorption structures on the Mn3O4(001) surface (Figure 4). In the first upright structure, two S−O bonds of SO3 form two O−Mn bonds with the surface, while the third S−O bond points away from it, with a ΔEads of −1.424 eV. The second flat structure evolves from the first, where the S atom additionally bonds with a nearby surface O atom, still with the third S−O bond slightly oriented away from the surface. This second structure has a more negative ΔEads of −2.301 eV, indicating that SO3 adsorption is more competitive compared to NH3, NO, and SO2 on Mn3O4(001).

2.2.2. Formation of SOx-NH and SOx-NH2 Complexes

Among the four stages of NH3-SCR on Mn3O4(001), the N−N bond formation to create intermediates NH2NO* (E-R path) and NHNO* (L-H path) is crucial. However, SOx species in the reaction atmosphere compete with NO for adsorption sites on Mn3O4(001), and also for binding with NH2* and NH*, potentially hindering N−N bonding and disrupting the NH3-SCR process. Figure 5 illustrates the binding configurations of SO2 and SO3 with NH2* and NH*. For the formation of NONH2*, the adsorption energy of NO binding to the N atom of NH2* is −0.515 eV. SO2 binds more readily to NH2* and occupies the N site with an adsorption energy of −1.342 eV, forming a more stable SO2-NH2* structure with a S−N bond length of 1.922 Å. For the formation of NONH*, the adsorption energy of NO binding to the N atom of NH* is −1.610 eV. SO2 also binds to NH* to form SO2-NH* with a S−N bond length of 1.701 Å and an adsorption energy of −1.536 eV. Consequently, while the formation of NONH* in the L-H path is less affected by SO2 adsorption, the formation of NONH2* in the E-R path is significantly impacted by SO2 binding to NH2*. Combining this with the ΔEads values of NO (−1.864 eV) and SO2 (−1.451 eV) on Mn3O4(001), it can be concluded that the poisoning effects of SO2 on the surface adsorption sites and the formation of key intermediates are limited, proving that the Mn3O4 surface has certain anti-SO2 poisoning properties for the NH3-SCR process. The adsorption energies of SO3 binding to the N atom of NH2* and NH* (Figure 5c,f) were calculated to be −2.304 and −2.803 eV, respectively, indicating a much higher binding strength compared with SO2. Combining with the ΔEads of SO3 (−2.301 eV) on Mn3O4(001), we found that the poisoning effects of SO3 on the surface adsorption sites and the formation of key intermediates are substantially strong.

2.2.3. Formation of NH4HSO4

The conversion of SO2 to SO3 requires a reaction temperature above 400 °C. In contrast, the working temperature for Mn3O4, the low-temperature denitrification catalyst, is generally below 300 °C. Consequently, within the current catalytic system utilizing Mn3O4, SO2 does not convert to SO3, and thus it cannot further transform into NH4HSO4. The formation of NH4HSO4 primarily results from the existing SO3. In the gas-phase mechanism for NH4HSO4 formation [10], SO3 reacts with H2O to form H2SO4, which then binds with NH3 to produce NH4HSO4. For the most stable adsorption configuration of SO3 on Mn3O4(001), however, the calculation of adsorbed SO3 and gas-phase H2O to form H2SO4 was not convergent. Instead, we found that vertically adsorbed SO3 can sequentially react with gas-phase H2O and NH3 to yield NH4HSO4 (Figure 6), with corresponding free energy changes of −0.094 and −0.193 eV. For the newly formed adsorption structure of NH4HSO4, we observed that one O−H bond has broken, with the dissociated H atom relocating to an adjacent surface O atom, thereby creating a H* Brønsted acid site. The O−H distance is 2.010 Å, which falls within the typical range for a classic O−H hydrogen bond (1.5~2.2 Å), indicating a weak H−bond interaction. This H−bond helps stabilize NH4HSO4 on the Mn3O4 surface. The continuous exothermic nature of NH4HSO4 formation suggests thermodynamic feasibility and further poisoning of the Mn3O4 surface.

3. Computation Methods

DFT calculations were performed with the Vienna ab initio simulation package (VASP) [76,77] using the projector augmented wave method (PAW) [43,78] to describe the interaction between core and valence electrons. The exchange correlation effect was processed using generalized gradient approximation developed by Perdew–Burke–Ernzerhof (GGA-PBE) [79]. Plane waves were included for the electronic wave functions up to a cutoff energy of 400 eV. The convergence criteria for the energy and force were set at 10−5 eV and −0.03 eV/Å, respectively. The Gaussian smearing method with a width of 0.05 eV was used to speed up the convergence of electronic structure optimization. All calculations were performed with spin polarization. The lattice constants of trimanganese tetroxide (Mn3O4) were fitted by the Birch–Murnaghan equation of state, and the Brillouin zones were sampled with 7 × 7 × 7 k-point meshes [80]. The lattice constant of Mn3O4 was calculated to be 6.286 Å, in good agreement with the experimental value (6.345 Å) [67]. This calculated lattice constant was subsequently used in all calculations.
As shown in Figure 7, the Mn3O4(001) surface was built with a 2 × 2 periodic slab model, consisting of nine atomic layers and a vacuum gap of 15 Å. Full geometry optimizations were performed for all relevant adsorbates and the uppermost four atomic layers of the Mn3O4(001) slab without symmetry restriction, while the bottom five atomic layers were fixed in the positions at the calculated lattice constant. The reciprocal space was sampled by a grid of (3 × 3 × 1) k-points generated automatically using the Monkhorst–Pack method [80]. The (001) surface was chosen because it is the most favored and involves the lowest cleavage energy (2.33 J·m−2) and surface energy (1.40 J·m−2) compared with other crystallographic planes, such as (100), (101), and (110) [81]. The (001) surface consists of two alternating terminations, denoted as Mntet and Mnoct-O. The Mntet termination is composed of tetrahedral Mn2+ cations. The Mnoct-O termination comprises Mn3+ with octahedral coordination and O2- anions. The Mntet layer terminated (001) surface is not stable and will undergo reconstruction, similar to the case on the Fe-rich Fe3O4(001) surface, while the Mnoct-O layer terminated (001) surface remains stable and exhibits high catalytic activity [82,83]. Therefore, the Mnoct-O termination was adopted in this work.
The adsorption energy (ΔEads) was calculated using the following formula:
ΔEads = Eadsorbate/substrateEadsorbateEsubstrate
In the formula, Eadsorbate and Esubstrate are the total energies of the free adsorbate molecule and the substrate, respectively. The Eadsorbate/substrate is the total energy of the adsorbate molecule adsorbed on the substrate. In this definition, the negative value of ΔEads corresponds to the stable adsorption.
Without considering the contribution of thermal internal energy, the Gibbs free energy change of each elementary step is determined as
ΔG = ΔE + ΔEZPETΔS
where ΔE is the reaction energy of each elementary step and ΔEZPE and ΔS are the zero point energy (ZPE) difference and the reaction entropy difference between the product and reactant of each elementary step, respectively. In this work, the vibrational entropies were considered and the vibrational frequencies (ν) were calculated by the finite difference method [68]. The ZPE of each adsorbed intermediate can be expressed as E ZPE = i 1 2 h v i , and the vibrational entropy contribution (Sνib) was calculated by the following standard equation from statistical mechanics [69]:
S vib = R i ( Θ vi T exp ( Θ vi T ) 1 ln [ 1 exp ( Θ vi T ) ] )
where R is the gas constant, T is the absolute temperature (500 K), and Θνi is the characteristic vibrational temperature, which equals hνi/kB. Here, h is Planck’s constant, νi is the vibrational frequency, and kB is Boltzmann’s constant. For each gas phase species, the entropy contribution is obtained from NIST database [84]. This method has been successfully applied to describe redox reactions of nitrogen and oxygen successfully [85].

4. Conclusions

DFT calculations were performed to study the intrinsic mechanisms of NH3-SCR and SOx poisoning on Mn3O4(001). Both NH3 and NO adsorb atop the surface Mn site via N−Mn bonding, and the corresponding adsorption energies were calculated to be −0.971 and −1.864 eV, respectively. The Lewis acid site (surface Mn site) is the principal active center for NH3 and NO adsorption, while the H* Brønsted acid site does not favorably interact with either NH3 or NO, demonstrating the importance of the Lewis acid sites in facilitating the overall SCR process, while highlighting the limited involvement of the Brønsted acid sites in the key reaction steps. The L-H mechanism is more favorable than the E-R mechanism because of the lower free energy profile. The key intermediates of the N−N bonding are the NH2HO* in the E−R path and NHNO* in the L−H path, and the N−N bonding is easier for the L−H path thermodynamically. The potential-determining step is the surface H transfer step for both the E−R and L−H paths, in which the cleavage of one O−H bond occurs accompanied by the formation of adsorbed H2OL. In contrast to the traditional L-H mechanism in which gaseous NO is first oxidized to form adsorbed nitrites or nitrates and then react with adsorbed NHx species to produce H2O and N2, a potential L-H pathway is identified in this work, which involves gaseous NO first adsorbing and then reacting with NH* to generate the key intermediate NHNO*, followed by the formation of H2O and N2. This new L-H pathway is more efficient as it bypasses the NO oxidation step and is more selective for N2 formation by avoiding N2O production.
Based on the exploration of the surface NH3-SCR process, the poisoning mechanisms of SOx on Mn3O4(001) are initially revealed. SOx species in the reaction atmosphere compete with NO for adsorption sites on Mn3O4(001) and also for binding with NH2* and NH*, potentially hindering N−N bonding and disrupting the NH3-SCR process. Compared with NO, SO2 exhibits a lower adsorption strength on Mn3O4(001) with the adsorption energy of −1.451 eV, indicating a small impact of SO2 on NO adsorption. For the formation of NONH2*, the adsorption energy of NO binding to NH2* is only −0.515 eV, while SO2 binds more readily to NH2* and occupies the N site with an adsorption energy of −1.342 eV, forming a more stable SO2-NH2* structure. For the formation of NONH*, the adsorption energy of NO binding to NH* is −1.610 eV, while SO2 can also bind to NH*, forming SO2-NH* with an adsorption energy of −1.536 eV. Consequently, the formation of NONH2* in the E-R path can be significantly impacted by SO2 binding to NH2*, while the formation of NONH* in the L-H path is less affected by SO2 adsorption to NH*. The poisoning effects of SO2 on the surface adsorption sites and the formation of key intermediates are limited, proving that the Mn3O4 surface has certain anti-SO2 poisoning properties for the NH3-SCR process. With a more negative adsorption energy of −2.301 eV, SO3 adsorption is more competitive compared to NH3, NO, and SO2 on Mn3O4(001). The adsorption energies of SO3 binding to the N atom of NH2* and NH* were calculated to be −2.304 and −2.803 eV, respectively, indicating a much higher binding strength compared with SO2. In addition, NH4HSO4, as a poison to the Mn3O4 surface, can be generated from SO3 via the reaction with NH3 and H2O. Compared with SO2, SO3 exhibits substantially stronger poisoning effects, which include occupying adsorption sites, hindering the formation of key intermediates, and leading to the formation of ammonium bisulfate. This study provides detailed insights into the NH3-SCR reaction mechanisms, as well as the SOx poisoning mechanisms, thereby guiding the rational design of more efficient and durable Mn-based SCR catalysts.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/catal15030241/s1. Figure S1: The adsorption of NH3 at the H* site to form ammonium species (NH4*). on Mn3O4(001).

Author Contributions

Conceptualization, H.Z. and X.G.; methodology, Y.F.; software, W.G.; validation, D.L.; investigation, Z.L., X.W., and X.L.; resources, W.G.; writing—original draft preparation, Z.L.; writing—review and editing, H.Z.; visualization, Z.L.; supervision, H.R.; project administration, X.G.; funding acquisition, H.Z., X.Z., and H.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by State Key Laboratory of Petroleum Molecular & Process Engineering (RIPP, SINOPEC), 36800000-23-ZC0699-0042, the National Natural Science Foundation of China, 22072182, 22473114 and U23B208, and the Shandong Provincial Natural Science Foundation of China, ZR2023MB034.

Data Availability Statement

The data presented in this study are available on request from the corresponding authors due to privacy.

Conflicts of Interest

Authors Xiaoxin Zhang and Xiaoxiao Gongwas employed by the company State Key Laboratory of Petroleum Molecular & Process Engineering. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Feng, X.; Zhu, J.; Song, K.; Zeng, J.; Zhou, X.; Guo, X.; Lin, K.; Zhang, C.; Xie, C.; Shi, J.-W. Insight into the Reasons for Enhanced NH3-SCR Activity and SO2 Tolerance of Mn-Co Layered Oxides. Sep. Purif. Technol. 2024, 336, 126285. [Google Scholar] [CrossRef]
  2. Liu, W.; Yang, Z.; Luan, X.; Zhai, Y.; Zhang, J.; Wang, L.; Wang, Z. Optimizing the Catalytic Performance of MnCo2O4 Spinel Catalysts for the NH3-SCR: Structure-Activity Relationships and Reaction Mechanism. Appl. Surf. Sci. 2024, 674, 160915. [Google Scholar] [CrossRef]
  3. Chen, W.; Zou, R.; Wang, X. Toward an Atomic-Level Understanding of the Catalytic Mechanism of Selective Catalytic Reduction of NOx with NH3. ACS Catal. 2022, 12, 14347–14375. [Google Scholar] [CrossRef]
  4. Shi, Z.; Peng, Q.; E, J.; Xie, B.; Wei, J.; Yin, R.; Fu, G. Mechanism, Performance and Modification Methods for NH3-SCR Catalysts: A Review. Fuel 2023, 331, 125885. [Google Scholar] [CrossRef]
  5. Park, E.D. Recent Progress on Low-Temperature Selective Catalytic Reduction of NOx with Ammonia. Molecules 2024, 29, 4506. [Google Scholar] [CrossRef]
  6. Guan, B.; Jiang, H.; Wei, Y.; Liu, Z.; Wu, X.; Lin, H.; Huang, Z. Density Functional Theory Researches for Atomic Structure, Properties Prediction, and Rational Design of Selective Catalytic Reduction Catalysts: Current Progresses and Future Perspectives. Mol. Catal. 2021, 510, 111704. [Google Scholar] [CrossRef]
  7. Liu, C.; Shi, J.-W.; Gao, C.; Niu, C. Manganese Oxide-Based Catalysts for Low-Temperature Selective Catalytic Reduction of NOx with NH3: A Review. Appl. Catal. Gen. 2016, 522, 54–69. [Google Scholar] [CrossRef]
  8. Han, L.; Cai, S.; Gao, M.; Hasegawa, J.; Wang, P.; Zhang, J.; Shi, L.; Zhang, D. Selective Catalytic Reduction of NOx with NH3 by Using Novel Catalysts: State of the Art and Future Prospects. Chem. Rev. 2019, 119, 10916–10976. [Google Scholar] [CrossRef]
  9. Paolucci, C.; Khurana, I.; Parekh, A.A.; Li, S.; Shih, A.J.; Li, H.; Di Iorio, J.R.; Albarracin-Caballero, J.D.; Yezerets, A.; Miller, J.T.; et al. Dynamic Multinuclear Sites Formed by Mobilized Copper Ions in NOxSelective Catalytic Reduction. Science 2017, 357, 898–903. [Google Scholar] [CrossRef]
  10. Wang, X.; Du, X.; Liu, S.; Yang, G.; Chen, Y.; Zhang, L.; Tu, X. Understanding the Deposition and Reaction Mechanism of Ammonium Bisulfate on a Vanadia SCR Catalyst: A Combined DFT and Experimental Study. Appl. Catal. B Environ. 2020, 260, 118168. [Google Scholar] [CrossRef]
  11. Zhu, M.; Lai, J.-K.; Tumuluri, U.; Wu, Z.; Wachs, I.E. Nature of Active Sites and Surface Intermediates during SCR of NO with NH3 by Supported V2O5–WO3 /TiO2 Catalysts. J. Am. Chem. Soc. 2017, 139, 15624–15627. [Google Scholar] [CrossRef]
  12. Lai, J.-K.; Wachs, I.E. A Perspective on the Selective Catalytic Reduction (SCR) of NO with NH3 by Supported V2O5–WO3/TiO2 Catalysts. ACS Catal. 2018, 8, 6537–6551. [Google Scholar] [CrossRef]
  13. Liu, J.; Wei, Y.; Li, P.-Z.; Zhang, P.; Su, W.; Sun, Y.; Zou, R.; Zhao, Y. Experimental and Theoretical Investigation of Mesoporous MnO2 Nanosheets with Oxygen Vacancies for High-Efficiency Catalytic DeNOx. ACS Catal. 2018, 8, 3865–3874. [Google Scholar] [CrossRef]
  14. Meng, D.; Zhan, W.; Guo, Y.; Guo, Y.; Wang, L.; Lu, G. A Highly Effective Catalyst of Sm-MnOx for the NH3-SCR of NOx at Low Temperature: Promotional Role of Sm and Its Catalytic Performance. ACS Catal. 2015, 5, 5973–5983. [Google Scholar] [CrossRef]
  15. Gao, F.; Tang, X.; Yi, H.; Zhao, S.; Li, C.; Li, J.; Shi, Y.; Meng, X. A Review on Selective Catalytic Reduction of NOx by NH3 over Mn–Based Catalysts at Low Temperatures: Catalysts, Mechanisms, Kinetics and DFT Calculations. Catalysts 2017, 7, 199. [Google Scholar] [CrossRef]
  16. Zhang, S.; Zhang, B.; Liu, B.; Sun, S. A Review of Mn-Containing Oxide Catalysts for Low Temperature Selective Catalytic Reduction of NOx with NH3: Reaction Mechanism and Catalyst Deactivation. RSC Adv. 2017, 7, 26226–26242. [Google Scholar] [CrossRef]
  17. Yang, Y.; Bian, X.; Xie, F.; Bai, Y.; Wang, J. Research Progress in the Composition and Performance of Mn-Based Low-Temperature Selective Catalytic Reduction Catalysts. Appl. Sci. 2024, 14, 10198. [Google Scholar] [CrossRef]
  18. Zhang, J.; Ma, Z.; Cao, A.; Yan, J.; Wang, Y.; Yu, M.; Hu, L.; Pan, S. Research Progress of Mn-Based Low-Temperature SCR Denitrification Catalysts. RSC Adv. 2024, 14, 32583–32601. [Google Scholar] [CrossRef] [PubMed]
  19. Xu, J.; Shen, H.; Zhou, X.; Zheng, T.; Guo, F.; Zhang, Q.; Duan, M. Enhancement Low-Temperature NH3-SCR Activity of the Fe-Mn-Mo/TiO2 Catalyst and Its DFT Calculations and Kinetics. Mol. Catal. 2023, 551, 113657. [Google Scholar] [CrossRef]
  20. Guo, R.; Qin, B.; Wei, L.; Yin, T.; Zhou, J.; Pan, W. Recent Progress of Low-Temperature Selective Catalytic Reduction of NOx with NH3 over Manganese Oxide-Based Catalysts. Phys. Chem. Chem. Phys. 2022, 24, 6363–6382. [Google Scholar] [CrossRef] [PubMed]
  21. Zhang, Z.; Li, J.; Tian, J.; Zhong, Y.; Zou, Z.; Dong, R.; Gao, S.; Xu, W.; Tan, D. The Effects of Mn-Based Catalysts on the Selective Catalytic Reduction of NOx with NH3 at Low Temperature: A Review. Fuel Process. Technol. 2022, 230, 107213. [Google Scholar] [CrossRef]
  22. Zhang, Y.; Han, B.; Chen, Y.; Xia, K.; Gao, Q.; Zhou, C. Understanding the Mechanism of Selective Catalytic Reduction on Spinel TiMn2O4(001) Surface. Mol. Catal. 2022, 518, 112070. [Google Scholar] [CrossRef]
  23. Qi, G.; Yang, R.T. Low-Temperature Selective Catalytic Reduction of NO with NH3 over Iron and Manganese Oxides Supported on Titania. Appl. Catal. B Environ. 2003, 44, 217–225. [Google Scholar] [CrossRef]
  24. Li, Y.; Li, Y.; Wang, P.; Hu, W.; Zhang, S.; Shi, Q.; Zhan, S. Low-Temperature Selective Catalytic Reduction of NOx with NH3 over MnFeOx Nanorods. Chem. Eng. J. 2017, 330, 213–222. [Google Scholar] [CrossRef]
  25. Yang, S.; Wang, C.; Li, J.; Yan, N.; Ma, L.; Chang, H. Low Temperature Selective Catalytic Reduction of NO with NH3 over Mn–Fe Spinel: Performance, Mechanism and Kinetic Study. Appl. Catal. B Environ. 2011, 110, 71–80. [Google Scholar] [CrossRef]
  26. Liu, Z.; Chen, C.; Zhao, J.; Yang, L.; Sun, K.; Zeng, L.; Pan, Y.; Liu, Y.; Liu, C. Study on the NO2 Production Pathways and the Role of NO2 in Fast Selective Catalytic Reduction DeNOx at Low-Temperature over MnOx/TiO2 Catalyst. Chem. Eng. J. 2020, 379, 122288. [Google Scholar] [CrossRef]
  27. Liu, Z.; Wang, M.; Liu, S.; Chen, Z.; Yang, L.; Sun, K.; Chen, Y.; Zeng, L.; Wang, W.; Zhao, J.; et al. Design of Assembled Composite of Mn3O4@Graphitic Carbon Porous Nano-Dandelions: A Catalyst for Low–Temperature Selective Catalytic Reduction of NOx with Remarkable SO2 Resistance. Appl. Catal. B Environ. 2020, 269, 118731. [Google Scholar] [CrossRef]
  28. Liu, Z.; Sun, G.; Chen, C.; Sun, K.; Zeng, L.; Yang, L.; Chen, Y.; Wang, W.; Liu, B.; Lu, Y.; et al. Fe-Doped Mn3O4 Spinel Nanoparticles with Highly Exposed Feoct –O–Mntet Sites for Efficient Selective Catalytic Reduction (SCR) of NO with Ammonia at Low Temperatures. ACS Catal. 2020, 10, 6803–6809. [Google Scholar] [CrossRef]
  29. Tang, X.; Li, J.; Sun, L.; Hao, J. Origination of N2O from NO Reduction by NH3 over β-MnO2 and α-Mn2O3. Appl. Catal. B Environ. 2010, 99, 156–162. [Google Scholar] [CrossRef]
  30. Qi, G.; Yang, R.T.; Chang, R. MnOx-CeO2 Mixed Oxides Prepared by Co-Precipitation for Selective Catalytic Reduction of NO with NH3 at Low Temperatures. Appl. Catal. B Environ. 2004, 51, 93–106. [Google Scholar] [CrossRef]
  31. Meng, D.; Xu, Q.; Jiao, Y.; Guo, Y.; Guo, Y.; Wang, L.; Lu, G.; Zhan, W. Spinel Structured CoaMnbOx Mixed Oxide Catalyst for the Selective Catalytic Reduction of NOx with NH3. Appl. Catal. B Environ. 2018, 221, 652–663. [Google Scholar] [CrossRef]
  32. Wan, Y.; Zhao, W.; Tang, Y.; Li, L.; Wang, H.; Cui, Y.; Gu, J.; Li, Y.; Shi, J. Ni-Mn Bi-Metal Oxide Catalysts for the Low Temperature SCR Removal of NO with NH3. Appl. Catal. B Environ. 2014, 148–149, 114–122. [Google Scholar] [CrossRef]
  33. Fan, Z.; Shi, J.-W.; Gao, C.; Gao, G.; Wang, B.; Wang, Y.; He, C.; Niu, C. Gd-Modified MnOx for the Selective Catalytic Reduction of NO by NH3: The Promoting Effect of Gd on the Catalytic Performance and Sulfur Resistance. Chem. Eng. J. 2018, 348, 820–830. [Google Scholar] [CrossRef]
  34. Zhang, L.; Cui, S.; Guo, H.; Ma, X.; Luo, X. The Influence of K+ Cation on the MnOx-CeO2/TiO2 Catalysts for Selective Catalytic Reduction of NOx with NH3 at Low Temperature. J. Mol. Catal. Chem. 2014, 390, 14–21. [Google Scholar] [CrossRef]
  35. Liu, J.; Guo, R.; Li, M.; Sun, P.; Liu, S.; Pan, W.; Liu, S.; Sun, X. Enhancement of the SO2 Resistance of Mn/TiO2 SCR Catalyst by Eu Modification: A Mechanism Study. Fuel 2018, 223, 385–393. [Google Scholar] [CrossRef]
  36. Ye, B.; Lee, M.; Jeong, B.; Kim, J.; Lee, D.H.; Baik, J.M.; Kim, H.-D. Partially Reduced Graphene Oxide as a Support of Mn-Ce/TiO2 Catalyst for Selective Catalytic Reduction of NOx with NH3. Catal. Today 2019, 328, 300–306. [Google Scholar] [CrossRef]
  37. Li, J.; Chang, H.; Ma, L.; Hao, J.; Yang, R.T. Low-Temperature Selective Catalytic Reduction of NOx with NH3 over Metal Oxide and Zeolite Catalysts—A Review. Catal. Today 2011, 175, 147–156. [Google Scholar] [CrossRef]
  38. Fu, M.; Li, C.; Lu, P.; Qu, L.; Zhang, M.; Zhou, Y.; Yu, M.; Fang, Y. A Review on Selective Catalytic Reduction of NOx by Supported Catalysts at 100–300 °C—Catalysts, Mechanism, Kinetics. Catal. Sci. Technol. 2014, 4, 14–25. [Google Scholar] [CrossRef]
  39. Chen, C.; Wang, Y.; Li, J.; Tian, F.; Chen, W.; Feng, C.; Pan, Y.; Liu, Y. In Situ Construction of Heteroatom F-Doped Mn3O4 Spinel Catalysts with Robust Activity and SO2 Resistance for NH3-SCR at Low Temperature. Appl. Catal. B Environ. 2023, 338, 123086. [Google Scholar] [CrossRef]
  40. Chen, C.; Feng, C.; Wang, Y.; Li, J.; Liu, Z.; Wang, W.; Pan, Y.; Liu, Y. Design of Robust Co-Doped Mn3O4 Spinel Catalysts for Selective Catalytic Reduction of NO with NH3 at Low Temperatures. Appl. Surf. Sci. 2022, 602, 154384. [Google Scholar] [CrossRef]
  41. Yang, M.; Yuan, H.; Wang, H.; Hu, P. Insights into the Selective Catalytic Reduction of NO by NH3 over Mn3O4(110): A DFT Study Coupled with Microkinetic Analysis. Sci. China Chem. 2018, 61, 457–467. [Google Scholar] [CrossRef]
  42. Xiong, S.; Peng, Y.; Wang, D.; Huang, N.; Zhang, Q.; Yang, S.; Chen, J.; Li, J. The Role of the Cu Dopant on a Mn3O4 Spinel SCR Catalyst: Improvement of Low-Temperature Activity and Sulfur Resistance. Chem. Eng. J. 2020, 387, 124090. [Google Scholar] [CrossRef]
  43. Kapteijn, F.; Singoredjo, L.; Andreini, A.; Moulijn, J.A. Activity and Selectivity of Pure Manganese Oxides in the Selective Catalytic Reduction of Nitric Oxide with Ammonia. Appl. Catal. B Environ. 1994, 3, 173–189. [Google Scholar] [CrossRef]
  44. Kang, M.; Park, E.D.; Kim, J.M.; Yie, J.E. Manganese Oxide Catalysts for NOx Reduction with NH3 at Low Temperatures. Appl. Catal. Gen. 2007, 327, 261–269. [Google Scholar] [CrossRef]
  45. Xie, J.; Fang, D.; He, F.; Chen, J.; Fu, Z.; Chen, X. Performance and Mechanism about MnOx Species Included in MnOx/TiO2 Catalysts for SCR at Low Temperature. Catal. Commun. 2012, 28, 77–81. [Google Scholar] [CrossRef]
  46. Busca, G.; Lietti, L.; Ramis, G.; Berti, F. Chemical and Mechanistic Aspects of the Selective Catalytic Reduction of NOx by Ammonia over Oxide Catalysts: A Review. Appl. Catal. B Environ. 1998, 18, 1–36. [Google Scholar] [CrossRef]
  47. Forzatti, P. Present Status and Perspectives in De-NOx SCR Catalysis. Appl. Catal. Gen. 2001, 222, 221–236. [Google Scholar] [CrossRef]
  48. Marbán, G.; Fuertes, A.B. Kinetics of the Low-Temperature Selective Catalytic Reduction of NO with NH3 over Activated Carbon ®ber Composite-Supported Iron Oxides. Catal. Lett. 2002, 84, 13–19. [Google Scholar] [CrossRef]
  49. Marbán, G.; Valdés-Solís, T.; Fuertes, A.B. Mechanism of Low Temperature Selective Catalytic Reduction of NO with NH3 over Carbon-Supported Mn3O4: Active Phase and Role of Surface NO Species. Phys Chem Chem Phys 2004, 6, 453–464. [Google Scholar] [CrossRef]
  50. Marban, G. Mechanism of Low-Temperature Selective Catalytic Reduction of NO with BNH3 over Carbon-Supported Mn3O4Role of Surface NH3 Species: SCR Mechanism. J. Catal. 2004, 226, 138–155. [Google Scholar] [CrossRef]
  51. Wu, Z.; Jiang, B.; Liu, Y.; Wang, H.; Jin, R. DRIFT Study of Manganese/Titania-Based Catalysts for Low-Temperature Selective Catalytic Reduction of NO with NH3. Environ. Sci. Technol. 2007, 41, 5812–5817. [Google Scholar] [CrossRef] [PubMed]
  52. Hu, H.; Cai, S.; Li, H.; Huang, L.; Shi, L.; Zhang, D. Mechanistic Aspects of deNO x Processing over TiO2 Supported Co–Mn Oxide Catalysts: Structure–Activity Relationships and In Situ DRIFTs Analysis. ACS Catal. 2015, 5, 6069–6077. [Google Scholar] [CrossRef]
  53. Hu, H.; Cai, S.; Li, H.; Huang, L.; Shi, L.; Zhang, D. In Situ DRIFTs Investigation of the Low-Temperature Reaction Mechanism over Mn-Doped Co3O4 for the Selective Catalytic Reduction of NOx with NH3. J. Phys. Chem. C 2015, 119, 22924–22933. [Google Scholar] [CrossRef]
  54. Suárez, S.; Martín, J.A.; Yates, M.; Avila, P.; Blanco, J. N2O Formation in the Selective Catalytic Reduction of NOx with NH3 at Low Temperature on CuO-Supported Monolithic Catalysts. J. Catal. 2005, 229, 227–236. [Google Scholar] [CrossRef]
  55. Ilchenko, N. Catalytic Oxidation of Ammonia, I. Reaction Kinetics and Mechanism. J. Catal. 1975, 39, 57–72. [Google Scholar] [CrossRef]
  56. Singoredjo, L.; Korver, R.; Kapteijn, F.; Moulijn, J. Alumina Supported Manganese Oxides for the Low-Temperature Selective Catalytic Reduction of Nitric Oxide with Ammonia. Appl. Catal. B Environ. 1992, 1, 297–316. [Google Scholar] [CrossRef]
  57. Madia, G.; Koebel, M.; Elsener, M.; Wokaun, A. Side Reactions in the Selective Catalytic Reduction of NOx with Various NO2 Fractions. Ind. Eng. Chem. Res. 2002, 41, 4008–4015. [Google Scholar] [CrossRef]
  58. Ciardelli, C.; Nova, I.; Tronconi, E.; Chatterjee, D.; Bandl-Konrad, B.; Weibel, M.; Krutzsch, B. Reactivity of NO/NO2–NH3 SCR System for Diesel Exhaust Aftertreatment: Identification of the Reaction Network as a Function of Temperature and NO2 Feed Content. Appl. Catal. B Environ. 2007, 70, 80–90. [Google Scholar] [CrossRef]
  59. He, Z.; Wang, Y.; Liu, Y.; Lian, L.; Kong, D.; Zhao, Y. Recent Advances in Sulfur Poisoning of Selective Catalytic Reduction (SCR) Denitration Catalysts. Fuel 2024, 365, 131126. [Google Scholar] [CrossRef]
  60. Yang, G.; Zhao, H.; Luo, X.; Shi, K.; Zhao, H.; Wang, W.; Chen, Q.; Fan, H.; Wu, T. Promotion Effect and Mechanism of the Addition of Mo on the Enhanced Low Temperature SCR of NOx by NH3 over MnOx/γ-Al2O3 Catalysts. Appl. Catal. B Environ. 2019, 245, 743–752. [Google Scholar] [CrossRef]
  61. Gao, F.; Tang, X.; Yi, H.; Li, J.; Zhao, S.; Wang, J.; Chu, C.; Li, C. Promotional Mechanisms of Activity and SO2 Tolerance of Co- or Ni-Doped MnOx-CeO2 Catalysts for SCR of NOx with NH3 at Low Temperature. Chem. Eng. J. 2017, 317, 20–31. [Google Scholar] [CrossRef]
  62. Yan, Q.; Chen, S.; Zhang, C.; Wang, Q.; Louis, B. Synthesis and Catalytic Performance of Cu1Mn0.5Ti0.5O Mixed Oxide as Low-Temperature NH3-SCR Catalyst with Enhanced SO2 Resistance. Appl. Catal. B Environ. 2018, 238, 236–247. [Google Scholar] [CrossRef]
  63. Deng, S.; Meng, T.; Xu, B.; Gao, F.; Ding, Y.; Yu, L.; Fan, Y. Advanced MnOx /TiO2 Catalyst with Preferentially Exposed Anatase {001} Facet for Low-Temperature SCR of NO. ACS Catal. 2016, 6, 5807–5815. [Google Scholar] [CrossRef]
  64. Jin, R.; Liu, Y.; Wang, Y.; Cen, W.; Wu, Z.; Wang, H.; Weng, X. The Role of Cerium in the Improved SO2 Tolerance for NO Reduction with NH3 over Mn-Ce/TiO2 Catalyst at Low Temperature. Appl. Catal. B Environ. 2014, 148–149, 582–588. [Google Scholar] [CrossRef]
  65. Yu, J.; Guo, F.; Wang, Y.; Zhu, J.; Liu, Y.; Su, F.; Gao, S.; Xu, G. Sulfur Poisoning Resistant Mesoporous Mn-Base Catalyst for Low-Temperature SCR of NO with NH3. Appl. Catal. B Environ. 2010, 95, 160–168. [Google Scholar] [CrossRef]
  66. Wei, L.; Cui, S.; Guo, H.; Ma, X.; Zhang, L. DRIFT and DFT Study of Cerium Addition on SO2 of Manganese-Based Catalysts for Low Temperature SCR. J. Mol. Catal. Chem. 2016, 421, 102–108. [Google Scholar] [CrossRef]
  67. Jarosch, D. Crystal Structure Refinement and Reflectance Measurements of Hausmannite, Mn3O4. Mineral. Petrol. 1987, 37, 15–23. [Google Scholar] [CrossRef]
  68. Vegge, T.; Rasmussen, T.; Leffers, T.; Pedersen, O.B.; Jacobsen, K.W. Atomistic Simulations of Cross-Slip of Jogged Screw Dislocations in Copper. Philos. Mag. Lett. 2001, 81, 137–144. [Google Scholar] [CrossRef]
  69. Howalt, J.G.; Bligaard, T.; Rossmeisl, J.; Vegge, T. DFT Based Study of Transition Metal Nano-Clusters for Electrochemical NH3 Production. Phys. Chem. Chem. Phys. 2013, 15, 7785. [Google Scholar] [CrossRef]
  70. Yang, Y.; Liu, J.; Liu, F.; Wang, Z.; Ding, J.; Huang, H. Reaction Mechanism for NH3-SCR of NOx over CuMn2O4 Catalyst. Chem. Eng. J. 2019, 361, 578–587. [Google Scholar] [CrossRef]
  71. Zhang, L.; Cui, S.; Guo, H.; Ma, X.; Lu, W. Density Function Theoretical and Experimental Study of NH3+NO Adsorptions on MnO/TiO2 Surface. Comput. Mater. Sci. 2016, 112, 238–244. [Google Scholar] [CrossRef]
  72. Liu, B.; Liu, J.; Xin, L.; Zhang, T.; Xu, Y.; Jiang, F.; Liu, X. Unraveling Reactivity Descriptors and Structure Sensitivity in Low-Temperature NH3-SCR Reaction over CeTiOxCatalysts: A Combined Computational and Experimental Study. ACS Catal. 2021, 11, 7613–7636. [Google Scholar] [CrossRef]
  73. Alemany, L.J. Characterization and Composition of Commercial V2O5-WO3-TiO2 SCR Catalysts. Appl. Catal. B Environ. 1996, 10, 299–311. [Google Scholar] [CrossRef]
  74. Ramis, G.; Busca, G.; Bregani, F.; Forzatti*, P. Fourier Transform-Infrared Study of the Adsorption and Coadsorption of Nitric Oxide, Nitrogen Dioxide and Ammonia on Vanadia-Titania and Mechanism of Selective Catalytic Reduction. Appl. Catal. 1990, 64, 259–278. [Google Scholar] [CrossRef]
  75. Lietti, L.; Svachula, J.; Forzatti, P.; Busca, G.; Ramis, G.; Bregani, P. Surface and Catalytic Properties of Vanadia-Titania and Tungsta-Titania Systems in the Selective Catalytic Reduction of Nitrogen Oxides. Catal. Today 1993, 17, 131–139. [Google Scholar] [CrossRef]
  76. Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for Ab Initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  77. Kresse, G.; Furthmüller, J. Efficiency of Ab-Initio Total Energy Calculations for Metals and Semiconductors Using a Plane-Wave Basis Set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  78. Kresse, G.; Joubert, D. From Ultrasoft Pseudopotentials to the Projector Augmented-Wave Method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  79. Perdew, J.P.; Burke, K.; Wang, Y. Generalized Gradient Approximation for the Exchange-Correlation Hole of a Many-Electron System. Phys. Rev. B 1996, 54, 16533–16539. [Google Scholar] [CrossRef]
  80. Monkhorst, H.J.; Pack, J.D. Special Points for Brillouin-Zone Integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  81. Garcês Gonçalves, P.R., Jr.; De Abreu, H.A.; Duarte, H.A. Stability, Structural, and Electronic Properties of Hausmannite (Mn3O4) Surfaces and Their Interaction with Water. J. Phys. Chem. C 2018, 122, 20841–20849. [Google Scholar] [CrossRef]
  82. Liu, S.; Liu, L.; Cheng, Z.; Zhu, J.; Yu, R. Surface Structures of Mn3 O4 and the Partition of Oxidation States of Mn. J. Phys. Chem. Lett. 2021, 12, 5675–5681. [Google Scholar] [CrossRef] [PubMed]
  83. Zhao, S.; Liu, Y.; Wu, Y.; Zhang, H.; Zhang, D.; Zhang, Y.; Deng, H.; Sun, G.; Li, H.; Shan, W.; et al. Elucidating Atomic Structure and Reconstruction of Mn3O4(0 0 1) Surface. Appl. Surf. Sci. 2025, 680, 161339. [Google Scholar] [CrossRef]
  84. Johnson, R. NIST 101. Computational Chemistry Comparison and Benchmark Database. 1999. Available online: http://cccbdb.nist.gov (accessed on 22 January 2025).
  85. Wu, G.; Zelenay, P. Nanostructured Nonprecious Metal Catalysts for Oxygen Reduction Reaction. Acc. Chem. Res. 2013, 46, 1878–1889. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The most stable adsorption structures of NH3 and NO on Mn3O4(001). (a) NH3 binding at the atop-Mn site, (b) NO binding at the atop Mn site via the O atom, (c) NO binding at the bridge-Mn site, and (d) NO binding at the atop Mn site via the N atom. Red, purple, blue, and white balls represent the O, Mn, N, and H atoms, respectively.
Figure 1. The most stable adsorption structures of NH3 and NO on Mn3O4(001). (a) NH3 binding at the atop-Mn site, (b) NO binding at the atop Mn site via the O atom, (c) NO binding at the bridge-Mn site, and (d) NO binding at the atop Mn site via the N atom. Red, purple, blue, and white balls represent the O, Mn, N, and H atoms, respectively.
Catalysts 15 00241 g001
Figure 2. Gibbs free energy diagram for the NH3-SCR process on Mn3O4(001). The red path denotes the “E-R” mechanism, and the blue path is the potential “L-H” mechanism proposed in this work.
Figure 2. Gibbs free energy diagram for the NH3-SCR process on Mn3O4(001). The red path denotes the “E-R” mechanism, and the blue path is the potential “L-H” mechanism proposed in this work.
Catalysts 15 00241 g002
Figure 3. Optimized structures of intermediates for the NH3-SCR process on Mn3O4(001). (ag) represent the intermediates involved in the “E-R” path, which are (NH2*+H*), (NONH2*+H*), (NH2NO*+H*), (trans-NONH*+2H*), (ONN*+3H*), (ONN*+H2OL*+H*), and (OV+ONN*+H*). (hj) denote the intermediates involved in the side pathways, which are (NH*+2H*), (N*+3H*), and (3H*), respectively. (kt) represent the intermediates involved in the “L-H” pathway, which are (NO*+NH3*), (NO*+NH2*+H*), (NO*+NH*+2H*), (NONH*+2H*), (cis-NHNO*+2H*), (iso-NHNO*+2H*), (iso-NHNO*+H2OL*), (HNN*+H2O*), (HNN*), and (H*), respectively. Red, purple, blue, and white balls represent the O, Mn, N, and H atoms, respectively.
Figure 3. Optimized structures of intermediates for the NH3-SCR process on Mn3O4(001). (ag) represent the intermediates involved in the “E-R” path, which are (NH2*+H*), (NONH2*+H*), (NH2NO*+H*), (trans-NONH*+2H*), (ONN*+3H*), (ONN*+H2OL*+H*), and (OV+ONN*+H*). (hj) denote the intermediates involved in the side pathways, which are (NH*+2H*), (N*+3H*), and (3H*), respectively. (kt) represent the intermediates involved in the “L-H” pathway, which are (NO*+NH3*), (NO*+NH2*+H*), (NO*+NH*+2H*), (NONH*+2H*), (cis-NHNO*+2H*), (iso-NHNO*+2H*), (iso-NHNO*+H2OL*), (HNN*+H2O*), (HNN*), and (H*), respectively. Red, purple, blue, and white balls represent the O, Mn, N, and H atoms, respectively.
Catalysts 15 00241 g003
Figure 4. The most stable adsorption structures of SO2 and SO3 on Mn3O4(001). (a) SO2 binding at the atop-Mn site via the S atom, (b) SO2 binding at the bridge Mn site via the two O atoms, (c) SO2 binding at the hollow site, (d) SO3 binding at the bridge Mn site via the two O atoms, and (e) SO3 binding at the hollow site. Red, purple, blue and yellow balls represent the O, Mn, N, and S atoms, respectively.
Figure 4. The most stable adsorption structures of SO2 and SO3 on Mn3O4(001). (a) SO2 binding at the atop-Mn site via the S atom, (b) SO2 binding at the bridge Mn site via the two O atoms, (c) SO2 binding at the hollow site, (d) SO3 binding at the bridge Mn site via the two O atoms, and (e) SO3 binding at the hollow site. Red, purple, blue and yellow balls represent the O, Mn, N, and S atoms, respectively.
Catalysts 15 00241 g004
Figure 5. (ac) The respective adsorption configurations of NO, SO2 and SO3 on the substrate of (NH2*+H*). (df) The respective adsorption configurations of NO, SO2 and SO3 on the substrate of (NH*+2H*). Red, purple, blue, yellow, and white balls represent the O, Mn, N, S and H atoms, respectively.
Figure 5. (ac) The respective adsorption configurations of NO, SO2 and SO3 on the substrate of (NH2*+H*). (df) The respective adsorption configurations of NO, SO2 and SO3 on the substrate of (NH*+2H*). Red, purple, blue, yellow, and white balls represent the O, Mn, N, S and H atoms, respectively.
Catalysts 15 00241 g005
Figure 6. Free energy profile depicting the reaction pathway for the formation of NH4HSO4, accompanied by the adsorption configurations of reaction intermediates. Red, purple, blue, yellow, and white balls represent the O, Mn, N, S and H atoms, respectively.
Figure 6. Free energy profile depicting the reaction pathway for the formation of NH4HSO4, accompanied by the adsorption configurations of reaction intermediates. Red, purple, blue, yellow, and white balls represent the O, Mn, N, S and H atoms, respectively.
Catalysts 15 00241 g006
Figure 7. The 2 × 2 Mn3O4(001) slab model with two different side views from [010] (a) and [100] (b) directions. Red and purple balls represent O and Mn atoms, respectively.
Figure 7. The 2 × 2 Mn3O4(001) slab model with two different side views from [010] (a) and [100] (b) directions. Red and purple balls represent O and Mn atoms, respectively.
Catalysts 15 00241 g007
Table 1. Adsorption sites, adsorption energies (ΔEads, eV), and bond lengths (d, Å) for adsorbed reactant molecules on Mn3O4(001).
Table 1. Adsorption sites, adsorption energies (ΔEads, eV), and bond lengths (d, Å) for adsorbed reactant molecules on Mn3O4(001).
MoleculeSiteBond LengthΔEads
NH3atop-MndN-Mn = 2.088−0.971
NOatop-Mn
(O-Mn) a
dO-Mn = 2.115−0.375
bridge-MndN-Mn = 1.758
dO-Mn = 2.241
−1.431
atop-Mn
(N-Mn) a
dN-Mn = 1.671−1.864
SO2atop-MndS-Mn = 2.763−0.284
bridge-MndO-Mn = 1.955, 1.949−0.833
hollowdS-O = 1.662
dO-Mn = 1.983, 2.021
−1.451
SO3bridge-MndO-Mn = 1.842, 1.839−1.424
hollowdS-O =1.597
dO-Mn = 2.005, 2.027
−2.301
a “O-Mn” denotes that NO binds to the Mn atom via the O atom, while “N-Mn” refers to the N-Mn bonding.
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

Zhu, H.; Liu, Z.; Zhang, X.; Fan, Y.; Wang, X.; Liu, D.; Li, X.; Gong, X.; Guo, W.; Ren, H. Selective Catalytic Reduction of NO by NH3 and SOx Poisoning Mechanisms on Mn3O4 Catalysts: A Density Functional Investigation. Catalysts 2025, 15, 241. https://doi.org/10.3390/catal15030241

AMA Style

Zhu H, Liu Z, Zhang X, Fan Y, Wang X, Liu D, Li X, Gong X, Guo W, Ren H. Selective Catalytic Reduction of NO by NH3 and SOx Poisoning Mechanisms on Mn3O4 Catalysts: A Density Functional Investigation. Catalysts. 2025; 15(3):241. https://doi.org/10.3390/catal15030241

Chicago/Turabian Style

Zhu, Houyu, Zhennan Liu, Xiaoxin Zhang, Yucheng Fan, Xin Wang, Dongyuan Liu, Xiaohan Li, Xiaoxiao Gong, Wenyue Guo, and Hao Ren. 2025. "Selective Catalytic Reduction of NO by NH3 and SOx Poisoning Mechanisms on Mn3O4 Catalysts: A Density Functional Investigation" Catalysts 15, no. 3: 241. https://doi.org/10.3390/catal15030241

APA Style

Zhu, H., Liu, Z., Zhang, X., Fan, Y., Wang, X., Liu, D., Li, X., Gong, X., Guo, W., & Ren, H. (2025). Selective Catalytic Reduction of NO by NH3 and SOx Poisoning Mechanisms on Mn3O4 Catalysts: A Density Functional Investigation. Catalysts, 15(3), 241. https://doi.org/10.3390/catal15030241

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