Next Article in Journal
A Review on Thermal Conversion of Plant Oil (Edible and Inedible) into Green Fuel Using Carbon-Based Nanocatalyst
Next Article in Special Issue
Catalytic Properties of Double Substituted Lanthanum Cobaltite Nanostructured Coatings Prepared by Reactive Magnetron Sputtering
Previous Article in Journal
Versatile Sulfathiazole-Functionalized Magnetic Nanoparticles as Catalyst in Oxidation and Alkylation Reactions
Previous Article in Special Issue
H2O and/or SO2 Tolerance of Cu-Mn/SAPO-34 Catalyst for NO Reduction with NH3 at Low Temperature
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of Low Temperature NH3-SCR for Removal of NOx

by
Devaiah Damma
1,
Padmanabha R. Ettireddy
2,
Benjaram M. Reddy
3 and
Panagiotis G. Smirniotis
1,*
1
Chemical Engineering, College of Engineering and Applied Science, University of Cincinnati, Cincinnati, OH 45221-0012, USA
2
Cummins Inc., Columbus, IN 47201, USA
3
Catalysis and Fine Chemicals Department, CSIR-Indian Institute of Chemical Technology (IICT), Uppal Road, Hyderabad, Telangana 500007, India
*
Author to whom correspondence should be addressed.
Catalysts 2019, 9(4), 349; https://doi.org/10.3390/catal9040349
Submission received: 6 March 2019 / Revised: 3 April 2019 / Accepted: 4 April 2019 / Published: 10 April 2019
(This article belongs to the Special Issue Emissions Control Catalysis)

Abstract

:
The importance of the low-temperature selective catalytic reduction (LT-SCR) of NOx by NH3 is increasing due to the recent severe pollution regulations being imposed around the world. Supported and mixed transition metal oxides have been widely investigated for LT-SCR technology. However, these catalytic materials have some drawbacks, especially in terms of catalyst poisoning by H2O or/and SO2. Hence, the development of catalysts for the LT-SCR process is still under active investigation throughout seeking better performance. Extensive research efforts have been made to develop new advanced materials for this technology. This article critically reviews the recent research progress on supported transition and mixed transition metal oxide catalysts for the LT-SCR reaction. The review covered the description of the influence of operating conditions and promoters on the LT-SCR performance. The reaction mechanism, reaction intermediates, and active sites are also discussed in detail using isotopic labelling and in situ FT-IR studies.

1. Introduction

The non-renewable fossil fuels are continuing to remain the dominant energy source in power plants and automobiles to satisfy the ever-growing energy demands. However, the combustion of fossil fuels mainly generates nitrogen oxide (NOx) pollutants (NO, NO2, and N2O and their derivatives) which can cause acid rain, photochemical smog, ozone depletion, and eutrophication problems [1,2,3,4]. Due to the negative impacts of NOx, the mitigation of NOx emissions is of paramount importance for environmental protection. Several technologies are available to reduce NOx emissions by using catalytic materials and among them, selective catalytic reduction of NOx with NH3 (NH3-SCR) has been widely applied due to its high NOx removal efficiency [5,6,7]. Usually, the flue gas temperature of the industrial process is as low as 300 °C and, thus, the SCR catalyst must be active in the low-temperature regime (100–300 °C). V2O5–WO3(MoO3)/TiO2 is the typical and efficient catalyst and has been commercialized for NH3-SCR technology for medium temperature process [8,9]. However, this catalyst has some intrinsic drawbacks such as narrow and high working temperature window (350–400 °C), and low N2 selectivity in the high-temperature range [3,10,11]. Therefore, many researchers continue to develop highly active catalysts for low-temperature NH3-SCR in a wide temperature window.
With this perspective, several transition metal oxide-based catalysts have been extensively investigated for low-temperature NH3-SCR reaction due to their excellent redox properties, low price, and high thermodynamic stability. Especially, the easy gain and loss of electrons in the d shell of the transition metal ions could be responsible for the facile redox properties [12,13,14]. For example, the Cr/TiO2 [15], Cr-MnOx [8], Fe-MnOx [16], Mn/TiO2 [17,18], FexTiOy [19], MnOx/CeO2 [20], and Cu/TiO2 [21] catalysts were shown to exhibit good SCR activity in the low temperature range. In our earlier work, we investigated the low-temperature NH3-SCR in the presence of excess O2 on the TiO2 supported V, Cr, Mn, Fe, Co, Ni, and Cu oxides and found the catalytic performance decreased in the following order of Mn > Cu ≥ Cr ≫ Co > Fe ≫ V ≫ Ni [16]. Particularly, manganese-containing catalysts have attracted much attention due to its variable valence states and excellent redox ability [2,5]. In the recent past, we published a series of papers on Mn-based SCR catalysts that showed a highly promising deNOx potential in the low-temperature region [5,22,23,24,25,26]. However, these catalysts are very sensitive to the presence of SO2 in the feed and exhibit lower N2 selectivity [8,27,28,29]. Hence, the development of catalysts that show both good low-temperature activity and high SO2/H2O durability is of great importance for the NH3-SCR reaction. In general, there are two plausible strategies available to enhance low-temperature NH3-SCR performance. One strategy is to modify the transition metal oxide with one or multiple metal oxides, which could enhance the active sites for the reaction by inducing the synergistic effect [30,31,32,33]. The other approach is to synthesize the supported materials to disperse the transition metal-based oxides which can increase the activity by metal-support interactions [26,34,35,36,37]. Recently, many supported and mixed transition metal catalyst formulations have been studied to improve the low-temperature SCR performance, as well as resistance to SO2/H2O.
In this study, we systematically reviewed the recent advancements in developing the transition metal-based catalysts for low-temperature NH3-SCR reaction. This review also demonstrated the action of different promoters and supports on the catalytic performance and SO2/H2O tolerance of the transition metal-based catalysts in NH3-SCR of NOx. The reported catalysts were divided into four categories, such as binary, ternary/multi, supported single, and supported binary/multi-transition metal-based catalysts.

2. Binary Transition Metal-Based Catalysts

Various transition-metal oxides have been proved to be active for the NH3-SCR at low-temperature. However, the catalytic performance on single transition metal oxide is far from satisfactory due to their low specific surface area and thermal instability [38,39,40,41]. The addition of dopants is a common method to improve the drawbacks associated with pure transition metal oxide. Hence, much progress has done to improve the low-temperature SCR activity of transition metal oxides by mixing or doping with other metal oxides. In recent years, Mn, Fe, Co, Ni, and Cu-based binary oxide catalysts have been extensively studied for low-temperature NH3-SCR reaction due to their attractive catalytic performance [19,33,41,42,43,44,45,46,47,48,49]. Particularly, Mn-based binary oxides are popular and proven to be effective catalysts for low-temperature NH3-SCR reaction [42,50,51]. Recently, Xin et al. [52] designed bifunctional Va-MnOx (where a represents the molar ratios of V / (V + Mn)) catalysts composed of Mn2O3 and Mn2V2O7 phases that significantly improved both NOx conversion and N2 selectivity in comparison with Mn2O3 at low-temperature (Figure 1). Although Mn2V2O7 showed an excellent N2 selectivity, the NOx conversion is much lower on it. Especially, above 90% NOx conversion and 80% N2 selectivity was observed in the temperature region of 120–240 °C over the V0.05-MnOx catalyst. The V0.05-MnOx catalyst also found to be exhibit higher NOx conversion to N2 as compared to the mechanically mixed Mn2O3 + Mn2V2O7 sample which has the same component content to V0.05-MnOx (Figure 1). This finding indicated that the synergism between Mn2O3 and Mn2V2O7 exists in the chemically prepared V0.05-MnOx rather than the mechanically mixed Mn2O3 + Mn2V2O7 sample. Moreover, the mechanically mixed Mn2O3 + Mn2V2O7 sample showed higher activity in comparison to mechanically mixed MoO3 + Mn2V2O7 sample, suggesting that the presence of Mn2O3 phase in the catalyst is necessary for NH3-SCR reaction. In conjunction with in situ IR characterization and DFT (density functional theory) calculation results, the authors concluded that the Mn2O3 phase of the catalyst could activate NH3 into NH2 intermediate, which then transferred to the Mn2V2O7 phase of the catalyst and reacted with gaseous NO into NH2NO. Finally, the generated NH2NO intermediate on the Mn2V2O7 phase exclusively decomposed to the N2 rather than the undesired byproduct, N2O, which is formed due to the deep oxidation of adsorbed NH3 on Mn2O3.
Han and co-workers [53] fabricated triple-shelled NiMn2O4 hollow spheres (Figure 2a,b) by using a solvothermal method and tested their ability for low-temperature NH3-SCR reaction. As shown in Figure 2c, the prepared NiMn2O4 hollow spheres (NiMn2O4-S) showed the best catalytic activity with NOx conversion of above 90% over a wide temperature range from 100 °C to 225 °C as compared to the NiMn2O4 nanoparticles (NiMn2O4-P). The triple-layer shell structure of the NiMn2O4-S catalyst generates a larger surface area (165.3 m2 g‒1) that exposes more active sites (such as surface Mn4+ and surface adsorbed oxygen species), which are responsible for its superior activity. Additionally, the NiMn2O4-S catalyst displayed outstanding stability and good tolerance to H2O and SO2 (Figure 2d).
Gao et al. [54] investigated the low-temperature NH3-SCR reaction over the hydroxyl-containing Me-Mn binary oxides (Me = Co, Ni) prepared by a combined complexation–esterification method. It was found that the NOx conversion decreased in the order of Mn3O4-Co3O4-OH (Co-MnOx binary oxide) > Mn2O3-NiMnO3-OH (Ni-MnOx binary oxide) > Mn2O3-OH, while the N2 selectivity increased in the sequence of Mn3O4-Co3O4-OH < Mn2O3-OH < Mn2O3-NiMnO3-OH. Although the Co and Ni elements in the catalysts delay the poisoning of SO2 as compared to MnOx sample, the Co-MnOx and Ni-MnOx binary oxides are deactivated by SO2 over the postponement due to the formation of metal sulfate and ammonia hydrogensulfite species. In another study, Sun and co-workers [55] prepared Mn0.66M0.33Ox catalysts (M = Fe, Zn, Cu) and a series of FeαMn1−αOx (α = 1, 0.25, 0.33, 0.50, 0 mol%) catalysts and examined for NH3-SCR at low-temperatures. The results demonstrated that the Fe0.33Mn0.66Ox catalyst displayed the superior NH3-SCR activity (NOx removal efficiency > 90%) in a wide temperature range (75–225 °C) among the Cu0.33Mn0.66Ox, Zn0.33Mn0.66Ox, and FeαMn1−αOx (α = 1, 0.25, 0.50, 0 mol%) catalysts. The authors proposed that the distortion of the catalyst structure by Fe doping could play a key role in improving the NH3-SCR performance over the Fe0.33Mn0.66Ox catalyst.
Rare-earth metal oxides have been frequently adopted to modify the MnOx as an efficient low-temperature NH3-SCR catalyst due to their incomplete 4f and empty 5d orbitals [50,51,56]. Fan et al. [57] synthesized Gd-modified MnOx catalysts with Gd/Mn molar ratio of 0.05, 0.1, and 0.3 to improve the catalytic performance and sulfur resistance in the NH3-SCR reaction at low-temperature. The MnGdO-2 catalyst (the mole ratio of Gd/Mn = 0.1) found to show the optimal NO conversion and N2 selectivity among the investigated catalysts. The addition of a proper amount of Gd into MnOx could enhance the concentrations of surface Mn4+ and chemisorbed oxygen species, and increase the amount and the strength of surface acid sites, which lead to better low-temperature catalytic performance than the others. Furthermore, the MnGdO-2 catalyst had an excellent tolerance to SO2/H2O as compared to pure MnOx sample (Figure 3). Their results demonstrate that the doping of Gd could restrains the transformation of MnO2 to Mn2O3 and the generation of MnSO4, obstructs the decrease in Lewis acid sites and the increase in Brønsted acid sites, and eases the competitive adsorption between the NO and SO2 and, thus, improves the resistance to SO2.
Li et al. [58] developed hollow MnOx-CeO2 binary nanotubes as efficient low-temperature NH3-SCR catalysts via an interfacial oxidation-reduction process using KMnO4 aqueous solution and Ce(OH)CO3 nanorod as both template and reducing agent without any other intermediate. They reported that the MnOx-CeO2 hollow nanotube catalyst with 3.75 g of Ce(OH)CO3 template (denoted it as MnOx-CeO2-B) exhibited outstanding performance with more than 96% NOx conversion in the temperature range of 100–180 °C. The best activity of the MnOx-CeO2-B catalyst was due to its ample number of surface Mn4+ and O species, and hollow and porous structures that provide abundant Lewis acid sites and large surface area. Additionally, MnOx-CeO2-B catalyst showed an excellent resistance to H2O and SO2 (Figure 4) and especially, the great SO2 tolerance was ascribed to the hierarchically porous and hollow structure that inhibits the deposition of ammonium sulfate species, and the doping of ceria that acts as an SO2 trap to limit sulfation of the main active phase.
Fe-based binary catalysts have also been studied as NH3-SCR catalysts due to their high activity, excellent resistance to H2O and SO2, outstanding environmentally friendly performance, lower cost, and higher abundance [59,60,61,62]. Mu et al. [63] synthesized a series of vanadium-doped Fe2O3 catalysts and evaluated the effect of V on the low-temperature NH3-SCR activity of hematite. The NH3-SCR activity and N2 selectivity are greatly enhanced after the incorporation of vanadium into Fe2O3 and the Fe0.75V0.25Oδ catalyst with a Fe/V mole ratio of 3/1 showed the best catalytic performance over a wide temperature window and strong resistance to H2O and SO2. They found that the charge transfer from Fe to V due to the electron inductive effect between Fe and V which could enhance the redox ability and surface acidity thereby superior NH3-SCR activity at low-temperature. The in situ DRIFTS and kinetic studies suggested that the SCR reaction followed the Langmuir–Hinshelwood mechanism below 200 °C, while an Eley−Rideal mechanism dominated at and above 200 °C. Li and co-workers [64] reported novel iron titanium (CT-FeTi) catalyst, prepared by a CTAB-assisted process, showing good deNOx efficiency and H2O resistance at low-temperature as compared to the FeTi catalyst that prepared without adding CTAB. The addition of CTAB during the CT-FeTi catalyst preparation not only promotes to form the uniform mesoporous structure to avoid being excessively enlarged in the presence of H2O, but also enhances the adsorption of bridging nitrate and NH3 species on Lewis acid sites. Thus, the authors concluded that the CTAB acted as a “structural” and “chemical” promoter in improving the NH3-SCR activity and H2O resistance at low-temperature.
Recently, Co-based spinel catalysts have shown to exhibit a remarkable low-temperature NH3-SCR activity, N2 selectivity, and tolerance to SO2/H2O [30,65,66,67]. Meng et al. [48] synthesized a highly efficient CoaMnbOx (where a/b is the molar ratio of Co/Mn) mixed oxide catalysts and investigated the effects of the Co/Mn molar ratio on the low-temperature NH3-SCR reaction. The CoaMnbOx mixed oxides showed higher NH3-SCR activity than either MnOx or CoOx alone due to their improved redox properties and surface acid sites by the synergistic effects between the Co and Mn species. Particularly, the catalyst with Co/Mn molar ratio of 7:3 (Co7Mn3Ox) exhibited the greatest activity (>80% NOx conversion) in a temperature window of 116–285 °C as compared to the catalysts with Co/Mn molar ratio of 5:5 (Co5Mn5Ox) and 3:7 (Co3Mn7Ox). They considered that the high NO + O2 adsorption ability and enhanced redox properties of the Co7Mn3Ox catalyst, emerging from its MnCo2O4.5 spinel phase and higher surface area, were beneficial to augment the NH3-SCR performance by forming nitrate species on the catalyst surface. Although Co7Mn3Ox catalyst had better resistance to H2O/SO2 than the Co3Mn7Ox and MnOx, the tolerance to SO2 poisoning still need to be improved for practical use. Nevertheless, it was found that the deactivated Co7Mn3Ox, Co3Mn7Ox, and MnOx catalysts in SO2 stream can be regenerated simply by washing with water. Based on their results, the authors also proposed the NH3-SCR reaction mechanism over the Co7Mn3Ox catalyst, which is shown in Scheme 1. The reaction was initiated by adsorption and activation of gaseous oxygen on oxygen vacancies (symbol □), which was then transformed into lattice oxygen O2− (Step 1); This lattice oxygen was diffused to the catalyst surface and then it had become surface active oxygen (O*) (Step 2); Gaseous NO was adsorbed and subsequently reacted with O* to form NO2/NO3 intermediates (Step 3); Meanwhile, NH3 was activated to −NH2 and NH4+ species by Mn4+ (Step 4); Finally, NO2/NO3 intermediates reacted with the NH species to produce reaction products, N2, and H2O (Step 5); By the electron transfer from Mn3+ to Co3+ (Step 6); the catalyst was recovered to its original state (Step 7); Thus, the synergistic effect between the Co and Mn plays a key role in improving the NH3-SCR activity over Co7Mn3Ox catalyst.
Mesoporous materials have been proved as promising catalysts for NH3-SCR reaction since they can facilitate to promote effective diffusion of reactants towards the active sites [30,65,66,68]. With this perspective, Hu et al. [47] developed mesoporous 3D nanosphere-like Mn-Co-O catalysts through a template-free approach and evaluated for low-temperature NH3-SCR reaction. It was found that the synthesized Mn-Co-O samples showed excellent NH3-SCR activity in a broad working temperature window of 75 to 325 °C (NOx conversion above 80%). They ascribed this outstanding performance to the strong and abundant acid sites, the strong adsorption of NOx, robust redox properties, the formation of more oxygen vacancies and metal-metal interactions between the cobalt and manganese species.
Besides Mn, Fe, and Co oxides, CuOx has also been considered in the bimetallic catalyst formulations for low-temperature NH3-SCR reaction [69,70]. For instance, Ali et al. [71] reported the Cux-Nb1.1-x (x = 0.45, 0.35, 0.25, 0.15) bimetal oxides catalysts and found that the Cu/Nb ratio was crucial in enhancing the NH3-SCR activity. As shown in Figure 5a,b, all binary Cux-Nb1.1-x oxides exhibited significantly higher activity than the CuOx and Nb2O5, and among the Cux-Nb1.1-x samples, Cu0.25-Nb0.85 catalyst displayed the best performance in a wide temperature window of 180–330 °C (>90% NO conversion). Even at a high GHSV of 105,000 h−1, the optimal Cu0.25-Nb0.85 catalyst showed a good NO removal efficiency (above 90% NO conversion) from 210 °C to 360 °C (Figure 5c). Although the SO2/H2O streams in the feed gas have some adverse impact on Cu0.25-Nb0.85, still the catalyst showed excellent resistance to SO2/H2O with reversible deactivation (Figure 5d). The superior NH3-SCR performance and SO2/H2O tolerance of Cu0.25-Nb0.85 catalyst were attributed to its high acid amount and NO adsorption capacity.

3. Ternary and Multi-Transition Metal-Based Catalysts

The catalytic performance of single transition metal oxides can also be improved by mixing with two or multi other metal oxides. Thus, transition metal oxides are widely reported to fabricate ternary- or multi-metal-based low-temperature NH3-SCR catalysts, which could improve the catalytic activity by the enlarged synergetic interactions [14,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87]. Fang et al. [88] investigated the low-temperature NH3-SCR reaction over the Fe0.3Mn0.5Zr0.2 catalyst and found that it showed an excellent deNOx activity with 100% NO conversion in the temperature range of 200–360 °C as compared to the Fe0.5Zr0.5 and Mn0.5Zr0.5 samples. Moreover, the Fe0.3Mn0.5Zr0.2 catalyst had outstanding stability and good tolerance to SO2 (Figure 6), which they attributed to the strong interactions among Fe, Mn, and Zr species. However, the durability of the catalyst in the presence of both SO2 and H2O need to be tested to investigate its feasibility in practical use.
Guo and co-workers [89] studied the effect of Sb doping on the activity of MnTiOx catalyst for NH3-SCR reaction. The results showed that Sb modification has greatly improved the NH3-SCR performance of MnSbTiOx catalysts in comparison to the MnTiOx and SbTiOx samples. Particularly, the MnSbTiOx-0.2 (Sb/Mn molar ratio = 0.2) catalyst exhibited the best activity with above 90% NOx conversion in the temperature range of 138–367 °C as it had good adsorption and activation properties for NH3 and NOx reactants in SCR. It can be seen from Figure 7 that the addition of Sb dramatically improved the SO2 and H2O resistance of MnTiOx catalyst. Although the NOx conversion over the MnSbTiOx-0.2 catalyst slightly decreased in the presence of SO2 and H2O, it recovered to almost the original level after stopping the SO2/H2O supply.
Shi et al. [90] synthesized a series of NiyCo1-yMn2Ox microspheres (MSs) (y = 0.1, 0.3, 0.5, 0.7, 0.9) for NH3-SCR using a hydrothermal method. It was observed that the activity of all ternary MSs was greater than binary CoMn2Ox and NiMn2Ox, and Ni0.7Co0.3Mn2Ox showed the best NH3-SCR performance among the NiyCo1-yMn2Ox catalysts. Although the Ni0.7Co0.3Mn2Ox catalyst exhibited good resistance to H2O, it had poor SO2 tolerance which needs to be improved. Wu and co-workers [91] compared the DeNOx performance of MnO2/CoAl-LDO and CoMnAl-LDO mixed metal oxides prepared from CoAl-MnO2-LDH and CoMnAl-LDH templates by ion-exchange/redox reaction and hexamethylenetetramine (HMT) hydrolysis methods, respectively. The CoAl-MnO2-LDH showed higher NH3-SCR activity in a broad temperature window of 90–300 °C (Figure 8a) as well as better stability and SO2/H2O resistance (Figure 8b) than the CoMnAl-LDO, which was attributed to its larger specific surface area, stronger redox ability, more quantitative acid sites, and abundant active components.
Leng and collaborators [92] synthesized Mn0.2TiOx, Ce0.3TiOx and series of MnaCe0.3TiOx (a = 0.1, 0.2, 0.3) catalysts and investigated their applicability for low-temperature NH3-SCR reaction. They demonstrated that the low-temperature NH3-SCR activity of MnaCe0.3TiOx was greatly improved after incorporation of Mn, and the Mn0.1Ce0.3TiOx catalyst displayed the best performance (with 100% NO conversion and above 90% N2 selectivity) in the temperature range of 175–400 °C even at high GHSV of 80,000 h−1. The outstanding performance of Mn0.1Ce0.3TiOx catalyst in NH3-SCR resulted from its enhanced acidity and chemisorbed oxygen, and suitable redox property derived from Ce3+ + Mn4+ ↔ Ce4+ + Mn3+ reaction. Furthermore, the NO conversion over the Mn0.1Ce0.3TiOx decreased and stabilized at 82% after the introduction of 100 ppm SO2 and 6% H2O and restored to almost 100% NO conversion after stopping the supply of SO2 and H2O (Figure 9), suggesting that the catalyst had excellent resistance to SO2/H2O and the effects were reversible.
Ali et al. [93] developed a series of Nb-promoted Fex-Nb0.5-x-Ce0.5 (x = 0.45, 0.4, 0.35) oxides for NH3-SCR. The best activity (>90% NO conversion and near 100% N2 selectivity) in the broad temperature window of 180–400 °C as well as excellent SO2/H2O resistance (Figure 10) was observed for Fe0.4-Nb0.1-Ce0.5 catalyst. The authors considered the strong interaction among Nb, Fe, and Ce oxides leading to the enhancement of BET surface area, redox ability, acid amount, and NO adsorption capacity, which could be responsible for the outstanding performance of the catalyst.
Sun and co-workers [94] reported the multimetallic Sm- and/or Zr-doped MnOx-TiO2 catalysts for NH3-SCR reaction. As shown in Figure 11, the Sm and Zr co-doped MnOx-TiO2 (MSZTOx) catalyst had better activity (≈100% NO conversion and >95% N2 selectivity) in a wide temperature range (125–275 °C) with an excellent H2O/SO2 tolerance than the MSTOx (MnOx-SmOx-TiO2), MZTOx (MnOx-ZrOx-TiO2) and MTOx (MnOx-TiO2) catalysts. The authors claimed that the enhanced redox properties and acidic sites play a crucial role in improving the NH3-SCR performance of MSZTOx catalyst. Yan and co-workers [95] fabricated Cu0.5Mg1.5Mn0.5Al0.5Ox catalyst from layered double hydroxides and found that it showed better activity in a wide temperature range together with superior SO2 and H2O tolerance than conventional Mn/γ-Al2O3. The improved performance of Cu0.5Mg1.5Mn0.5Al0.5Ox was attributed to the high specific surface area, high reducibility of MnO2 and CuO species, an abundance of acid sites, and the good dispersion of MnO2 and CuO species.
Chen et al. [96] investigated the NH3-SCR reaction over a series of Co0.2CexMn0.8-xTi10 (x = 0, 0.05, 0.15, 0.25, 0.35, and 0.40) oxides catalysts and observed that the Co0.2Ce0.35Mn0.45Ti10 catalyst exhibited the best catalytic performance with 100% NO conversion and over 91% N2 selectivity in a broad temperature window of 180–390 °C. Although NOx conversion decreased to some extent after introducing SO2 and H2O, the Co0.2Ce0.35Mn0.45Ti10 catalyst showed excellent resistance to SO2/H2O with reversible inhibition effect (Figure 12). It was concluded that the interactions among Ce, Co, Mn, and Ti oxides led to more surface Brønsted acid and Lewis acid sites, NOx adsorption sites and modest redox ability which could play a crucial role to improve the NH3-SCR activity of Co0.2Ce0.35Mn0.45Ti10.

4. Supported Single Transition Metal-Based Catalysts

Support materials have proved to be highly beneficial for enhancing the activity and durability of catalysts as they possess high surface area and good thermal stability. In virtue of the fine dispersion of the active component on the surface of the support and the synergistic effect between active component and support, supported catalysts exhibits improved NH3-SCR performance than the unsupported transition metal oxide. Therefore, a lot of attention has been focused to increase the de-NOx efficiency by dispersing the transition metal oxides over different support materials such as TiO2, Al2O3, SiO2, carbon nanotubes (CNTs), etc. TiO2 supported transition metal oxides, especially manganese oxides, have been widely reported as promising catalysts for NH3-SCR reaction at low temperature [17,24,97,98,99,100,101,102,103,104,105]. Smirniotis and co-workers [17,106,107] first reported the transition metal oxides (V, Cr, Mn, Fe, Co, Ni, and Cu) supported on Hombikat TiO2 for NH3-SCR reaction at low-temperature. Among the investigated samples, the Mn/Hombikat TiO2 catalyst found to exhibit the highest activity even in the presence of water. They also studied the effect of different supports on the NH3-SCR performance and observed that the Mn/Hombikat TiO2 (anatase, high surface area) had the best activity as compared to the Kemira TiO2 (rutile), Degussa P25 TiO2 (anatase, rutile), Aldrich TiO2 (anatase, low surface area), Puralox γ-Al2O3, Aldrich SiO2 supported Mn catalysts. It was concluded that the Lewis acidity, redox behavior, and a high surface concentration of MnO2 could play a key role in improving the NH3-SCR activity. Later, they investigated the effect of Mn loading on the NH3-SCR performance of Mn/Hombikat TiO2 and reported that the catalyst with 16.7 wt% Mn had optimal activity and excellent tolerance to H2O during 10 days of the reaction [108]. In another work, they also proposed the NH3-SCR reaction mechanism over the Mn/TiO2 catalyst using transient isotopic labeled and in-situ FT-IR studies. As shown in Figure 13, the reaction proceeds via a Mars-van-Krevelen-like mechanism, in which NH3 and NO species were first adsorbed onto the Mn4+ sites (Lewis acid sites), followed by the formation of nitrosamide and azoxy intermediate species. Finally, these intermediates converted into N2 and H2O products [24].
In the aspect of catalyst structure design, Smirniotis and co-workers [26] developed a series of manganese confined titania nanotube (Mn/TNT-X) catalysts using different TiO2 precursors (X = Ishihara (I), Kemira (K), Degussa P25 (P25), Sigma–Aldrich (SA), Hombikat (H), TiO2 synthesized from titanium oxysulfate (TOS)) for low-temperature NH3-SCR reaction. As can be observed from the Figure 14, all Mn/TNT-X catalysts exhibited an excellent NOx conversion in broad temperature window, and especially, Mn(0.25)/TNT-H sample obtained the superior activity in the temperature range of 100–300 °C as compared to other catalysts. They believed that the better performance of Mn(0.25)/TNT-H catalyst was due to the high surface area (421 m2/g) of the TNT-H support, and high dispersion of active components. The Mn(0.25)/TNT-H catalyst also showed greater catalytic performance than the conventional Mn-loaded titania nanoparticles (Mn/TiO2), suggesting that the unique multiwall nanotube with open-ended structure could be advantageous to promote the reaction. Besides, the Mn (0.25)/TNT-H displayed outstanding tolerance to 10 vol% H2O in the feed (Figure 15), which might be attributed to the preferential existence of highly active and redox potential pairs of Mn4+ and Mn3+ in the tubular framework.
Recently, Boningari et al. [109] extended this work by comparing the NH3-SCR activity of various metal oxide confined titania nanotubes M/TNT (M = Mn, Cu, Ce, Fe, V, Cr, and Co) based on the Hombikat TiO2 support. As shown in Figure 16, the Mn-, V-, Cr-, and Cu-oxide confined titania nanotubes had excellent low-temperature activity and meanwhile, vanadium oxide confined titania nanotubes showed a broad operation temperature window for the NH3-SCR reaction.
Sheng et al. [100] synthesized core-shell MnOx/TiO2 nanorod catalyst, showed high activity, stability, and N2 selectivity in NH3-SCR. They concluded that the abundant mesopores, Lewis-acid sites, and high redox capability could be beneficial to improve catalytic performance. Although the MnOx/TiO2 catalyst exhibited excellent resistance to H2O, it was deactivated in the presence of SO2 and SO2/H2O. Jia et al. [110] reported the low-temperature NH3-SCR efficiency of MnOx/TiO2, MnOx/ZrO2, and MnOx/ZrO2-TiO2 catalysts, and found that MnOx/ZrO2-TiO2 obtained good activity at a temperature of 80–360 °C and excellent resistance to H2O at 200 °C. However, all the catalysts showed poor tolerance to SO2 and SO2/H2O that caused irreversible deactivation. Similar findings were also observed by Zhang et al. [111] over the Mn/Ti, Mn/Zr, and Mn/Ti-Zr catalysts, in which Mn/Ti-Zr sample exhibited an excellent NH3-SCR performance in a wide temperature range due to its high surface area, Lewis acid sites, and surface Mn4+ ions.
Carbon nanotubes (CNTs) have been reported as promising catalyst support for NH3-SCR catalysis due to their excellent stability and unique electronic and structural properties [36,112,113,114,115]. Qu and co-workers [116] reported that the NH3-SCR performance of Fe2O3 was dramatically enhanced when it supported on CNTs (Figure 17a). It was concluded that the large surface area, fine dispersion of Fe2O3, and interaction between Fe2O3 and CNTs were important factors to improve the NH3-SCR activity. In addition, the Fe2O3/CNTs catalyst showed an excellent tolerance to H2O/SO2. Interestingly, SO2 stream in the feed had promoting effect on the NO conversion (Figure 17b), which could be attributed to the increased acid sites for NH3 adsorption and activation on the catalyst surface in presence of SO2. Bai et al. [117] developed CNTs supported copper oxide catalysts, and found that the 10 wt% CuO/CNTs showed good NH3-SCR activity and excellent stability at 200 °C. The 10 wt% CuO/CNTs also had greater performance in comparison to 10 wt% CuO/TiO2. However, it exhibited poor resistance to SO2 and moderate tolerance to H2O.

5. Supported Binary and Multi Transition Metal-Based Catalysts

Given that the dispersion of two active components on support enhances the active sites further, researchers have been widely reported the supported binary transition metal-based oxides to improve the performance and SO2/H2O tolerance in NH3-SCR reaction. With this perspective, several composites, such as MnCe/CNTs [118], Mn-Fe/TiO2 [119], MnOx-CeO2/graphene [120], Mg-MnOx/TiO2 [121], CeOx-MnOx/TiO2-graphene [122], Fe-Mn/Al2O3 [123], Mn-Fe/W-Ti [124], MnOx-CeO2/TiO2-1%NG (NG = N-doped grapheme) [125], Mn-Ce/CeAPSO-34 [126], etc., were investigated for the NH3-SCR reaction at low-temperature. Smirniotis et al. [22] studied the promotional effect of co-doped metals (Cr, Fe, Co, Ni, Cu, Zn, Ce, and Zr) on the NH3-SCR performance of Mn/TiO2. As shown in Figure 18, except Zn and Zr, all other co-doped metals had a positive impact on the activity of Mn/TiO2, and particularly, the Mn-Ni/TiO2 exhibited the highest NO conversion and N2 selectivity among the other titania-supported bimetallic catalysts.
They also investigated the influence of Ni loading on the activity of Mn/TiO2 catalyst, and found that the 5wt%Mn-2wt%Ni/TiO2 (Mn-Ni(0.4)/TiO2, where Ni/Mn = 0.4) had the optimal activity with complete NO conversion at the temperature range of 200–250 °C (Figure 19a) and outstanding stability even in the presence of 10 vol% water (Figure 19b,c) [5,23]. The enhanced reducibility of manganese oxide and dominant phase of MnO2 claimed to be responsible for the best activity and stability of Mn–Ni/TiO2 catalyst [5,22,23]. In another study, they compared the de-NOx performance of high surface texture hydrated titania and Hombikat TiO2 supported Mn‒Ce bimetallic catalysts, and observed that the Mn–Ce/TiO2 (Hombikat) showed the better activity and excellent resistance to H2O (Figure 20). The superior performance could be attributed to the enhancement in reduction potential of active components, broadening of acid sites distribution, and the promotion of Mn4+/Mn3+, Ce3+/Ce4+ ratios including surface labile oxygen and small pore openings [25].
Xu and co-workers [127] reported Ce-Mn/TiO2 catalysts with different Ce loadings, and the Ce(20)-Mn/TiO2 found to show high activity with >90% NO conversion in the temperature range of 140–260 °C (Figure 21a). Their SO2 tolerance results showed that the resistance ability was decreased in the order of Ce(20)-Mn/TiO2 > Ce(30)-Mn/TiO2 > Ce(10)-Mn/TiO2 (Figure 21b). Although the Ce(20)-Mn/TiO2 catalyst had reasonable resistance to 100 ppm SO2 at different reaction temperatures (Figure 21c), it exhibited moderate tolerance to SO2 poisoning when added higher than 100 ppm SO2 to the reaction feed (Figure 21d). They ascribed the good SO2 resistance of Ce(20)-Mn/TiO2 to the widely distributed elements of Mn and Ce which in turn led to the inability of the sulfate material to remain on the surface.
Lin et al. [128] synthesized Me-Fe/TiO2 (SD) catalyst via an aerosol-assisted deposition method, showing an excellent NH3-SCR performance and good tolerance to SO2/H2O as compared to its counterparts prepared by co-precipitation and wet impregnation methods. The authors concluded that the enhanced surface reducibility and adsorption ability of NH3/NOx of Mn-Fe/TiO2 (SD) catalyst could be responsible for its superior activity. Lee and co-workers [129] investigated the poisoning effect of SO2 as metal sulfate and/or ammonium sulfate deposits on the low-temperature NH3-SCR activity of MnFe/TiO2 catalysts. They found that the metal sulfates had a more serious deactivation effect than that of ammonium salts on the MnFe/TiO2 catalysts. Their results showed that metal sulfates poisoning resulted in lower crystallinity, lower specific surface area, a lower ratio of Mn4+/Mn3+, higher surface acidity, and more chemisorbed oxygen, which in turn led to an adverse effect on the NH3-SCR activity of the catalyst. Mu et al. [130] prepared Fe-Mn/Ti catalyst by ethylene glycol-assisted impregnation method, showing high NH3-SCR efficiency over a broad temperature window (100−325 °C) and outstanding tolerance to sulfur poisoning. The formation of the Fe-O-Ti structure with strong interaction strengthened the electronic inductive effect and increased the ratio of surface chemisorption oxygen, thereby the enhancement of NOx adsorption capacity and NO oxidation performance, which could be beneficial to improve the NH3-SCR activity.
Liu and co-workers [131] reported that the addition of Eu had noticeably improved the NH3-SCR performance of Mn/TiO2 catalyst even after sulfation process under SCR conditions (Figure 22a). However, both the Mn/TiO2 and MnEu/TiO2 catalysts showed poor activity when they sulfated only with SO2 + O2 (Figure 22a). Further, the MnEu/TiO2 catalyst found to show better SO2 tolerance as compared to the Mn/TiO2 (Figure 22b). Their results revealed that Eu modification could inhibit the formation of surface sulfate species on the Mn/TiO2 catalyst during the NH3-SCR in the presence of SO2, which could be the reason for improved SO2 resistance.
Sun et al. [132] investigated the NH3-SCR activity over the Nb-doped Mn/TiO2 catalysts with different Nb/Mn molar ratios, and found that the MnNb/TiO2-0.12 (where Nb/Mn = 0.12) catalyst had optimal NOx conversion and N2 selectivity in the temperature range of 100–400 °C (Figure 23a,b). The optimal MnNb/TiO2-0.12 catalyst also exhibited greater SO2 resistance than Mn/TiO2 catalyst (Figure 23c). The incorporation of Nb into Mn/TiO2 catalyst led to increase surface acidity and reducibility as well as generate more surface Mn4+ and chemisorbed oxygen species along with more NO2, which results in the better NH3-SCR activity. In situ DRIFT studies over the Mn/TiO2 and MnNb/TiO2-0.12 catalysts disclosed that the NH3-SCR took place through Eley–Rideal mechanism even in presence of SO2, in which the reaction mainly occurred between adsorbed NO2 and gaseous NH3. Hence, it was concluded that the higher SO2 tolerance of the MnNb/TiO2-0.12 catalyst could be due to the existence of more adsorbed NO2 on its surface.
In another study, they reported the Mo-modified Mn/TiO2 catalysts, exhibiting improved NH3-SCR activity from 50 to 400 °C in comparison to Mn/TiO2 catalyst. Particularly, the optimal MnMo/TiO2-0.04 (where molar ratio of Mo/Mn = 0.04) catalyst better tolerance to SO2 poisoning compared with Mn/TiO2 (Figure 24) [133].
Fan and co-workers [134] fabricated ordered mesoporous titania supported CuO and MnO2 composites (CuO/MnO2-mTiO2) through a facile acetic acid-assisted one-pot synthesis approach, showing high deNOx efficiency (>90% NO conversion) and N2 selectivity (>95%) in a wide operating temperature range of 120–300 °C. They considered that the superior NH3-SCR performance could be attributed to the unique structure and highly integrated mesoporous TiO2 supported by the multicomponent system with high surface areas, accessible and homogenously dispersed CuO and MnO2 with multivalent nature and good redox activity. Although the CuO/MnO2-mTiO2 catalyst had good tolerance to H2O, the resistance to SO2 and H2O/SO2 poisoning, as well as high space velocity (GHSV), still need to be enhanced for practical use. Li et al. [135] synthesized fly ash-derived SBA-15 mesoporous molecular sieves supported Fe and/or Mn catalysts, and reported that Fe-Mn/SBA-15 catalyst showed notably greater NH3-SCR activity than Mn/SBA-15 or Fe/SBA-15 in the temperature range of 150–250 °C. Moreover, the Fe-Mn/SBA-15 catalyst exhibited good time-on-stream stability (200 h) and water tolerance at 200 °C. The high metal dispersion, Mn4+/Mn3+ ratio, the concentration of adsorbed oxygen, and the redox activity are important features to enhance the NH3-SCR performance of the Fe-Mn/SBA-15 catalyst. In their subsequent study, the authors investigated the mechanisms of NO reduction and N2O formation using in-situ DRIFT and transient reaction studies and proposed a possible denitration mechanism over the Fe-Mn/SBA-15 catalyst which is shown in Figure 25. The NH3-SCR reaction over the Fe-Mn/SBA-15 catalyst proceeded through Langmuir-Hinshelwood, Eley-Rideal, and Mars-van Krevelen mechanisms. Their results also revealed that a large amount of nitrate thereby N2O being produced over the Fe-Mn/SBA-15 during the reaction due to its strong oxidation ability, low acidity, and high basicity, which resulted in the lower N2 selectivity [136].
Tang and co-workers [137] reported that Mn2CoO4/reduced graphene oxide (Mn2CoO4/rGO) catalyst with an optimal amount of CoCl2·6H2O of 0.3 (millimole) showed excellent NH3-SCR activity and stability at low-temperature due to its large specific surface area, abundant Lewis acid sites, and special three-dimensional architecture. When 100 ppm SO2 added to reaction feed, the NOx conversion over the optimal catalyst decreased significantly (96% to 53%), and the NOx conversion was recovered to original level by water-washing after stopping the supply of SO2. However, the decreased activity (100% to 82% NOx conversion) in the presence of H2O was restored to the original level after removing H2O from the feed gas. Wang et al. [138] investigated the honeycomb cordierite-based Mn-Ce/Al2O3 catalyst for NH3-SCR reaction and found that it showed good activity and reasonable resistance to SO2/H2O. The catalyst deactivation in the presence of SO2 was ascribed to the deposition of ammonium hydrogen sulfate and sulfated CeO2 on the catalyst surface during the NH3-SCR process.
Meng et al. [139] synthesized a novel CuAlOx/CNTs (CNTs = carbon nanotubes) catalyst by facile one-step carbothermal reduction decomposition method for low-temperature NH3-SCR. The CuAlOx/CNTs catalyst was found to exhibit higher NOx conversion (>80%) and N2 selectivity (>90%) than the CuAlOx in the temperature range of 180–300 °C (Figure 26a). They concluded that more favorable formation of Cu+ active sites, better dispersion of active CuO species and higher surface adsorbed oxygen were beneficial to enhance the NH3-SCR activity of CuAlOx/CNTs catalyst. As shown in Figure 26b, the CuAlOx/CNTs catalyst displayed excellent resistance to SO2/H2O at 240 °C during the NH3-SCR. The authors attributed this outstanding SO2/H2O tolerance to the presence of CNTs that could promote the reaction of NH4HSO4 and NO continuously to avoid the formation and accumulation of excess ammonium sulfate salts on the catalyst surface. Li group [140] reported a series of ultra-low content copper-modified TiO2/CeO2 catalysts and observed that the catalyst with a Cu/Ce molar ratio of 0.005 exhibited the high NH3-SCR performance and good tolerance to SO2. Their characterization results disclosed that the addition of Cu into TiO2/CeO2 lead to enhance the Brønsted acid sites, amount of surface adsorbed oxygen and Ce3+ species, redox, and surface acidic properties, which in turn improve the NH3-SCR activity.
Recently, supported multi-metal oxide catalysts have been considered as the very promising candidates for low-temperature NH3-SCR reaction because of the enlarged synergetic catalysis effects of different components as well as improved metal-support interactions [141,142,143,144,145]. Wang and co-workers [146] reported a series of Nb modified Cu-Ce-Ti mixed oxide (NbyCCT, where y represented the atomic ratio of Nb to Ti) catalysts for low-temperature NH3-SCR reaction. It was found that NbyCCT catalysts demonstrated a better activity than the Cu-Ce-Ti (CCT) and Ce-Ti (CT) samples (Figure 27a). Among all the NbyCCT catalysts, Nb0.05CCT showed a higher NO conversion (>90%) in a broad temperature range of 180–360 °C under the GHSV of 40,000 h−1 (Figure 27a). Results indicated that the incorporation of Nb to Cu-Ce-Ti led to strong interactions among the active phases that increased the oxygen vacancies and inhibited the over-oxidation of NH3, which in turn improved the NH3-SCR activity and N2 selectivity in a wide temperature window. DRIFTS studies revealed that the introduction of Nb promoted the generation of NO2, which could improve the activity via “fast” SCR reaction process (Langmuir–Hinshelwood reaction pathway). As shown in Figure 27b, the optimal Nb0.05CCT catalyst exhibited higher resistance to SO2/H2O as compared to the Nb free catalyst. Li et al. [147] investigated the effect of Ho doping on the NH3-SCR performance and the SO2/H2O resistance of Mn-Ce/TiO2 catalyst. Among the catalysts tested, the catalyst with Ho/Ti of 0.1 (Mn0.4Ce0.07Ho0.1/TiO2) showed the best performance with >90% NO conversion in the temperature range of 150–220 °C, which was attributed to high concentration of chemisorbed oxygen, surface Mn4+/Mn3+ ratio, and acidity, as well as large specific surface area. Although the Mn0.4Ce0.07Ho0.1/TiO2 showed higher resistance to SO2 and H2O than the Mn0.4Ce0.07/TiO2 catalyst, it was deactivated some extent in presence of SO2/H2O which is irreversible.
Lu group [148] synthesized a series of activated coke (AC) supported FexCoyCezOm catalysts for low-temperature NH3-SCR, and found that the 3%Fe0.6Co0.2Ce0.2O1.57/AC catalyst had the best activity at 250–350 °C and good tolerance to H2O/SO2 at 250 °C. The superior performance of the catalyst was ascribed to the co-participation of Fe, Co, and Ce species with different valence states, high concentration of chemisorbed oxygen, well dispersed active components, increase of weak acid sites, good redox properties of metallic oxides, and abundant functional groups on the catalyst surface. Their mechanistic and kinetic studies also indicated that the enhanced active sites for the adsorption of NO and NH3, and the redox cycle among Fe, Co and Ce were responsible for the improved activity. Zhao et al. [149] reported a series of Mn-Ce-V-WOx/TiO2 composite oxide catalysts, exhibiting greater NH3-SCR activity than the TiO2 supported single-component catalysts (Figure 28a,b). Particularly, the catalyst with a molar ratio of active components/TiO2 = 0.2 showed the best performance (>90% NO conversion) from 150 to 400 °C (Figure 28a). As shown in Figure 28c, the optimal Mn-Ce-V-WOx/TiO2 (molar ratio of Mn-Ce-V-WOx/TiO2 = 0.2) showed excellent stability and outstanding tolerance to H2O/SO2 at 250 °C. The authors concluded that the better performance of Mn-Ce-V-WOx/TiO2 mainly attributed to the variety of valence states of the four active components and their high oxidation-reduction ability.

6. Conclusions

Since the emission standards for NOx are becoming more stringent to keep our atmosphere clean, the widespread use of fossil fuel in automobiles and industries require advanced catalytic materials for NOx emission control. The low-temperature SCR of NOx with NH3 would be a promising solution to mitigate the NOx emissions from mobile and stationary sources. Hence, the development of efficient catalysts for the low-temperature NH3-SCR with high deNOx activity, N2 selectivity, and high resistance toward SO2/H2O poisonings is the subject of increasing interest in the field of environmental catalysis. Transition metal-based oxide catalysts have drawn much attention for low-temperature NH3-SCR due to their excellent redox properties, high activity, durability, and relatively low manufacturing costs. In this review, we have summarized the recent progress in the low-temperature NH3-SCR technology over the various transition metal-based catalysts. Over the past decades, significant research efforts have been made to improve the de-NOx efficiency and SO2/H2O tolerance of transition metal-based oxides in NH3-SCR at low-temperatures. Various transition metal-based mixed oxides with and without support have been extensively studied for NH3-SCR reaction and, particularly, MnOx-based catalyst formulations have caught much attention because of their excellent de-NOx efficiency at low-temperatures. The modification of transition metal oxides by doping with other metal oxides led to high redox ability and acidic sites, and consequently, better NH3-SCR performance at low-temperature. The loading of single and multi-transition metal-based oxides on the surface of supports (TiO2, TiO2 nanotubes, carbon nanotubes, etc.) could also enhance the NOx conversion and N2 selectivity in NH3-SCR reaction by the fine dispersion of active component/s and its/their strong interaction with the support. The choice of metal loading and the support could play a key role in the catalytic function of the supported transition metal-based catalysts. The synergistic redox interaction between the active components of mixed metal oxide/supported metal oxide catalysts was also found to be an important factor to design the efficient denitration catalysts. In spite of the significant progress on the SO2/H2O tolerance of the catalysts, the durability of catalysts in the presence of both SO2 and H2O still needs to be improved. Most transition metal-based catalysts suffered from low resistance when the reaction feed contains both SO2 and H2O streams simultaneously. Hence, researchers have continuously explored the different options of transition metal-based mixed oxides and active transition metal/s-support combinations in order to develop the better NH3-SCR catalysts in terms of SO2/H2O tolerance at low-temperature. The understanding of the inhibition mechanism of SO2 and H2O could be a promising strategy to develop high SO2/H2O resistance catalysts for NH3-SCR reaction. However, the SO2/H2O inhibition mechanism was not very clear that needs to be investigated deeply. Especially, the design of transition metal-based catalysts with a combination of high NOx conversion and N2 selectivity in a wide operation temperature window and good resistance to SO2/H2O have attracted paramount attention, but it is still challenging task. The scope of NH3-SCR research is quite vast and a large number of improvements need to be achieved in the near future.

Author Contributions

The original draft was prepared by D.D., reviewed and edited by D.D., P.G.S., B.M.R, and P.R.E.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. 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 2014, 148–149, 114–122. [Google Scholar] [CrossRef]
  2. Boningari, T.; Smirniotis, P.G. Impact of nitrogen oxides on the environment and human health: Mn-based materials for the NOx abatement. Curr. Opin. Chem. Eng. 2016, 13, 133–141. [Google Scholar] [CrossRef]
  3. Xin, Y.; Li, Q.; Zhang, Z. Zeolitic materials for DeNOx selective catalytic reduction. ChemCatChem 2018, 10, 29–41. [Google Scholar] [CrossRef]
  4. Damma, D.; Boningari, T.; Ettireddy, P.R.; Reddy, B.M.; Smirniotis, P.G. Direct decomposition of NOx over TiO2 supported transition metal oxides at low temperatures. Ind. Eng. Chem. Res. 2018, 57, 16615–16621. [Google Scholar] [CrossRef]
  5. Thirupathi, B.; Smirniotis, P.G. Nickel-doped Mn/TiO2 as an efficient catalyst for the low-temperature SCR of NO with NH3: Catalytic evaluation and characterizations. J. Catal. 2012, 288, 74–83. [Google Scholar] [CrossRef]
  6. Chen, L.; Si, Z.; Wu, X.; Weng, D. DRIFT study of CuO−CeO2−TiO2 mixed oxides for NOx reduction with NH3 at low temperatures. ACS Appl. Mater. Interfaces 2014, 6, 8134–8145. [Google Scholar] [CrossRef]
  7. Chen, M.; Yang, J.; Liu, Y.; Li, W.; Fan, J.; Ran, X.; Teng, W.; Sun, Y.; Zhang, W.-X.; Li, G.; et al. TiO2 interpenetrating networks decorated with SnO2 nanocrystals: Enhanced activity of selective catalytic reduction of NO with NH3. J. Mater. Chem. A 2015, 3, 1405–1409. [Google Scholar] [CrossRef]
  8. Chen, Z.; Yang, Q.; Li, H.; Li, X.; Wang, L.; Tsang, S.C. Cr–MnOx mixed-oxide catalysts for selective catalytic reduction of NOx with NH3 at low temperature. J. Catal. 2010, 276, 56–65. [Google Scholar] [CrossRef]
  9. Boningari, T.; Koirala, R.; Smirniotis, P.G. Low-temperature selective catalytic reduction of NO with NH3 over V/ZrO2 prepared by flame-assisted spray pyrolysis: Structural and catalytic properties. Appl. Catal. B 2012, 127, 255–264. [Google Scholar] [CrossRef]
  10. Chen, L.; Li, J.; Ge, M. DRIFT study on cerium-tungsten/titiania catalyst for selective catalytic reduction of NOx with NH3. Environ. Sci. Technol. 2010, 44, 9590–9596. [Google Scholar] [CrossRef]
  11. Liu, F.; Asakura, K.; He, H.; Shan, W.; Shi, X.; Zhang, C. Influence of sulfation on iron titanate catalyst for the selective catalytic reduction of NOx with NH3. Appl. Catal. B 2011, 103, 369–377. [Google Scholar] [CrossRef]
  12. Cai, S.; Zhang, D.; Zhang, L.; Huang, L.; Li, H.; Gao, R.; Shi, L.; Zhang, J. Comparative study of 3D ordered macroporous Ce0.75Zr0.2M0.05O2−δ (M = Fe, Cu, Mn, Co) for selective catalytic reduction of NO with NH3. Catal. Sci. Technol. 2014, 4, 93–101. [Google Scholar] [CrossRef]
  13. Wang, X.; Wen, W.; Su, Y.; Wang, R. Influence of transition metals (M = Co, Fe and Mn) on ordered mesoporous CuM/CeO2 catalysts and applications in selective catalytic reduction of NOx with H2. RSC Adv. 2015, 5, 63135–63141. [Google Scholar] [CrossRef]
  14. Liu, J.; Li, X.; Zhao, Q.; Ke, J.; Xiao, H.; Lv, X.; Liu, S.; Tadé, M.; Wang, S. Mechanistic investigation of the enhanced NH3-SCR on cobalt-decorated Ce-Ti mixed oxide: In situ FTIR analysis for structure-activity correlation. Appl. Catal. B 2017, 200, 297–308. [Google Scholar] [CrossRef]
  15. Schneider, H.; Maciejewski, M.; Köhler, K.; Wokaun, A.; Baiker, A. Chromia supported on titania, VI Properties of different chromium oxide phases in the catalytic reduction of NO by NH3 studied by in situ diffuse reflectance FTIR spectroscopy. J. Catal. 1995, 157, 312–320. [Google Scholar] [CrossRef]
  16. Chen, Z.; Wang, F.; Li, H.; Yang, Q.; Wang, L.; Li, X. Low-temperature selective catalytic reduction of NOx with NH3 over Fe‒Mn mixed-oxide catalysts containing Fe3Mn3O8 phase. Ind. Eng. Chem. Res. 2012, 51, 202–212. [Google Scholar] [CrossRef]
  17. Smirniotis, P.G.; Peña, D.A.; Uphade, B.S. Low-temperature selective catalytic reduction (SCR) of NO with NH3 by using Mn, Cr, and Cu oxides supported on hombikat TiO2. Angew. Chem. Int. Ed. 2001, 40, 2479–2482. [Google Scholar] [CrossRef]
  18. Wu, Z.; Jiang, B.; Liu, Y. Effect of transition metals addition on the catalyst of manganese/titania for low-temperature selective catalytic reduction of nitric oxide with ammonia. Appl. Catal. B 2008, 79, 347–355. [Google Scholar] [CrossRef]
  19. Liu, F.; He, H.; Zhang, C.; Feng, Z.; Zheng, L.; Xie, Y.; Hu, T. Selective catalytic reduction of NO with NH3 over iron titanate catalyst: Catalytic performance and characterization. Appl. Catal. B 2010, 96, 408–420. [Google Scholar] [CrossRef]
  20. Yao, X.; Kong, T.; Chen, L.; Ding, S.; Yang, F.; Dong, L. Enhanced low-temperature NH3-SCR performance of MnOx/CeO2 catalysts by optimal solvent effect. Appl. Surf. Sci. 2017, 420, 407–415. [Google Scholar] [CrossRef]
  21. Sreekanth, P.M.; Peña, D.A.; Smirniotis, P.G. Titania supported bimetallic transition metal oxides for low-temperature SCR of NO with NH3. Ind. Eng. Chem. Res. 2006, 45, 6444–6449. [Google Scholar] [CrossRef]
  22. Thirupathi, B.; Smirniotis, P.G. Co-doping a metal (Cr, Fe, Co, Ni, Cu, Zn, Ce, and Zr) on Mn/TiO2 catalyst and its effect on the selective reduction of NO with NH3 at low-temperatures. Appl. Catal. B 2011, 110, 195–206. [Google Scholar] [CrossRef]
  23. Thirupathi, B.; Smirniotis, P.G. Effect of nickel as dopant in Mn/TiO2 catalysts for the low-temperature selective reduction of NO with NH3. Catal. Lett. 2011, 141, 1399–1404. [Google Scholar] [CrossRef]
  24. Ettireddy, P.R.; Ettireddy, N.; Boningari, T.; Pardemann, R.; Smirniotis, P.G. Investigation of the selective catalytic reduction of nitric oxide with ammonia over Mn/TiO2 catalysts through transient isotopic labeling and in situ FT-IR studies. J. Catal. 2012, 292, 53–63. [Google Scholar] [CrossRef]
  25. Boningari, T.; Ettireddy, P.R.; Somogyvari, A.; Liu, Y.; Vorontsov, A.; McDonald, C.A.; Smirniotis, P.G. Influence of elevated surface texture hydrated titania on Ce-doped Mn/TiO2 catalysts for the low-temperature SCR of NOx under oxygen-rich conditions. J. Catal. 2015, 325, 145–155. [Google Scholar] [CrossRef]
  26. Pappas, D.K.; Boningari, T.; Boolchand, P.; Smirniotis, P.G. Novel manganese oxide confined interweaved titania nanotubes for the low-temperature selective catalytic reduction (SCR) of NOx by NH3. J. Catal. 2016, 334, 1–13. [Google Scholar] [CrossRef]
  27. 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 2010, 95, 160–168. [Google Scholar] [CrossRef]
  28. Cai, S.; Zhang, D.; Shi, L.; Xu, J.; Zhang, L.; Huang, L.; Li, H.; Zhang, J. Porous Ni–Mn oxide nanosheets in situ formed on nickel foam as 3D hierarchical monolith de-NOx catalysts. Nanoscale 2014, 6, 7346–7353. [Google Scholar] [CrossRef] [PubMed]
  29. 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]
  30. Zhang, L.; Shi, L.; Huang, L.; Zhang, J.; Gao, R.; Zhang, D. Rational design of high-performance DeNOx catalysts based on MnxCo3−xO4 nanocages derived from metal−organic frameworks. ACS Catal. 2014, 4, 1753–1763. [Google Scholar] [CrossRef]
  31. Li, L.; Wu, Y.; Hou, X.; Chu, B.; Nan, B.; Qin, Q.; Fan, M.; Sun, C.; Li, B.; Dong, L.; et al. Investigation of two-phase intergrowth and coexistence in Mn−Ce−Ti−O catalysts for the selective catalytic reduction of NO with NH3: Structure−activity relationship and reaction mechanism. Ind. Eng. Chem. Res. 2019, 58, 849–862. [Google Scholar] [CrossRef]
  32. Gao, F.; Tang, X.; Yi, H.; Zhao, S.; Wang, J.; Gu, T. Improvement of activity, selectivity and H2O&SO2-tolerance of micromesoporous CrMn2O4 spinel catalyst for low-temperature NH3-SCR of NOx. Appl. Surf. Sci. 2019, 466, 411–424. [Google Scholar]
  33. Li, X.J.; Du, Y.; Guo, X.M.; Wang, R.N.; Hou, B.H.; Wu, X. Synthesis of a novel NiMnTi mixed metal oxides from LDH precursor and its catalytic application for selective catalytic reduction of NOx with NH3. Catal. Lett. 2019, 149, 456–464. [Google Scholar] [CrossRef]
  34. 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 2019, 245, 743–752. [Google Scholar] [CrossRef]
  35. Wang, X.; Wu, W.; Chen, Z.; Wang, R. Bauxite-supported transition metal oxides: Promising low-temperature and SO2-tolerant catalysts for selective catalytic reduction of NOx. Sci. Rep. 2015, 5, 9766. [Google Scholar] [CrossRef]
  36. Wang, Z.-Y.; Guo, R.-T.; Shi, X.; Pan, W.-G.; Liu, J.; Sun, X.; Liu, S.-W.; Liu, X.-Y.; Qin, H. The enhanced performance of Sb-modified Cu/TiO2 catalyst for selective catalytic reduction of NOx with NH3. Appl. Surf. Sci. 2019, 475, 334–341. [Google Scholar] [CrossRef]
  37. Hu, H.; Cai, S.; Li, H.; Huang, L.; Shi, L.; Zhang, D. Mechanistic aspects of deNOx 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]
  38. Meng, B.; Zhao, Z.; Chen, Y.; Wang, X.; Li, Y.; Qiu, J. Low-temperature synthesis of Mn-based mixed metal oxides with novel fluffy structures as efficient catalysts for selective reduction of nitrogen oxides by ammonia. Chem. Commun. 2014, 50, 12396–12399. [Google Scholar] [CrossRef]
  39. Qiu, M.; Zhan, S.; Zhu, D.; Yu, H.; Shi, Q. NH3-SCR performance improvement of mesoporous Sn modified Cr-MnOx catalysts at low temperatures. Catal. Today 2015, 258, 103–111. [Google Scholar] [CrossRef]
  40. Shi, J.-W.; Gao, C.; Liu, C.; Fan, Z.; Gao, G.; Niu, C. Porous MnOx for low-temperature NH3-SCR of NOx: The intrinsic relationship between surface physicochemical property and catalytic activity. J. Nanopart. Res. 2017, 19, 194–204. [Google Scholar] [CrossRef]
  41. Husnain, N.; Wang, E.; Li, K.; Anwar, M.T.; Mehmood, A.; Gul, M.; Li, D.; Mao, J. Iron oxide-based catalysts for low-temperature selective catalytic reduction of NOx with NH3. Rev. Chem. Eng. 2019, 35, 239–264. [Google Scholar] [CrossRef]
  42. Liu, F.; Shan, W.; Lian, Z.; Xie, L.; Yang, W.; He, H. Novel MnWOx catalyst with remarkable performance for low temperature NH3-SCR of NOx. Catal. Sci. Technol. 2013, 3, 2699–2707. [Google Scholar] [CrossRef]
  43. Fan, G.; Shi, J.-W.; Gao, C.; Gao, G.; Wang, B.; Niu, C. Rationally designed porous MnOx−FeOx nanoneedles for low-temperature selective catalytic reduction of NOx by NH3. ACS Appl. Mater. Interfaces 2017, 9, 16117–16127. [Google Scholar] [CrossRef] [PubMed]
  44. Gao, G.; Shi, J.-W.; Fan, G.; Gao, C.; Niu, C. MnM2O4 microspheres (M = Co, Cu, Ni) for selective catalytic reduction of NO with NH3: Comparative study on catalytic activity and reaction mechanism via in-situ diffuse reflectance infrared Fourier transform spectroscopy. Chem. Eng. J. 2017, 325, 91–100. [Google Scholar] [CrossRef]
  45. 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]
  46. Gao, C.; Shi, J.-W.; Fan, Z.; Gao, G.; Niu, C. Sulfur and water resistance of Mn-based catalysts for low-temperature selective catalytic reduction of NOx: A review. Catalysts 2018, 8, 11. [Google Scholar] [CrossRef]
  47. Hu, X.; Huang, L.; Zhang, J.; Li, H.; Zha, K.; Shi, L.; Zhang, D. Facile and template-free fabrication of mesoporous 3D nanosphere-like MnxCo3‒xO4 as highly effective catalysts for low temperature SCR of NOx with NH3. J. Mater. Chem. A 2018, 6, 2952–2963. [Google Scholar] [CrossRef]
  48. 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 2018, 221, 652–663. [Google Scholar] [CrossRef]
  49. Yan, Q.; Nie, Y.; Yang, R.; Cui, Y.; O’Hare, D.; Wang, Q. Highly dispersed CuyAlOx mixed oxides as superior low-temperature alkali metal and SO2 resistant NH3-SCR catalysts. Appl. Catal. A 2017, 538, 37–50. [Google Scholar] [CrossRef]
  50. 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]
  51. Sun, P.; Guo, R.-T.; Liu, S.-M.; Wang, S.-X.; Pan, W.-G.; Li, M.-Y. The enhanced performance of MnOx catalyst for NH3-SCR reaction by the modification with Eu. Appl. Catal. A 2017, 531, 129–138. [Google Scholar] [CrossRef]
  52. Xin, Y.; Li, H.; Zhang, N.; Li, Q.; Zhang, Z.; Cao, X.; Hu, P.; Zheng, L.; Anderson, J.A. Molecular-level insight into selective catalytic reduction of NOx with NH3 to N2 over a highly efficient bifunctional Va-MnOx catalyst at low temperature. ACS Catal. 2018, 8, 4937–4949. [Google Scholar] [CrossRef]
  53. Han, Y.; Mu, J.; Li, X.; Gao, J.; Fan, S.; Tan, F.; Zhao, Q. Triple-shelled NiMn2O4 hollow spheres as an efficient catalyst for low-temperature selective catalytic reduction of NOx with NH3. Chem. Commun. 2018, 54, 9797–9800. [Google Scholar] [CrossRef] [PubMed]
  54. Gao, F.; Tang, X.; Yi, H.; Zhao, S.; Wang, J.; Shi, Y.; Meng, X. Novel Co– or Ni–Mn binary oxide catalysts with hydroxyl groups for NH3–SCR of NOx at low temperature. Appl. Surf. Sci. 2018, 443, 103–113. [Google Scholar] [CrossRef]
  55. Sun, W.; Li, X.; Zhao, Q.; Mu, J.; Chen, J. Fe–Mn mixed oxide catalysts synthesized by one-step urea-precipitation method for the selective catalytic reduction of NOx with NH3 at low temperatures. Catal. Lett. 2018, 148, 227–234. [Google Scholar] [CrossRef]
  56. Kwon, D.W.; Nam, K.B.; Hong, S.C. The role of ceria on the activity and SO2 resistance of catalysts for the selective catalytic reduction of NOx by NH3. Appl. Catal. B 2015, 166–167, 37–44. [Google Scholar] [CrossRef]
  57. Fan, G.; 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]
  58. Li, C.; Tang, X.; Yi, H.; Wang, L.; Cui, X.; Chu, C.; Li, J.; Zhang, R.; Yu, Q. Rational design of template-free MnOx-CeO2 hollow nanotube as de-NOx catalyst at low temperature. Appl. Surf. Sci. 2018, 428, 924–932. [Google Scholar] [CrossRef]
  59. Liu, F.; Yu, Y.; He, H. Environmentally-benign catalysts for the selective catalytic reduction of NOx from diesel engines: Structure–activity relationship and reaction mechanism aspects. Chem. Commun. 2014, 50, 8445–8463. [Google Scholar] [CrossRef]
  60. Zhan, S.; Qiu, M.; Yang, S.; Zhu, D.; Yu, H.; Li, Y. Facile preparation of MnO2 doped Fe2O3 hollow nanofibers for low temperature SCR of NO with NH3. J. Mater. Chem. A 2014, 2, 20486–20493. [Google Scholar] [CrossRef]
  61. Li, Y.; Wan, Y.; Li, Y.; Zhan, S.; Guan, Q.; Tian, Y. Low-temperature selective catalytic reduction of NO with NH3 over Mn2O3-doped Fe2O3 hexagonal microsheets. ACS Appl. Mater. Interfaces 2016, 8, 5224–5233. [Google Scholar] [CrossRef]
  62. Yan, L.; Liu, Y.; Zha, K.; Li, H.; Shi, L.; Zhang, D. Scale-activity relationship of MnOx-FeOy nanocage catalysts derived from prussian blue analogues for low-temperature NO reduction: Experimental and DFT studies. ACS Appl. Mater. Interfaces 2017, 9, 2581–2593. [Google Scholar] [CrossRef] [PubMed]
  63. Mu, J.; Li, X.; Sun, W.; Fan, S.; Wang, X.; Wang, L.; Qin, M.; Gan, G.; Yin, Z.; Zhang, D. Inductive effect boosting catalytic performance of advanced Fe1−xVxOδ catalysts in low-temperature NH3 selective catalytic reduction: Insight into the structure, interaction, and mechanisms. ACS Catal. 2018, 8, 6760–6774. [Google Scholar] [CrossRef]
  64. Li, Y.; Han, X.; Hou, Y.; Guo, Y.; Liu, Y.; Cui, Y.; Huang, Z. Role of CTAB in the improved H2O resistance for selective catalytic reduction of NO with NH3 over iron titanium catalyst. Chem. Eng. J. 2018, 347, 313–321. [Google Scholar] [CrossRef]
  65. Qiu, M.; Zhan, S.; Yu, H.; Zhu, D. Low-temperature selective catalytic reduction of NO with NH3 over ordered mesoporous MnxCo3−xO4 catalyst. Catal. Commun. 2015, 62, 107–111. [Google Scholar] [CrossRef]
  66. Qiu, M.; Zhan, S.; Yu, H.; Zhu, D.; Wang, S. Facile preparation of ordered mesoporous MnCo2O4 for low-temperature selective catalytic reduction of NO with NH3. Nanoscale 2015, 7, 2568–2577. [Google Scholar] [CrossRef]
  67. 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]
  68. Qiao, J.; Wang, N.; Wang, Z.; Sun, W.; Sun, K. Porous bimetallic Mn2Co1Ox catalysts prepared by a one-step combustion method for the low temperature selective catalytic reduction of NOx with NH3. Catal. Commun. 2015, 72, 111–115. [Google Scholar] [CrossRef]
  69. Kang, M.; Park, E.D.; Kim, J.M.; Yie, J.E. Cu–Mn mixed oxides for low temperature NO reduction with NH3. Catal. Today 2006, 111, 236–241. [Google Scholar] [CrossRef]
  70. Guo, R.-T.; Zhen, W.-L.; Pan, W.-G.; Zhou, Y.; Hong, J.-N.; Xu, H.-J.; Jin, Q.; Ding, C.-G.; Guo, S.-Y. Effect of Cu doping on the SCR activity of CeO2 catalyst prepared by citric acid method. J. Ind. Eng. Chem. 2014, 20, 1577–1580. [Google Scholar] [CrossRef]
  71. Ali, S.; Chen, L.; Li, Z.; Zhang, T.; Li, R.; Bakhtiar, S.U.H.; Leng, X.; Yuan, F.; Niu, X.; Zhu, Y. Cux-Nb1.1-x (x = 0.45, 0.35, 0.25, 0.15) bimetal oxides catalysts for the low temperature selective catalytic reduction of NO with NH3. Appl. Catal. B 2018, 236, 25–35. [Google Scholar] [CrossRef]
  72. Yan, Q.; Chen, S.; Zhang, C.; Wang, Q.; Louis, B. Synthesis and catalytic performance of Cu1Mn0.5Ti0.5Ox mixed oxide as low-temperature NH3-SCR catalyst with enhanced SO2 resistance. Appl. Catal. B 2018, 238, 236–247. [Google Scholar] [CrossRef]
  73. Liu, J.; Li, X.; Li, R.; Zhao, Q.; Ke, J.; Xiao, H.; Wang, L.; Liu, S.; Tadé, M.; Wang, S. Facile synthesis of tube-shaped Mn-Ni-Ti solid solution and preferable Langmuir-Hinshelwood mechanism for selective catalytic reduction of NOx by NH3. Appl. Catal. A 2018, 549, 289–301. [Google Scholar] [CrossRef]
  74. Wang, C.; Yu, F.; Zhu, M.; Wang, X.; Dan, J.; Zhang, J.; Cao, P.; Dai, B. Microspherical MnO2-CeO2-Al2O3 mixed oxide for monolithic honeycomb catalyst and application in selective catalytic reduction of NOx with NH3 at 50–150 °C. Chem. Eng. J. 2018, 346, 182–192. [Google Scholar] [CrossRef]
  75. Cheng, K.; Liu, B.; Song, W.; Liu, J.; Chen, Y.; Zhao, Z.; Wei, Y. Effect of Nb promoter on the structure and performance of iron titanate catalysts for the selective catalytic reduction of NO with NH3. Ind. Eng. Chem. Res. 2018, 57, 7802–7810. [Google Scholar] [CrossRef]
  76. Yan, Q.; Chen, S.; Qiu, L.; Gao, Y.; O’Hare, D.; Wang, Q. The synthesis of CuyMnzAl1−zOx mixed oxide as a low-temperature NH3-SCR catalyst with enhanced catalytic performance. Dalton Trans. 2018, 47, 2992–3004. [Google Scholar] [CrossRef] [PubMed]
  77. Geng, Y.; Shan, W.; Yang, S.; Liu, F. W-modified Mn−Ti mixed oxide catalyst for the selective catalytic reduction of NO with NH3. Ind. Eng. Chem. Res. 2018, 57, 9112–9119. [Google Scholar] [CrossRef]
  78. 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]
  79. Xiong, Z.-B.; Hu, Q.; Liu, D.-Y.; Wu, C.; Zhou, F.; Wang, Y.-Z.; Jin, J.; Lu, C.-M. Influence of partial substitution of iron oxide by titanium oxide on the structure and activity of iron–cerium mixed oxide catalyst for selective catalytic reduction of NOx with NH3. Fuel 2016, 165, 432–439. [Google Scholar] [CrossRef]
  80. Wang, X.; Li, X.; Zhao, Q.; Sun, W.; Tade, M.; Liu, S. Improved activity of W-modified MnOx–TiO2 catalysts for the selective catalytic reduction of NO with NH3. Chem. Eng. J. 2016, 288, 216–222. [Google Scholar] [CrossRef]
  81. Zamudio, M.A.; Russo, N.; Fino, D. Low temperature NH3 selective catalytic reduction of NOx over substituted MnCr2O4 spinel-oxide catalysts. Ind. Eng. Chem. Res. 2011, 50, 6668–6672. [Google Scholar] [CrossRef]
  82. Zhang, T.; Qiu, F.; Chang, H.; Peng, Y.; Li, J. Novel W-modified SnMnCeOx catalyst for the selective catalytic reduction of NOx with NH3. Catal. Commun. 2017, 100, 117–120. [Google Scholar] [CrossRef]
  83. Ali, S.; Chen, L.; Yuan, F.; Li, R.; Zhang, T.; Bakhtiar, S.U.H.; Leng, X.; Niu, X.; Zhu, Y. Synergistic effect between copper and cerium on the performance of Cux-Ce0.5-x-Zr0.5 (x = 0.1–0.5) oxides catalysts for selective catalytic reduction of NO with ammonia. Appl. Catal. B 2017, 210, 223–234. [Google Scholar] [CrossRef]
  84. Wei, Y.; Fan, H.; Wang, R. Transition metals (Co, Zr, Ti) modified iron-samarium oxide as efficient catalysts for selective catalytic reduction of NOx at low-temperature. Appl. Surf. Sci. 2018, 459, 63–73. [Google Scholar] [CrossRef]
  85. Gao, X.; Du, X.-S.; Cui, L.-W.; Fu, Y.-C.; Luo, Z.-Y.; Cen, K.-F. A Ce–Cu–Ti oxide catalyst for the selective catalytic reduction of NO with NH3. Catal. Commun. 2010, 12, 255–258. [Google Scholar] [CrossRef]
  86. Xiong, Z.B.; Wu, C.; Hu, Q.; Wang, Y.Z.; Jin, J.; Lu, C.M.; Guo, D.X. Promotional effect of microwave hydrothermal treatment on the low-temperature NH3-SCR activity over iron-based catalyst. Chem. Eng. J. 2016, 286, 459–466. [Google Scholar] [CrossRef]
  87. Cheng, K.; Song, W.; Cheng, Y.; Zheng, H.; Wang, L.; Liu, J.; Zhao, Z.; Wei, Y. Enhancing the low temperature NH3-SCR activity of FeTiOx catalysts via Cu doping: A combination of experimental and theoretical study. RSC Adv. 2018, 8, 19301–19309. [Google Scholar] [CrossRef]
  88. Fang, N.; Guo, J.; Shu, S.; Luo, H.; Chu, Y.; Li, J. Enhancement of low-temperature activity and sulfur resistance of Fe0.3Mn0.5Zr0.2 catalyst for NO removal by NH3-SCR. Chem. Eng. J. 2017, 325, 114–123. [Google Scholar] [CrossRef]
  89. Guo, R.-T.; Sun, X.; Liu, J.; Pan, W.-G.; Li, M.-Y.; Liu, S.-M.; Sun, P.; Liu, S.-W. Enhancement of the NH3-SCR catalytic activity of MnTiOx catalyst by the introduction of Sb. Appl. Catal. A 2018, 558, 1–8. [Google Scholar] [CrossRef]
  90. Shi, J.-W.; Gao, G.; Fan, Z.; Gao, C.; Wang, B.; Wang, Y.; Li, Z.; He, C.; Niu, C. NiyCo1-yMn2Ox microspheres for the selective catalytic reduction of NOx with NH3: The synergetic effects between Ni and Co for improving low-temperature catalytic performance. Appl. Catal. A 2018, 560, 1–11. [Google Scholar] [CrossRef]
  91. Wu, X.; Feng, Y.; Du, Y.; Liu, X.; Zou, C.; Li, Z. Enhancing DeNOx performance of CoMnAl mixed metal oxides in low-temperature NH3-SCR by optimizing layered double hydroxides (LDHs) precursor template. Appl. Surf. Sci. 2019, 467–468, 802–810. [Google Scholar] [CrossRef]
  92. Leng, X.; Zhang, Z.; Li, Y.; Zhang, T.; Ma, S.; Yuan, F.; Niu, X.; Zhu, Y. Excellent low temperature NH3-SCR activity over MnaCe0.3TiOx (a = 0.1–0.3) oxides: Influence of Mn addition. Fuel Process. Technol. 2018, 181, 33–43. [Google Scholar] [CrossRef]
  93. Ali, S.; Li, Y.; Zhang, T.; Bakhtiar, S.U.H.; Leng, X.; Li, Z.; Yuan, F.; Niu, X.; Zhu, Y. Promotional effects of Nb on selective catalytic reduction of NO with NH3 over Fex-Nb0.5-x-Ce0.5 (x = 0.45, 0.4, 0.35) oxides catalysts. Mol. Catal. 2018, 461, 97–107. [Google Scholar] [CrossRef]
  94. Sun, C.; Liu, H.; Chen, W.; Chen, D.; Yu, S.; Liu, A.; Dong, L.; Feng, S. Insights into the Sm/Zr co-doping effects on N2 selectivity and SO2 resistance of a MnOx-TiO2 catalyst for the NH3-SCR reaction. Chem. Eng. J. 2018, 347, 27–40. [Google Scholar] [CrossRef]
  95. Yan, Q.; Chen, S.; Zhang, C.; O’Hare, D.; Wang, Q. Synthesis of Cu0.5Mg1.5Mn0.5Al0.5Ox mixed oxide from layered double hydroxide precursor as highly efficient catalyst for low-temperature selective catalytic reduction of NOx with NH3. J. Colloid Interface Sci. 2018, 526, 63–74. [Google Scholar] [CrossRef] [PubMed]
  96. Chen, L.; Yuan, F.; Li, Z.; Niu, X.; Zhu, Y. Synergistic effect between the redox property and acidity on enhancing the low temperature NH3-SCR activity for NOx removal over the Co0.2CexMn0.8-xTi10 (x = 0–0.40) oxides catalysts. Chem. Eng. J. 2018, 354, 393–406. [Google Scholar] [CrossRef]
  97. Peña, D.A.; Uphade, B.S.; Reddy, E.P.; Smirniotis, P.G. Identification of surface species on titania-supported manganese, chromium, and copper oxide low-temperature SCR catalysts. J. Phys. Chem. B 2004, 108, 9927–9936. [Google Scholar] [CrossRef]
  98. Zhuang, K.; Qiu, J.; Tang, F.; Xu, B.; Fan, Y. The structure and catalytic activity of anatase and rutile titania supported manganese oxide catalysts for selective catalytic reduction of NO by NH3. Phys. Chem. Chem. Phys. 2011, 13, 4463–4469. [Google Scholar] [CrossRef]
  99. Boningari, T.; Pappas, D.K.; Ettireddy, P.R.; Kotrba, A.; Smirniotis, P.G. Influence of SiO2 on M/TiO2 (M = Cu, Mn, and Ce) formulations for low-temperature selective catalytic reduction of NOx with NH3: Surface properties and key components in relation to the activity of NOx reduction. Ind. Eng. Chem. Res. 2015, 54, 2261–2273. [Google Scholar] [CrossRef]
  100. Sheng, Z.; Ma, D.; Yu, D.; Xiao, X.; Huang, B.; Yang, L.; Wang, S. Synthesis of novel MnOx@TiO2 core-shell nanorod catalyst for low-temperature NH3-selective catalytic reduction of NOx with enhanced SO2 tolerance. Chin. J. Catal. 2018, 39, 821–830. [Google Scholar] [CrossRef]
  101. Cimino, S.; Totarella, G.; Tortorelli, M.; Lisi, L. Combined poisoning effect of K+ and its counter-ion (Cl or NO3−) on MnOx/TiO2 catalyst during the low temperature NH3-SCR of NO. Chem. Eng. J. 2017, 330, 92–101. [Google Scholar] [CrossRef]
  102. Fang, D.; Xie, J.; Hu, H.; Yang, H.; He, F.; Fu, Z. Identification of MnOx species and Mn valence states in MnOx/TiO2 catalysts for low temperature SCR. Chem. Eng. J. 2015, 271, 23–30. [Google Scholar] [CrossRef]
  103. Huang, J.; Huang, H.; Liu, L.; Jiang, H. Revisit the effect of manganese oxidation state on activity in low-temperature NO-SCR. Mol. Catal. 2018, 446, 49–57. [Google Scholar] [CrossRef]
  104. Yang, S.; Qi, F.; Xiong, S.; Dang, H.; Liao, Y.; Wong, P.K.; Li, J. MnOx supported on Fe–Ti spinel: A novel Mn based low temperature SCR catalyst with a high N2 selectivity. Appl. Catal. B 2016, 181, 570–580. [Google Scholar] [CrossRef]
  105. 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. A 2016, 522, 54–69. [Google Scholar] [CrossRef]
  106. Peña, D.A.; Uphade, B.S.; Smirniotis, P.G. TiO2-supported metal oxide catalysts for low-temperature selective catalytic reduction of NO with NH3, I. Evaluation and characterization of first row transition metals. J. Catal. 2004, 221, 421–431. [Google Scholar] [CrossRef]
  107. Smirniotis, P.G.; Sreekanth, P.M.; Peña, D.A.; Jenkins, R.G. Manganese oxide catalysts supported on TiO2, Al2O3, and SiO2: A comparison for low-temperature SCR of NO with NH3. Ind. Eng. Chem. Res. 2006, 45, 6436–6443. [Google Scholar] [CrossRef]
  108. Ettireddy, P.R.; Ettireddy, N.; Mamedov, S.; Boolchand, P.; Smirniotis, P.G. Surface characterization studies of TiO2 supported manganese oxide catalysts for low temperature SCR of NO with NH3. Appl. Catal. B 2007, 76, 123–134. [Google Scholar] [CrossRef]
  109. Boningari, T.; Pappas, D.K.; Smirniotis, P.G. Metal oxide-confined interweaved titania nanotubes M/TNT (M = Mn, Cu, Ce, Fe, V, Cr, and Co) for the selective catalytic reduction of NOx in the presence of excess oxygen. J. Catal. 2018, 365, 320–333. [Google Scholar] [CrossRef]
  110. Jia, B.; Guo, J.; Luo, H.; Shu, S.; Fang, N.; Li, J. Study of NO removal and resistance to SO2 and H2O of MnOx/TiO2, MnOx/ZrO2 and MnOx/ZrO2–TiO2. Appl. Catal. A 2018, 553, 82–90. [Google Scholar] [CrossRef]
  111. Zhang, B.; Zhang, S.; Liu, B.; Shen, H.; Li, L. High N2 selectivity in selective catalytic reduction of NO with NH3 over Mn/Ti–Zr catalysts. RSC Adv. 2018, 8, 12733–12741. [Google Scholar] [CrossRef]
  112. Han, J.; Zhang, D.; Maitarad, P.; Shi, L.; Cai, S.; Li, H.; Huang, L.; Zhang, J. Fe2O3 nanoparticles anchored in situ on carbon nanotubes via an ethanol-thermal strategy for the selective catalytic reduction of NO with NH3. Catal. Sci. Technol. 2015, 5, 438–446. [Google Scholar] [CrossRef]
  113. Pourkhalil, M.; Moghaddam, A.Z.; Rashidi, A.; Towfighi, J.; Mortazavi, Y. Preparation of highly active manganese oxides supported on functionalized MWNTs for low temperature NOx reduction with NH3. Appl. Surf. Sci. 2013, 279, 250–259. [Google Scholar] [CrossRef]
  114. Pan, S.; Luo, H.; Li, L.; Wei, Z.; Huang, B. H2O and SO2 deactivation mechanism of MnOx/MWCNTs for low-temperature SCR of NOx with NH3. J. Mol. Catal. A 2013, 377, 154–161. [Google Scholar] [CrossRef]
  115. Fang, C.; Zhang, D.; Shi, L.; Gao, R.; Li, H.; Ye, L.; Zhang, J. Highly dispersed CeO2 on carbon nanotubes for selective catalytic reduction of NO with NH3. Catal. Sci. Technol. 2013, 3, 803–811. [Google Scholar] [CrossRef]
  116. Qu, Z.; Miao, L.; Wang, H.; Fu, Q. Highly dispersed Fe2O3 on carbon nanotubes for low-temperature selective catalytic reduction of NO with NH3. Chem. Commun. 2015, 51, 956–958. [Google Scholar] [CrossRef] [PubMed]
  117. Bai, S.; Li, H.; Wang, L.; Guan, Y.; Jiang, S. The properties and mechanism of CuO modified carbon nanotube for NOx removal. Catal. Lett. 2014, 144, 216–221. [Google Scholar] [CrossRef]
  118. Zhang, D.; Zhang, L.; Shi, L.; Fang, C.; Li, H.; Gao, R.; Huang, L.; Zhang, J. In situ supported MnOx–CeOx on carbon nanotubes for the low-temperature selective catalytic reduction of NO with NH3. Nanoscale 2013, 5, 1127–1136. [Google Scholar] [CrossRef]
  119. Putluru, S.S.R.; Schill, L.; Jensen, A.D.; Siret, B.; Tabaries, F.; Fehrmann, R. Mn/TiO2 and Mn–Fe/TiO2 catalysts synthesized by deposition precipitation‒promising for selective catalytic reduction of NO with NH3 at low temperatures. Appl. Catal. B 2015, 165, 628–635. [Google Scholar] [CrossRef]
  120. You, X.; Sheng, Z.; Yu, D.; Yang, L.; Xiao, X.; Wang, S. Influence of Mn/Ce ratio on the physicochemical properties and catalytic performance of graphene supported MnOx-CeO2 oxides for NH3-SCR at low temperature. Appl. Surf. Sci. 2017, 423, 845–854. [Google Scholar] [CrossRef]
  121. Fang, D.; He, F.; Liu, X.; Qi, K.; Xie, J.; Li, F.; Yu, C. Low temperature NH3-SCR of NO over an unexpected Mn-based catalyst: Promotional effect of Mg doping. Appl. Surf. Sci. 2018, 427, 45–55. [Google Scholar] [CrossRef]
  122. Lu, X.; Song, C.; Jia, S.; Tong, Z.; Tang, X.; Teng, Y. Low-temperature selective catalytic reduction of NOX with NH3 over cerium and manganese oxides supported on TiO2–graphene. Chem. Eng. J. 2015, 260, 776–784. [Google Scholar] [CrossRef]
  123. Wang, X.; Wu, S.; Zou, W.; Yu, S.; Gui, K.; Dong, L. Fe-Mn/Al2O3 catalysts for low temperature selective catalytic reduction of NO with NH3. Chin. J. Catal. 2016, 37, 1314–1323. [Google Scholar] [CrossRef]
  124. Shi, J.; Zhang, Z.; Chen, M.; Zhang, Z.; Shangguan, W. Promotion effect of tungsten and iron co-addition on the catalytic performance of MnOx/TiO2 for NH3-SCR of NOx. Fuel 2017, 210, 783–789. [Google Scholar] [CrossRef]
  125. Zhao, C.; Wu, Y.; Liang, H.; Chen, X.; Tang, J.; Wang, X. N-doped graphene and TiO2 supported manganese and cerium oxides on low-temperature selective catalytic reduction of NOx with NH3. J. Adv. Ceram. 2018, 7, 197–206. [Google Scholar] [CrossRef]
  126. Fan, Y.; Ling, W.; Huang, B.; Dong, L.; Yu, C.; Xi, H. The synergistic effects of cerium presence in the framework and the surface resistance to SO2 and H2O in NH3-SCR. J. Ind. Eng. Chem. 2017, 56, 108–119. [Google Scholar] [CrossRef]
  127. Xu, Q.; Yang, W.; Cui, S.; Street, J.; Luo, Y. Sulfur resistance of Ce-Mn/TiO2 catalysts for low-temperature NH3–SCR. R. Soc. Open Sci. 2018, 5, 171846. [Google Scholar] [CrossRef] [PubMed]
  128. Lin, L.-Y.; Lee, C.-Y.; Zhang, Y.-R.; Bai, H. Aerosol-assisted deposition of Mn-Fe oxide catalyst on TiO2 for superior selective catalytic reduction of NO with NH3 at low temperatures. Catal. Commun. 2018, 111, 36–41. [Google Scholar] [CrossRef]
  129. Lee, T.; Bai, H. Metal Sulfate poisoning effects over MnFe/TiO2 for selective catalytic reduction of NO by NH3 at low temperature. Ind. Eng. Chem. Res. 2018, 57, 4848–4858. [Google Scholar] [CrossRef]
  130. Mu, J.; Li, X.; Sun, W.; Fan, S.; Wang, X.; Wang, L.; Qin, M.; Gan, G.; Yin, Z.; Zhang, D. Enhancement of low-temperature catalytic activity over a highly dispersed Fe−Mn/Ti catalyst for selective catalytic reduction of NOx with NH3. Ind. Eng. Chem. Res. 2018, 57, 10159–10169. [Google Scholar] [CrossRef]
  131. Liu, J.; Guo, R.-T.; Li, M.-Y.; Sun, P.; Liu, S.-M.; Pan, W.-G.; Liu, S.-W.; 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]
  132. Sun, P.; Huang, S.-X.; Guo, R.-T.; Li, M.-Y.; Liu, S.-M.; Pan, W.-G.; Fu, Z.-G.; Liu, S.-W.; Sun, X.; Liu, J. The enhanced SCR performance and SO2 resistance of Mn/TiO2 catalyst by the modification with Nb: A mechanistic study. Appl. Surf. Sci. 2018, 447, 479–488. [Google Scholar] [CrossRef]
  133. Sun, X.; Guo, R.-T.; Liu, J.; Fu, Z.-G.; Liu, S.-W.; Pan, W.-G.; Shi, X.; Qin, H.; Wang, Z.-Y.; Liu, X.-Y. The enhanced SCR performance of Mn/TiO2 catalyst by Mo modification: Identification of the promotion mechanism. Int. J. Hydrogen Energy 2018, 43, 16038–16048. [Google Scholar] [CrossRef]
  134. Fan, J.; Lv, M.; Luo, W.; Ran, X.; Deng, Y.; Zhang, W.-X.; Yang, J. Exposed metal oxide active sites on mesoporous titania channels: A promising design for low-temperature selective catalytic reduction of NO with NH3. Chem. Commun. 2018, 54, 3783–3786. [Google Scholar] [CrossRef] [PubMed]
  135. Li, G.; Wang, B.; Wang, H.; Ma, J.; Xu, W.Q.; Li, Y.; Han, Y.; Sun, Q. Fe and/or Mn oxides supported on fly ash-derived SBA-15 for low-temperature NH3-SCR. Catal. Commun. 2018, 108, 82–87. [Google Scholar] [CrossRef]
  136. Li, G.; Wang, B.; Wang, Z.; Li, Z.; Sun, Q.; Xu, W.Q.; Li, Y. Reaction mechanism of low-temperature selective catalytic reduction of NOx over Fe-Mn oxides supported on fly-ash-derived SBA-15 molecular sieves: Structure−activity relationships and in situ DRIFT analysis. J. Phys. Chem. C 2018, 122, 20210–20231. [Google Scholar] [CrossRef]
  137. Tang, X.; Li, C.; Yi, H.; Wang, L.; Yu, Q.; Gao, F.; Cui, X.; Chu, C.; Li, J.; Zhang, R. Facile and fast synthesis of novel Mn2CoO4@rGO catalysts for the NH3-SCR of NOx at low temperature. Chem. Eng. J. 2018, 333, 467–476. [Google Scholar] [CrossRef]
  138. Wang, C.; Zhang, C.; Zhao, Y.; Yan, X.; Cao, P. Poisoning effect of SO2 on honeycomb cordierite-based Mn–Ce/Al2O3 catalysts for NO reduction with NH3 at low temperature. Appl. Sci. 2018, 8, 95. [Google Scholar] [CrossRef]
  139. Meng, H.; Liu, J.; Du, Y.; Hou, B.; Wu, X.; Xie, X. Novel Cu-based oxides catalyst from one-step carbothermal reduction decomposition method for selective catalytic reduction of NO with NH3. Catal. Commun. 2019, 119, 101–105. [Google Scholar] [CrossRef]
  140. Li, L.; Zhang, L.; Ma, K.; Zou, W.; Cao, Y.; Xiong, Y.; Tang, C.; Dong, L. Ultra-low loading of copper modified TiO2/CeO2 catalysts for low-temperature selective catalytic reduction of NO by NH3. Appl. Catal. B 2017, 207, 366–375. [Google Scholar] [CrossRef]
  141. Zhu, Y.; Zhang, Y.; Xiao, R.; Huang, T.; Shen, K. Novel holmium-modified Fe-Mn/TiO2 catalysts with a broad temperature window and high sulfur dioxide tolerance for low-temperature SCR. Catal. Commun. 2017, 88, 64–67. [Google Scholar] [CrossRef]
  142. Nam, K.B.; Kwon, D.W.; Hong, S.C. DRIFT study on promotion effects of tungsten-modified Mn/Ce/Ti catalysts for the SCR reaction at low-temperature. Appl. Catal. A 2017, 542, 55–62. [Google Scholar] [CrossRef]
  143. Zhao, B.; Ran, R.; Guo, X.; Cao, L.; Xu, T.; Chen, Z.; Wu, X.; Si, Z.; Weng, D. Nb-modified Mn/Ce/Ti catalyst for the selective catalytic reduction of NO with NH3 at low temperature. Appl. Catal. A 2017, 545, 64–71. [Google Scholar] [CrossRef]
  144. Li, Y.; Li, G.; Lu, Y.; Hao, W.; Wei, Z.; Liu, J.; Zhang, Y. Denitrification performance of non-pitch coal-based activated coke by the introduction of MnOx–CeOx–M (FeOx, CoOx) at low temperature. Mol. Catal. 2018, 445, 21–28. [Google Scholar] [CrossRef]
  145. Pan, Y.; Shen, Y.; Jin, Q.; Zhu, S. Promotional effect of Ba additives on MnCeOx/TiO2 catalysts for NH3-SCR of NO at low temperature. J. Mater. Res. 2018, 33, 2414–2422. [Google Scholar] [CrossRef]
  146. Wang, X.; Liu, Y.; Ying, Q.; Yao, W.; Wu, Z. The superior performance of Nb-modified Cu-Ce-Ti mixed oxides for the selective catalytic reduction of NO with NH3 at low temperature. Appl. Catal. A 2018, 562, 19–27. [Google Scholar] [CrossRef]
  147. Li, W.; Zhang, C.; Li, X.; Tan, P.; Zhou, A.; Fang, Q.; Chen, G. Ho-modified Mn–Ce/TiO2 for low-temperature SCR of NOx with NH3: Evaluation and characterization. Chin. J. Catal. 2018, 39, 1653–1663. [Google Scholar] [CrossRef]
  148. Lu, P.; Li, R.; Xing, Y.; Li, Y.; Zhu, T.; Yue, H.; Wu, W. Low temperature selective catalytic reduction of NOX with NH3 by activated coke loaded with FexCoyCezOm: The enhanced activity, mechanism and kinetics. Fuel 2018, 233, 188–199. [Google Scholar] [CrossRef]
  149. Zhao, X.; Mao, L.; Dong, G. Mn-Ce-V-WOx/TiO2 SCR catalysts: Catalytic activity, stability and interaction among catalytic oxides. Catalysts 2018, 8, 76. [Google Scholar] [CrossRef]
Figure 1. (a) NOx conversion and (b) N2 selectivity for Va-MnOx, Mn2O3, Mn2V2O7, and reference samples. Reprinted from Reference [52]. Copyright 2018, with Permission from American Chemical Society.
Figure 1. (a) NOx conversion and (b) N2 selectivity for Va-MnOx, Mn2O3, Mn2V2O7, and reference samples. Reprinted from Reference [52]. Copyright 2018, with Permission from American Chemical Society.
Catalysts 09 00349 g001
Figure 2. (a) Scheme of the preparation of triple shelled NiMn2O4 hollow spheres and (b) TEM image of NiMn2O4-S; (c) NOx conversion over the NiMn2O4 catalysts and (d) durability tests of the NiMn2O4-S catalyst at 150 °C. Reaction conditions: [NO] = [NH3] = 500 ppm, [O2] = 5 vol%, [SO2] = 100 ppm (when used), [H2O] = 5 vol% (when used), balanced with Ar, GHSV = 68,000 h‒1. Adapted from Reference [53]. Copyright 2018, with Permission from Royal Society of Chemistry.
Figure 2. (a) Scheme of the preparation of triple shelled NiMn2O4 hollow spheres and (b) TEM image of NiMn2O4-S; (c) NOx conversion over the NiMn2O4 catalysts and (d) durability tests of the NiMn2O4-S catalyst at 150 °C. Reaction conditions: [NO] = [NH3] = 500 ppm, [O2] = 5 vol%, [SO2] = 100 ppm (when used), [H2O] = 5 vol% (when used), balanced with Ar, GHSV = 68,000 h‒1. Adapted from Reference [53]. Copyright 2018, with Permission from Royal Society of Chemistry.
Catalysts 09 00349 g002
Figure 3. The (a) resistance to water vapor poisoning test and (b) resistance to sulfur poisoning test. (Reaction condition: [NO] = [NH3] = 500 ppm, [O2] = 5 vol%, balanced with N2, [H2O] = 5 vol%, [SO2] = 100 ppm, and GHSV = 36,000 h−1). Reprinted from Reference [57]. Copyright 2018, with Permission from Elsevier.
Figure 3. The (a) resistance to water vapor poisoning test and (b) resistance to sulfur poisoning test. (Reaction condition: [NO] = [NH3] = 500 ppm, [O2] = 5 vol%, balanced with N2, [H2O] = 5 vol%, [SO2] = 100 ppm, and GHSV = 36,000 h−1). Reprinted from Reference [57]. Copyright 2018, with Permission from Elsevier.
Catalysts 09 00349 g003
Figure 4. H2O tolerance and SO2 tolerance of the MnOx-CeO2-B hollow nanotube. Reaction conditions: [NOx] = [NH3] = 1000 ppm, [O2] = 5%, N2 as balance gas, and GHSV = 30,000 h−1. Reprinted from Reference [58]. Copyright 2018, with Permission from Elsevier.
Figure 4. H2O tolerance and SO2 tolerance of the MnOx-CeO2-B hollow nanotube. Reaction conditions: [NOx] = [NH3] = 1000 ppm, [O2] = 5%, N2 as balance gas, and GHSV = 30,000 h−1. Reprinted from Reference [58]. Copyright 2018, with Permission from Elsevier.
Catalysts 09 00349 g004
Scheme 1. The proposed mechanism of the NH3-SCR reaction over the Co7Mn3Ox catalyst and the synergetic catalytic effect between Mn and Co cations. Reprinted from Reference [48]. Copyright 2018, with Permission from Elsevier.
Scheme 1. The proposed mechanism of the NH3-SCR reaction over the Co7Mn3Ox catalyst and the synergetic catalytic effect between Mn and Co cations. Reprinted from Reference [48]. Copyright 2018, with Permission from Elsevier.
Catalysts 09 00349 sch001
Figure 5. (a) NO conversion; (b) N2 selectivity over Cux-Nb1.1-x (x = 0.45, 0.35, 0.25, 0.15) as a function of temperature under a GHSV of 35,000 h−1; (c) effect of GHSV on NO conversion over Cu0.25-Nb0.85, and (d) effect of SO2, H2O, and SO2 + H2O on NO conversion over Cu0.25-Nb0.85 at 200 °C under a GHSV of 35,000 h-1. (Reaction conditions: [NO] = [NH3] = 1000 ppm, [O2] = 3% and N2 balance), and Effect of SO2, H2O, and SO2 + H2O on NO conversion over Cu0.25-Nb0.85 at 200 °C under a GHSV of 35,000 h−1. Reprinted from Reference [71]. Copyright 2018, with Permission from Elsevier.
Figure 5. (a) NO conversion; (b) N2 selectivity over Cux-Nb1.1-x (x = 0.45, 0.35, 0.25, 0.15) as a function of temperature under a GHSV of 35,000 h−1; (c) effect of GHSV on NO conversion over Cu0.25-Nb0.85, and (d) effect of SO2, H2O, and SO2 + H2O on NO conversion over Cu0.25-Nb0.85 at 200 °C under a GHSV of 35,000 h-1. (Reaction conditions: [NO] = [NH3] = 1000 ppm, [O2] = 3% and N2 balance), and Effect of SO2, H2O, and SO2 + H2O on NO conversion over Cu0.25-Nb0.85 at 200 °C under a GHSV of 35,000 h−1. Reprinted from Reference [71]. Copyright 2018, with Permission from Elsevier.
Catalysts 09 00349 g005
Figure 6. (a) NO removal efficiency of Fe0.3Mn0.5Zr0.2 as a function of time at 200 °C and (b) effect of SO2 on the NO removal over Fe0.3Mn0.5Zr0.2 at 200 °C. Reprinted from Reference [88]. Copyright 2017, with Permission from Elsevier.
Figure 6. (a) NO removal efficiency of Fe0.3Mn0.5Zr0.2 as a function of time at 200 °C and (b) effect of SO2 on the NO removal over Fe0.3Mn0.5Zr0.2 at 200 °C. Reprinted from Reference [88]. Copyright 2017, with Permission from Elsevier.
Catalysts 09 00349 g006
Figure 7. SO2 resistance of MnTiOx and MnSbTiOx-0.2 catalysts at 150 °C. Reaction conditions: 600 ppm NO, 600 ppm NH3, 5% O2, 5% H2O, 100 ppm SO2, balance Ar, GHSV = 108,000 h−1. Reprinted from Reference [89]. Copyright 2018, with Permission from Elsevier.
Figure 7. SO2 resistance of MnTiOx and MnSbTiOx-0.2 catalysts at 150 °C. Reaction conditions: 600 ppm NO, 600 ppm NH3, 5% O2, 5% H2O, 100 ppm SO2, balance Ar, GHSV = 108,000 h−1. Reprinted from Reference [89]. Copyright 2018, with Permission from Elsevier.
Catalysts 09 00349 g007
Figure 8. (a) NOx conversion and N2 selectivity over MnO2/CoAl-LDO and CoMnAl-LDO catalysts prepared by calcination at 500 °C; and (b) the stability and SO2/H2O resistance test (inset) of MnO2/CoAl-LDO and CoMnAl-LDO catalysts at 240 °C. Reaction conditions: [NH3] = 600 ppm, [NO] = 600 ppm, [O2] = 5 vol%, 100 ppm SO2 (when used), 10 vol% H2O. (when used) balanced by N2. Reprinted from Reference [91]. Copyright 2019, with Permission from Elsevier.
Figure 8. (a) NOx conversion and N2 selectivity over MnO2/CoAl-LDO and CoMnAl-LDO catalysts prepared by calcination at 500 °C; and (b) the stability and SO2/H2O resistance test (inset) of MnO2/CoAl-LDO and CoMnAl-LDO catalysts at 240 °C. Reaction conditions: [NH3] = 600 ppm, [NO] = 600 ppm, [O2] = 5 vol%, 100 ppm SO2 (when used), 10 vol% H2O. (when used) balanced by N2. Reprinted from Reference [91]. Copyright 2019, with Permission from Elsevier.
Catalysts 09 00349 g008
Figure 9. Effect of H2O and SO2 on NO conversion over the Mn0.1Ce0.3TiOx catalyst at 200 °C (1000 ppm NO, 1000 ppm NH3, 3% O2, balance N2, GHSV = 40,000 h−1). Reprinted from Reference [92]. Copyright 2018, with Permission from Elsevier.
Figure 9. Effect of H2O and SO2 on NO conversion over the Mn0.1Ce0.3TiOx catalyst at 200 °C (1000 ppm NO, 1000 ppm NH3, 3% O2, balance N2, GHSV = 40,000 h−1). Reprinted from Reference [92]. Copyright 2018, with Permission from Elsevier.
Catalysts 09 00349 g009
Figure 10. Effect of SO2, H2O, and SO2 + H2O on NO conversion over the Fe0.4-Nb0.1-Ce0.5 catalyst at 220 °C. Reprinted from Reference [93]. Copyright 2018, with Permission from Elsevier.
Figure 10. Effect of SO2, H2O, and SO2 + H2O on NO conversion over the Fe0.4-Nb0.1-Ce0.5 catalyst at 220 °C. Reprinted from Reference [93]. Copyright 2018, with Permission from Elsevier.
Catalysts 09 00349 g010
Figure 11. (a) NO conversions and (b) N2 selectivities of the catalysts in the NH3-SCR reaction as a function of temperature; (c) SO2 (100 ppm) resistance tests, and (d) H2O + SO2 (2.5 vol%, 100 ppm) resistance tests at 200 °C over the catalysts. Reproduced from Reference [94]. Copyright 2018, with Permission from Elsevier.
Figure 11. (a) NO conversions and (b) N2 selectivities of the catalysts in the NH3-SCR reaction as a function of temperature; (c) SO2 (100 ppm) resistance tests, and (d) H2O + SO2 (2.5 vol%, 100 ppm) resistance tests at 200 °C over the catalysts. Reproduced from Reference [94]. Copyright 2018, with Permission from Elsevier.
Catalysts 09 00349 g011
Figure 12. Effects of H2O and/or SO2 on NOx conversion over the Co0.2Ce0.35Mn0.45Ti10 catalyst. Reproduced from Reference [96]. Copyright 2018, with Permission from Elsevier.
Figure 12. Effects of H2O and/or SO2 on NOx conversion over the Co0.2Ce0.35Mn0.45Ti10 catalyst. Reproduced from Reference [96]. Copyright 2018, with Permission from Elsevier.
Catalysts 09 00349 g012
Figure 13. Plausible SCR mechanism over the surface of Mn/TiO2 catalyst. Adapted from Reference [24]. Copyright 2012, with Permission from Elsevier.
Figure 13. Plausible SCR mechanism over the surface of Mn/TiO2 catalyst. Adapted from Reference [24]. Copyright 2012, with Permission from Elsevier.
Catalysts 09 00349 g013
Figure 14. Catalytic evaluation of the Mn(0.25)/TNT-X (X = Hombikat, Ishihara, P25 Degussa, Kemira, Sigma–Aldrich, and Titania oxysulfate) family of catalyst for the SCR of NOx by NH3, in the presence of 900 ppm NO, 100 ppm NO2, 1000 ppm NH3, 10 vol% O2 with He balance under a GHSV of 50,000 h‒1 in the temperature range from 100–300 °C. Reproduced from Reference [26]. Copyright 2016, with Permission from Elsevier.
Figure 14. Catalytic evaluation of the Mn(0.25)/TNT-X (X = Hombikat, Ishihara, P25 Degussa, Kemira, Sigma–Aldrich, and Titania oxysulfate) family of catalyst for the SCR of NOx by NH3, in the presence of 900 ppm NO, 100 ppm NO2, 1000 ppm NH3, 10 vol% O2 with He balance under a GHSV of 50,000 h‒1 in the temperature range from 100–300 °C. Reproduced from Reference [26]. Copyright 2016, with Permission from Elsevier.
Catalysts 09 00349 g014
Figure 15. Influence of inlet water concentrations (10 vol%) on NOx conversion in the SCR reaction over Mn(0.25)/TNT-H catalyst at 140 °C; feed: NO = 900 ppm, NO2 = 100 ppm, NH3/NOx (ANR) = 1.0, O2 = 10 vol%, He carrier gas, GHSV = 50,000 h‒1. Reproduced from Reference [26]. Copyright 2016, with Permission from Elsevier.
Figure 15. Influence of inlet water concentrations (10 vol%) on NOx conversion in the SCR reaction over Mn(0.25)/TNT-H catalyst at 140 °C; feed: NO = 900 ppm, NO2 = 100 ppm, NH3/NOx (ANR) = 1.0, O2 = 10 vol%, He carrier gas, GHSV = 50,000 h‒1. Reproduced from Reference [26]. Copyright 2016, with Permission from Elsevier.
Catalysts 09 00349 g015
Figure 16. Catalytic activity evaluation of metal oxides confined titania (made of Hombikat titania) nanotube catalytic formulations M/TNT where M = Mn, Cu, Ce, Fe, V, Cr, and Co for the selective catalytic reduction of NOx by NH3 in the presence of 900 ppm NO, 100 ppm NO2, 1000 ppm NH3, 10 vol% O2 in He balance, under a GHSV = 50,000 h‒1. Reprinted from Reference [109]. Copyright 2018, with Permission from Elsevier.
Figure 16. Catalytic activity evaluation of metal oxides confined titania (made of Hombikat titania) nanotube catalytic formulations M/TNT where M = Mn, Cu, Ce, Fe, V, Cr, and Co for the selective catalytic reduction of NOx by NH3 in the presence of 900 ppm NO, 100 ppm NO2, 1000 ppm NH3, 10 vol% O2 in He balance, under a GHSV = 50,000 h‒1. Reprinted from Reference [109]. Copyright 2018, with Permission from Elsevier.
Catalysts 09 00349 g016
Figure 17. (a) NOx conversion as a function of temperature over different catalysts and (b) SO2/H2O tolerance of the Fe2O3/CNTs catalyst. Reprinted from Reference [116]. Copyright 2015, with Permission from Royal Society of Chemistry.
Figure 17. (a) NOx conversion as a function of temperature over different catalysts and (b) SO2/H2O tolerance of the Fe2O3/CNTs catalyst. Reprinted from Reference [116]. Copyright 2015, with Permission from Royal Society of Chemistry.
Catalysts 09 00349 g017
Figure 18. N2 selectivity and catalytic performance of Mn-M′/TiO2 anatage (M′ = Cr, Fe, Co, Ni, Cu, Zn, Zr, and Ce) catalysts: NH3 = 400 ppm; NO = 400 ppm; O2 = 2.0 vol%; GHSV = 50,000 h−1; catalyst wt. = 0.1 g; reaction temperature = 200 °C. Reproduced from Reference [22]. Copyright 2011, with Permission from Elsevier.
Figure 18. N2 selectivity and catalytic performance of Mn-M′/TiO2 anatage (M′ = Cr, Fe, Co, Ni, Cu, Zn, Zr, and Ce) catalysts: NH3 = 400 ppm; NO = 400 ppm; O2 = 2.0 vol%; GHSV = 50,000 h−1; catalyst wt. = 0.1 g; reaction temperature = 200 °C. Reproduced from Reference [22]. Copyright 2011, with Permission from Elsevier.
Catalysts 09 00349 g018
Figure 19. (a) Influence of Ni/Mn atomic ratio on NO conversion in the SCR reaction at a temperature range (160–240 °C) over Mn-Ni/TiO2 catalysts (XNO% = conversion of NO at 6 h on stream); (b) SCR of NO with NH3 at 200 °C temperature over Mn/TiO2 and Mn-Ni/TiO2 catalysts; (c) Influence of inlet water concentrations (10 vol%) on NO conversion in the SCR reaction over Mn–Ni(0.4)/TiO2 catalyst at 200 °C (GHSV = 50,000 h‒1; feed: NO = 400 ppm, NH3 = 400 ppm, O2 = 2 vol%, He carrier gas, catalyst = 0.1 g, total flow = 140 mL min‒1). Reprinted from Reference [5]. Copyright 2012, with Permission from Elsevier.
Figure 19. (a) Influence of Ni/Mn atomic ratio on NO conversion in the SCR reaction at a temperature range (160–240 °C) over Mn-Ni/TiO2 catalysts (XNO% = conversion of NO at 6 h on stream); (b) SCR of NO with NH3 at 200 °C temperature over Mn/TiO2 and Mn-Ni/TiO2 catalysts; (c) Influence of inlet water concentrations (10 vol%) on NO conversion in the SCR reaction over Mn–Ni(0.4)/TiO2 catalyst at 200 °C (GHSV = 50,000 h‒1; feed: NO = 400 ppm, NH3 = 400 ppm, O2 = 2 vol%, He carrier gas, catalyst = 0.1 g, total flow = 140 mL min‒1). Reprinted from Reference [5]. Copyright 2012, with Permission from Elsevier.
Catalysts 09 00349 g019
Figure 20. Influence of inlet water concentrations (7 vol%) on NOx conversion in the SCR reaction over Mn–Ce(5.1)/TiO2-Hk (Hombikat) catalyst at 175 °C; feed: NO = 900 ppm, NO2 = 100 ppm NH3/NOx (ANR) = 1.0, O2 = 10 vol%, He carrier gas, catalyst. 0.08 g, GHSV. 80,000 h‒1. Reprinted from Reference [25]. Copyright 2015, with Permission from Elsevier.
Figure 20. Influence of inlet water concentrations (7 vol%) on NOx conversion in the SCR reaction over Mn–Ce(5.1)/TiO2-Hk (Hombikat) catalyst at 175 °C; feed: NO = 900 ppm, NO2 = 100 ppm NH3/NOx (ANR) = 1.0, O2 = 10 vol%, He carrier gas, catalyst. 0.08 g, GHSV. 80,000 h‒1. Reprinted from Reference [25]. Copyright 2015, with Permission from Elsevier.
Catalysts 09 00349 g020
Figure 21. (a) Catalytic activity of Ce-Mn/TiO2 catalyst for NH3–SCR. Catalysts were loaded with 10%, 20% and 30% Ce and denoted as Ce(10), Ce(20) and Ce(30), respectively. Pure TiO2 was also used for comparison; (b) The effect of various Ce concentrations using a Ce-Mn/TiO2 catalyst on SO2 resistance; (c) The effects of reaction temperature on NO conversion of the Ce(20)-Mn/TiO2 catalyst in the presence of SO2. The above three types of reactions were performed at: 500 ppm NO, 500 ppm NH3, SO2 100 ppm, 3% O2, N2 balance. gas, GHSV = 10 000 h−1; and (d) the effects of SO2 concentration on NO conversion of Ce(20)-Mn/TiO2 catalysts (T = 180°C, 500 ppm NO, 500 ppm NH3, 3%O2, N2 balance gas, GHSV = 10 000 h−1). Adapted from Reference [127]. We thank the Royal Society Open Science for this Contribution.
Figure 21. (a) Catalytic activity of Ce-Mn/TiO2 catalyst for NH3–SCR. Catalysts were loaded with 10%, 20% and 30% Ce and denoted as Ce(10), Ce(20) and Ce(30), respectively. Pure TiO2 was also used for comparison; (b) The effect of various Ce concentrations using a Ce-Mn/TiO2 catalyst on SO2 resistance; (c) The effects of reaction temperature on NO conversion of the Ce(20)-Mn/TiO2 catalyst in the presence of SO2. The above three types of reactions were performed at: 500 ppm NO, 500 ppm NH3, SO2 100 ppm, 3% O2, N2 balance. gas, GHSV = 10 000 h−1; and (d) the effects of SO2 concentration on NO conversion of Ce(20)-Mn/TiO2 catalysts (T = 180°C, 500 ppm NO, 500 ppm NH3, 3%O2, N2 balance gas, GHSV = 10 000 h−1). Adapted from Reference [127]. We thank the Royal Society Open Science for this Contribution.
Catalysts 09 00349 g021
Figure 22. (a) NH3-SCR activities of the fresh and sulfated catalysts; and (b) SO2 tolerances of Mn/TiO2 and MnEu/TiO2 in NH3-SCR reaction. Reprinted from Reference [131]. Copyright 2018, with Permission from Elsevier.
Figure 22. (a) NH3-SCR activities of the fresh and sulfated catalysts; and (b) SO2 tolerances of Mn/TiO2 and MnEu/TiO2 in NH3-SCR reaction. Reprinted from Reference [131]. Copyright 2018, with Permission from Elsevier.
Catalysts 09 00349 g022
Figure 23. (a) SCR activities and (b) N2 selectivities over different catalyst samples as a function of reaction temperature. Reaction conditions: [NO] = [NH3] = 600 ppm, [O2] = [H2O] = 5%, balance Ar, GHSV = 108,000 h‒1; (c) Effect of SO2 on the SCR activities over Mn/TiO2 and MnNb/TiO2-0.12 catalyst samples. Reaction conditions: [NO] = [NH3] = 600 ppm, [O2] = [H2O] = 5%, [SO2] = 100 ppm, balance Ar, GHSV = 108,000 h‒1, reaction temperature = 150 °C. Reproduced from Reference [132]. Copyright 2018, with Permission from Elsevier.
Figure 23. (a) SCR activities and (b) N2 selectivities over different catalyst samples as a function of reaction temperature. Reaction conditions: [NO] = [NH3] = 600 ppm, [O2] = [H2O] = 5%, balance Ar, GHSV = 108,000 h‒1; (c) Effect of SO2 on the SCR activities over Mn/TiO2 and MnNb/TiO2-0.12 catalyst samples. Reaction conditions: [NO] = [NH3] = 600 ppm, [O2] = [H2O] = 5%, [SO2] = 100 ppm, balance Ar, GHSV = 108,000 h‒1, reaction temperature = 150 °C. Reproduced from Reference [132]. Copyright 2018, with Permission from Elsevier.
Catalysts 09 00349 g023
Figure 24. SO2 tolerance of Mn/TiO2 and MnMo/TiO2-0.04 catalysts at 150 °C, Reaction conditions: 600 ppm NO, 600 ppm NH3, 100 ppm SO2, 5% O2, balance Ar, GHSV = 108,000 h‒1. Reprinted from Reference [133]. Copyright 2018, with Permission from Elsevier.
Figure 24. SO2 tolerance of Mn/TiO2 and MnMo/TiO2-0.04 catalysts at 150 °C, Reaction conditions: 600 ppm NO, 600 ppm NH3, 100 ppm SO2, 5% O2, balance Ar, GHSV = 108,000 h‒1. Reprinted from Reference [133]. Copyright 2018, with Permission from Elsevier.
Catalysts 09 00349 g024
Figure 25. Low-temperature NH3-SCR reaction mechanism on Fe-Mn/SBA-15 catalyst. Adapted from Reference [136]. Copyright 2018, with Permission from American Chemical Society.
Figure 25. Low-temperature NH3-SCR reaction mechanism on Fe-Mn/SBA-15 catalyst. Adapted from Reference [136]. Copyright 2018, with Permission from American Chemical Society.
Catalysts 09 00349 g025
Figure 26. (a) NH3-SCR activity and N2 selectivity as a function of temperature from 150 °C to 330 °C; and (b) SO2/H2O resistance test of CuAlOx/CNTs catalyst at 240 °C. Reaction conditions: 600 ppm NH3, 600 ppm NO, 5.0 vol% O2, 100 ppm SO2 (when used), 10 vol% H2O (when used) balanced by N2 with a GHSV was 45,000 h−1. Reprinted from Reference [139]. Copyright 2019, with Permission from Elsevier.
Figure 26. (a) NH3-SCR activity and N2 selectivity as a function of temperature from 150 °C to 330 °C; and (b) SO2/H2O resistance test of CuAlOx/CNTs catalyst at 240 °C. Reaction conditions: 600 ppm NH3, 600 ppm NO, 5.0 vol% O2, 100 ppm SO2 (when used), 10 vol% H2O (when used) balanced by N2 with a GHSV was 45,000 h−1. Reprinted from Reference [139]. Copyright 2019, with Permission from Elsevier.
Catalysts 09 00349 g026
Figure 27. (a) NH3-SCR activities of different catalysts {Reaction conditions: [NH3] = [NO] = 600 ppm, [O2] = 3%, [H2O] = 5 vol%, N2 as balance. GHSV = 40 000 h−1}; and (b) SO2 and H2O resistance of the catalysts at 250 °C in SCR reaction process {Reaction conditions: [NH3] = [NO] = 600 ppm, [O2] = 3%, [SO2] = 50 ppm, [H2O] = 5 vol%, N2 balance, GHSV = 40 000 h−1}. Reproduced from Reference [146]. Copyright 2018, with Permission from Elsevier.
Figure 27. (a) NH3-SCR activities of different catalysts {Reaction conditions: [NH3] = [NO] = 600 ppm, [O2] = 3%, [H2O] = 5 vol%, N2 as balance. GHSV = 40 000 h−1}; and (b) SO2 and H2O resistance of the catalysts at 250 °C in SCR reaction process {Reaction conditions: [NH3] = [NO] = 600 ppm, [O2] = 3%, [SO2] = 50 ppm, [H2O] = 5 vol%, N2 balance, GHSV = 40 000 h−1}. Reproduced from Reference [146]. Copyright 2018, with Permission from Elsevier.
Catalysts 09 00349 g027
Figure 28. (A) Selective catalytic reduction (SCR) activity of Mn-Ce-V-WOx/TiO2 composite catalysts with molar ratio of active components/TiO2 at different values; (a) 0.1; (b) 0.2; (c) 0.3; (d) 0.6; (B) SCR activity of V2O5/TiO2, WO3/TiO2, MnO2/TiO2, CeO2/TiO2 and TiO2. Reaction conditions: [NO] = [NH3] = 1500 ppm, O2 = 3%, gas hourly space velocity (GHSV) = 40,000 h‒1; and (C) the lifetime of Mn-Ce-V-WOx/TiO2 catalyst with molar ratio of 0.2 at 250 °C: inset (a-c) H2O and SO2 resistance at 250 °C. Reaction conditions: [NO] = [NH3] = 1500 ppm, [O2] = 3%, [H2O] = 5%, [SO2] = 100 ppm, GHSV = 40,000 h‒1. Adapted from Reference [149].
Figure 28. (A) Selective catalytic reduction (SCR) activity of Mn-Ce-V-WOx/TiO2 composite catalysts with molar ratio of active components/TiO2 at different values; (a) 0.1; (b) 0.2; (c) 0.3; (d) 0.6; (B) SCR activity of V2O5/TiO2, WO3/TiO2, MnO2/TiO2, CeO2/TiO2 and TiO2. Reaction conditions: [NO] = [NH3] = 1500 ppm, O2 = 3%, gas hourly space velocity (GHSV) = 40,000 h‒1; and (C) the lifetime of Mn-Ce-V-WOx/TiO2 catalyst with molar ratio of 0.2 at 250 °C: inset (a-c) H2O and SO2 resistance at 250 °C. Reaction conditions: [NO] = [NH3] = 1500 ppm, [O2] = 3%, [H2O] = 5%, [SO2] = 100 ppm, GHSV = 40,000 h‒1. Adapted from Reference [149].
Catalysts 09 00349 g028

Share and Cite

MDPI and ACS Style

Damma, D.; Ettireddy, P.R.; Reddy, B.M.; Smirniotis, P.G. A Review of Low Temperature NH3-SCR for Removal of NOx. Catalysts 2019, 9, 349. https://doi.org/10.3390/catal9040349

AMA Style

Damma D, Ettireddy PR, Reddy BM, Smirniotis PG. A Review of Low Temperature NH3-SCR for Removal of NOx. Catalysts. 2019; 9(4):349. https://doi.org/10.3390/catal9040349

Chicago/Turabian Style

Damma, Devaiah, Padmanabha R. Ettireddy, Benjaram M. Reddy, and Panagiotis G. Smirniotis. 2019. "A Review of Low Temperature NH3-SCR for Removal of NOx" Catalysts 9, no. 4: 349. https://doi.org/10.3390/catal9040349

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