Next Article in Journal
Efficient Removal of Eriochrome Black T (EBT) Dye and Chromium (Cr) by Hydrotalcite-Derived Mg-Ca-Al Mixed Metal Oxide Composite
Next Article in Special Issue
Materials Design for N2O Capture: Separation in Gas Mixtures
Previous Article in Journal
Nickel Nanoparticles Decorated on Glucose-Derived Carbon Spheres as a Novel, Non-Palladium Catalyst for Epoxidation of Olefin
Previous Article in Special Issue
Assessment of a Euro VI Step E Heavy-Duty Vehicle’s Aftertreatment System
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Phosphotungstic Acid-Modified MnOx for Selective Catalytic Reduction of NOx with NH3

1
Shanghai Key Laboratory of Atmospheric Particle Pollution and Prevention (LAP3), Department of Environmental Science and Engineering, Fudan University, Shanghai 200433, China
2
Research Institute of Applied Catalysis, School of Chemical and Environmental Engineering, Shanghai Institute of Technology, Shanghai 201418, China
3
Shanghai Institute of Pollution Control and Ecological Security, Shanghai 200092, China
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(10), 1248; https://doi.org/10.3390/catal12101248
Submission received: 28 September 2022 / Revised: 10 October 2022 / Accepted: 12 October 2022 / Published: 16 October 2022
(This article belongs to the Special Issue Catalytic Removal and Resource Utilization of NOx)

Abstract

:
H3PW12O40-modified MnOx catalysts (denoted as Mn-HPW) were used for NOx elimination with co-fed NH3. The optimal Mn-HPW0.02 catalyst exhibited over 90% NOx conversion at 90–270 °C. The incorporation of HPW increased the amount of Lewis acid sites of the catalyst for adsorbing NH3, and accelerated the reaction between the adsorbed NH3 species and gas-phase NOx, thus, increasing the low-temperature catalytic activity. The oxidation ability of the Mn catalyst was decreased due to the addition of HPW, thus, mitigating the overoxidation of the adsorbed NH3 species and improving the de-NOx activity and N2 selectivity in the high-temperature region. DRIFT results revealed that the NH3 species on Lewis and Brønsted acid sites, bridged nitrate, and bidentate nitrate were important species/intermediates for the reaction. NH3-SCR over the Mn and Mn-HPW0.02 catalysts obeyed the Eley–Rideal and Langmuir–Hinshelwood mechanisms, simultaneously, at 120 °C.

1. Introduction

Nitrogen oxides (NOx) contribute to environmental problems, including photochemical smog and acid rain. Moreover, NOx facilitates the formation of secondary aerosols and fine particulate matters [1]. Hence, it is important to use efficient technology to eliminate NOx. Among various de-NOx technologies, selective catalytic reduction of NOx using NH3 (NH3-SCR) has been employed extensively, and its decisive factor is the performance of catalysts [2,3,4,5].
Mn-based catalysts have been researched extensively for NH3-SCR [6,7,8,9]. However, MnOx usually has low surface areas and can achieve high NOx conversion only in narrow temperature windows [5,10]. In order to improve the catalytic efficiency, MnOx can be deposited onto TiO2 [11,12,13], Al2O3 [14,15,16], carbon [17,18,19], and zeolites [20,21,22] that can enhance the dispersion of MnOx and lead to synergistic effects between MnOx and the supports. Additionally, the SCR activity of MnOx is promoted by modification with metal oxides, such as CeOx [23,24,25,26], FeOx [27,28,29], CoOx [9,30,31], NiOx [32,33,34], and WOx [35,36]. For instance, Xu et al. [35] loaded MnOx-CeO2 onto WO3-ZrO2, and the optimal catalyst exhibited above 80% NO conversion at 150–380 °C. The de-NOx activity, N2 selectivity, and SO2 tolerance on MnFe/Ti can be enhanced by doping W [36]. The generation of Mn-O-W and Fe-O-W bonds facilitates the transport of electrons and oxygen, thus, accelerating the SCR reaction. Liu et al. [37] prepared ternary (Mn-Ce-Ti) and binary (Mn-Ti and Ce-Ti) catalysts using a hydrothermal method. The results of de-NOx activity evaluation illustrate that Mn-Ce-Ti exhibits excellent NH3-SCR activity and a strong resistance against H2O and SO2, which is mainly attributed to the dual redox cycle (Mn4+ + Ce3+ ↔ Mn3+ + Ce4+, Mn4+ + Ti3+ ↔ Mn3+ + Ti4+) and the amorphous structure. Additionally, metal–organic frameworks (MOFs)-derived metal oxides possess high porosity and large specific surface areas, profiting from the introduction and dispersion of Bronsted acid sites, thus, enhancing the SCR activity of the Mn-based catalysts [30,38,39,40].
Tungstophosphoric acid (H3PW12O40, HPW) is composed of bridging metal atoms and oxygen atoms, and exhibits good redox properties, strong acidity, and good thermal stability, which is favorable for the SCR reaction. Thus, HPW-modified SCR catalysts such as HPW/Fe2O3 [41,42,43,44], HPW/CeO2 [45,46,47], HPW-Cu-Ti [48], HPW-Fe-Ti [48], HPW-V2O5/TiO2 [49], and HPW-Mn-Ce-Co [50] have been developed. For instance, Wu et al. [44] prepared HPW/Fe2O3 by impregnation, and they found that 100% NO conversion at 240–460 °C can be attained on HPW/Fe2O3-350-0.5. The introduction of HPW enriches the number of acid sites, thus, enhancing NH3 adsorption. Meanwhile, the oxidation ability of Fe2O3 is decreased by adding HPW, thereby inhibiting the overoxidation of NH3. Geng et al. [46] developed HPW/CeO2, and they reported that HPW could allow for NH3 adsorption, while CeO2 could activate the adsorbed NH3 and NO species. Thus, superior catalytic activity was obtained on HPW/CeO2. Tang et al. [50] discovered that adsorption and activation of NH3 were boosted with the introduction of HPW to Mn-Ce-Co-O, and HPW can weaken the influence of SO2 on active bidentate nitrate, and restrain the formation of metal sulfate, thereby achieving good SO2 resistance of the catalyst. However, to the best of our knowledge, HPW-modified MnOx for NH3-SCR was not reported.
Herein, Mn-HPW composite catalysts were prepared and the effects of adding HPW to the catalytic performance were evaluated. Furthermore, the correlation between SCR performance and physicochemical properties was researched in detail. The evolution process of NH3- and NOx-derived species over the Mn and Mn-HPW0.02 catalysts was monitored at 60–300 °C using DRIFT, and the possible reaction mechanisms at 120 °C were revealed.

2. Results and Discussion

2.1. Textural and Structural Properties

2.1.1. XRD

XRD patterns of the Mn and Mn-HPWx samples are given in Figure 1. The Mn catalyst shows MnO2 (PDF # 44-0141), Mn5O8 phases (PDF # 39-1218), and Mn2O3 (PDF # 41-1442) [51,52,53], corresponding to the standard pattern in Figure S1. As shown in Figure S2, HPW calcinated at 400 °C shows the peaks of H3PW12O40·6H2O (PDF # 50-0304), coinciding with the phenomenon reported by Geng et al. [46] and Wu et al. [44]. For all Mn-HPWx catalysts, the peak intensities of Mn2O3 and Mn5O8 decline gradually with the increase of x, and the characteristic peak of HPW is not detected due to the lower content [47]. However, as the x is higher than 0.01, weak peaks of the WO3 species (PDF # 41-1315) and the MnWO4 phase (PDF # 13-0434) are detected, due to the partial decomposition of HPW during calcination [54,55]. Characteristic peaks of Mn-HPW0.02, calcinated at different temperatures, are shown in Figure S3, and the phases of MnWO4 and WO3 are distinct in Mn-HPW0.02 calcinated at 650 °C.

2.1.2. FTIR

FTIR spectra of HPW calcinated at different temperatures are shown in Figure 2. The evident bands of Keggin anion vibration appear at 1080, 980, 890, and 795 cm−1 on HPW before calcination and HPW calcinated at 400 °C (Figure 2 in (a,b), and these bands can be ascribed to P-Oa (1080 cm−1) in the central tetrahedra PO4, W=Od (980 cm−1) in terminal oxygens linked to a lone tungsten atom, W-Ob-W (890 cm−1) connecting two different W3O13 groups, and the W-Oc-W (795 cm−1) vibrations of the same W3O13 groups, respectively [54,55]. However, the bands of W=Od, W-Ob-W, and W-Oc-W are not detected in HPW calcinated 650 °C (Figure 2 in (c), due to the destruction of the HPW structure. For Mn-HPW0.02 before calcination, Mn-HPW0.02 calcinated at 200 °C, and Mn-HPW0.02 calcined at 400 °C (Figure 2 in (d–f), the characteristic bands of PO4, W=Od, and W-Ob-W are evident, although the positions migrate to lower wavenumbers. Especially, the bands of W-Oc-W are invisible. These phenomena reveal that the structure of HPW is influenced, to some extent, during the preparation. Mn-HPW0.02 calcinated at 650 °C shows no obvious bands of HPW, indicating the decomposition of HPW at 650 °C (Figure 2 in (g).

2.1.3. N2 Adsorption-Desorption

The textural properties of Mn, Mn-HPWx, and HPW are summarized in Table 1. The specific BET surface areas (SBET) of HPW and the Mn catalyst are 1.6 and 33.7 m2/g, respectively. The SBET of Mn-HPWx increases with x, reaching a maximum (69.1 m2/g) at x = 0.02, and then decreases when the x increases further. The value of the total pore volume increases only by 0.03–0.04 cm3/g as x is less than 0.02, while the total pore volume (0.14 cm3/g) is equal to the value of the Mn catalyst when x = 0.02–0.04. These data reveal that the influence of introducing HPW on the total pore volume over the Mn catalyst is negligible. For the average pore diameter over various samples, with the increase in the HPW content, the value first decreases and then increases after reaching the minimum on Mn-HPW0.02, which is the opposite trend of the value of SBET. N2 adsorption/desorption isotherms are shown in Figure S4. All the catalysts exhibit a representative IV-type isotherm, accompanied by an obvious H3-type hysteretic loop. The adsorption curve presents an upward tendency as the p/p0 is higher than 0.6, revealing the existence of mesopores in all investigated catalysts. The ascending trend of the curves is obvious with p/p0 higher than 0.8, meaning the presence of the macropores [56,57]. In the p/p0 range of 0.6–0.8, the N2 adsorption capacities on Mn-HPWx (x = 0.005–0.02) are larger than those on Mn, and the hysteretic loop is clearer, suggesting that the formation of more mesopores is facilitated by adding HPW, thus, increasing the values of SBET and decreasing the average pore diameter. As the amount of HPW is higher than 0.02, the N2 adsorption capacities with the p/p0 higher than 0.8 are enhanced, demonstrating the formation of more macropores, thus, enlarging the average pore diameter.

2.1.4. Morphology Structure

The SEM images of HPW, Mn, and Mn-HPW are exhibited in Figure 3. HPW presents a large bulk (Figure 3e), thus, leading to the small value of SBET (coinciding with the results of Table 1). The rough, irregular, and agglomerated spherical particles can be observed on Mn (Figure 3b) and Mn-HPW0.02 (Figure 3c). Additionally, for Mn-HPW0.02, the combination of HPW and Mn(NO3)2 can restrain the thermal growth of MnOx and HPW during the calcination process, inducing the lower crystallinity of MnOx and HPW (evidenced by the results of XRD), the smaller particles in Figure 3c, and the larger value of SBET. The EDX elemental mapping of the Mn catalyst in Figure S5a–c displays that the Mn and O elements distribute well. The same situation is also obtained on Mn-HPW0.02, in which W and P atoms are well dispersed (Figure S5d–h). TEM images of HPW, Mn, and Mn-HPW0.02 are shown in Figure 4. The clear lattice fringes (d = 0.35 nm) of HPW can be observed on the HPW catalyst (Figure 4b). The lattice fringes of Mn2O3 and Mn5O8 are shown on the Mn catalyst (Figure 4d). For Mn-HPW0.02 in Figure 4f,h, the lattice fringes of WO3 and MnWO4 are detected, and these are in accordance with the results of XRD.

2.2. Catalytic Performance

Initially, the effect of phosphomolybdic acid (HPMo), tungstosilicic acid (HSiW), and silicomolybdic acid (HSiMo), on the catalytic performance of the Mn catalyst for the NH3-SCR reaction was evaluated at 60–300 °C (Figure S6), and the molar ratio of Mn to heteropoly acid was set as 1 to 0.02. Among all investigated catalysts, Mn-HPW exhibits the broadest temperature window with 90–270 °C (>90% NOx conversion). Therefore, the Mn-HPW catalyst is chosen for further investigation in this study.
Figure 5 shows the de-NOx activities of Mn-HPWx catalysts with a variety of molar ratios of HPW to Mn. HPW exhibits ~15% de-NOx activity at 60–300 °C in Figure 5a. Mn shows higher activity at 60–300 °C, and 98% NOx can be converted at 150 °C. The SCR activities on Mn-HPWx catalysts are improved. For instance, at 60–180 °C (Figure 5b), the NOx conversion enhances monotonously as the HPW/Mn molar ratio increases from 0.005 to 0.02 and then drops as the HPW/Mn molar ratio increases further. The maximum SCR activity can be obtained on Mn-HPW0.02, in which case 96% NOx conversion can be achieved at 90–180 °C. At 180–300 °C (Figure 5a), the activities on Mn-HPWx are much higher than those on Mn, and the amount of doping HPW has a negligible effect on NOx conversion. Based on the above results, Mn-HPW0.02 exhibits optimal de-NOx activities. Additionally, Mn-HPW0.02 exhibits higher N2 selectivity (Figure S7) and better SO2 resistance (Figure S8).
The de-NOx activity over the Mn catalyst can be enhanced by the modification with metal oxides, such as NbOx [58], CoOx [30], SmOx [53], FeOx [28], EuOx [59], CuOx [60], NiOx [61], and CeOx [62,63]. Table S1 lists some representative Mn-based catalysts for NH3-SCR. For instance, more than 90% NOx can be converted at 125–350 °C on MnxCo3-xO4 nanocages under relatively milder conditions (0.40 g catalyst, (NO) = 500 ppm, flow rate = 210 mL/min) [30]. Above 90% NOx at 125–225 °C can be achieved on (Cu1.0Mn2.0)1-δO4 under conditions (0.15 g catalyst, (NO) = 500 ppm, flow rate = 200 mL/min) [60] similar to the conditions in our work. For comparison, Mn-HPW0.02 in our work showed > 90% NOx conversion at 90–270 °C (0.15 g catalyst, (NO) = 500 ppm, flow rate = 200 mL/min). Hence, this catalyst was selected for further research.

2.3. NO and NH3 Oxidation

The oxidation of NO to NO2 is considered to be an essential reaction in the fast SCR reaction (2NH3 + NO + NO2 → 2N2 + 3H2O). Thereby, catalytic NO oxidation over the Mn and Mn-HPW0.02 catalysts was performed (Figure 6a). The Mn catalyst presents a relatively high conversion of NO to NO2 at 60–300 °C, while NO oxidation can be suppressed on Mn-HPW0.02, illustrating the oxidation capacity is impaired on the Mn catalyst by doping HPW. Although the conversion for NO oxidation on Mn-HPW0.02 is lower than that on Mn, the NOx conversion on Mn-HPW0.02 is much higher than that on Mn at low temperatures (Figure 5a), demonstrating that the improvement of the de-NOx activity is not closely connected with fast SCR reaction over Mn-HPW0.02 in the low-temperature region, and similar situations associated with Mn-Nb [58] and Co-Ce-Mn-Ti catalysts [64] were also reported.
The decrease in NH3-SCR activities at high temperatures has to do with the oxidation of NH3. Thus, catalytic NH3 oxidation on Mn and Mn-HPW0.02 at 60–300 °C was studied (Figure 6b). NH3 oxidation can be ignored below 150 °C, and no by-products, such as N2O, NO, and NO2, can be detected in Figure S7. As the temperature is more than 150 °C, NH3 conversion on both catalysts increases rapidly and reaches ~98% at 210 °C (Figure 6b). The yield of the by-products on Mn-HPW0.02 is much less than those on Mn at 180–270 °C in Figure S9. Obviously, the introduction of HPW can undermine the unselective oxidation of NH3, promoting the conversion of NH3 to N2 (4NH3 + 3O2 → 2N2 + 6H2O), and improving the N2 selectivity.

2.4. H2-TPR

The H2-TPR data of the Mn, Mn-HPWx, and HPW catalysts are shown in Figure 7. For HPW calcinated at 400 °C, no significant reduction peak of the tungsten species can be observed at 100–650 °C, because the decomposition temperature of HPW to WO3 is 597 °C and the reduction temperature of WO3 species is 828 °C [43,46]. The Mn catalyst shows two major reduction peaks at 150–360 °C (peak I) and 360–550 °C (peak II). Peak I is linked to the reduction of MnO2 to Mn2O3, whereas peak II is involved in the reduction of Mn2O3 to Mn3O4 and Mn3O4 to MnO [25,50,65,66]. The reduction behavior changes, evidently, with the introduction of HPW as shown in Table S2. As the doping amount of HPW is increased, the areas of two reduction peaks decrease (Table S2). Meanwhile, the reduction temperature of the peak II peak shifts towards the higher temperature due to the suppression of electronic transmission between gaseous H2 and MnOx after adding HPW, and Geng et al. reported a similar situation [43]. It can be concluded that the addition of HPW weakens the reducibility of the MnOx, consistent with the results of NO oxidation (Figure 4a) and NH3 oxidation (Figure 4b). The moderate reducibility can suppress NH3 unselective catalytic oxidization, thus, enhancing the conversion of NOx and lessening the formation of by-product N2O for NH3-SCR [44,50,58]. In this case, Mn-HPW0.02 with moderate reducibility shows the optimal conversion of NOx and the higher N2 selectivity in SCR.

2.5. XPS

XPS spectra of Mn 2p, W 4f, P 2p, and O 1s on Mn and Mn-HPWx are illustrated in Figure 8. In Figure 8a, the peaks of Mn 2p3/2 and Mn 2p1/2 are inspected at ~642.0 eV and ~654.0 eV, respectively, and the Mn 2p3/2 spectra are deconvoluted and fitted by Gaussian-Lorentz into three subpeaks, including Mn2+ (640.8 ± 0.1 eV), Mn3+ (642.0 ± 0.2 eV), and Mn4+ (643.5 ± 0.1 eV) [63,67,68]. The W 4f spectra on Mn-HPWx in Figure 8b can be divided into two W6+ peaks at 35.3 ± 0.1 eV (W 4f7/2) and 37.4 ± 0.1 eV (W 4f5/2), respectively [44,69,70]. A broad peak at 133.5 ± 0.1 eV is attributed to P5+ in the P 2p spectra (Figure 8c) [54,71]. The surface atomic concentration of Mn element over Mn-HPWx decreases with the increase in the content of HPW in Table 2, while the W and P concentrations enrich significantly. It is recognized that the Mn species with different valence states can influence the electron transfer and redox characteristics of the catalysts. For instance, Mn4+ and Mn3+ species are considered the active sites to adsorb and activate the reactants during NH3-SCR, and these species exhibit excellent redox properties, playing a vital role in the redox reaction (Mn3+ ↔ Mn4+ and Mn3+ ↔ Mn2+) [52,66]. The relative concentrations of Mn4+ and Mn3+ on Mn and Mn-HPWx are listed in Table 2 and they decline gradually with the increase in HPW content owing to the electron transfer between W6+ in HPW and Mn4+/Mn3+ [36,50]. Since the high valence Mn species tend to facilitate the deep dehydrogenation of the adsorbed NH3 species [50], the moderate decrease in the concentrations of Mn4+ species appropriately weakens the redox ability of MnOx after adding HPW (identical to the results of H2-TPR, Figure 7), and then alleviates the over-oxidation of NH3 (Figure 6b) and increases the N2 selectivity [72].
Figure 8d displays the O 1s spectra of Mn-HPWx. All curves involve two main peaks at 530.1 ± 0.3 eV (lattice oxygen (Oα) and 531.3 ± 0.2 eV (surface chemisorbed oxygen (Oβ)) [72,73,74]. The binding energy of Oα (529.8 eV) on the Mn catalyst moves to higher energy (530.4 eV) with the increase in the amount of HPW. This is related to the fact that HPW is composed of bridging metal atoms and oxygen atoms, and the lattice oxygen atoms in HPW are different from those in MnOx [50,54]. From Table 2, increasing the HPW amount allows the oxygen atomic concentrations to enhance, whereas the relative concentrations of Oβ with high mobility fluctuate slightly in the range of 17–18%, disclosing that the Oβ concentration is not a main factor of controlling the de-NOx activity.

2.6. NH3 Adsorption

NH3-TPD profiles of Mn and Mn-HPWx are shown in Figure S10. Two desorption stages can be observed for all catalysts at 100–250 °C and 300–500 °C, corresponding to the desorption of NH3 adsorbed on the weak acid sites and on the medium-strong acid sites, respectively [12,75]. After the Mn catalyst is doped with HPW, changes in the peak areas and the peak temperatures of NH3 desorption are not obvious, illustrating that the effect of doping HPW on the acid number and acid strength is not obvious.
DRIFT spectra were recorded to further investigate the intermediates formed by adsorbed NH3. The spectra of the NH3-derived species on the Mn and Mn-HPW0.02 catalysts were collected at 60–300 °C (Figure 9). A couple of adsorbed NH3 species form on the Mn surface after exposing NH3 for 60 min (Figure 9a). The bands at 1220, 1290, and 1600 cm−1 are attributed to the coordinated NH3 bonded to the Lewis acid sites (L acid sites) [30,63,76], and the bands at 1445 and 1680 cm−1 are assigned to the NH4+ species adsorbed at the Brønsted acid sites (B acid sites) [12,19,30,72,76]. The bands of 1220, 1290, and 1600 cm−1 on L acids sites and the bands of 1445 and 1680 cm−1 display higher intensities at the low temperature, whereas these band intensities decrease gradually with rising desorption temperature. The band at 1330 cm−1 appears at 300 °C, and it is ascribed to the partial oxidation/deformation of adsorbed ammonia species by the active surface oxygen species [15,27,74]. The bands at 1220, 1445, 1600, and 1680 cm−1 are also detected on Mn-HPW0.02 in Figure 9b, and the variation trends of band intensities are close to those on Mn from 60 to 300 °C. Especially, the ammonia over Mn-HPW0.02 is adsorbed on L acid sites (1220 cm−1), while the prevailing ammonia species over Mn is NH4+ on B acid sites (1445 cm−1), demonstrating that the introduction of HPW can promote the adsorption of more ammonia species on L acid sites. Moreover, no oxidation band (1330 cm−1) of the adsorption NH3 species can be detected at 300 °C. These phenomena illustrate that the oxidation ability of the Mn catalyst is weakened via adding an appropriate amount of HPW, thus protecting the adsorbed ammonia species from oxidation, in accordance with the results of the NH3 oxidation activities (Figure 6b). The results mean that more ammonia species participate in the reaction with NOx and increase the conversion of NOx.

2.7. NOx Adsorption

Figure 10 exhibits the NO-TPD curves of the Mn, Mn-HPW0.005, Mn-HPW0.02, and Mn-HPW0.04 catalysts. The NO desorption can be divided into two regions at 50–500 °C. The peak at 50–200 °C is ascribed to the desorption of physically absorbed NO and the decomposition of monodentate nitrite/nitrate species, while that at 200–500 °C is ascribed to the decomposition of bridged and bidentate nitrate species [77,78,79]. The amount of desorbed NO was calculated according to the NO-TPD curves. In contrast to the Mn catalyst (the desorption area of NO on Mn is set as 1.00), the peak areas of desorbed NO over Mn-HPWx increase evidently, flowing the sequence of Mn-HPW0.02 (1.96) > Mn-HPW0.005 (1.68) > Mn-HPW0.04 (1.48) > Mn (1.00), meaning that the introduction of HPW enhances the NO adsorption capacity.
The intermediates of adsorbed NOx complexes were studied with in situ DRIFT (Figure 11). After NO and O2 were added to IR cell, the NOx-derived species on Mn in Figure 11a appeared mainly at 1290, 1343, 1540, and 1627 cm−1, and they can be attributed to monodentate nitrate (1290 cm−1) [11,72,80], monodentate nitrite (1343 cm−1) [27,73,81,82], bidentate nitrate (1540 cm−1) [19,53,83,84,85], and bridged nitrate (1627 cm−1) [19,36,53,86], respectively. The intensities of the adsorbed NOx-derived species become weaker and weaker due to the decomposition of the nitrate and/or nitrite species as the temperature increases. The bands at 1290 and 1343 cm−1 are hardly detected at 210 °C, demonstrating that these species are weakly adsorbed. With respect to Mn-HPW0.02, the bands at 1270 cm−1 (monodentate nitrate), 1540 cm−1, and 1627 cm−1 are also observed in Figure 11b. As shown in Figure 11, bidentate nitrate (1540 cm−1) is the leading species on Mn, while monodentate nitrate (1270 cm−1) is dominant on Mn-HPW0.02. Moreover, the bands of monodentate nitrate (1270 cm−1) with the greater intensity on Mn-HPW0.02 are still clearly visible at 300 °C. The observation means that the introduction of HPW affects both the existence mode and stability of the nitrates.

2.8. Reaction Mechanism

2.8.1. Reaction between NO + O2 and Pre-Adsorbed NH3

At 120 °C, DRIFT spectra were acquired to reveal the reactivity of adsorbed ammonia species with NO + O2. Coordinated NH3 (1227 and 1600 cm−1) and NH4+ (1445 cm−1) exist on the surface after Mn was pre-treated in NH3 for 60 min in Figure 12a. With the mixture gas of NO and O2 passing into IR cell, the band intensities of the adsorbed NH3 species decrease gradually within 40 min and do not vanish completely (Figure S11a). This observation reveals that the reaction between the NH3 species on L and B acid sites and gaseous NOx follows the Eley–Rideal (E-R) route [87]. Especially, the band intensity of NH4+ at 1445 cm−1 barely changes in the first 10 min after the introduction of NO + O2, and then increases slight within 20 min due to the formation of NH4NO2 [88]. After the exposure to the mixture of NO and O2 is longer than 40 min, the adsorbed NOx-derived species, i.e., monodentate nitrate (1255 cm−1), bidentate nitrate (1530 cm−1), and bridged nitrate (1627 cm−1) emerge on Mn surface, and the bands of these species partially overlap with those of the unreacted NH3 species. For instance, the emergence of bidentate nitrate at 1530 cm−1 influences the intensity of NH4+ at 1445 cm−1. The spectra on Mn-HPW0.02 are demonstrated in Figure 12b. The bands of NH3 species on L acid sites (1220 and 1600 cm−1) and B acid sites (1680 cm−1) diminish quickly and then disappear as NO and O2 were employed for 30 min in Figure S11b. Subsequently, the bands of monodentate nitrate (1270 cm−1) and bridged nitrate (1627 cm−1) become apparent. In Figure S11a,b, NH3 species on Mn and Mn-HPW0.02 begin to be consumed as the NO and O2 are passed for 20 min and 5 min, respectively, indicating that HPW can enhance the reactivity of ammonia species. Additionally, the intensity of NH3 species on L acid sites (1220/1227 cm−1) decreases faster than that on B acid sites, illustrating that NH3 species on L acid sites possess higher reactivity. For the NH3 pre-adsorbed catalysts, Mn-HPW0.02 shows a more coordinated NH3 species than Mn. Therefore, Mn-HPW0.02 exhibits excellent SCR activity in the low-temperature region, which coincides with the results in Figure 5a.

2.8.2. Reaction between NH3 and Pre-Adsorbed NO + O2

The reaction DRIFT spectra of the pre-adsorbed NO + O2 species with NH3 at 120 °C on Mn and Mn-HPW0.02 are presented in Figure 13. The bands of the monodentate nitrate (1280 cm−1), bidentate nitrate (1540 cm−1), and bridged nitrate (1630 cm−1) form on the Mn surface after pretreated in NO and O2 for 60 min in Figure 13a. The dependence of the intensities of the NOx-derived species on exposure time to NH3 is also shown in Figure S12a. As the introduction time of NH3 is less than 10 min, the change of band at 1280 cm−1 is ignorable, while the band intensities at 1540 and 1630 cm−1 attenuate progressively, and the band at 1630 cm−1 vanishes after exposing to NH3 for 20 min. These demonstrate that monodentate nitrate (1280 cm−1) is inactive in NH3-SCR, whereas the reaction of bidentate nitrate (1540 cm−1) and/or bridged nitrate (1630 cm−1) with adsorbed ammonia species obey the Langmuir–Hinshelwood (L-H) pathway [74,82,87]. As the NH3 continues to enter the system, the bands adsorbed on L acid sites (1240, 1290, and 1600 cm−1), B acid sites (1445 cm−1), and the oxidation/deformation of adsorbed ammonia species (1350 cm−1) become visible on the Mn surface. Additionally, the bands of adsorbed NH3 species at 1290 cm−1 cover those of monodentate nitrate (1280 cm−1), and the bands of NH4+ at 1445 cm−1 overlap with those of bidentate nitrate (1540 cm−1). For the spectra of Mn-HPW0.02, the NOx-derived species containing monodentate nitrate (1270 cm−1), bidentate nitrate (1550 cm−1), and bridged nitrate (1630 cm−1) are formed (Figure S12b and Figure 13b). With the introduction of NH3, the variation tendencies of these band intensities are similar to those on Mn, confirming that the L-H mechanism is also appropriate for Mn-HPW0.02. Notably, compared with the reaction rate of adsorbed NH3 species with NOx (Figure S11b), the reaction rate of adsorbed NOx-derived species with NH3 (Figure S12b) is slower. Furthermore, the introduction of HPW promotes the adsorption of more NH3 species on L acid sites. Thus, the E-R mechanism on Mn-HPW0.02 plays a dominant role and further enhances the performance for NH3-SCR reaction.

3. Materials and Methods

3.1. Catalyst Preparation

H3PW12O40-modified MnOx catalysts were prepared as follows. First, 12.0 g aqueous Mn(NO3)2 solution (50%, Sinopharm, Shanghai, China), 4.2 g oxalic acid dihydrate (OA, ≥ 99.0%, Adamas, Shanghai, China), 2.1 g ethylene glycol (EG, ≥ 99.0%, Adamas), and the required amounts of tungstophosphoric acid (H3PW12O40, HPW, 99%, Adamas) were added into an agate mortar. The sample was transferred into a beaker after hand grinding for 30 min, and subjected to ultrasonication for 30 min. The slurry was dried at 110 °C for 12 h, and then calcinated at 400 °C for 4 h to obtain Mn-HPWx, in which x represents the HPW/Mn molar ratio (0.005–0.04) and the molar ratio of Mn:OA:EG is 1:1:1. Moreover, the mono-component Mn catalyst was also prepared with the same process without adding HPW. HPW was prepared as a reference with a molar ratio of OA:EG:HPW = 1:1:0.02.

3.2. Characterization

X-ray diffraction data were obtained by a Bruker D8 Advance instrument (Karlsruhe, Germany). Textural properties of catalysts were analyzed at −196 °C on a Micromeritics ASAP-2020 HD88 adsorption apparatus (Norcross, GA, USA). The morphologies and EDX mapping data of samples were gained on ZEISS Gemini 300 (Oberkochen, Germany). TEM images were observed with a JEM-2100plus TEM instrument (Tokyo, Japan).
Temperature-programmed reduction with hydrogen (H2-TPR) and temperature-programmed desorption of NH3 (NH3-TPD) were performed by a thermal conductivity detector. First, a certain amount of catalyst was pretreated by N2 at 300 °C for 30 min and then cooled down to 30 °C. In H2-TPR experiment, 10% H2/N2 (50 mL/min) was introduced to 50 mg of catalyst, and then, the reduction experiments proceeded to 650 °C at 5 °C/min. For NH3-TPD, 100 mg sample was saturated with 10% NH3/N2 for 30 min at 30 °C followed by being flushed with N2 for 30 min at 100 °C, and subsequently, heated to 500 °C at 5 °C/min.
X-ray photoelectron spectroscopy (XPS) measurements were performed with an ESCALAB 250 Xi spectrometer (Thermo Fisher Scientific, Waltham, MA, USA).
Temperature-programmed desorption of NO (NO-TPD) was conducted on a mass spectrometer (Pfeiffer Vacuum Quadstar, 32-bit, Aßlar, Germany). First, 100 mg catalyst was pretreated in He (30 mL/min) at 300 °C for 30 min. Subsequently, the sample was flushed with 500 ppm NO/N2 and 5 vol % O2 for 1 h after cooling down to 30 °C. He gas was passed for 30 min at 50 °C, and then the NO-TPD test was started from 50 to 650 °C at 5 °C/min.
In-situ DRIFT data were obtained on a Nicolet 6700 spectrometer (Waltham, MA, USA) with 64 accumulated scans and a resolution of 4 cm−1. The sample was treated at 300 °C in N2 flow for 30 min. Subsequently, the background spectra were collected at the corresponding temperatures under N2 flow during the cooling process and were subtracted from each sample spectra in the test. For the experiments of NH3 or NOx adsorption at different temperatures, the catalysts were pretreated under the condition of 500 ppm NH3 or 500 ppm NO + 5% O2 for 60 min at 30 °C, followed by purging with N2 for 30 min. DRFIT spectra of adsorbed NH3- or NOx-derived species are collected at the corresponding temperature under the N2 atmosphere. With regards to the experiments of transient reaction, the catalysts were pretreated with 500 ppm NH3 or 500 ppm NO + 5% O2 at 120 °C for 60 min, and the spectra were recorded at different times. Subsequently, the samples were purged by N2 at 120 °C for 30 min, and finally, the mixture gas of 500 ppm NO + 5% O2 + or 500 ppm NH3 was introduced into the IR cell to react with the NH3- or NOx-derived species, collecting the spectra at different times.

3.3. Catalytic Activity Test

The activities were evaluated in a fixed quartz reactor. The reaction gas was composed of 500 ppm NO, 500 ppm NH3, 5% O2, 50 ppm SO2 (when used), and balanced N2. The gas flow rate was kept at 200 mL/min (WHSV = 80,000 mL g−1 h−1). The concentration of NO and NO2 in the reaction gas was determined with a flue gas analyzer (Testo 340, Lenzkirch, Germany) and the concentration of N2O was obtained by a gas chromatograph (Panna A91, China). NOx conversion and N2 selectivity were calculated [52,89,90]:
NO x   conversion   ( % ) = [ NO x ] inlet [ NO x ] outlet [ NO x ] inlet   ×   100 %
N 2   selectivity = ( 1 2 × [ N 2 O ] outlet [ NO x ] inlet + [ NH 3 ] inlet )   ×   100 %  
The individual NO (NH3) oxidation reaction was performed on the same devices without the introduction of NH3 (NO). A detector (LY500-NH3-H2) was employed to record the concentration of NH3 in effluent gas, and NO and NH3 oxidation conversion were calculated:
NO   conversion   ( % ) = [ NO ] inlet [ NO ] outlet [ NO ] inlet   ×   100 %
NH 3   conversion   ( % ) = [ NH 3 ] inlet [ NH 3 ] outlet [ NH 3 ] inlet   ×   100 %

4. Conclusions

HPW-modified MnOx catalysts were prepared for NOx removal. The effect of the HPW doping amount on the activities over the Mn catalyst was evaluated. Additionally, the correlation between physicochemical properties and NH3-SCR performance, and reaction mechanisms were surveyed via several characterizations. The conclusions can be drawn.
(1) The Mn catalyst modified by HPW possesses good NH3-SCR activity and above 90% NOx conversion at 90–270 °C can be obtained on Mn-HPW0.02 along with relatively higher N2 selectivity.
(2) The NH3 species on Lewis and Brønsted acid sites can take part in NH3-SCR, and the reactivity of NH3 species on Lewis acid sites is higher than those on Brønsted acid sites. The amount of Lewis acid sites increases obviously after adding HPW to the Mn catalyst, and the reaction of adsorbed NH3 species with gas-phase NOx is accelerated on Mn-HPW0.02, thus, boosting low-temperature activities.
(3) The reduction temperatures of the MnOx species move towards the higher temperature and the concentration of Mn4+ decreases, which weakens moderately the oxidation abilities of the Mn catalyst modified by HPW and protects the adsorbed NH3 species from the overoxidation. Therefore, NOx elimination and N2 selectivity are improved at high temperatures.
(4) Among the NOx-derived species, the bridged nitrate and bidentate nitrate on Mn and Mn-HPW0.02 have the higher reactivity, reacting with adsorbed NH3 species, while monodentate nitrate species fail to do so. At 120 °C, both E-R and L-H routes are suitable for the Mn and Mn-HPW0.02 catalysts for NH3-SCR, and the E-R mechanism is dominant on Mn-HPW0.02.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12101248/s1, Figure S1: XRD patterns of the Mn samples and the standard patterns of MnO2, Mn5O8, and Mn2O3 phases; Figure S2: XRD patterns of HPW calcinated at 400 °C and 650 °C; the standard patterns of HPW and WO3 phases; Figure S3: XRD patterns of Mn-HPW0.02 before calcination, Mn-HPW0.02 calcinated at 200 °C, 400 °C, and 650 °C; the standard patterns of MnWO4 phases; Figure S4: N2 adsorption–desorption isotherms on the Mn and Mn-HPWx catalysts; Figure S5: EDX mapping of Mn (a–c) and Mn-HPW0.02 (d–h); Figure S6: NH3-SCR activity on Mn, Mn-HPW, Mn-HPMo, Mn-HSiW, and Mn-HSiMo at 60–300 °C; Figure S7: N2 selectivity on the Mn and Mn-HPW0.02 catalysts at 60–300 °C; Figure S8: Effect of 50 ppm SO2 on NH3-SCR activity over the Mn and Mn-HPW0.02 catalysts at 150 °C; Figure S9: Yield of N2O, NO, and NO2 on Mn and Mn-HPW0.02 at 60-300 °C; Figure S10: NH3-TPD profiles of the Mn and Mn-HPWx catalysts; Figure S11: Dependence of the band intensities of NH3-derived species on time over Mn (a) and Mn-HPW0.02 (b); Figure S12: Dependence of the band intensities of the NOx-derived species on time over Mn (a) and Mn-HPW0.02 (b); Table S1: Representative Mn-based catalysts decorated by different assistant for NH3-SCR; Table S2: Reduction temperature and relative ratio in H2-TPR curves.

Author Contributions

Conceptualization, H.X. and X.G.; methodology, H.X.; validation, D.M., T.M. and J.Y.; formal analysis, H.X.; investigation, H.X. and X.G.; resources, D.M. and T.M.; data curation, H.X.; writing—original draft preparation, H.X.; writing—review and editing, H.X., X.G., D.M. and Z.M.; visualization, H.X.; supervision, X.G. and Z.M.; project administration, D.M.; funding acquisition, T.M. and Z.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Opening Project of Shanghai Key Laboratory of Atmospheric Particle Pollution and Prevention (LAP3) (FDLAP21007).

Data Availability Statement

Data are available from the authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Huang, R.J.; Zhang, Y.L.; Bozzetti, C.; Ho, K.F.; Cao, J.J.; Han, Y.M.; Daellenbach, K.R.; Slowik, J.G.; Platt, S.M.; Canonaco, F.; et al. High secondary aerosol contribution to particulate pollution during haze events in China. Nature 2014, 514, 218–222. [Google Scholar] [CrossRef] [Green Version]
  2. Han, L.P.; Cai, S.X.; Gao, M.; Hasegawa, J.Y.; Wang, P.L.; Zhang, J.P.; Shi, L.Y.; Zhang, D.S. Selective catalytic reduction of NOx with NH3 by using novel catalysts: State of the art and future prospects. Chem. Rev. 2019, 119, 10916–10976. [Google Scholar] [CrossRef]
  3. Xu, G.Y.; Guo, X.L.; Cheng, X.X.; Yu, J.; Fang, B.Z. A review of Mn-based catalysts for low-temperature NH3-SCR: NOx removal and H2O/SO2 resistance. Nanoscale 2021, 13, 7052–7080. [Google Scholar] [CrossRef]
  4. Wei, L.G.; Guo, R.T.; Zhou, J.; Qin, B.; Chen, X.; Bi, Z.X.; Pan, W.G. Chemical deactivation and resistance of Mn-based SCR catalysts for NOx removal from stationary sources. Fuel 2022, 316, 123438. [Google Scholar] [CrossRef]
  5. Guo, R.T.; Qin, B.; Wei, L.G.; Yin, T.Y.; Zhou, J.; Pan, W.G. Recent progress of low-temperature selective catalytic reduction of NOx with NH3 over manganese oxide-based catalysts. Phys. Chem. Chem. Phys. 2022, 24, 6363–6382. [Google Scholar] [CrossRef]
  6. Zhang, L.; Zhang, D.S.; Zhang, J.P.; Cai, S.X.; Fang, C.; Huang, L.; Li, H.R.; Gao, R.H.; Shi, L.Y. Design of meso-TiO2@MnOx-CeOx/CNTs with a core-shell structure as DeNOx catalysts: Promotion of activity, stability and SO2-tolerance. Nanoscale 2013, 5, 9821–9829. [Google Scholar] [CrossRef] [PubMed]
  7. Zhang, S.G.; Zhang, B.L.; Liu, B.; Sun, S.L. 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] [Green Version]
  8. Huang, C.Y.; Guo, R.T.; Pan, W.G.; Sun, X.; Liu, S.W.; Liu, J.; Wang, Z.Y.; Shi, X. SCR of NOx by NH3 over MnFeOx@TiO2 catalyst with a core-shell structure: The improved K resistance. J. Energy Inst. 2019, 92, 1364–1378. [Google Scholar] [CrossRef]
  9. Shi, Y.R.; Yi, H.H.; Gao, F.Y.; Zhao, S.Z.; Xie, Z.L.; Tang, X.L. Facile synthesis of hollow nanotube MnCoOx catalyst with superior resistance to SO2 and alkali metal poisons for NH3-SCR removal of NOx. Sep. Purif. Technol. 2021, 265, 118517. [Google Scholar] [CrossRef]
  10. Liu, C.; Shi, J.-W.; Gao, C.; Niu, C.M. Manganese oxide-based catalysts for low-temperature selective catalytic reduction of NOx with NH3: A review. Appl. Catal. A Gen. 2016, 522, 54–69. [Google Scholar] [CrossRef]
  11. Jiang, B.Q.; Lin, B.L.; Li, Z.G.; Zhao, S.; Chen, Z.Y. Mn/TiO2 catalysts prepared by ultrasonic spray pyrolysis method for NOx removal in low-temperature SCR reaction. Colloids Surf. A Physicochem. Eng. Asp. 2020, 586, 124210. [Google Scholar] [CrossRef]
  12. Wei, L.; Wang, Z.W.; Liu, Y.; Guo, G.S.; Dai, H.X.; Cui, S.P.; Deng, J.G. Support promotion effect on the SO2 and K+ co-poisoning resistance of MnO2/TiO2 for NH3-SCR of NO. J. Hazard. Mater. 2021, 416, 126117. [Google Scholar] [CrossRef] [PubMed]
  13. Kang, H.; Wu, M.X.; Li, S.Y.; Wei, C.H.; Chen, X.P.; Chen, J.J.; Jing, F.L.; Chu, W.; Liu, Y.F. Converting poisonous sulfate species to an active promoter on TiO2 predecorated MnOx catalysts for the NH3-SCR reaction. ACS Appl. Mater. Interfaces 2021, 13, 61237–61247. [Google Scholar] [CrossRef]
  14. Cao, F.; Xiang, J.; Su, S.; Wang, P.Y.; Sun, L.S.; Hu, S.; Lei, S.Y. The activity and characterization of MnOx-CeO2-ZrO2/γ-Al2O3 catalysts for low temperature selective catalytic reduction of NO with NH3. Chem. Eng. J. 2014, 243, 347–354. [Google Scholar] [CrossRef]
  15. Cao, F.; Su, S.; Xiang, J.; Wang, P.Y.; Hu, S.; Sun, L.S.; Zhang, A.C. The activity and mechanism study of Fe-Mn-Ce/γ-Al2O3 catalyst for low temperature selective catalytic reduction of NO with NH3. Fuel 2015, 139, 232–239. [Google Scholar] [CrossRef]
  16. Yang, G.; Zhao, H.T.; Luo, X.; Shi, K.Q.; Zhao, H.B.; Wang, W.K.; Chen, Q.H.; Fan, H.; Wu, T. Promotion effect and mechanism of the addition of Mo on the enhanced low temperature SCR of NOx by NH3 over MnOx/γ-Al2O3 catalysts. Appl. Catal. B Environ. 2019, 245, 743–752. [Google Scholar] [CrossRef] [Green Version]
  17. Wang, L.S.; Huang, B.C.; Su, Y.X.; Zhou, G.Y.; Wang, K.L.; Luo, H.C.; Ye, D.Q. Manganese oxides supported on multi-walled carbon nanotubes for selective catalytic reduction of NO with NH3: Catalytic activity and characterization. Chem. Eng. J. 2012, 192, 232–241. [Google Scholar] [CrossRef]
  18. Zhou, Y.H.; Su, B.X.; Ren, S.; Chen, Z.C.; Su, Z.H.; Yang, J.; Chen, L.; Wang, M.M. Nb2O5-modified Mn-Ce/AC catalyst with high ZnCl2 and SO2 tolerance for low-temperature NH3-SCR of NO. J. Environ. Chem. Eng. 2021, 9, 106323. [Google Scholar] [CrossRef]
  19. Yang, J.; Ren, S.; Su, B.X.; Wang, M.M.; Chen, L.; Liu, Q.C. Understanding the dual-acting of iron and sulfur dioxide over Mn-Fe/AC catalysts for low-temperature SCR of NO. Mol. Catal. 2022, 519, 112150. [Google Scholar] [CrossRef]
  20. Lou, X.R.; Liu, P.; Li, J.; Li, Z.; He, K. Effects of calcination temperature on Mn species and catalytic activities of Mn/ZSM-5 catalyst for selective catalytic reduction of NO with ammonia. Appl. Surf. Sci. 2014, 307, 382–387. [Google Scholar] [CrossRef]
  21. Li, J.; Guo, J.X.; Shi, X.K.; Wen, X.R.; Chu, Y.H.; Yuan, S.D. Effect of aluminum on the catalytic performance and reaction mechanism of Mn/MCM-41 for NH3-SCR reaction. Appl. Surf. Sci. 2020, 534, 147592. [Google Scholar] [CrossRef]
  22. Lu, P.; Ye, L.; Yan, X.H.; Fang, P.; Chen, X.B.; Chen, D.S.; Cen, C.P. Impact of toluene poisoning on MnCe/HZSM-5 SCR catalyst. Chem. Eng. J. 2021, 414, 128838. [Google Scholar] [CrossRef]
  23. Qi, G.; Yang, R.T.; Chang, R. MnOx-CeO2 mixed oxides prepared by co-precipitation for selective catalytic reduction of NO with NH3 at low temperatures. Appl. Catal. B Environ. 2004, 51, 93–106. [Google Scholar] [CrossRef]
  24. Liu, Z.M.; Yi, Y.; Zhang, S.X.; Zhu, T.L.; Zhu, J.Z.; Wang, J.G. Selective catalytic reduction of NOx with NH3 over Mn-Ce mixed oxide catalyst at low temperatures. Catal. Today 2013, 216, 76–81. [Google Scholar] [CrossRef]
  25. Li, W.M.; Liu, H.D.; Chen, Y.F. Fabrication of MnOx-CeO2-based catalytic filters and their application in low-temperature selective catalytic reduction of NO with NH3. Ind. Eng. Chem. Res. 2020, 59, 12657–12665. [Google Scholar] [CrossRef]
  26. Fang, X.; Liu, Y.J.; Cheng, Y.; Cen, W.L. Mechanism of Ce-modified birnessite-MnO2 in promoting SO2 poisoning resistance for low-temperature NH3-SCR. ACS Catal. 2021, 11, 4125–4135. [Google Scholar] [CrossRef]
  27. Yang, S.J.; Wang, C.Z.; Li, J.H.; Yan, N.Q.; Ma, L.; Chang, H.Z. Low temperature selective catalytic reduction of NO with NH3 over Mn-Fe spinel: Performance, mechanism and kinetic study. Appl. Catal. B Environ. 2011, 110, 71–80. [Google Scholar] [CrossRef]
  28. Fan, Z.Y.; Shi, J.W.; Gao, C.; Gao, G.; Wang, B.R.; Niu, C.M. 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]
  29. Liu, Z.; Sun, G.X.; Chen, C.; Sun, K.; Zeng, L.Y.; Yang, L.Z.; Chen, Y.J.; Wang, W.H.; Liu, B.; Lu, Y.K.; et al. Fe-doped Mn3O4 spinel nanoparticles with highly exposed Feoct-O-Mntet sites for eficient selective catalytic reduction (SCR) of NO with ammonia at low temperatures. ACS Catal. 2020, 10, 6803–6809. [Google Scholar] [CrossRef]
  30. Zhang, L.; Shi, L.Y.; Huang, L.; Zhang, J.P.; Gao, R.H.; Zhang, D.S. 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. Zhu, W.J.; Tang, X.L.; Gao, F.Y.; Yi, H.H.; Zhang, R.C.; Wang, J.E.; Yang, C.; Ni, S.Q. The effect of non-selective oxidation on the Mn2Co1Ox catalysts for NH3-SCR: Positive and non-positive. Chem. Eng. J. 2020, 385, 123797. [Google Scholar] [CrossRef]
  32. Cai, S.X.; Zhang, D.S.; Shi, L.Y.; Xu, J.; Zhang, L.; Huang, L.; Li, H.R.; Zhang, J.P. 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]
  33. Wan, Y.P.; Zhao, W.R.; Tang, Y.; Li, L.; Wang, H.J.; Cui, Y.L.; Gu, J.L.; Li, Y.S.; Shi, J.L. Ni-Mn bi-metal oxide catalysts for the low temperature SCR removal of NO with NH3. Appl. Catal. B Environ. 2014, 148–149, 114–122. [Google Scholar] [CrossRef]
  34. Shi, Y.; Chu, Q.; Xiong, W.; Gao, J.S.; Huang, L.; Zhang, Y.Q.; Ding, Y. A new type bimetallic NiMn-MOF-74 as an efficient low-temperatures catalyst for selective catalytic reduction of NO by CO. Chem. Eng. Process. 2021, 159, 108232. [Google Scholar] [CrossRef]
  35. Xu, H.; Zhang, Q.L.; Qiu, C.T.; Lin, T.; Gong, M.C.; Chen, Y.Q. Tungsten modified MnOx-CeO2/ZrO2 monolith catalysts for selective catalytic reduction of NOx with ammonia. Chem. Eng. Sci. 2012, 76, 120–128. [Google Scholar] [CrossRef]
  36. Liu, L.J.; Su, S.; Chen, D.Z.; Shu, T.; Zheng, X.T.; Yu, J.Y.; Feng, Y.; Wang, Y.; Hu, S.; Xiang, J. Highly efficient NH3-SCR of NOx over MnFeW/Ti catalyst at low temperature: SO2 tolerance and reaction mechanism. Fuel 2022, 307, 121805. [Google Scholar] [CrossRef]
  37. Liu, Z.M.; Zhu, J.Z.; Li, J.H.; Ma, L.L.; Woo, S.I. Novel Mn-Ce-Ti mixed-oxide catalyst for the selective catalytic reduction of NOx with NH3. ACS Appl. Mater. Interfaces 2014, 6, 14500–14508. [Google Scholar] [CrossRef] [PubMed]
  38. Jiang, H.X.; Wang, C.X.; Wang, H.Q.; Zhang, M.H. Synthesis of highly efficient MnOx catalyst for low-temperature NH3-SCR prepared from Mn-MOF-74 template. Mater. Lett. 2016, 168, 17–19. [Google Scholar] [CrossRef]
  39. Yao, H.Y.; Cai, S.X.; Yang, B.; Han, L.P.; Wang, P.L.; Li, H.R.; Yan, T.T.; Shi, L.Y.; Zhang, D.S. In situ decorated MOF-derived Mn-Fe oxides on Fe mesh as novel monolithic catalysts for NOx reduction. New J. Chem. 2020, 44, 2357–2366. [Google Scholar] [CrossRef]
  40. Chen, R.Y.; Fang, X.Y.; Li, Z.G.; Liu, Z.M. Selective catalytic reduction of NOx with NH3 over a novel MOF- derived MnOx catalyst. Appl. Catal. A Gen. 2022, 643, 118754. [Google Scholar] [CrossRef]
  41. Ren, Z.Y.; Fan, H.; Wang, R. A novel ring-like Fe2O3-based catalyst: Tungstophosphoric acid modification, NH3-SCR activity and tolerance to H2O and SO2. Catal. Commun. 2017, 100, 71–75. [Google Scholar] [CrossRef]
  42. Ren, Z.Y.; Teng, Y.F.; Zhao, L.Y.; Wang, R. Keggin-tungstophosphoric acid decorated Fe2O3 nanoring as a new catalyst for selective catalytic reduction of NOx with ammonia. Catal. Today 2017, 297, 36–45. [Google Scholar] [CrossRef]
  43. Geng, Y.; Xiong, S.C.; Li, B.; Peng, Y.; Yang, S.J. Promotion of H3PW12O40 grafting on NOx abatement over γ-Fe2O3: Performance and reaction mechanism. Ind. Eng. Chem. Res. 2018, 57, 13661–13670. [Google Scholar] [CrossRef]
  44. Wu, J.H.; Jin, S.L.; Wei, X.D.; Gu, F.J.; Han, Q.; Lan, Y.X.; Qian, C.L.; Li, J.Q.; Wang, X.R.; Zhang, R.; et al. Enhanced sulfur resistance of H3PW12O40-modified Fe2O3 catalyst for NH3-SCR: Synergistic effect of surface acidity and oxidation ability. Chem. Eng. J. 2021, 412, 128712. [Google Scholar] [CrossRef]
  45. Weng, X.L.; Dai, X.X.; Zeng, Q.S.; Liu, Y.; Wu, Z.B. DRIFT studies on promotion mechanism of H3PW12O40 in selective catalytic reduction of NO with NH3. J. Colloid Interf. Sci. 2016, 461, 9–14. [Google Scholar] [CrossRef] [PubMed]
  46. Geng, Y.; Xiong, S.C.; Li, B.; Liao, Y.; Xiao, X.; Yang, S.J. H3PW12O40 grafted on CeO2: A high-performance catalyst for the selective catalytic reduction of NOx with NH3. Ind. Eng. Chem. Res. 2018, 57, 856–866. [Google Scholar] [CrossRef]
  47. Geng, Y.; Jin, K.; Mei, J.; Su, G.Y.; Ma, L.; Yang, S.J. CeO2 grafted with different heteropoly acids for selective catalytic reduction of NO with NH3. J. Hazard. Mater. 2020, 382, 121032. [Google Scholar] [CrossRef]
  48. Putluru, S.S.R.; Mossin, S.; Riisager, A.; Fehrmann, R. Heteropoly acid promoted Cu and Fe catalysts for the selective catalytic reduction of NO with ammonia. Catal. Today 2011, 176, 292–297. [Google Scholar] [CrossRef]
  49. Putluru, S.S.R.; Jensen, A.D.; Riisager, A.; Fehrmann, R. Heteropoly acid promoted V2O5/TiO2 catalysts for NO abatement with ammonia in alkali containing flue gases. Catal. Sci. Technol. 2011, 1, 631–637. [Google Scholar] [CrossRef]
  50. Tang, X.L.; Wang, C.Z.; Gao, F.Y.; Zhang, R.C.; Shi, Y.R.; Yi, H.H. Acid modification enhances selective catalytic reduction activity and sulfur dioxide resistance of manganese-cerium-cobalt catalysts: Insight into the role of phosphotungstic acid. J. Colloid Interf. Sci. 2021, 603, 291–306. [Google Scholar] [CrossRef]
  51. Xiao, X.; Sheng, Z.Y.; Yang, L.; Dong, F. Low-temperature selective catalytic reduction of NOx with NH3 over a manganese and cerium oxide/graphene composite prepared by a hydrothermal method. Catal. Sci. Technol. 2016, 6, 1507–1514. [Google Scholar] [CrossRef]
  52. Gao, F.Y.; Tang, X.L.; Yi, H.H.; Li, J.Y.; Zhao, S.Z.; Wang, J.G.; Chu, C.; Li, C.L. 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]
  53. Meng, D.M.; Zhan, W.C.; Guo, Y.; Guo, Y.L.; Wang, L.; Lu, G.Z. 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]
  54. Jalil, P.A.; Faiz, M.; Tabet, N.; Hamdan, N.M.; Hussain, Z. A study of the stability of tungstophosphoric acid, H3PW12O40, using synchrotron XPS, XANES, hexane cracking, XRD, and IR spectroscopy. J. Catal. 2003, 217, 292–297. [Google Scholar] [CrossRef]
  55. Essayem, N.; Holmqvist, A.; Gayraud, P.Y.; Vedrine, J.C.; Taarit, Y.B. In situ FTIR studies of the protonic sites of H3PW12O40 and its acidic cesium salts MxH3-xPW12O40. J. Catal. 2001, 197, 273–280. [Google Scholar] [CrossRef]
  56. Jiang, Y.; Liu, T.Y.; Gao, W.Q.; Ge, H.W.; Yang, Z.D.; Lin, R.Y.; Wang, X.W. Three-dimensionally ordered macroporous Ce-W-Nb oxide catalysts for selective catalytic reduction of NO with NH3. Chem. Eng. J. 2022, 433, 134576. [Google Scholar] [CrossRef]
  57. Zhang, X.; Liu, Y.X.; Deng, J.G.; Zhao, X.T.; Zhang, K.F.; Yang, J.; Han, Z.; Jiang, X.Y.; Dai, H.X. Three-dimensionally ordered macroporous Cr2O3-CeO2: High-performance catalysts for the oxidative removal of trichloroethylene. Catal. Today 2020, 339, 200–209. [Google Scholar] [CrossRef]
  58. Lian, Z.H.; Liu, F.D.; He, H.; Shi, X.Y.; Mo, J.S.; Wu, Z.B. Manganese-niobium mixed oxide catalyst for the selective catalytic reduction of NOx with NH3 at low temperatures. Chem. Eng. J. 2014, 250, 390–398. [Google Scholar] [CrossRef]
  59. 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 Gen. 2017, 531, 129–138. [Google Scholar] [CrossRef]
  60. Xiong, S.C.; Peng, Y.; Wang, D.; Huang, N.; Zhang, Q.F.; Yang, S.J.; Chen, J.J.; Li, J.H. The role of the Cu dopant on a Mn3O4 spinel SCR catalyst: Improvement of low-temperature activity and sulfur resistance. Chem. Eng. J. 2020, 387, 124090. [Google Scholar] [CrossRef]
  61. Wu, X.; Liu, L.L.; Liu, J.N.; Hou, B.H.; Du, Y.L.; Xie, X.M. NiMn mixed oxides with enhanced low-temperature deNOx performance: Insight into the coordinated decoration of MnOx by NiO phase via glycine combustion method. Appl. Catal. A Gen. 2021, 610, 117918. [Google Scholar] [CrossRef]
  62. Xiong, S.C.; Liao, Y.; Xiao, X.; Dang, H.; Yang, S.J. Novel effect of H2O on the low temperature selective catalytic reduction of NO with NH3 over MnOx-CeO2: Mechanism and kinetic Study. J. Phys. Chem. C 2015, 119, 4180–4187. [Google Scholar] [CrossRef]
  63. Wang, C.Z.; Gao, F.Y.; Ko, S.J.; Liu, H.H.; Yi, H.H.; Tang, X.L. Structural control for inhibiting SO2 adsorption in porous MnCe nanowire aerogel catalysts for low-temperature NH3-SCR. Chem. Eng. J. 2022, 434, 134729. [Google Scholar] [CrossRef]
  64. Chen, L.Q.; Yuan, F.L.; Li, Z.B.; Niu, X.Y.; Zhu, Y.J. Synergistic effect between the redox property and acidity on enhancing the low temperature NH3-SCR activity for NO removal over the Co0.2CexMn0.8-xTi10 (x = 0–0.40) oxides catalysts. Chem. Eng. J. 2018, 354, 393–406. [Google Scholar] [CrossRef]
  65. Kim, M.J.; Youn, J.R.; Lee, S.J.; Ryu, I.S.; Nam, S.C.; Jeong, S.K.; Jeon, S.K. Facile control of surface properties in CeO2-promoted Mn/TiO2 catalyst for low-temperature selective catalytic reduction of NO by NH3. J. Ind. Eng. Chem. 2022, 108, 438–448. [Google Scholar] [CrossRef]
  66. Yoon, W.; Kim, Y.; Kim, G.J.; Kim, J.-R.; Lee, S.; Han, H.; Park, G.H.; Chae, H.-J.; Kim, W.B. Boosting low temperature De-NOx performance and SO2 resistance over Ce-doped two dimensional Mn-Cr layered double oxide catalyst. Chem. Eng. J. 2022, 434, 134676. [Google Scholar] [CrossRef]
  67. Chen, L.; Yao, X.J.; Cao, J.; Yang, F.M.; Tang, C.J.; Dong, L. Effect of Ti4+ and Sn4+ co-incorporation on the catalytic performance of CeO2-MnOx catalyst for low temperature NH3-SCR. Appl. Surf. Sci. 2019, 476, 283–292. [Google Scholar] [CrossRef]
  68. Zhang, Y.P.; Li, G.B.; Wu, P.; Zhuang, K.; Shen, K.; Wang, S.; Huang, T.J. Effect of SO2 on the low-temperature denitrification performance of Ho-modified Mn/Ti catalyst. Chem. Eng. J. 2020, 400, 122597. [Google Scholar] [CrossRef]
  69. Li, C.X.; Shen, M.Q.; Yu, T.; Wang, J.Q.; Wang, J.; Zhai, Y.P. The mechanism of ammonium bisulfate formation and decomposition over V/WTi catalysts for NH3-selective catalytic reduction at various temperatures. Phys. Chem. Chem. Phys. 2017, 19, 15194–15206. [Google Scholar] [CrossRef]
  70. Xiong, Z.B.; Wang, W.; Li, J.; Huang, L.H.; Lu, W. The synergistic promotional effect of W doping and sulfate modification on the NH3-SCR activity of CeO2 catalyst. Mol. Catal. 2022, 522, 112250. [Google Scholar] [CrossRef]
  71. Liu, Y.P.; He, B.; Li, Q.; Liu, H.; Qiu, L.; Liu, J.; Xiang, W.; Liu, Y.X.; Wang, G.K.; Wu, Z.G.; et al. Relieving capacity decay and voltage fading of Li1.2Ni0.13Co0.13Mn0.54O2 by Mg2+ and PO43- dual doping. Mater. Res. Bull. 2020, 130, 110923. [Google Scholar] [CrossRef]
  72. Li, Y.F.; Hou, Y.Q.; Zhang, Y.Z.; Yang, Y.T.; Huang, Z.G. Confinement of MnOx@Fe2O3 core-shell catalyst with titania nanotubes: Enhanced N2 selectivity and SO2 tolerance in NH3-SCR process. J. Colloid Interf. Sci. 2022, 608, 2224–2234. [Google Scholar] [CrossRef] [PubMed]
  73. Xue, H.Y.; Guo, X.M.; Meng, T.; Guo, Q.S.; Mao, D.S.; Wang, S. Cu-ZSM-5 catalyst impregnated with Mn-Co oxide for the selected catalytic reduction of NO: Physicochemical property-catalytic activity relationship and in situ DRIFTS study for the reaction mechanism. ACS Catal. 2021, 11, 7702–7718. [Google Scholar] [CrossRef]
  74. Liu, Y.Y.; Gao, F.Y.; Ko, S.J.; Wang, C.Z.; Liu, H.H.; Tang, X.L.; Yi, H.H.; Zhou, Y.S. Superior catalytic performance within H2O-vapor of W-modified CoMn2O4/TiO2 catalyst for selective catalytic reduction of NOx with NH3. Chem. Eng. J. 2022, 434, 134770. [Google Scholar] [CrossRef]
  75. Fan, Z.Y.; Shi, J.W.; Gao, C.; Gao, G.; Wang, B.R.; Wang, Y.; He, C.; Niu, C.M. 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]
  76. Gao, E.H.; Feng, W.J.; Huang, B.; Zhu, J.L.; Wang, W.; Li, J.W.; He, Y. The enhanced resistance to Na+-poisoning of MnCoCrOx SCR catalyst by acidity regulation: The mechanism of sulfuric acid pretreatment. Mol. Catal. 2022, 518, 112084. [Google Scholar] [CrossRef]
  77. Hu, H.; Cai, S.X.; Li, H.R.; Huang, L.; Shi, L.Y.; Zhang, D.S. 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]
  78. Meng, D.M.; Xu, Q.; Jiao, Y.L.; Guo, Y.; Guo, Y.L.; Wang, L.; Lu, G.Z.; Zhan, W.C. Spinel structured CoaMnbOx mixed oxide catalyst for the selective catalytic reduction of NOx with NH3. Appl. Catal. B Environ. 2018, 221, 652–663. [Google Scholar] [CrossRef]
  79. Fan, A.D.; Jing, Y.; Guo, J.X.; Shi, X.K.; Yuan, S.D.; Li, J.J. Investigation of Mn doped perovskite La-Mn oxides for NH3-SCR activity and SO2/H2O resistance. Fuel 2022, 310, 122237. [Google Scholar] [CrossRef]
  80. Xu, Q.; Li, Z.Y.; Wang, L.; Zhan, W.C.; Guo, Y.L.; Guo, Y. Understand the role of redox property and NO adsorption over MnFeOx for NH3-SCR. Catal. Sci. Technol. 2022, 12, 2030–2041. [Google Scholar] [CrossRef]
  81. Hu, H.; Cai, S.X.; Li, H.R.; Huang, L.; Shi, L.Y.; Zhang, D.S. 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]
  82. Xue, H.Y.; Guo, X.M.; Meng, T.; Mao, D.S.; Ma, Z. NH3-SCR of NO over M/ZSM-5 (M = Mn, Co, Cu) catalysts: An in-situ DRIFTS study. Surf. Interfaces 2022, 29, 101722. [Google Scholar] [CrossRef]
  83. Qiu, L.; Pang, D.D.; Zhang, C.L.; Meng, J.J.; Zhu, R.S.; Ouyang, F. In situ IR studies of Co and Ce doped Mn/TiO2 catalyst for low-temperature selective catalytic reduction of NO with NH3. Appl. Surf. Sci. 2015, 357, 189–196. [Google Scholar] [CrossRef]
  84. Guo, R.T.; Wang, S.X.; Pan, W.G.; Li, M.Y.; Sun, P.; Liu, S.M.; Sun, X.; Liu, S.W.; Liu, J. Different poisoning effects of K and Mg on the Mn/TiO2 catalyst for selective catalytic reduction of NOx with NH3: A mechanistic study. J. Phys. Chem. C 2017, 121, 7881–7891. [Google Scholar] [CrossRef]
  85. Xie, S.Z.; Li, L.L.; Jin, L.J.; Wu, Y.H.; Liu, H.; Qin, Q.J.; Wei, X.L.; Liu, J.X.; Dong, L.H.; Li, B. Low temperature high activity of M (M = Ce, Fe, Co, Ni) doped M-Mn/TiO2 catalysts for NH3-SCR and in situ DRIFTS for investigating the reaction mechanism. Appl. Surf. Sci. 2020, 515, 146014. [Google Scholar] [CrossRef]
  86. Kijlstra, W.S.; Brands, D.S.; Poels, E.K.; Bliek, A. Mechanism of the selective catalytic reduction of NO by NH3 over MnOx/Al2O3 I. Adsorption and desorption of the single reaction components. J. Catal. 1997, 171, 208–218. [Google Scholar] [CrossRef]
  87. Wei, L.H.; Li, X.Y.; Mu, J.C.; Wang, X.Y.; Fan, S.Y.; Yin, Z.F.; Tadé, M.O.; Liu, S.M. Rationally tailored redox properties of a mesoporous Mn-Fe spinel nanostructure for boosting low-temperature selective catalytic reduction of NOx with NH3. ACS Sustain. Chem. Eng. 2020, 8, 17727–17739. [Google Scholar] [CrossRef]
  88. Sun, Q.; Gao, Z.X.; Wen, B.; Sachtler, W.M.H. Spectroscopic evidence for a nitrite intermediate in the catalytic reduction of NOx with ammonia on Fe/MFI. Catal. Lett. 2002, 78, 1–5. [Google Scholar] [CrossRef]
  89. Gao, F.Y.; Tang, X.L.; Yi, H.H.; Zhao, S.Z.; Wang, J.E.; Gu, T. Improvement of activity, selectivity and H2O & SO2-tolerance of micro-mesoporous CrMn2O4 spinel catalyst for low-temperature NH3-SCR of NOx. Appl. Surf. Sci. 2019, 466, 411–424. [Google Scholar] [CrossRef]
  90. Gao, F.Y.; Tang, X.L.; Yi, H.H.; Zhao, S.Z.; Wang, J.G.; Shi, Y.R.; Meng, X.M. 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]
Figure 1. XRD patterns of Mn and Mn-HPWx.
Figure 1. XRD patterns of Mn and Mn-HPWx.
Catalysts 12 01248 g001
Figure 2. FTIR spectra of HPW before calcinatation (a), HPW calcinated at 400 °C (b) and 650 °C (c); Mn-HPW0.02 before calcination (d), Mn-HPW0.02 calcinated at 200 °C (e), 400 °C (f), and 650 °C (g).
Figure 2. FTIR spectra of HPW before calcinatation (a), HPW calcinated at 400 °C (b) and 650 °C (c); Mn-HPW0.02 before calcination (d), Mn-HPW0.02 calcinated at 200 °C (e), 400 °C (f), and 650 °C (g).
Catalysts 12 01248 g002
Figure 3. SEM images of HPW (a), Mn (b), and Mn-HPW0.02 (c).
Figure 3. SEM images of HPW (a), Mn (b), and Mn-HPW0.02 (c).
Catalysts 12 01248 g003
Figure 4. TEM images of HPW (a,b), Mn (c,d), and Mn-HPW0.02 (eg).
Figure 4. TEM images of HPW (a,b), Mn (c,d), and Mn-HPW0.02 (eg).
Catalysts 12 01248 g004
Figure 5. NH3-SCR activity on HPW, Mn, and Mn-HPWx at 60–300 °C (a) and at 60–180 °C (b).
Figure 5. NH3-SCR activity on HPW, Mn, and Mn-HPWx at 60–300 °C (a) and at 60–180 °C (b).
Catalysts 12 01248 g005
Figure 6. Conversion of NO to NO2 (a) and NH3 conversion (b) on Mn and Mn-HPW0.02 at 60–300 °C.
Figure 6. Conversion of NO to NO2 (a) and NH3 conversion (b) on Mn and Mn-HPW0.02 at 60–300 °C.
Catalysts 12 01248 g006
Figure 7. H2-TPR profiles of Mn, Mn-HPWx, and HPW.
Figure 7. H2-TPR profiles of Mn, Mn-HPWx, and HPW.
Catalysts 12 01248 g007
Figure 8. XPS results of Mn 2p (a), W 4f (b), P 2p (c), and O 1s (d) over Mn and Mn-HPWx samples.
Figure 8. XPS results of Mn 2p (a), W 4f (b), P 2p (c), and O 1s (d) over Mn and Mn-HPWx samples.
Catalysts 12 01248 g008
Figure 9. DRIFT spectra of NH3-derived species on Mn (a), Mn-HPW0.02 (b) obtained by subtracting the corresponding background spectra. The catalysts were pretreated with 500 ppm NH3 for 60 min and then purging by N2 for 30 min at 30 °C.
Figure 9. DRIFT spectra of NH3-derived species on Mn (a), Mn-HPW0.02 (b) obtained by subtracting the corresponding background spectra. The catalysts were pretreated with 500 ppm NH3 for 60 min and then purging by N2 for 30 min at 30 °C.
Catalysts 12 01248 g009
Figure 10. NO-TPD profiles on Mn, Mn-HPW0.005, Mn-HPW0.02, and Mn-HPW0.04.
Figure 10. NO-TPD profiles on Mn, Mn-HPW0.005, Mn-HPW0.02, and Mn-HPW0.04.
Catalysts 12 01248 g010
Figure 11. DRIFT spectra of NOx-derived species on Mn (a) and Mn-HPW0.02 (b) obtained by subtracting the corresponding background spectra. The catalysts were pretreated with 500 ppm NO + 5% O2 for 60 min and then purging by N2 for 30 min at 30 °C.
Figure 11. DRIFT spectra of NOx-derived species on Mn (a) and Mn-HPW0.02 (b) obtained by subtracting the corresponding background spectra. The catalysts were pretreated with 500 ppm NO + 5% O2 for 60 min and then purging by N2 for 30 min at 30 °C.
Catalysts 12 01248 g011
Figure 12. DRIFT spectra obtained by subtracting the corresponding background spectra of NO + O2 reacted with pre-adsorbed NH3 species at 120 °C on Mn (a) and Mn-HPW0.02 (b).
Figure 12. DRIFT spectra obtained by subtracting the corresponding background spectra of NO + O2 reacted with pre-adsorbed NH3 species at 120 °C on Mn (a) and Mn-HPW0.02 (b).
Catalysts 12 01248 g012
Figure 13. DRIFT spectra obtained by subtracting the corresponding background spectra of of NH3 reacted with pre-adsorbed NOx-derived species at 120 °C on Mn (a) and Mn-HPW0.02 (b).
Figure 13. DRIFT spectra obtained by subtracting the corresponding background spectra of of NH3 reacted with pre-adsorbed NOx-derived species at 120 °C on Mn (a) and Mn-HPW0.02 (b).
Catalysts 12 01248 g013
Table 1. Surface area and pore structural characterizations of the Mn, Mn-HPWx, and HPW catalysts.
Table 1. Surface area and pore structural characterizations of the Mn, Mn-HPWx, and HPW catalysts.
SamplesBET Surface Area (m2/g)Total Pore Volume (cm3/g)Average Pore Diameter (nm)
Mn33.70.1417.0
Mn-HPW0.00554.90.1813.7
Mn-HPW0.0158.50.1711.6
Mn-HPW0.0269.10.148.3
Mn-HPW0.0355.70.1410.1
Mn-HPW0.0452.50.1412.4
HPW1.60.002.0
Table 2. Atomic surface concentrations on Mn and Mn-HPWx.
Table 2. Atomic surface concentrations on Mn and Mn-HPWx.
CatalystSurface Atomic Concentration 1 (at. %)Relative Ratio 2 (%)Relative Concentration 3 (at. %)
MnOWPMn4+/MnMn3+/MnMn2+/MnOβ/OMn4+Mn3+Mn2+Oβ
Mn30.751.2--26.752.221.135.18.216.06.518.0
Mn-HPW0.00527.452.82.01.226.651.821.632.27.314.25.917.0
Mn-HPW0.0126.154.13.11.425.951.522.632.36.813.45.917.5
Mn-HPW0.0224.654.04.01.724.652.323.132.36.112.85.717.5
Mn-HPW0.0321.654.55.32.124.353.122.632.35.311.44.917.6
Mn-HPW0.0421.255.26.12.324.253.422.432.35.111.34.817.8
1 The ratio of the Mn (or O, or W) atom to the sum of the measured elements (such as Mn, O, C, N, W, and P). 2 The ratio of the corresponding peak areas in the XPS spectrum. 3 The relative concentration was calculated using the relative ratio multiplied by the surface atomic concentration.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Xue, H.; Guo, X.; Mao, D.; Meng, T.; Yu, J.; Ma, Z. Phosphotungstic Acid-Modified MnOx for Selective Catalytic Reduction of NOx with NH3. Catalysts 2022, 12, 1248. https://doi.org/10.3390/catal12101248

AMA Style

Xue H, Guo X, Mao D, Meng T, Yu J, Ma Z. Phosphotungstic Acid-Modified MnOx for Selective Catalytic Reduction of NOx with NH3. Catalysts. 2022; 12(10):1248. https://doi.org/10.3390/catal12101248

Chicago/Turabian Style

Xue, Hongyan, Xiaoming Guo, Dongsen Mao, Tao Meng, Jun Yu, and Zhen Ma. 2022. "Phosphotungstic Acid-Modified MnOx for Selective Catalytic Reduction of NOx with NH3" Catalysts 12, no. 10: 1248. https://doi.org/10.3390/catal12101248

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