Next Article in Journal
3D-Printed Raney-Cu POCS as Promising New Catalysts for Methanol Synthesis
Next Article in Special Issue
Protection Effect of Ammonia on CeNbTi NH3-SCR Catalyst from SO2 Poisoning
Previous Article in Journal
Photocatalytic Degradation of Eriochrome Black-T Using BaWO4/MoS2 Composite
Previous Article in Special Issue
Contributions of Washcoat Components in Different Configurations to the NOX and Oxygen Storage Performance of LNT Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Abundant Oxygen Vacancies Induced by the Mechanochemical Process Boost the Low-Temperature Catalytic Performance of MnO2 in NH3-SCR

1
Key Laboratory of Resources Chemicals and Materials, Shenyang University of Chemical Technology, Ministry of Education, Shenyang 110142, China
2
State Key Laboratory of Multiphase Complex Systems, Institute of Process Engineering, Chinese Academy of Sciences, Beijing 100190, China
3
Institute of Sustainability for Chemicals, Energy and Environment, Agency for Science, Technology and Research (A*STAR), 1 Pesek Road, Jurong Island, Singapore 627833, Singapore
4
Institute of Industrial Chemistry and Energy Technology, Shenyang University of Chemical Technology, Shenyang 110142, China
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(10), 1291; https://doi.org/10.3390/catal12101291
Submission received: 22 September 2022 / Revised: 13 October 2022 / Accepted: 19 October 2022 / Published: 21 October 2022

Abstract

:
Manganese oxides (MnOx) have attracted particular attention in the selective catalytic reduction of NOx with NH3 (NH3-SCR) because of their excellent low-temperature activity. Herein, we prepared a highly efficient MnO2 (MnO2-M) catalyst through a facile ball milling-assisted redox strategy. MnO2-M shows a 90% NOx conversion in a wide operating temperature window of 75–200 °C under a gas hourly space velocity of 40,000 h−1, which is much more active than the MnO2 catalyst prepared by the redox method without the ball-milling process. Moreover, MnO2-M exhibits better H2O and SO2 resistance. The enhanced catalytic properties of MnO2-M originated from the high surface area, abundant oxygen vacancies, more acid sites, and higher Mn4+ content induced by the ball-milling process. In situ DRIFTS studies probed the reaction intermediates, and the SCR reaction was deduced to proceed via the typical Eley–Rideal mechanism. This work provides a facile method to enhance the catalytic performance of Mn-based catalysts for low-temperature denitrification and deep insights into the NH3-SCR reaction process.

1. Introduction

One of the most significant factors causing environmental pollution is NOx emitted by burning fossil fuels, leading to a series of environmental concerns such as ozone depletion, acid rain, and photochemical smog [1,2]. At present, selective catalytic reduction by NH3 (NH3-SCR) is one of the effective technologies to reduce NOx [3,4,5], typically using the V2O5-WO3/TiO2 and V2O5-MoO3/TiO2 catalysts [6,7]. However, the V-based catalysts have some disadvantages, e.g., the narrow working temperature window and biological toxicity. Additionally, their operating temperature is required to be above 300 ℃, which is energy-consuming [8,9]. Therefore, it is essential to develop novel non-V-based SCR catalysts with high activity at low temperatures [10,11,12].
Mn-based catalysts have been widely investigated for low-temperature NH3-SCR due to their high catalytic activity, good redox properties, relatively low toxicities, and low cost [13,14]. In the case of pure MnOx catalysts, their different valence states typically affect the catalytic activity. Yang et al. [15] found that MnO2 exhibited higher NO conversion at 75–250 °C than Mn2O3 and Mn3O4 because of the higher valence state of Mn in MnO2. In addition, the crystallinity and surface area of MnOx can also affect its catalytic performance. Tang et al. [16] found that amorphous MnOx with a higher specific area had higher activity than highly crystallized MnOx. Moreover, the NH3-SCR reaction involves the participation of reactive oxygen species, e.g., NO oxidation to NO2. Oxygen vacancies (OVs) are also helpful in improving the catalytic activity of MnOx [17]. For example, Liu et al. [18] reported that mesoporous α-MnO2 nanosheets with more OVs reached 100% NO conversion under a gas hourly space velocity (GHSV) of 700,000 h−1 at 100 °C.
Several strategies have been developed to generate OVs in metal oxides. For example, Chen et al. [19] employed aqueous NaBH4 to reduce Co3O4, and OVs were generated during the surface reduction process. Sun et al. [20] found that the post-acid-etching process effectively made OVs in MnOx. Hilpert et al. [21] prepared LaCrO3-δ with many OVs by exposing LaCrO3 to oxidizing and reducing atmospheres under 900–1100 °C operating conditions. Yim et al. [22] used ion sputtering and annealing reduction method to obtain rutile TiO2 with high OVs concentration. Although these approaches have proven effective, they are relatively complicated and costly. Increasing the OVs concentration of MnOx through simple and economical alternative routes to improve its catalytic performance is essential but challenging.
Herein, we report a facile ball milling-assisted redox strategy to obtain a high-performing MnO2 catalyst (MnO2-M) by introducing abundant OVs. In contrast to the MnO2 catalyst obtained by the conventional redox method (MnO2), MnO2-M exhibits much higher activity in the NH3-SCR reaction. This is mainly attributed to the abundant OVs, more acid sites, and higher Mn4+ content induced by the ball-milling process. The corresponding structure-activity relationship and reaction mechanism was elucidated by investigating the catalyst structure in detail. This work demonstrates significant progress for the practical applications of pure MnOx in NOx elimination at low temperatures.

2. Results and Discussion

2.1. Characterization of the Catalysts

2.1.1. XRD and PSD Analysis

Figure 1a shows the XRD patterns of the catalysts. In the pattern of MnO2-M, three prominent diffraction peaks appear at 37.5°, 42.0°, and 57.0°, which can be indexed as the (211), (301), and (431) planes of α-MnO2 (PDF#44-0141), respectively [23]. Compared with the MnO2-M catalyst, MnO2 shows more diffraction peaks. These diffraction peaks match well with the standard pattern of the α-MnO2 phase (PDF#44-0141), where the stronger diffraction peaks at 2θ = 28.8°, 37.5°, and 42.0° can be ascribed to the (310), (211), and (301) planes of the MnO2, respectively [24,25]. It should be noted that the XRD pattern of the MnO2-M catalyst shows significantly broadened diffraction peaks with poor crystallinity, indicating that its crystal growth is suppressed during the ball milling-assisted redox process. Indeed, M-MnO2 obtained by the ball-milling of MnO2 for 1 h also presents lower crystallinity and smaller particle size, indicating that the ball-milling process does affect the structure of the catalyst. The crystal sizes of the MnO2-M catalyst based on the Debye–Scherrer equation calculated from the strongest diffraction peak at 2θ = 37.5° is 4.92 nm, while that at 2θ = 28.8° calculated at 27.17 nm for the MnO2 catalyst (Table 1). The particle size of the catalysts was measured using a laser particle size analyzer, and the results are shown in Figure 1b. The size of MnO2 is centered at 1–20 µm. However, the MnO2-M catalyst has different distributions at 0.2–2, 2–6, and 6–20 µm, suggesting that both small and large particles coexist. The mixture of particles with different sizes benefits the formation of inter-particle pores.

2.1.2. N2 Adsorption Measurements

The BET surface area and total pore volume of the catalysts are listed in Table 1. The MnO2-M catalyst possesses a larger surface area (177.9 vs. 66.4 m2/g) and pore volume (0.335 vs. 0.149 cm3/g) than the MnO2 catalyst. The larger surface area may be because the MnO2-M catalyst contains abundant inter-particle pores, which can also be deduced from the PSD results. To interpret such a difference, the N2 adsorption–desorption isotherms and pore size distribution of the two catalysts are also shown in Figure 2. The two catalysts display the type IV adsorption isotherms with H3-type hysteresis, indicating the formation of a porous structure [26]. As seen in Figure 2b, the pore size distributions of the two catalysts obtained from nitrogen adsorption are both in the range of 5–30 nm. However, the peak intensity of the pore size distribution in the MnO2-M catalyst is much higher than in the MnO2 catalyst, which means that more pores, including large and small sizes, exist in the MnO2-M catalyst.

2.1.3. SEM Observation

Figure 3 depicts the typical SEM images of MnO2-M and MnO2. The MnO2-M catalyst exhibits irregular particles of varying sizes, while the MnO2 catalyst displays a unique waxberry-like shape with consistent dimensions. Compared with the MnO2 catalyst containing particles in close contact, the particles in the MnO2-M catalyst are in loose connection, which is more conducive to forming inter-particle pores. This is also evidenced by the PSD result (Figure 1b) and N2 adsorption result (Figure 2). In short, inter-particle pores can increase the specific surface area of the catalyst and provide more active sites between the catalyst and gas-phase reactants.

2.1.4. H2-TPR Analysis

It is well known that the redox property of catalysts is a critical factor for the NH3-SCR reaction [27]. Therefore, we investigated the reduction behavior of the catalysts by H2-TPR test, and the results are presented in Figure 4. For the two catalysts, there are two obvious reduction peaks located in the temperature range of 213–400 ℃ (I) and 400–533 ℃ (II), which can be assigned to the successive reduction of MnO2 → Mn2O3/Mn3O4 → MnO [28,29]. It should be noted that the reduction of MnO2-M catalyst starts at a lower temperature but ends up at a higher temperature. In contrast, the two reduction peaks of the MnO2 catalyst overlap, and the peaks are more concentrated. This difference between the two catalysts may be attributed to their different particle size distributions. Smaller particles in MnO2-M catalyst are easier to be reduced at a low temperature, while larger particles are more difficult to be reduced because of limited mass transfer. Therefore, the two reduction peaks of the MnO2-M catalyst have a larger temperature gap due to the varying particle size. By contrast, the uniform particle size in the MnO2 catalyst leads to a narrow reduction peak. This result indicates that the ball-milling process can create small particles and improve the reduction potentials at low temperature. Furthermore, quantitative analysis data of H2-TPR are listed in Table 2. Although the two catalysts in the H2-TPR experiments are reduced to green MnO, the experimental H2 consumption of the catalysts is significantly less than the theoretical consumption (11.5 mmol/g). This may be due to some manganese in the Mn3+ oxidation state in catalysts. In addition, the MnO2-M catalyst shows more H2 consumption than the MnO2 catalyst, which indicates that there exists more of an Mn4+ oxidation state than MnO2. It has been widely reported that Mn4+ is the most effective oxidation state for NOx conversion [30]. Therefore, the MnO2-M catalyst can be a significant advantage in NH3-SCR compared with MnO2.

2.1.5. XPS Analysis

Figure 5 shows the XPS Mn 2p and O 1s spectra of MnO2-M and MnO2 catalysts. The Mn 2p3/2 spectra can be separated into three characteristic peaks attributed to Mn2+ (641.0 eV), Mn3+ (642. 4 eV), and Mn4+ (643.5 eV), respectively (Figure 5a), by performing peak-fitting deconvolutions [31,32,33,34]. The Mn4+/(Mn3+ + Mn2+) atomic ratio (Table 3) was calculated from XPS spectra. The relative surface content of Mn4+/(Mn3+ + Mn2+) on MnO2-M (0.81) is much higher than that of MnO2 (0.56). Therefore, much more Mn4+ is exposed on the surface of the MnO2-M catalyst, which is beneficial for NO oxidation to NO2, leading to high activity in the NH3-SCR reaction at low temperatures.
Two peaks can be distinguished in the O 1s spectra of the catalysts (Figure 5b). The peak at 529.9 eV is assigned to the lattice oxygen (marked as Olatt), and the peak at 531.3 eV is ascribed to the surface adsorbed oxygen (marked as Oads, O22− or O belonging to defect-oxide or hydroxyl-like group) [35,36]. It is well known that the Oads species are more active than the Olatt species due to their higher mobility [37]. From Table 3, the Oads/Olatt ratio in MnO2-M (1.17) is much higher than that in MnO2 (0.60), suggesting that the ball-milling process could produce much more OVs in the MnO2-M catalyst. The above results show that the MnO2-M catalyst has a higher Mn4+/(Mn3+ + Mn2+) content and more surface oxygen species than the MnO2 catalyst, which is critical to the NH3-SCR reaction.

2.1.6. NH3-TPD Analysis

The surface acidity of the catalyst also plays a vital role in the low-temperature SCR of NO by NH3. Therefore, NH3-TPD was carried out to evaluate the acidity of the catalysts, as shown in Figure 6. The detailed ammonia desorption parameters of the catalysts are summarized in Table 4. From Figure 6, both MnO2-M and MnO2 catalysts exhibit one broad desorption peak below 300 ℃, which can be attributed to the subsequent desorption of NH3 coordinated with weak acid sites [36]. As shown in Table 4, the total amount of desorbed NH3 for MnO2-M (0.43 μmol g−1) is higher than that of MnO2 (0.32 μmol g−1), possibly due to the larger specific surface area of the MnO2-M catalyst. The superiority of acid sites on the MnO2-M catalyst could promote the adsorption and activation of NH3, which favors the SCR reaction [38].

2.2. Formation Mechanism of MnO2-M

Previously, studies reported that MnO2 with different morphology structures could be obtained by tailoring manganese precursors, such as MnCl2, MnSO4⋅H2O, Mn(NO3)2 H2O, and KMnO4 [39]. Here, we used Mn(NO3)2 solution and KMnO4 as precursors. Mn precursors underwent an oxidation-reduction reaction (2KMnO4 + 3Mn(NO3)2 + 2H2O = 5MnO2 +HNO3 + 2KNO3) to generate MnO2 sediment (Scheme 1). Ball-milling as a mechanical technique is widely used to grind powders into fine particles. When ball-milling was introduced into the catalyst preparation process, these Mn2+ and Mn7+ precursors could quickly react and convert to irregular MnO2-M particles with uneven particle size. The crystal cannot grow due to inadequate time, evidenced by SEM and PSD results. The irregular particles with variable sizes in MnO2-M benefit the formation of inter-particle pores, thereby increasing the surface area of the catalyst. The higher the surface area, the more active sites are available for NH3-SCR reaction. Considering the ball milling method is an environmentally friendly and cost-effective technique, the facile ball milling-assisted synthesized strategy provides the potential for the scaled-up production of MnOx catalysts in industrial applications.

2.3. Catalytic Performance

Figure 7a shows the NOx conversion as a function of temperature in the NH3-SCR reaction on MnO2-M and MnO2 catalysts. MnO2-M exhibits higher activity with NOx conversion of 68% at 50 °C and 100% at 100 °C. In contrast, only 49.8% NOx conversion is obtained at 50 °C on MnO2, and 100% NOx conversion is not achieved even at 200 °C. In addition, the N2 selectivity of the catalysts was evaluated, and the results are shown in Figure 7b. Both the catalysts exhibit an acceptable and similar N2 selectivity below 150 °C.
Based on the results of BET, H2-TPR, XPS, and NH3-TPD, the much higher NOx conversion on MnO2-M can be attributed to its higher specific surface area, more adsorbed surface oxygen, higher reducibility, and more acid sites. It should be noted that the preparation process of the two catalysts is similar, except that the latter uses ball milling. Therefore, it can be deduced that the ball-milling process may be the key to the optimization of catalyst structure and the improvement in the SCR performance. To verify this, we applied the ball milling method to the MnO2 catalyst and obtained the M-MnO2 catalyst. As shown in Figure 8, the M-MnO2 catalyst presents an NOx conversion of 55% at 50 °C, 78% at 75 °C, and 97% at 100 °C, suggesting a much better NH3-SCR performance than the pristine MnO2.
The resistance to H2O and SO2 under reaction conditions is important for low-temperature SCR catalysts. The H2O and SO2 tolerance tests of the MnO2-M and MnO2 were conducted at 150 °C. Figure 9a shows the H2O tolerance results of the catalysts. When 10 vol.% H2O is added to the reaction gas, the NOx conversion on the MnO2-M catalyst changes from nearly 100% at 4 h to 85.5% at 14 h, and that of MnO2 decreases from nearly 100% at 4 h to 76.9% at 14 h, suggesting that the MnO2-M catalyst shows better resistance to H2O than MnO2. When H2O is cut off, its catalytic activities recover quickly. In particular, the MnO2-M catalyst can be fully recovered to the original level. Therefore, the activity decreases in the two catalysts caused by H2O originated from competitive adsorption of H2O and NH3/NOx for active sites, but not the deactivation of active sites. As shown in Figure 9b, further introduction of SO2 to the reaction gas significantly decreased the initial NOx conversion over the two catalysts compared to the case of only H2O. In addition, the NOx conversions of the two catalysts decrease gradually along with reaction time. However, the NOx conversion on the MnO2-M catalyst is still much higher than that on MnO2. Moreover, when the H2O and SO2 are removed from the reaction gas, the NOx conversions of the MnO2-M and MnO2 catalysts cannot recover, which may be due to the formation of sulfating Mn species. Even so, the MnO2-M catalyst could achieve more than 88.4% NOx conversion after the process, implying that it has good resistance to H2O and SO2.

2.4. In Situ DRIFTS Studies

To obtain insights about the reaction mechanism, in situ DRIFTS studies were conducted. In situ DRIFTS spectra of MnO2-M and MnO2 catalysts exposed to NO + O2 at 50 °C as a function of time are shown in Figure 10. Four main bands over the two catalysts appear after introducing NO + O2 into the IR cell. For the MnO2-M catalyst, the band located at 1026 cm−1 is ascribed to monodentate nitrite, while those at 1539 and 1271 cm−1 are assigned to bidentate nitrate. The band at 1630 cm−1 is attributed to the NO2 ad-species (nitro group or adsorbed NO2 molecule) [40,41]. The intensity of these peaks becomes stronger with the increase in exposure time. On the MnO2 catalyst, the bands assigned to monodentate nitrite and the NO2 ad-species at 1031 and 1620 cm−1 are also detected. Additionally, the bands that are indexed to monodentate nitrates (1313 cm−1) and ionic nitrate (1462 cm−1) appear on the MnO2 catalyst. However, those bands corresponding to bidentate nitrate disappear. The above results show that more stable nitrate exists, e.g., the bidentate nitrate on the MnO2-M catalyst, which further proves that the formation of more OVs on the MnO2-M catalyst promotes NH3-SCR performance.
To investigate the nature of the adsorbed NH3 species and potential intermediates, in situ DRIFTS spectra of the two catalysts exposed in the NH3 atmosphere at 50 °C were collected, as shown in Figure 10c,d. The bands in 3450–3000 cm−1 can be assigned to N-H stretching vibration of ordinated NH3 adsorbed on Lewis sites [42,43]. The corresponding bending vibrations are observed at 1170/1178 cm−1 [42,43,44]. With the increasing adsorption time, the bands of the two catalysts increase until the adsorption is saturated. For the MnO2-M catalyst, the band at 1597 cm−1 is assigned to NH3 coordinately linked to the Lewis acid sites, while the band at 1388 cm−1 can be ascribed to monodentate nitrites. In Figure 10d, these bands show a little shift over the MnO2 catalyst, e.g., from 1597 cm−1 and 1388 cm−1 to 1608 cm−1 and 1398 cm−1, respectively. Compared with the MnO2 catalyst, a new band at 1641 cm−1 corresponding to bridging nitrates appears. This indicates that partial NH3 at Lewis acid sites can be oxidized to bidentate nitrate and monodentate nitrites. The MnO2-M catalyst possesses stronger acid intensity to absorb the ammonia species, further explaining the higher SCR activity of the MnO2-M catalyst.
To further investigate the SCR reaction mechanism over the two catalysts, in situ DRIFTS spectra in a flow of NO + O2 + NH3 were conducted at different temperatures. As shown in Figure 11, raising the temperature decreased the intensity of all the bands. For the MnO2-M catalyst at 50 ℃, the band at 1196 cm−1 is attributed to N-H bending vibrations of NH3 adsorbed on Lewis sites. The bands at 1319, 1602, and 3200–3400 cm−1 are attributed to ordinated NH3 adsorbed on Lewis sites. And the band at 1319 cm−1 shifts to lower wavenumbers with the increasing temperature, eventually reaching 1278 cm−1 at 300 ℃. The bands at 1394 cm−1 were attributed to monodentate nitrate [40,45]. In situ DRIFTS spectra of the MnO2 catalyst, as shown in Figure 11c, are similar to those obtained over the MnO2-M catalyst. However, the intensities of the bands on MnO2 are much lower than that on the MnO2-M catalyst, possibly due to the less active sites in the MnO2 catalyst.
The above results suggest that the adsorption of NH3 is dominated under the SCR reaction conditions for the two catalysts. The possible reaction mechanism over MnO2-M at low temperature can be described in Scheme 2. The NO molecules in the gas phase are adsorbed on the catalyst surface and react with O2 with OVs’ assistance to form NO2. Due to Lewis acidity, the NH3 molecules are adsorbed onto the MnO2-M surface and form NH3(ad). Then, the formed NH3(ad) can further react with NOx to generate N2 and H2O. Therefore, the SCR reaction on MnO2-M proceeds via the typical Eley–Rideal (ER) mechanism.

3. Experiment

3.1. Chemical Reagents

Manganese nitrate solution (Mn(NO3)2, 50% in H2O) was purchased from Shanghai Macklin Biochemical Co., Ltd., Shanghai, China and potassium permanganate (KMnO4) was purchased from Beijing Chemical Works Co., Ltd., Beijing, China. All the chemicals were analytical grade and used as received without further treatment. Deionized water was used in all experiments.

3.2. Preparation of Catalysts

The Mn-based catalyst was prepared by the ball milling-assisted redox method. Scheme 3 depicts the synthesis procedure of the catalyst. In brief, 3.64 g of KMnO4 and 6.17 g of Mn(NO3)2 were added to an agate jar (100 mL) along with some zirconia balls (5 mm diameter). Here, the ball to powder mass ratio was 10:1. The agate jar was placed in a high-speed ball milling apparatus (Planetary Ball Mill QM-3SP04, Nanjing Chishun Sclence And Technology Development Co., Ltd, Nanjing, China), and the reactants were ball-milling for 1 h at a frequency of 580 r/min. The resulting solid was washed several times with deionized water, dried overnight at 105 °C, and calcined at 400 °C for 3 h to obtain the Mn-based catalyst. The prepared catalyst was denoted as MnO2-M.
For comparison, the MnO2 catalyst was prepared without ball milling. First, 3.64 g of KMnO4 was dissolved in 100 mL of deionized water to obtain solution A. Next, 6.17 g of Mn(NO3)2 was dissolved in 100 mL of deionized water to obtain solution B. Subsequently, solution B was added dropwise into solution A under stirring at room temperature, and the obtained mixture was stirred for 4 h to form a slurry. Then, the slurry was filtered and washed several times with deionized water. The filter cake was dried overnight at 105 °C and calcined at 400 °C for 3 h. The obtained catalyst was named MnO2. For comparison with MnO2 and MnO2-M, part of the synthesized MnO2 catalyst was further ball-milled for 1 h, obtaining the reference catalyst named M-MnO2.

3.3. Characterization

X-ray diffraction (XRD) patterns were recorded using a Cu Kα radiation (λ = 1.5148 Å) range from 10.0 to 90.0° (X’Pert PRO MPD, PANalytical, Almelo, the Netherlands). The scanning electron micrograph (SEM) images were obtained at 5.0 kV (Regulus 8100, JEOL, Tokyo, Japan). The particle size distribution (PSD) was measured using a laser particle size analyzer (BT-9300Z, Bettersize Instruments Ltd., Dandong, China). The N2 adsorption-desorption isotherms were measured on a surface area and pore size analyzer (NOVA 3200e, Quantachrome, FL, USA). The specific surface areas of the catalysts were determined by using the Brunauer–Emmett–Teller (BET) method. The catalysts were degassed at 200 °C for 4 h and analyzed at −196 °C. The pore volume and pore size distribution were measured from the desorption branch using the Barrett–Joyner–Halenda (BJH) method.
H2 temperature-programmed reduction (H2-TPR) experiments were carried out on an automated chemisorption analyzer (ChemBET pulsar TPR/TPD, Quantachrome, FL, USA). 0.05 g of catalyst was loaded into a quartz U-tube, then the catalyst was degassed at 150 °C for 0.5 h under helium. When the temperature was cooled to 50 °C, the gas was changed to 10% H2/Ar. Finally, the catalyst was heated from 50 to 1000 °C at 10 °C min−1 under a H2/Ar flow (30 mL min−1). X-ray photoelectron spectra (XPS) were collected to analyze the surface chemical compositions of the catalysts (Model VG ESCALAB 250 spectrometer, Thermo Electron, London, UK). Temperature-programmed desorption of ammonia (NH3-TPD) experiments were carried out on a FT-IR instrument (Nicolet 380, Thermo Fisher Scientific, Waltham, MA, USA). Before the experiment, 100 mg of catalyst was treated in 5% O2/N2 at 300 °C for 30 min. After cooling to 50 °C in the same gas mixture, the catalyst was exposed to 500 ppm NH3/N2 for 30 min and then flushed in N2 gas flow for 30 min. The NH3 desorption was recorded in N2 while heating to 500 °C at a rate of 10 °C/min. In situ diffuse reflectance infrared Fourier transform spectroscopy (in situ DRIFTS) of the catalyst was carried out with an FTIR (Nicolet iS50, Thermo Fisher Scientific, Waltham, MA, USA) equipped with an in situ diffuse reflection chamber and MCT detector. The catalysts were pretreated at 300 °C in a 5% O2/N2 atmosphere for 30 min and then gradually cooled to 50 °C. Then, the catalysts were purged with the reaction atmosphere, and the spectra were collected as a function of time.

3.4. Catalytic Measurement

The NH3-SCR activity reaction was carried out in a fixed-bed reactor operating at atmospheric pressure. Before each test, 1.5 mL of catalyst was pretreated with N2 gas flow in the reactor at 300 °C for 1 h. The inlet gas comprised NO (500 ppm), NH3 (500 ppm), O2 (5%), and N2 (the balance). The GHSV was 40,000 h−1. A FT-IR spectrometer monitored the gaseous species (NO, NH3, NO2, and N2O) in the exhaust (Nicolet 380, Thermo Fisher Scientific, Waltham, MA, USA).
The NH3-SCR stability reaction was carried out in a fixed-bed reactor operating at atmospheric pressure. Before each test, 1.5 mL of catalyst was pretreated with N2 gas flow in the reactor at 300 °C for 1 h. The inlet gas comprised NO (500 ppm), NH3 (500 ppm), O2 (5%), SO2 (50 ppm, in use), H2O (10 vol.%, in use), and N2 (the balance). The GHSV was 40,000 h−1. The NOx concentrations in the exhaust were analyzed by a multiple gas analyzer (Testo 350, Testo SE & Co. KGaA, Baden-Württemberg, Germany). The NOx conversion and N2 selectivity were calculated as follows:
NO x   conversion   ( % ) = [ NO x ] in     [ NO x ] out     2   ×   [ N 2 O ] out [ NO x ] in   ×   100
  N 2   selectivity   ( % ) = ( 1   2   ×   [ N 2 O ] out [ NH 3 ] in     [ NH 3 ] out + [ NO x ] in     [ NO x ] out )   ×   100
The “in” and “out” subscripts represent the inlet and outlet gas concentrations in the steady state, respectively.

4. Conclusions

The ball milling-assisted redox method was successfully applied to prepare the MnO2-M catalyst with a high specific surface area. Compared with the MnO2 catalyst obtained by the redox method, the MnO2-M catalyst exhibits a higher catalytic activity and better resistance against H2O and SO2 in NH3-SCR reaction at low temperatures. Based on characterization analysis, more OVs can be manufactured, which is beneficial for NO oxidation and promotes the NH3-SCR reaction. Moreover, the MnO2-M catalyst has more acid sites, higher Mn4+ content, surface adsorbed oxygen, and active sites. All these features can contribute to the high NH3-SCR performance of the MnO2-M catalyst. In situ DRIFTS studies probed the reaction intermediates and predicted that the adsorption of NH3 is dominated under the SCR reaction via the typical Eley–Rideal mechanism.

Author Contributions

Conceptualization, B.J. and F.S.; methodology, Y.D., B.J., S.L. and F.S.; validation, Y.D.; formal analysis, Y.D.; investigation, Y.D.; resources, K.W.; data curation, Y.D. and S.L.; writing—original draft preparation, Y.D.; writing—review and editing, B.J., J.G. and F.S.; visualization, Y.D., B.J. and F.S.; supervision, B.J. and F.S.; project administration, F.S.; funding acquisition, F.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

We gratefully acknowledge the financial support from the Beijing Chenxi Environmental Engineering Co., Ltd.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liu, Z.; Liu, Y.; Li, Y.; Su, H.; Ma, L. WO3 promoted Mn–Zr mixed oxide catalyst for the selective catalytic reduction of NOx with NH3. Chem. Eng. J. 2016, 283, 1044–1050. [Google Scholar] [CrossRef]
  2. Anenberg, S.C.; Miller, J.; Minjares, R.; Du, L.; Henze, D.K.; Lacey, F.; Malley, C.S.; Emberson, L.; Franco, V.; Klimont, Z.; et al. Impacts and mitigation of excess diesel-related NOx emissions in 11 major vehicle markets. Nature 2017, 545, 467–471. [Google Scholar] [CrossRef] [PubMed]
  3. Xie, S.; Tan, W.; Li, Y.; Ma, L.; Ehrlich, S.N.; Deng, J.; Xu, P.; Gao, F.; Dong, L.; Liu, F. Copper Single Atom-Triggered Niobia–Ceria Catalyst for Efficient Low-Temperature Reduction of Nitrogen Oxides. ACS Catal. 2022, 12, 2441–2453. [Google Scholar] [CrossRef]
  4. Yao, X.; Zhang, L.; Li, L.; Liu, L.; Cao, Y.; Dong, X.; Gao, F.; Deng, Y.; Tang, C.; Chen, Z.; et al. Investigation of the structure, acidity, and catalytic performance of CuO/Ti0.95Ce0.05O2 catalyst for the selective catalytic reduction of NO by NH3 at low temperature. Appl. Catal. B Environ. 2014, 150–151, 315–329. [Google Scholar] [CrossRef]
  5. Forzatti, P.; Nova, I.; Tronconi, E. Enhanced NH3 Selective Catalytic Reduction for NOx Abatement. Angew. Chem. 2009, 121, 8516–8518. [Google Scholar] [CrossRef]
  6. Xu, L.; Wang, C.; Chang, H.; Wu, Q.; Zhang, T.; Li, J. New Insight into SO2 Poisoning and Regeneration of CeO2-WO3/TiO2 and V2O5-WO3/TiO2 Catalysts for Low-Temperature NH3-SCR. Environ. Sci. Technol. 2018, 52, 7064–7071. [Google Scholar] [CrossRef]
  7. Qiu, Y.; Liu, B.; Du, J.; Tang, Q.; Liu, Z.; Liu, R.; Tao, C. The monolithic cordierite supported V2O5–MoO3/TiO2 catalyst for NH3-SCR. Chem. Eng. J. 2016, 294, 264–272. [Google Scholar] [CrossRef]
  8. Chen, H.; Xia, Y.; Fang, R.; Huang, H.; Gan, Y.; Liang, C.; Zhang, J.; Zhang, W.; Liu, X. The effects of tungsten and hydrothermal aging in promoting NH3-SCR activity on V2O5/WO3-TiO2 catalysts. Appl. Surf. Sci. 2018, 459, 639–646. [Google Scholar] [CrossRef]
  9. Beretta, A.; Lanza, A.; Lietti, L.; Alcove Clave, S.; Collier, J.; Nash, M. An investigation on the redox kinetics of NH3-SCR over a V/Mo/Ti catalyst: Evidence of a direct role of NO in the re-oxidation step. Chem. Eng. J. 2019, 359, 88–98. [Google Scholar] [CrossRef]
  10. Liu, L.; Xu, K.; Su, S.; He, L.; Qing, M.; Chi, H.; Liu, T.; Hu, S.; Wang, Y.; Xiang, J. Efficient Sm modified Mn/TiO2 catalysts for selective catalytic reduction of NO with NH3 at low temperature. Appl. Catal. A Gen. 2020, 592, 117413. [Google Scholar] [CrossRef]
  11. Zhang, N.; Yan, H.; Li, L.; Wu, R.; Song, L.; Zhang, G.; Liang, W.; He, H. Use of rare earth elements in single-atom site catalysis: A critical review—Commemorating the 100th anniversary of the birth of Academician Guangxian Xu. J. Rare Earths 2021, 39, 233–242. [Google Scholar] [CrossRef]
  12. Jia, L.; Liu, J.; Huang, D.; Zhao, J.; Zhang, J.; Li, K.; Li, Z.; Zhu, W.; Zhao, Z.; Liu, J. Interface Engineering of a Bifunctional Cu-SSZ-13@CZO Core–Shell Catalyst for Boosting Potassium Ion and SO2 Tolerance. ACS Catal. 2022, 12, 11281–11293. [Google Scholar] [CrossRef]
  13. Chen, L.; Li, R.; Li, Z.; Yuan, F.; Niu, X.; Zhu, Y. Effect of Ni doping in NixMn1xTi10 (x = 0.1–0.5) on activity and SO2 resistance for NH3-SCR of NO studied with in situ DRIFTS. Catal. Sci. Technol. 2017, 7, 3243–3257. [Google Scholar] [CrossRef]
  14. Raja, S.; Alphin, M.S.; Sivachandiran, L.; Singh, P.; Damma, D.; Smirniotis, P.G. TiO2-carbon nanotubes composite supported MnOx-CuO catalyst for low-temperature NH3-SCR of NO: Investigation of SO2 and H2O tolerance. Fuel 2022, 307, 121886. [Google Scholar] [CrossRef]
  15. Yang, J.; Ren, S.; Zhou, Y.; Su, Z.; Yao, L.; Cao, J.; Jiang, L.; Hu, G.; Kong, M.; Yang, J.; et al. In situ IR comparative study on N2O formation pathways over different valence states manganese oxides catalysts during NH3–SCR of NO. Chem. Eng. J. 2020, 397, 125446. [Google Scholar] [CrossRef]
  16. Tang, X.; Hao, J.; Xu, W.; Li, J. Low temperature selective catalytic reduction of NO with NH3 over amorphous MnOx catalysts prepared by three methods. Catal. Commun. 2007, 8, 329–334. [Google Scholar] [CrossRef]
  17. Fang, X.; Liu, Y.; Cen, W.; Cheng, Y. Birnessite as a Highly Efficient Catalyst for Low-Temperature NH3-SCR: The Vital Role of Surface Oxygen Vacancies. Ind. Eng. Chem. Res. 2020, 59, 14606–14615. [Google Scholar] [CrossRef]
  18. Liu, J.; Wei, Y.; Li, P.-Z.; Zhang, P.; Su, W.; Sun, Y.; Zou, R.; Zhao, Y. Experimental and Theoretical Investigation of Mesoporous MnO2 Nanosheets with Oxygen Vacancies for High-Efficiency Catalytic DeNOx. ACS Catal. 2018, 8, 3865–3874. [Google Scholar] [CrossRef]
  19. Chen, H.; Yang, M.; Tao, S.; Chen, G. Oxygen vacancy enhanced catalytic activity of reduced Co3O4 towards p-nitrophenol reduction. Appl. Catal. B Environ. 2017, 209, 648–656. [Google Scholar] [CrossRef]
  20. Sun, M.; Li, W.; Zhang, B.; Cheng, G.; Lan, B.; Ye, F.; Zheng, Y.; Cheng, X.; Yu, L. Enhanced catalytic performance by oxygen vacancy and active interface originated from facile reduction of OMS-2. Chem. Eng. J. 2018, 331, 626–635. [Google Scholar] [CrossRef]
  21. Hilpert, K.; Steinbrech, R.W.; Boroomand, F.; Wessel, E.; Meschke, F.; Zuev, A.; Teller, O.; Nickel, H.; Singheiser, L. Defect formation and mechanical stability of perovskites based on LaCrO3 for solid oxide fuel cells (SOFC). J. Eur. Ceram. Soc. 2003, 23, 3009–3020. [Google Scholar] [CrossRef]
  22. Yim, C.M.; Pang, C.L.; Thornton, G. Oxygen vacancy origin of the surface band-gap state of TiO2(110). Phys. Rev. Lett. 2010, 104, 036806. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Wu, S.; Liu, H.; Huang, Z.; Xu, H.; Shen, W. O-vacancy-rich porous MnO2 nanosheets as highly efficient catalysts for propane catalytic oxidation. Appl. Catal. B Environ. 2022, 312, 121387. [Google Scholar] [CrossRef]
  24. Ji, J.; Lu, X.; Chen, C.; He, M.; Huang, H. Potassium-modulated δ-MnO2 as robust catalysts for formaldehyde oxidation at room temperature. Appl. Catal. B Environ. 2020, 260, 118210. [Google Scholar] [CrossRef]
  25. Chen, L.; Ren, S.; Xing, X.; Yang, J.; Li, J.; Yang, J.; Liu, Q. Effect of MnO2 crystal types on CeO2@MnO2 oxides catalysts for low-temperature NH3-SCR. J. Environ. Chem. Eng. 2022, 10, 108239. [Google Scholar] [CrossRef]
  26. Huang, N.; Qu, Z.; Dong, C.; Qin, Y.; Duan, X. Superior performance of α@β-MnO2 for the toluene oxidation: Active interface and oxygen vacancy. Appl. Catal. A Gen. 2018, 560, 195–205. [Google Scholar] [CrossRef]
  27. Wang, Y.; Li, Z.; Ding, Z.; Kang, N.; Fan, R.; Wang, Y.; Zhang, C.; Guo, X.; Wang, R. Effect of Ion-Exchange Sequences on Catalytic Performance of Cerium-Modified Cu-SSZ-13 Catalysts for NH3-SCR. Catalysts 2021, 11, 997. [Google Scholar] [CrossRef]
  28. Andreoli, S.; Deorsola, F.A.; Pirone, R. MnOx-CeO2 catalysts synthesized by solution combustion synthesis for the low-temperature NH3-SCR. Catal. Today 2015, 253, 199–206. [Google Scholar] [CrossRef]
  29. Li, S.; Huang, B.; Yu, C. A CeO2-MnOx core-shell catalyst for low-temperature NH3-SCR of NO. Catal. Commun. 2017, 98, 47–51. [Google Scholar] [CrossRef]
  30. Duan, X.; Dou, J.; Zhao, Y.; Khoshk Rish, S.; Yu, J. A Study on Mn-Fe Catalysts Supported on Coal Fly Ash for Low-Temperature Selective Catalytic Reduction of NOx in Flue Gas. Catalysts 2020, 10, 1399. [Google Scholar] [CrossRef]
  31. Chen, C.; Xie, H.; He, P.; Liu, X.; Yang, C.; Wang, N.; Ge, C. Comparison of low-temperature catalytic activity and H2O/SO2 resistance of the Ce-Mn/TiO2 NH3-SCR catalysts prepared by the reverse co-precipitation, co-precipitation and impregnation method. Appl. Surf. Sci. 2022, 571, 151285. [Google Scholar] [CrossRef]
  32. 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]
  33. Liu, S.; Ji, Y.; Xu, W.; Zhang, J.; Jiang, R.; Li, L.; Jia, L.; Zhong, Z.; Xu, G.; Zhu, T.; et al. Hierarchically interconnected porous MnxCo3-xO4 spinels for Low-temperature catalytic reduction of NO by CO. J. Catal. 2022, 406, 72–86. [Google Scholar] [CrossRef]
  34. Liu, Y.-z.; Guo, R.-t.; Duan, C.-p.; Wu, G.-l.; Miao, Y.-f.; Gu, J.-w.; Pan, W.-g. A highly effective urchin-like MnCrOx catalyst for the selective catalytic reduction of NOx with NH3. Fuel 2020, 271, 117667. [Google Scholar] [CrossRef]
  35. Xie, S.; Li, L.; Jin, L.; Wu, Y.; Liu, H.; Qin, Q.; Wei, X.; Liu, J.; Dong, L.; 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]
  36. Chen, Z.; Ren, S.; Wang, M.; Yang, J.; Chen, L.; Liu, W.; Liu, Q.; Su, B. Insights into samarium doping effects on catalytic activity and SO2 tolerance of MnFeOx catalyst for low-temperature NH3-SCR reaction. Fuel 2022, 321, 124113. [Google Scholar] [CrossRef]
  37. Yuan, H.; Sun, N.; Chen, J.; Jin, J.; Wang, H.; Hu, P. Insight into the NH3-Assisted Selective Catalytic Reduction of NO on β-MnO2(110): Reaction Mechanism, Activity Descriptor, and Evolution from a Pristine State to a Steady State. ACS Catal. 2018, 8, 9269–9279. [Google Scholar] [CrossRef]
  38. Meng, Z.; Yan, X.; Cui, M.; Hou, L.; Wu, W. Study on denitration performance of mineral catalyst prepared by high energy ball milling of rare earth concentrate supported Fe2O3 for NH3-SCR. Mater. Chem. Phys. 2022, 276, 125248. [Google Scholar] [CrossRef]
  39. Xie, Y.; Yu, Y.; Gong, X.; Guo, Y.; Guo, Y.; Wang, Y.; Lu, G. Effect of the crystal plane figure on the catalytic performance of MnO2 for the total oxidation of propane. CrystEngComm 2015, 17, 3005–3014. [Google Scholar] [CrossRef]
  40. Li, Q.; Gu, H.; Li, P.; Zhou, Y.; Liu, Y.; Qi, Z.; Xin, Y.; Zhang, Z. In situ IR studies of selective catalytic reduction of NO with NH3 on Ce-Ti amorphous oxides. Chin. J. Catal. 2014, 35, 1289–1298. [Google Scholar] [CrossRef]
  41. Wang, W.; McCool, G.; Kapur, N.; Yuan, G.; Shan, B.; Nguyen, M.; Graham, U.M.; Davis, B.H.; Jacobs, G.; Cho, K.; et al. Mixed-phase oxide catalyst based on Mn-mullite (Sm, Gd)Mn2O5 for NO oxidation in diesel exhaust. Science 2012, 337, 832–835. [Google Scholar] [CrossRef] [PubMed]
  42. 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] [Green Version]
  43. Kong, M.; Liu, Q.; Zhou, J.; Jiang, L.; Tian, Y.; Yang, J.; Ren, S.; Li, J. Effect of different potassium species on the deactivation of V2O5-WO3/TiO2 SCR catalyst: Comparison of K2SO4, KCl and K2O. Chem. Eng. J. 2018, 348, 637–643. [Google Scholar] [CrossRef]
  44. Zhu, M.; Lai, J.K.; Tumuluri, U.; Wu, Z.; Wachs, I.E. Nature of Active Sites and Surface Intermediates during SCR of NO with NH3 by Supported V2O5-WO3/TiO2 Catalysts. J. Am. Chem. Soc. 2017, 139, 15624–15627. [Google Scholar] [CrossRef] [PubMed]
  45. Hu, X.; Shi, Q.; Zhang, H.; Wang, P.; Zhan, S.; Li, Y. NH3-SCR performance improvement over Mo modified Mo(x)-MnOx nanorods at low temperatures. Catal. Today 2017, 297, 17–26. [Google Scholar] [CrossRef]
Figure 1. XRD patterns (a) and PSD curves (b) of the catalysts.
Figure 1. XRD patterns (a) and PSD curves (b) of the catalysts.
Catalysts 12 01291 g001
Figure 2. N2 adsorption-desorption isotherms (a) and pore size distribution curves (b) of the catalysts.
Figure 2. N2 adsorption-desorption isotherms (a) and pore size distribution curves (b) of the catalysts.
Catalysts 12 01291 g002
Figure 3. SEM images of MnO2-M (a,b) and MnO2 (c,d).
Figure 3. SEM images of MnO2-M (a,b) and MnO2 (c,d).
Catalysts 12 01291 g003
Figure 4. H2-TPR curves of the catalysts.
Figure 4. H2-TPR curves of the catalysts.
Catalysts 12 01291 g004
Figure 5. XPS spectra: Mn 2p (a) and O 1s (b) of the catalysts.
Figure 5. XPS spectra: Mn 2p (a) and O 1s (b) of the catalysts.
Catalysts 12 01291 g005
Figure 6. NH3-TPD curves of the catalysts.
Figure 6. NH3-TPD curves of the catalysts.
Catalysts 12 01291 g006
Scheme 1. Schematic illustration of the preparation of catalysts.
Scheme 1. Schematic illustration of the preparation of catalysts.
Catalysts 12 01291 sch001
Figure 7. Catalytic performances of the catalysts NOx conversion (a) and N2 selectivity (b) at different temperature (reaction conditions: [NO] = 500 ppm, [NH3] = 500 ppm, [O2] = 5%, and N2 as balance gas, GHSV = 40,000 h−1).
Figure 7. Catalytic performances of the catalysts NOx conversion (a) and N2 selectivity (b) at different temperature (reaction conditions: [NO] = 500 ppm, [NH3] = 500 ppm, [O2] = 5%, and N2 as balance gas, GHSV = 40,000 h−1).
Catalysts 12 01291 g007
Figure 8. Comparison of NOx conversion of MnO2, MnO2-M, and M-MnO2 catalysts at different temperature (reaction conditions: [NO] = 500 ppm, [NH3] = 500 ppm, [O2] = 5%, and N2 as balance gas, GHSV = 40,000 h−1).
Figure 8. Comparison of NOx conversion of MnO2, MnO2-M, and M-MnO2 catalysts at different temperature (reaction conditions: [NO] = 500 ppm, [NH3] = 500 ppm, [O2] = 5%, and N2 as balance gas, GHSV = 40,000 h−1).
Catalysts 12 01291 g008
Figure 9. The effect of H2O on NOx conversion over the catalysts (a) and the effect of SO2 and H2O on NOx conversion at 150 °C (b) (reaction conditions: [NO] = 500 ppm, [NH3] = 500 ppm, [O2] = 5%, and N2 as balance gas, [H2O(g)] = 10 vol.% (in use), [SO2] = 50 ppm (in use), GHSV = 40,000 h−1).
Figure 9. The effect of H2O on NOx conversion over the catalysts (a) and the effect of SO2 and H2O on NOx conversion at 150 °C (b) (reaction conditions: [NO] = 500 ppm, [NH3] = 500 ppm, [O2] = 5%, and N2 as balance gas, [H2O(g)] = 10 vol.% (in use), [SO2] = 50 ppm (in use), GHSV = 40,000 h−1).
Catalysts 12 01291 g009
Figure 10. In situ DRIFTS spectra of adsorption of NO + O2 at 50 °C on MnO2-M (a) and MnO2 (b). In situ DRIFTS spectra of NH3 adsorption at 50 °C on MnO2-M (c) and MnO2 (d).
Figure 10. In situ DRIFTS spectra of adsorption of NO + O2 at 50 °C on MnO2-M (a) and MnO2 (b). In situ DRIFTS spectra of NH3 adsorption at 50 °C on MnO2-M (c) and MnO2 (d).
Catalysts 12 01291 g010
Figure 11. In situ DRIFTS spectra of adsorption of NO + O2 + NH3 at a different temperature over MnO2-M (a) and MnO2 (b).
Figure 11. In situ DRIFTS spectra of adsorption of NO + O2 + NH3 at a different temperature over MnO2-M (a) and MnO2 (b).
Catalysts 12 01291 g011
Scheme 2. The main reaction mechanism of the NH3-SCR over the MnO2-M catalyst.
Scheme 2. The main reaction mechanism of the NH3-SCR over the MnO2-M catalyst.
Catalysts 12 01291 sch002
Scheme 3. Synthesis procedures for the preparation of the MnO2-M catalyst.
Scheme 3. Synthesis procedures for the preparation of the MnO2-M catalyst.
Catalysts 12 01291 sch003
Table 1. Physical properties of the catalysts.
Table 1. Physical properties of the catalysts.
CatalystSBET i
(m2 g−1)
VP ii
(cm3 g−1)
Lattice Parameter iii (Å)Average Crystallite Size iv (nm)
a = bc
MnO2-M177.90.3359.79042.85444.92
MnO266.40.1499.81562.856127.17
i Surface area derived from the BET equation. ii Pore volume was obtained from the volume of nitrogen adsorbed at a relative pressure of 0.99. iii XRD results for the MnO2 phase in the catalysts, calculated by the Bragg equation. iv Estimated from the XRD diffraction peak (2θ = 37.5° for MnO2-M and 28.8° for MnO2) using the Debye–Scherrer equation.
Table 2. H2 adsorption amount of the catalysts.
Table 2. H2 adsorption amount of the catalysts.
CatalystH2 Consumption i
(mmol g−1)
Total H2
Consumption i
(mmol g−1)
Theoretical H2
Consumption i
(mmol g−1)
III
MnO2-M4.853.918.7611.50
MnO23.912.276.1811.50
i Calculated based on the H2-TPR results.
Table 3. Surface compositions of the catalysts by XPS analysis.
Table 3. Surface compositions of the catalysts by XPS analysis.
CatalystMn4+/(Mn3+ + Mn2+)Oads/Olatt
MnO2-M0.811.17
MnO20.560.60
Table 4. Quantitative analysis of NH3-TPD over the catalysts.
Table 4. Quantitative analysis of NH3-TPD over the catalysts.
CatalystSurface Acidity i (μmol g−1)
MnO2-M0.43
MnO20.32
i Calculated based on the NH3-TPD results.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Dong, Y.; Jin, B.; Liu, S.; Gao, J.; Wang, K.; Su, F. Abundant Oxygen Vacancies Induced by the Mechanochemical Process Boost the Low-Temperature Catalytic Performance of MnO2 in NH3-SCR. Catalysts 2022, 12, 1291. https://doi.org/10.3390/catal12101291

AMA Style

Dong Y, Jin B, Liu S, Gao J, Wang K, Su F. Abundant Oxygen Vacancies Induced by the Mechanochemical Process Boost the Low-Temperature Catalytic Performance of MnO2 in NH3-SCR. Catalysts. 2022; 12(10):1291. https://doi.org/10.3390/catal12101291

Chicago/Turabian Style

Dong, Yuanyuan, Baofang Jin, Shaomian Liu, Jiajian Gao, Kangjun Wang, and Fabing Su. 2022. "Abundant Oxygen Vacancies Induced by the Mechanochemical Process Boost the Low-Temperature Catalytic Performance of MnO2 in NH3-SCR" Catalysts 12, no. 10: 1291. https://doi.org/10.3390/catal12101291

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