Next Article in Journal
NiO, Fe2O3, and MoO3 Supported over SiO2 Nanocatalysts for Asphaltene Adsorption and Catalytic Decomposition: Optimization through a Simplex–Centroid Mixture Design of Experiments
Next Article in Special Issue
VOC Removal from Manure Gaseous Emissions with UV Photolysis and UV-TiO2 Photocatalysis
Previous Article in Journal
Synthesis of DHA/EPA Ethyl Esters via Lipase-Catalyzed Acidolysis Using Novozym® 435: A Kinetic Study
Previous Article in Special Issue
Analysis of Research Status of CO2 Conversion Technology Based on Bibliometrics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Promoting Effect of Ti Species in MnOx-FeOx/Silicalite-1 for the Low-Temperature NH3-SCR Reaction

1
Beijing Key Laboratory for Solid Waste Utilization and Management, Department of Energy and Resources Engineering, College of Engineering, Peking University, Beijing 100871, China
2
State Key Joint Laboratory of Environment Simulation and Pollution Control, School of Environment, Tsinghua University, Beijing 100084, China
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(5), 566; https://doi.org/10.3390/catal10050566
Submission received: 28 April 2020 / Revised: 12 May 2020 / Accepted: 13 May 2020 / Published: 19 May 2020
(This article belongs to the Special Issue Catalysis for the Removal of Gas-Phase Pollutants)

Abstract

:
Manganese and iron oxides catalysts supported on silicalite-1 and titanium silicalite-1 (TS-1) are synthesized by the wet impregnation method for the selective catalytic reduction (SCR) of NOx with NH3 (NH3-SCR), respectively. The optimized catalyst demonstrates an increased NOx conversion efficiency of 20% below 150 °C, with a space velocity of 18,000 h−1, which can be attributed to the incorporation of Ti species. The presence of Ti species enhances surface acidity and redox ability of the catalyst without changing the structure of supporter. Moreover, further researches based on in situ NH3 adsorption reveal that Lewis acid sites linked to Mn4+ on the surface have a huge influence on the improvement of denitration efficiency of the catalyst at low temperatures.

1. Introduction

Nitrogen oxides (NOx) bring out a series of environmental problems, such as photochemical pollution, acid rain and haze, which seriously threaten human health [1,2]. Nowadays, selective catalytic reduction (SCR) with NH3 has been proved to be an effective technology for NOx emission control, which has been widely used in coal-fired power plants [3,4,5]. Commercial V2O5-WO3/TiO2 catalysts are very suitable for power plants, mainly due to the flue gas there, with a high temperature (>300 °C), which can satisfy the required reaction temperature of the catalyst. However, there are still a large number of untreated NOx derived from industrial activities such as cement kilns, steel sintering and waste incinerators, where the temperature of flue gas is too low (<200 °C) to suit for these commercial catalysts [6,7]. Therefore, it is urgent to develop novel high-efficient catalysts, which have high catalytic activities at low temperatures.
Manganese-based catalysts are proved to be excellent denitration catalysts at low temperatures among the commonly used transition metal oxide catalysts [8,9,10]. Furthermore, numerous metals, like Fe, Ce and Cu, are used as active components to modify manganese-based catalysts, and for that purpose, the single manganese oxides can only maintain high denitrification activities in a narrow temperature window, accompanied by several side reactions [11,12,13,14]. In addition, proper supporters can not only provide a huge surface to disperse the active components, but can also supply abundant adsorption sites for the heterogeneous reactions [15,16]. Both Al2O3 and TiO2 have been widely investigated as common supporters in the last few decades. However, the limited specific surface area, as well as poor water and sulfur resistance of the catalysts, hinder their further developments in industrial applications [17,18]. At the same time, carbon-based materials are also applied as supporters of catalysts, because of their large surface area, unique porous structure and abundant oxygen functional groups. Nevertheless, the poor thermal stability and oxidizability of carbon-based materials are a potential safety hazard in industrial applications [19,20,21]. In recent years, zeolites have gained widespread attention as supporters in SCR studies, due to their high thermal stability, unique porous structure and absence of harmful substances [22,23,24]. Wang et al. [25] found that the optimized Mn-Fe/ZSM achieved 100% NOx conversion efficiencies from 150 °C to 250 °C. The excellent catalytic performance can be attributed to the addition of iron, which increases the levels of Mn4+ and oxygen vacancy. Nevertheless, the presence of Al species in ZSM-5 makes contributions to its strong hydrophilicity, which hinders its applications in industrial denitration, because the flue gas is usually humid. However, silicalite-1 is reported to exhibit excellent water resistance in an NH3-SCR reaction, which can be attributed to the decrease in hydrophilicity due to the absence of heteroatoms [26,27]. However, compared with ZSM-5, the surface acidity of silicalite-1 is a little bit weaker, which is the key to the low-temperature denitration. To date, Ti atoms have been reported to replace a small amount of Si atoms for improving the surface acidity and maintaining excellent water resistance, which can benefit the low-temperature denitration [28].
Herein, we introduce novel high-efficient denitration catalysts with highly dispersed MnOx and FeOx nanoparticles supported on silicalite-1 and titanium silicalite-1 (TS-1) respectively. Compared to the catalyst supported on silicalite-1, the catalyst supported on TS-1 demonstrates superior NOx conversion efficiencies in the NH3-SCR reaction at low temperatures, which can be attributed to the incorporation of Ti species, without changing the structure of zeolites. Moreover, further researches on the surface properties of catalysts including their acidity, redox ability and in-situ adsorption, reveal that the presence of Ti species promotes the formation of Mn4+ on the surface, thereby increasing the acidity and redox ability of the catalyst, which are beneficial for the low-temperature denitration.

2. Results and Discussion

2.1. Chemical Compositions and Textural Properties of Samples

The chemical compositions of the catalysts are summarized in Table 1, based on the results of ICP-OES. The mass ratios of active metal oxides are close to the initial feed ratios. The contents of Mn and Fe species make no significant difference among all the catalysts.
The X-ray diffraction (XRD) patterns of the supporters are shown in Figure 1a. All of the supporters display obvious characteristic peaks at 7.9°, 8.8°, 23.1°, 23.8° and 24.3°, which can be attributed to a typical Mobil Five Instructure (MFI) structure. The pattern of anatase at 25.4° does not appear in TS-1-30, while it is found in TS-1-20, which indicates that the excessive Ti species form a small amount of anatase in TS-1-20. The existing form of Ti species in the supporters is further confirmed by Fourier transform infrared (FT-IR). As shown in Figure 2, the bands at 433, 545–546, 799–803, 1074–1075 and 1223–1225 cm−1 can be assigned to the characteristic peaks of MFI structure; the bands at 957–958 cm−1 can be the evidence for the incorporation of Ti species into frameworks [29]. Moreover, it can be seen from Figure 1b that there is no peak relevant to metal oxides in the supported catalysts, due to the extremely low metal loading of the catalysts. However, all the supported catalysts possess clear peaks that are consistent with corresponded supporters, which reveals that the supported catalysts remain the same structure with the supporters after the impregnation process [25,30].
N2 adsorption/desorption isotherms of the samples are illustrated in Figure 3, and the corresponding textural properties of the samples are summarized in Table 2. The isotherms can be classified as type I, indicating that all the samples have microporous structures [31,32]. The specific surface area, total pore volume and average pore diameter of Silicalite-1 are 454.31 m2/g, 0.32 cm3/g and 2.85 nm, respectively. When Ti species are introduced into the structure of zeolites, both TS-1-30 and TS-1-20 show an increase of 10% in specific surface area, total pore volume and average pore diameter, which can be attributed to the difference of atom size between Ti and Si [29]. After loading MnOx and FeOx, both specific surface area and total pore volume have a decrease of 10%, while average pore diameter remains constant, indicating that most of MnOx and FeOx species coexist on the surface, without changing the textural properties of supporters, as proposed in XRD analysis [30].
The morphologies of the catalysts are further investigated by SEM and transmission electron microscopy (TEM), with the results illustrated in Figure 4. The SEM images of catalysts present a typical MFI structure with a particle size of about 150 nm. However, the incorporation of Ti species results in some nanoscale folds on the surface of the catalysts, which can be attributed to the difference of atom size between Ti and Si. The active metal oxides are highly dispersed on the surface of catalysts in the form of small particles from the TEM images. As shown in Figure 5, the element mapping by energy-disperse X-ray spectroscopy (EDS) is applied to further identify the compositions of these small particles on the surface of the catalysts (EDS mappings of the supporters are shown in Figure S1). It can be found from Figure 5 that Mn, Fe and Ti species are all uniformly dispersed around the supporters. Besides, the results of Mn and Fe species are similar, indicating that Mn and Fe species may coexist in the form of FeMnOx, due to a strong interaction between metal oxides [33,34,35]. In summary, the introduction of Ti species does not change the textural properties of catalysts, but the surface of the catalysts is wrinkled, due to the difference of atom size between Ti and Si, which may provide better active sites for the catalytic reaction.

2.2. Surface Constituent and Chemical States of Samples

Figure 6 displays the XPS spectra of Mn 2p, Fe 2p and O 1s for different catalysts, and the results are summarized in Table 3. It can be seen from Figure 6a that the Mn 2p spectra consist of two major peaks that are assigned to Mn 2p3/2 (peak around 642 eV) and 2p1/2 (peak around 654 eV). The Mn 2p3/2 peak can be fitted by three peaks (around 642 eV, 643 eV and 645 eV), which can be assigned to Mn2+, Mn3+ and Mn4+, respectively [36,37]. Figure 6b displays the spectra of Fe 2p, which contains two main peaks relevant to Fe 2p3/2 (peak around 710 eV) and 2p1/2 (peak around 725 eV). The Fe 2p3/2 can be separated into two independent peaks (around 713 eV, 710.7 eV), which can be assigned to Fe2+ and Fe3+ [13,38,39]. Regarding the oxygen species, O 1s spectra are shown in Figure 6c. The spectra can be separated into two peaks that related to the lattice oxygen (peak around 530 eV, labeled as Oα) and surface adsorbed oxygen (peak around 533 eV, labeled as Oβ) [40,41], respectively.
It is broadly reported that Mn4+ species play a major role in low-temperature NH3-SCR reactions [37,42]. Hence, the ratios of Mn4+/(Mn4+ + Mn3+ + Mn2+) over the catalysts are calculated and listed in Table 4, together with the ratios of Fe3+/(Fe3+ + Fe2+) and Oβ/(Oα+Oβ), which are related to the catalytic performance of catalysts [30,43,44]. The Mn4+/(Mn4+ + Mn3+ + Mn2+) ratio of Mn3Fe2/TS-1-30 is 34.6%, which is a significant increase compared to that of Mn3Fe2/Silicalite-1 (25.0%). The increase of the Mn4+ ratio can be attributed to the redox reactions between Mn and Ti species such as Mn3+ + Ti4+ = Mn4+ + Ti3+, Mn2+ + 2Ti4+ = Mn4+ + 2Ti3+ [40]. However, the ratio of Mn4+ remains unchanged in Mn3Fe2/TS-1-20, which can be attributed to the formation of anatase on the supporter. As mentioned above, the excessive Ti species generate anatase on the surface of the supporter, which is reported to inhibit the electron transfer between Mn species and the supporter [42]. The Fe3+/(Fe3+ + Fe2+) ratio displays a slight decrease when the Ti species are introduced into the supporter, which can be attributed to redox equilibrium between Fe and Ti species like Fe3+ + Ti3+ = Fe2+ + Ti4+ [45]. According to the previously reported literature, surface adsorbed oxygen species (Oβ) show more reactive than lattice oxygen (Oα) in the low-temperature NH3-SCR reaction, due to their outstanding mobility [46,47]. As shown in Table 4, the incorporation of Ti species increases the Oβ/(Oα + Oβ) ratio, while the excessive Ti species consume the surface oxygen mainly because of the formation of anatase. In summary, Mn3Fe2/TS-1-30 tends to show the optimal catalytic performance, due to its highest ratio of Mn4+ and surface adsorbed oxygen species.

2.3. Redox Properties of Samples

The redox properties of the catalysts are closely related to the catalytic performance in the NH3-SCR reaction. Therefore, H2 temperature-programmed reduction (H2-TPR) is performed to characterize the redox properties of the catalysts and the results are exhibited in Figure 7 and Table 5. All the catalysts display three reduction peaks between 400 and 700 °C. The reduction peaks centered at 430 °C can be correlated with the simultaneous reduction of MnO2 and Fe2O3 (i.e., MnO2 → Mn2O3, Mn2O3 → Mn3O4 and Fe2O3 → FeO) [25,48]. The peaks centered at 580 °C can be assigned to the reduction of Mn3O4, while the peaks above 660 °C can be relevant to the reduction of FeO [48]. As shown in Table 5, when Ti species are introduced into the catalysts, the total area of the reduction peaks reaches 776 a.u., which has an increase of 20% compared to that in Mn3Fe2/Silicalite-1. Since the area of the reduction peak corresponds to the H2 consumption, the redox properties of the catalysts are significantly enhanced after the incorporation of Ti species, which is beneficial to promote the NH3-SCR reaction at low temperatures.

2.4. Surface Acidity of Samples

The surface acidity is another important factor in low-temperature SCR reaction. Hence, NH3 temperature-programmed desorption (NH3-TPD) is used to analyze the surface acidity of the catalysts, and the results are exhibited in Figure 8. For Mn3Fe2/Silicalite-1, it displays two desorption peaks labeled as Ⅰ and Ⅱ, which correspond to weak (<200 °C) and medium-strong (200~400 °C) acid sites, respectively [43,49]. However, more desorption peaks can be observed in Mn3Fe2/TS-1-30 and Mn3Fe2/TS-1-20, which are related to medium-strong acid sites [43,49], indicating that the new Lewis acid sites are related with the incorporation of Ti species and thus benefit the adsorption of NH3. The total acid amounts of these catalysts are calculated from TPD results and listed in Table 6. It shows that the acid amounts of Mn3Fe2/TS-1-30 and Mn3Fe2/TS-1-20 are larger than that of Mn3Fe2/Silicalite-1. Mn3Fe2/TS-1-30, especially, exhibits a remarkable increase (about 50%) in the acid amount, compared to that of Mn3Fe2/Silicalite-1. In summary, Ti species improve the surface acidity of the catalysts, which is beneficial for the adsorption of NH3, and finally enhances the catalytic performance in the low-temperature SCR reaction.
Figure 9 displays the in situ diffuse reflectance infrared Fourier transform (in situ DRIFTS) of NH3 adsorption in a temperature range between 60 and 330 °C, to further distinguish the Brønsted acid sites (labeled as B acid) and Lewis acid sites (labeled as L acid) on the surface of the catalysts. It is widely reported that the bands at 1602–1605 cm−1, 1225–1227 cm−1, 1142–1163 cm−1 and 1059–1064 cm−1 can be attributed to asymmetric and symmetric bending vibrations of the coordinated NH3 linked to L acid sites [50,51]. The bands at 1428–1448 cm−1 and 1737–1748 cm−1 can be assigned to the NH4+ species on B acid sites [52]. The bands at 1550–1554 cm−1 and 1302–1324 cm−1 can be attributed to NH2 species, which are transformed from the coordinated NH3 [53,54]. Moreover, with the increase of temperature, the bands linked to B acid sites are found to decrease more noticeable than these linked to L acid sites, indicating that the bands assigned to L acid sites are more stable at a higher temperature [37]. Besides, by comparing the spectra of Mn3Fe2/TS-1-30 (20) with that of Mn3Fe2/Silicalite-1, the intensity of bands at 1602–1605 cm−1 and 1428–1448 cm−1 is enhanced after the incorporation of Ti species, which indicates that Mn3Fe2/TS-1-30 (20) can provide more adsorbed NH3 for SCR reaction than Mn3Fe2/Silicalite-1, which can benefit the catalytic activities.

2.5. Catalytic Performance of Samples

Figure 10a shows the catalytic activities of the catalysts under different reaction temperatures. It can be seen from the figure that Mn3Fe2/TS-1-30 exhibits a NOx conversion efficiency of 90% at 120 °C, which has an increase of almost 30% compared to that of Mn3Fe2/Silicalite-1 at the same temperature. In addition, both Mn3Fe2/TS-1-30 and Mn3Fe2/TS-1-20 show an obvious improvement in catalytic activities at low temperatures, especially below 150 °C. The optimized catalysts, Mn3Fe2/TS-1-30 can maintain steady NOx conversion efficiencies above 80% in the temperature range between 110 and 230 °C, with a space velocity of 18,000 h−1. Based on the characterization of the catalysts’ physical and chemical properties mentioned above, the improvement of low-temperature denitration activity is mainly attributed to the incorporation of Ti species, which enhances the redox properties of the catalysts and provides more surface acid sites for the adsorption of NH3.
Since zeolites are reported to be sensitive to water vapor [26,28], the H2O resistance of the catalysts is further studied at the temperature of 180 °C. As shown in Figure 10b, when 10 vol.% H2O is introduced into the reaction, Mn3Fe2/Silicalite-1 displays a decrease of about 5% in NOx conversion efficiency, while Mn3Fe2/TS-1-30 remains unaffected by water vapor. It is widely accepted that the water deactivation is mainly because of the competitive adsorption between NH3 and water vapor [16,55], and the presence of Ti species further strengthens the hydrophobicity of the catalyst [29,56], thereby improving the H2O resistance of the catalysts.

3. Materials and Methods

3.1. Preparation of Samples

Silicalite-1 and TS-1 were synthesized by a hydrothermal template method, according to the previous literature [57,58]. The tetrapropylammonium hydroxide (TPAOH, 25 wt.% aqueous solution, Beijing Innochem Technology Co., Ltd., China) was served as a microporous template, while ethyl silicate (TEOS, Tianjin Guangfu Fine Chemical Research Institute, China) and tetrabutyl titanate (TBOT, Beijing Tongguang Fine Chemical Company, China) were served as the source of silicon and titanium, respectively. The molar composition of the synthesis solution was: SiO2:TiO2:TPAOH:H2O = 1:(0~0.05):0.2:7. Firstly, TEOS was added dropwise into 12.8 g aqueous solution of TPAOH at 40 °C with constant magnetic stirring while TBOT was dissolved in 10 g isopropanol. Secondly, the hydrolysis solutions of TEOS and TBOT were mixed and then heated up to 80 °C to distill off isopropanol. Then, the mixture was transferred into a stainless steel autoclave and then kept in an oven at 160 °C for 48 h. Afterwards, the white precipitation was washed by deionized water and ethanol for three times, respectively. Finally, the products were dried at 80 °C overnight, followed by calcination at 550 °C in the air for 6 h. The prepared samples without titanium labeled as Silicalite-1, while others with feeding n (Si/Ti) of 30, 20 were labeled as TS-1-30 and TS-1-20, respectively.
MnOx-FeOx catalysts were synthesized through the wet impregnation method. Manganese nitrate (Mn (NO3)2, 50 wt.% aqueous solution, Saen Chemical Technology (Shanghai) Co., Ltd., China) and iron nitrate nonahydrate (Fe (NO3)3·9H2O, 99.9%, Beijing Tongguang Fine Chemical Company, China) were dissolved in deionized water as precursors and zeolites including Silicalite-1 and TS-1 were immersed into solution as supporters. The mass ratio of the mixture was: Mn:Fe:supporter (Silicalite-1 or TS-1) = 3:2:100, which is favorable for NH3-SCR reaction according to the previous-reported researches [25,30]. After magnetic stirring at room temperature for 2 h, the mixture was transferred into an oven at 110 °C overnight, and followed by calcination at 500 °C in the air for 4 h. The samples were labeled as Mn3Fe2/Silicalite-1, Mn3Fe2/TS-1-30 and Mn3Fe2/TS-1-20, respectively.

3.2. Characterization

X-ray diffraction (XRD) patterns of as-prepared samples were collected by a powder X-ray diffractometer (Rigaku Dmax-2400, RIGAKU, Tokyo, Japan) with Cu-Kα target. Fourier transform infrared (FT-IR) spectra were collected on a FT-IR spectrophotometer (Nicolet iS50, ThermoFisher, Waltham, MA, USA). The textural properties of as-prepared samples were investigated using a nitrogen adsorption apparatus (ASAP2020, Micromeritics, Norcross, GA, USA) at 77 K. Specific surface area, pore volume and pore size were calculated by Brunauer-Emmett-Teller (BET), single point and Barrett-Joyner-Halenda (BJH) methods, respectively. Surface morphologies of as-prepared samples were carried out on a field emission scanning electron microscope (Merlin Compact, ZEISS, Oberkochen, Germany) operating at 10 kV. Transmission electron microscopy (TEM) images of as-prepared samples were screened on a transmission electron microscope (JEM-2100F, JEOL, Tokyo, Japan) operating at 200 kV. The chemical composition of as-prepared samples was measured by an inductively coupled plasma-atomic emission spectrometer (Thermo iCAP6000 ICP-OES, ThermoFisher, Waltham, MA, USA). The surface chemical species of as-prepared samples were characterized by an X-ray photoelectron spectroscopy (AXIS Supra, Kratos Analytical Ltd., Manchester, UK) with Al Kα radiation. H2 temperature-programmed reduction (H2-TPR) and NH3 temperature-programmed desorption (NH3-TPD) experiments were carried out on a chemisorption analyzer (ChemBET Pulsar TPR/TPD, Quantachrome, Boynton Beach, FL, USA). In situ diffuse reflectance infrared Fourier transform (in situ DRIFTS) of NH3 adsorption were obtained from a FT-IR spectrophotometer (Nicolet 6700, Thermo Fisher, USA), by accumulating 32 scans with a resolution of 4 cm−1. The samples were firstly pretreated under a high purified N2 stream at 400 °C for 1 h to remove physicsorbed water and other impurities. Then the background spectra were obtained at target temperatures during the cooling procedure. Subsequently, the samples were exposed to a stream of NH3/N2 (1 vol.% NH3) at room temperature, followed by purging with N2 for 30 min. The desorption of NH3 studies were measured by heating pre-adsorbed samples, and the spectra were recorded at stepped target temperatures by eliminating the corresponding background reference.

3.3. NH3-SCR Activity Measurements

The NH3-SCR catalytic performances were tested in a fixed-bed quartz reactor (6 mm of internal diameter) in the temperature range from 90 to 270 °C, with a GHSV of 18,000 h−1. The reaction gas mixture was consisted of 500 ppm NO, 500 ppm NH3, 3 vol.% O2, 10 vol.% H2O (if applied) and N2 in balance. Firstly, 200 mg catalyst was pretreated with N2 at 200 °C for 30 min to eliminate physisorbed water. The concentrations of NO and NO2 were obtained through a flue gas analyzer (Testo 350 Pro, Testo, Black forest, Germany) when the catalytic reaction substantially reached a steady state at every target temperature. Moreover, H2O resistance of the catalysts for low-temperature NH3-SCR reaction was evaluated at the target temperature. NOx conversion efficiency was calculated according to the following formula:
NO x   conversion   efficiency % = 1 NO x out NO x in × 100 %
NO x = NO + NO 2
where NO x in and NO x out refer to inlet and outlet concentrations of NOx at steady state, respectively.

4. Conclusions

In this work, silicalite-1 and TS-1 were applied as supporters to load MnOx and FeOx nanoparticles as active components for the low-temperature NH3-SCR reaction. Mn3Fe2/TS-1-30 maintains NOx conversion efficiencies above 80% when the range of temperature is from 110 to 230 °C. Moreover, Mn3Fe2/TS-1-30 can maintain a denitration efficiency of 94% in the presence of 10 vol.% H2O at 180 °C for over 12 h. Its excellent catalytic performance and H2O resistance can be attributed to having much more surface acid sites and the stronger redox ability of its active components, which can be explained by the incorporation of Ti species. Therefore, the superior denitration ability and H2O resistance of Mn3Fe2/TS-1-30 will make it a practical catalyst for industrial largescale denitration at low temperatures.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/10/5/566/s1, Figure S1: STEM images and EDS mappings of the supporters: (a) TS-1-30; (b) TS-1-20.

Author Contributions

Conceptualization, J.G. and X.W.; formal analysis, J.G., R.D. and Y.C.; writing—original draft preparation, J.G.; writing—review and editing, J.G., R.D., W.C. and X.W.; supervision, X.W.; project administration, L.L.; funding acquisition, L.L. and X.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Grant No. 51672006 and 51472006), the Ministry of Land and Resources Public Welfare Industry Research Project (No. 201511062-02) and the National key research and development plan (No. 2018YFC1901505).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Busca, G.; Lietti, L.; Ramis, G.; Berti, F. Chemical and mechanistic aspects of the selective catalytic reduction of NOx by ammonia over oxide catalysts: A review. Appl. Catal. B Environ. 1998, 18, 1–36. [Google Scholar] [CrossRef]
  2. Wang, X.; Liu, Y.; Wu, Z. Highly active NbOPO4 supported Cu-Ce catalyst for NH3-SCR reaction with superior sulfur resistance. Chem. Eng. J. 2020, 382, 122941. [Google Scholar] [CrossRef]
  3. Paolucci, C.; Khurana, I.; Parekh, A.A.; Li, S.; Shih, A.J.; Li, H.; Di Iorio, J.R.; Albarracin-Caballero, J.D.; Yezerets, A.; Miller, J.T.; et al. Dynamic multinuclear sites formed by mobilized copper ions in NOx selective catalytic reduction. Science 2017, 357, 898–903. [Google Scholar] [CrossRef] [Green Version]
  4. Gillot, S.; Tricot, G.; Vezin, H.; Dacquin, J.-P.; Dujardin, C.; Granger, P. Development of stable and efficient CeVO4 systems for the selective reduction of NOx by ammonia: Structure-activity relationship. Appl. Catal. B Environ. 2017, 218, 338–348. [Google Scholar] [CrossRef]
  5. Mou, X.; Zhang, B.; Li, Y.; Yao, L.; Wei, X.; Su, D.S.; Shen, W. Rod-Shaped Fe2O3 as an Efficient Catalyst for the Selective Reduction of Nitrogen Oxide by Ammonia. Angew. Chem. Int. Ed. 2012, 51, 2989. [Google Scholar] [CrossRef]
  6. Chen, L.; Si, Z.; Wu, X.; Weng, D. DRIFT study of CuO-CeO2-TiO2 mixed oxides for NOx reduction with NH3 at low temperatures. ACS Appl. Mater. Interfaces 2014, 6, 8134–8145. [Google Scholar] [CrossRef]
  7. Gao, F.; Tang, X.; Yi, H.; Li, J.; Zhao, S.; Wang, J.; Chu, C.; Li, C. Promotional mechanisms of activity and SO2 tolerance of Co- or Ni-doped MnOx-CeO2 catalysts for SCR of NOx with NH3 at low temperature. Chem. Eng. J. 2017, 317, 20–31. [Google Scholar] [CrossRef]
  8. Fang, C.; Zhang, D.; Cai, S.; Zhang, L.; Huang, L.; Li, H.; Maitarad, P.; Shi, L.; Gao, R.; Zhang, J. Low-temperature selective catalytic reduction of NO with NH3 over nanoflaky MnOx on carbon nanotubes in situ prepared via a chemical bath deposition route. Nanoscale 2013, 5, 9199. [Google Scholar] [CrossRef] [PubMed]
  9. Chen, Y.; Zhang, Z.; Liu, L.; Mi, L.; Wang, X. In situ DRIFTS studies on MnOx nanowires supported by activated semi-coke for low temperature selective catalytic reduction of NOx with NH3. Appl. Surf. Sci. 2016, 366, 139–147. [Google Scholar] [CrossRef]
  10. Zhang, C.; Chen, T.; Liu, H.; Chen, D.; Xu, B.; Qing, C. Low temperature SCR reaction over Nano-Structured Fe-Mn Oxides: Characterization, performance, and kinetic study. Appl. Surf. Sci. 2018, 457, 1116–1125. [Google Scholar] [CrossRef]
  11. Yang, J.; Ren, S.; Zhang, T.; Su, Z.; Long, H.; Kong, M.; Yao, L. Iron doped effects on active sites formation over activated carbon supported Mn-Ce oxide catalysts for low-temperature SCR of NO. Chem. Eng. J. 2020, 379, 122398. [Google Scholar] [CrossRef]
  12. Pan, W.-G.; Hong, J.-N.; Guo, R.-T.; Zhen, W.-L.; Ding, H.-L.; Jin, Q.; Ding, C.-G.; Guo, S.-Y. Effect of support on the performance of Mn–Cu oxides for low temperature selective catalytic reduction of NO with NH3. J. Ind. Eng. Chem. 2014, 20, 2224–2227. [Google Scholar] [CrossRef]
  13. Chen, Z.; Wang, F.; Li, H.; Yang, Q.; Wang, L.; Li, X. Low-Temperature Selective Catalytic Reduction of NOx with NH3 over Fe–Mn Mixed-Oxide Catalysts Containing Fe3Mn3O8 Phase. Ind. Eng. Chem. Res. 2011, 51, 202–212. [Google Scholar] [CrossRef]
  14. Liu, Z.; Yi, Y.; Zhang, S.; Zhu, T.; Zhu, J.; Wang, J. 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]
  15. Li, J.H.; Chang, H.Z.; Ma, L.; Hao, J.M.; Yang, R.T. Low-temperature selective catalytic reduction of NOx with NH3 over metal oxide and zeolite catalysts-A review. Catal. Today 2011, 175, 147–156. [Google Scholar] [CrossRef]
  16. Liu, C.; Shi, J.-W.; Gao, C.; Niu, C. Manganese oxide-based catalysts for low-temperature selective catalytic reduction of NOx with NH3: A review. Appl. Catal. A Gen. 2016, 522, 54–69. [Google Scholar] [CrossRef]
  17. Wu, S.; Zhang, L.; Wang, X.; Zou, W.; Cao, Y.; Sun, J.; Tang, C.; Gao, F.; Deng, Y.; Dong, L. Synthesis, characterization and catalytic performance of FeMnTiOx mixed oxides catalyst prepared by a CTAB-assisted process for mid-low temperature NH3-SCR. Appl. Catal. A Gen. 2015, 505, 235–242. [Google Scholar] [CrossRef]
  18. Cao, F.; Su, S.; Xiang, J.; Wang, P.; Hu, S.; Sun, L.; Zhang, A. 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]
  19. Chen, Y.; Wang, J.; Yan, Z.; Liu, L.; Zhang, Z.; Wang, X. Promoting effect of Nd on the reduction of NO with NH3 over CeO2 supported by activated semi-coke: An in situ DRIFTS study. Catal. Sci. Technol. 2015, 5, 2251–2259. [Google Scholar] [CrossRef]
  20. Wang, J.; Yan, Z.; Liu, L.; Chen, Y.; Zhang, Z.; Wang, X. In situ DRIFTS investigation on the SCR of NO with NH3 over V2O5 catalyst supported by activated semi-coke. Appl. Surf. Sci. 2014, 313, 660–669. [Google Scholar] [CrossRef]
  21. Su, Y.; Fan, B.; Wang, L.; Liu, Y.; Huang, B.; Fu, M.; Chen, L.; Ye, D. MnOx supported on carbon nanotubes by different methods for the SCR of NO with NH3. Catal. Today 2013, 201, 115–121. [Google Scholar] [CrossRef]
  22. Lou, X.; 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]
  23. Seo, C.-K.; Choi, B.; Kim, H.; Lee, C.-H.; Lee, C.-B. Effect of ZrO2 addition on de-NOx performance of Cu-ZSM-5 for SCR catalyst. Chem. Eng. J. 2012, 191, 331–340. [Google Scholar] [CrossRef]
  24. Lv, G.; Bin, F.; Song, C.; Wang, K.; Song, J. Promoting effect of zirconium doping on Mn/ZSM-5 for the selective catalytic reduction of NO with NH3. Fuel 2013, 107, 217–224. [Google Scholar] [CrossRef]
  25. Wang, T.; Wan, Z.; Yang, X.; Zhang, X.; Niu, X.; Sun, B. Promotional effect of iron modification on the catalytic properties of Mn-Fe/ZSM-5 catalysts in the Fast SCR reaction. Fuel Process. Technol. 2018, 169, 112–121. [Google Scholar] [CrossRef]
  26. Du, T.; Qu, H.; Liu, Q.; Zhong, Q.; Ma, W. Synthesis, activity and hydrophobicity of Fe-ZSM-5@silicalite-1 for NH3-SCR. Chem. Eng. J. 2015, 262, 1199–1207. [Google Scholar] [CrossRef]
  27. Vu, D.V.; Miyamoto, M.; Nishiyama, N.; Ichikawa, S.; Egashira, Y.; Ueyama, K. Catalytic activities and structures of silicalite-1/H-ZSM-5 zeolite composites. Microporous Mesoporous Mater. 2008, 115, 106–112. [Google Scholar] [CrossRef]
  28. Serrano, D.P.; Calleja, G.; Botas, J.A.; Gutierrez, F.J. Characterization of adsorptive and hydrophobic properties of silicalite-1, ZSM-5, TS-1 and Beta zeolites by TPD techniques. Sep. Purif. Technol. 2007, 54, 1–9. [Google Scholar] [CrossRef]
  29. Li, G.; Wang, X.; Guo, X.; Liu, S.; Zhao, Q.; Bao, X.; Lin, L. Titanium species in titanium silicalite TS-1 prepared by hydrothermal method. Mater. Chem. Phys. 2001, 71, 195–201. [Google Scholar] [CrossRef]
  30. Mu, W.; Zhu, J.; Zhang, S.; Guo, Y.; Su, L.; Li, X.; Li, Z. Novel proposition on mechanism aspects over Fe–Mn/ZSM-5 catalyst for NH3-SCR of NOx at low temperature: Rate and direction of multifunctional electron-transfer-bridge and in situ DRIFTs analysis. Catal. Sci. Technol. 2016, 6, 7532–7548. [Google Scholar] [CrossRef]
  31. Xia, Y.; Zhan, W.; Guo, Y.; Guo, Y.; Lu, G. Fe-Beta zeolite for selective catalytic reduction of NOx with NH3: Influence of Fe content. Chin. J. Catal. 2016, 37, 2069–2078. [Google Scholar] [CrossRef]
  32. Xu, W.; Zhang, G.; Chen, H.; Zhang, G.; Han, Y.; Chang, Y.; Gong, P. Mn/beta and Mn/ZSM-5 for the low-temperature selective catalytic reduction of NO with ammonia: Effect of manganese precursors. Chin. J. Catal. 2018, 39, 118–127. [Google Scholar] [CrossRef]
  33. Cheng, K.; Liu, J.; Zhao, Z.; Wei, Y.; Jiang, G.; Duan, A. Direct synthesis of V–W–Ti nanoparticle catalysts for selective catalytic reduction of NO with NH3. RSC Adv. 2015, 5, 45172–45183. [Google Scholar] [CrossRef]
  34. Hu, H.; Cai, S.; Li, H.; Huang, L.; Shi, L.; Zhang, D. In Situ DRIFTs Investigation of the Low-Temperature Reaction Mechanism over Mn-Doped Co3O4 for the Selective Catalytic Reduction of NOx with NH3. J. Phys. Chem. C 2015, 119, 22924–22933. [Google Scholar] [CrossRef]
  35. Liu, F.; He, H.; Ding, Y.; Zhang, C. Effect of manganese substitution on the structure and activity of iron titanate catalyst for the selective catalytic reduction of NO with NH3. Appl. Catal. B Environ. 2009, 93, 194–204. [Google Scholar] [CrossRef]
  36. Reddy, A.; Gopinath, C.; Chilukuri, S. Selective ortho-methylation of phenol with methanol over copper manganese mixed-oxide spinel catalysts. J. Catal. 2006, 243, 278–291. [Google Scholar] [CrossRef]
  37. Xiong, Y.; Tang, C.; Yao, X.; Zhang, L.; Li, L.; Wang, X.; Deng, Y.; Gao, B.; Dong, L. Effect of metal ions doping (M=Ti4+, Sn4+) on the catalytic performance of MnOx/CeO2 catalyst for low temperature selective catalytic reduction of NO with NH3. Appl. Catal. A Gen. 2015, 495, 206–216. [Google Scholar] [CrossRef]
  38. Zhu, L.; Zhang, L.; Qu, H.; Zhong, Q. A study on chemisorbed oxygen and reaction process of Fe–CuOx/ZSM-5 via ultrasonic impregnation method for low-temperature NH3-SCR. J. Mol. Catal. A Chem. 2015, 409, 207–215. [Google Scholar] [CrossRef]
  39. Tan, P. Active phase, catalytic activity, and induction period of Fe/zeolite material in nonoxidative aromatization of methane. J. Catal. 2016, 338, 21–29. [Google Scholar] [CrossRef]
  40. Li, L.; Wu, Y.; Hou, X.; Chu, B.; Nan, B.; Qin, Q.; Fan, M.; Sun, C.; Li, B.; Dong, L.; et al. Investigation of Two-Phase Intergrowth and Coexistence in Mn–Ce–Ti–O Catalysts for the Selective Catalytic Reduction of NO with NH3: Structure–Activity Relationship and Reaction Mechanism. Ind. Eng. Chem. Res. 2018, 58, 849–862. [Google Scholar] [CrossRef]
  41. Meng, D.; Zhan, W.; Guo, Y.; Guo, Y.; Wang, L.; Lu, G. A Highly Effective Catalyst of Sm-MnOx for the NH3-SCR of NOx at Low Temperature: Promotional Role of Sm and Its Catalytic Performance. ACS Catal. 2015, 5, 5973–5983. [Google Scholar] [CrossRef]
  42. Yao, X.; Chen, L.; Cao, J.; Chen, Y.; Tian, M.; Yang, F.; Sun, J.; Tang, C.; Dong, L. Enhancing the deNO performance of MnOx/CeO2-ZrO2 nanorod catalyst for low-temperature NH3-SCR by TiO2 modification. Chem. Eng. J. 2019, 369, 46–56. [Google Scholar] [CrossRef]
  43. You, X.; Sheng, Z.; Yu, D.; Yang, L.; Xiao, X.; Wang, S. Influence of Mn/Ce ratio on the physicochemical properties and catalytic performance of graphene supported MnOx-CeO2 oxides for NH3-SCR at low temperature. Appl. Surf. Sci. 2017, 423, 845–854. [Google Scholar] [CrossRef]
  44. Zha, K.; Cai, S.; Hu, H.; Li, H.; Yan, T.; Shi, L.; Zhang, D. In Situ DRIFTs Investigation of Promotional Effects of Tungsten on MnOx-CeO2/meso-TiO2 Catalysts for NOx Reduction. J. Phys. Chem. C 2017, 121, 25243–25254. [Google Scholar] [CrossRef]
  45. Kang, M.; Park, E.D.; Kim, J.M.; Yie, J.E. Cu–Mn mixed oxides for low temperature NO reduction with NH3. Catal. Today 2006, 111, 236–241. [Google Scholar] [CrossRef]
  46. Chen, J.; Chen, X.; Chen, X.; Xu, W.; Xu, Z.; Jia, H.; Chen, J. Homogeneous introduction of CeOy into MnOx-based catalyst for oxidation of aromatic VOCs. Appl. Catal. B Environ. 2018, 224, 825–835. [Google Scholar] [CrossRef]
  47. Du, J.; Qu, Z.; Dong, C.; Song, L.; Qin, Y.; Huang, N. Low-temperature abatement of toluene over Mn-Ce oxides catalysts synthesized by a modified hydrothermal approach. Appl. Surf. Sci. 2018, 433, 1025–1035. [Google Scholar] [CrossRef]
  48. Lin, L.-Y.; Lee, C.-Y.; Zhang, Y.-R.; Bai, H. Aerosol-assisted deposition of Mn-Fe oxide catalyst on TiO2 for superior selective catalytic reduction of NO with NH3 at low temperatures. Catal. Commun. 2018, 111, 36–41. [Google Scholar] [CrossRef]
  49. Peng, Y.; Wang, D.; Li, B.; Wang, C.; Li, J.; Crittenden, J.; Hao, J. Impacts of Pb and SO2 Poisoning on CeO2-WO3/TiO2-SiO2 SCR Catalyst. Environ. Sci. Technol. 2017, 51, 11943–11949. [Google Scholar] [CrossRef]
  50. Liu, Y.; Gu, T.; Weng, X.; Wang, Y.; Wu, Z.; Wang, H. DRIFT Studies on the Selectivity Promotion Mechanism of Ca-Modified Ce-Mn/TiO2 Catalysts for Low-Temperature NO Reduction with NH3. J. Phys. Chem. C 2012, 116, 16582–16592. [Google Scholar] [CrossRef]
  51. Ramis, G.; Yi, L.; Busca, G. Ammonia activation over catalysts for the selective catalytic reduction of NOx and the selective catalytic oxidation of NH3. An FT-IR study. Catal. Today 1996, 28, 373–380. [Google Scholar] [CrossRef]
  52. Lin, S.D.; Gluhoi, A.C.; Nieuwenhuys, B.E. Ammonia oxidation over Au/MO/γ-Al2O3—activity, selectivity and FTIR measurements. Catal. Today 2004, 90, 3–14. [Google Scholar] [CrossRef]
  53. Ettireddy, P.R.; Ettireddy, N.; Boningari, T.; Pardemann, R.; Smirniotis, P.G. Investigation of the selective catalytic reduction of nitric oxide with ammonia over Mn/TiO2 catalysts through transient isotopic labeling and in situ FT-IR studies. J. Catal. 2012, 292, 53–63. [Google Scholar] [CrossRef]
  54. Liu, Z.; Zhu, J.; Li, J.; Ma, 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]
  55. Zhang, S.; Zhang, B.; Liu, B.; Sun, S. A review of Mn-containing oxide catalysts for low temperature selective catalytic reduction of NOx with NH3: Reaction mechanism and catalyst deactivation. RSC Adv. 2017, 7, 26226–26242. [Google Scholar] [CrossRef] [Green Version]
  56. Chen, H.; Zhang, R.; Bao, W.; Wang, H.; Wang, Z.; Wei, Y. Effective catalytic abatement of indoor formaldehyde at room temperature over TS-1 supported platinum with relatively low content. Catal. Today 2019. [Google Scholar] [CrossRef]
  57. Kuperman, A.; Nadimi, S.; Oliver, S.; Ozin, G.A.; Garces, J.M.; Olken, M.M. Non-aqueous Synthesis of Giant Crystal of Zeolites and Molecular Sieves. Nature 1993, 365, 239–242. [Google Scholar] [CrossRef]
  58. Zuo, Y.; Liu, M.; Zhang, T.; Hong, L.; Guo, X.; Song, C.; Chen, Y.; Zhu, P.; Jaye, C.; Fischer, D. Role of pentahedrally coordinated titanium in titanium silicalite-1 in propene epoxidation. RSC Adv. 2015, 5, 17897–17904. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction (XRD) patterns of supporters (a) and catalysts (b).
Figure 1. X-ray diffraction (XRD) patterns of supporters (a) and catalysts (b).
Catalysts 10 00566 g001
Figure 2. Fourier transform infrared (FT-IR) spectra of supporters.
Figure 2. Fourier transform infrared (FT-IR) spectra of supporters.
Catalysts 10 00566 g002
Figure 3. N2 adsorption/desorption isotherm of supporters (a) and catalysts (b).
Figure 3. N2 adsorption/desorption isotherm of supporters (a) and catalysts (b).
Catalysts 10 00566 g003
Figure 4. SEM and transmission electron microscopy (TEM) images of catalysts: (a,d) Mn3Fe2/Silicalite-1; (b,e) Mn3Fe2/TS-1-30; (c,f) Mn3Fe2/TS-1-20.
Figure 4. SEM and transmission electron microscopy (TEM) images of catalysts: (a,d) Mn3Fe2/Silicalite-1; (b,e) Mn3Fe2/TS-1-30; (c,f) Mn3Fe2/TS-1-20.
Catalysts 10 00566 g004
Figure 5. STEM images and energy-disperse X-ray spectroscopy (EDS) mappings of catalysts: (a) Mn3Fe2/Silicalite-1; (b) Mn3Fe2/TS-1-30; (c) Mn3Fe2/TS-1-20.
Figure 5. STEM images and energy-disperse X-ray spectroscopy (EDS) mappings of catalysts: (a) Mn3Fe2/Silicalite-1; (b) Mn3Fe2/TS-1-30; (c) Mn3Fe2/TS-1-20.
Catalysts 10 00566 g005
Figure 6. XPS spectra of catalysts: (a) Mn 2p; (b) Fe 2p; (c) O 1s.
Figure 6. XPS spectra of catalysts: (a) Mn 2p; (b) Fe 2p; (c) O 1s.
Catalysts 10 00566 g006
Figure 7. H2 temperature-programmed reduction (H2-TPR) profiles of catalysts.
Figure 7. H2 temperature-programmed reduction (H2-TPR) profiles of catalysts.
Catalysts 10 00566 g007
Figure 8. NH3 temperature-programmed desorption (NH3-TPD) profiles of catalysts.
Figure 8. NH3 temperature-programmed desorption (NH3-TPD) profiles of catalysts.
Catalysts 10 00566 g008
Figure 9. In situ diffuse reflectance infrared Fourier transform (in situ DRIFTS) of NH3 adsorption of depending on reaction temperatures from 60 to 330 °C: (a) Mn3Fe2/Silicalite-1; (b) Mn3Fe2/TS-1-30; (c) Mn3Fe2/TS-1-20.
Figure 9. In situ diffuse reflectance infrared Fourier transform (in situ DRIFTS) of NH3 adsorption of depending on reaction temperatures from 60 to 330 °C: (a) Mn3Fe2/Silicalite-1; (b) Mn3Fe2/TS-1-30; (c) Mn3Fe2/TS-1-20.
Catalysts 10 00566 g009
Figure 10. The NH3-selective catalytic reduction (SCR) activity of catalysts at the variation of temperature (a); H2O resistance performance of catalysts at 180 °C (b).
Figure 10. The NH3-selective catalytic reduction (SCR) activity of catalysts at the variation of temperature (a); H2O resistance performance of catalysts at 180 °C (b).
Catalysts 10 00566 g010
Table 1. Chemical compositions of samples.
Table 1. Chemical compositions of samples.
SampleThe Mass Fraction of Metal Elements/wt.%
MnFeTi
Mn3Fe2/Silicalite-13.181.87-
Mn3Fe2/TS-1-303.171.972.23
Mn3Fe2/TS-1-203.111.943.56
Table 2. Textural data of samples.
Table 2. Textural data of samples.
SampleBET Surface Area (m2/g)Pore Volume (cm3/g)Pore Diameter (nm)
Silicalite-1454.310.322.85
TS-1-30485.430.373.08
TS-1-20471.390.373.16
Mn3Fe2/Silicalite-1393.640.272.76
Mn3Fe2/TS-1-30422.140.323.08
Mn3Fe2/TS-1-20423.220.312.96
Table 3. Surface atom concentrations of samples.
Table 3. Surface atom concentrations of samples.
SampleAtomic Concentration/at%
SiOMnFeTi
Mn3Fe2/TS-1-2037.3761.780.530.130.19
Mn3Fe2/TS-1-3037.3361.860.500.150.16
Mn3Fe2/Silicalite-137.0762.220.520.19-
Table 4. The chemical states and relative concentration ratios for different elements.
Table 4. The chemical states and relative concentration ratios for different elements.
SamplesAtomic Ratio/%
Mn4+/(Mn4+ + Mn3+ + Mn2+)Fe3+/(Fe3+ + Fe2+)Oβ/(Oα + Oβ)
Mn3Fe2/TS-1-2025.659.194.37
Mn3Fe2/TS-1-3034.662.995.01
Mn3Fe2/Silicalite-125.065.593.01
Table 5. The peak temperature and H2 consumption of samples from H2-TPR.
Table 5. The peak temperature and H2 consumption of samples from H2-TPR.
SamplesPeak Temperature/°CH2 Consumption/a.u.
TTTSSSStotal
Mn3Fe2/TS-1-20432574662356231150737
Mn3Fe2/TS-1-30439585669400236140776
Mn3Fe2/Silicalite-143660866743696119651
Table 6. The peak temperature and acid amount of samples from NH3-TPD.
Table 6. The peak temperature and acid amount of samples from NH3-TPD.
SamplesPeak Temperature/°CAcid Amount/a.u.
TTTTSSSSStotal
Mn3Fe2/TS-1-20130308368-3736588-526
Mn3Fe2/TS-1-3013131440649039912612311659
Mn3Fe2/Silicalite-1128317--259213--472

Share and Cite

MDPI and ACS Style

Gu, J.; Duan, R.; Chen, W.; Chen, Y.; Liu, L.; Wang, X. Promoting Effect of Ti Species in MnOx-FeOx/Silicalite-1 for the Low-Temperature NH3-SCR Reaction. Catalysts 2020, 10, 566. https://doi.org/10.3390/catal10050566

AMA Style

Gu J, Duan R, Chen W, Chen Y, Liu L, Wang X. Promoting Effect of Ti Species in MnOx-FeOx/Silicalite-1 for the Low-Temperature NH3-SCR Reaction. Catalysts. 2020; 10(5):566. https://doi.org/10.3390/catal10050566

Chicago/Turabian Style

Gu, Jialiang, Rudi Duan, Weibin Chen, Yan Chen, Lili Liu, and Xidong Wang. 2020. "Promoting Effect of Ti Species in MnOx-FeOx/Silicalite-1 for the Low-Temperature NH3-SCR Reaction" Catalysts 10, no. 5: 566. https://doi.org/10.3390/catal10050566

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