Next Article in Journal
Effects of Alkali Metals on Nickel/Alumina Catalyzed Ethanol Dry Reforming
Next Article in Special Issue
Advanced Synthesis and Characterization of Vanadia/Titania Catalysts through a Molecular Approach
Previous Article in Journal
Photocatalytic Porous Silica-Based Granular Media for Organic Pollutant Degradation in Industrial Waste-Streams
Previous Article in Special Issue
Microwave-Assisted Solvothermal Synthesis of Fe3O4/CeO2 Nanocomposites and Their Catalytic Activity in the Imine Formation from Benzyl Alcohol and Aniline
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Precursor Effect on Mn-Fe-Ce/TiO2 Catalysts for Selective Catalytic Reduction of NO with NH3 at Low Temperatures

1
Center for Catalysis and Sustainable Chemistry, Department of Chemistry, Building 207, Technical University of Denmark, DK-2800 Kgs. Lyngby, Denmark
2
Combustion and Harmful Emission Control Research Center, Department of Chemical and Biochemical Engineering, Building 229, Technical University of Denmark, DK-2800 Kgs. Lyngby, Denmark
3
LAB SA, 259 Avenue Jean Jaurés, 69364 CEDEX 07 Lyon, France
*
Author to whom correspondence should be addressed.
Catalysts 2021, 11(2), 259; https://doi.org/10.3390/catal11020259
Submission received: 20 January 2021 / Revised: 5 February 2021 / Accepted: 9 February 2021 / Published: 15 February 2021

Abstract

:
Preparation of Mn/TiO2, Mn-Fe/TiO2, and Mn-Fe-Ce/TiO2 by the deposition-precipitation (DP) method can afford very active catalysts for low-temperature NH3-SCR (selective catalytic reduction of NO with NH3). The effect of precursor choice (nitrate vs. acetate) of Mn, Fe, and Ce on the physiochemical properties including thermal stability and the resulting SCR activity were investigated. The resulting materials were characterized by N2-Physisorption, XRD (Powder X-ray diffraction), XPS (X-ray photoelectron spectroscopy), H2-TPR (temperature-programmed reduction with hydrogen), and the oxidation of NO to NO2 measured at 300 °C. Among all the prepared catalysts 5MnAce/Ti, 25Mn0.75AceFe0.25Nit/Ti, and 25Mn0.75AceFe0.20NitCe0.05Ace/Ti showed superior SCR activity at low temperature. The superior activity of the latter two materials is likely attributable to the presence of amorphous active metal oxide phases (manganese-, iron- and cerium-oxide) and the ease of the reduction of metal oxides on TiO2. Enhanced ability to convert NO to NO2, which can promote fast-SCR like pathways, could be another reason. Cerium was found to stabilize amorphous manganese oxide phases when exposed to high temperatures.

1. Introduction

NH3-SCR (selective catalytic reduction of NOx with NH3) with V2O5-WO3/TiO2 as the catalyst is used successfully in stationary plants [1,2,3,4,5]. The support of choice is TiO2 (anatase form) due to its higher surface area relative to rutile phase and the fact that SO3 does not deteriorate the TiO2 support. The commercial catalyst exhibits high selectivity and activity in the NH3-SCR of NO at 300–400 °C [3,4,5]. In order to operate at this temperature, the SCR unit is usually installed at a high dust position.
However, by placing the SCR unit at the high dust position, the catalyst’s life is shortened due to the high content of ash with alkali metals in the flue gas [6,7,8]. Therefore, the tail-end position, which is located behind the SO2/SO3-removing unit is attractive. Decreased erosion and fouling at the low dust level also increases the catalyst’s lifetime [9]. In order to avoid costly reheating of the flue gas to around 350 °C, tail-end placement necessitates the SCR catalyst to be significantly more active than the vanadia-tungsta based one.
In recent years, a large number of research articles on NH3-SCR of NO at low-temperature have been published. Among the reported catalysts, Mn/TiO2 based formulations are the most promising [10,11]. Furthermore, bimetallic Mn catalysts showed higher activities and selectivities. Hence, Mn-Fe/TiO2 [12,13,14,15], MnOx-CeO2 [16], and Mn-Ce/TiO2 [17,18] have been reported to be highly active bimetallic catalysts for NH3-SCR of NO at low temperatures. Recently, we reported highly active low temperature Mn-Fe/TiO2 catalysts prepared by deposition-precipitation using ammonia carbamate as a precipitating agent [19].
The low-temperature SCR activity of the MnOx catalysts depends on the precursor, preparation method, and metal loading. Kapteijn et al. [20] reported that a Mn/Al2O3 catalyst was more active when prepared with Mn-acetate than with Mn-nitrate. Likewise, Li et al. [21] concluded that a Mn-acetate derived Mn/TiO2 catalyst had better activity than its Mn-nitrate based version. However, Peña et al. [22] showed that a Mn/TiO2 catalyst prepared from manganese nitrate and calcined at 400 °C had better activity at lower temperatures than a catalyst obtained from manganese acetate. Detailed investigation of the precursor effect on more active formulations like Mn-Fe/TiO2 and Mn-Fe-Ce/TiO2 catalysts has not been reported.
The optimum Mn loading of low temperature Mn/TiO2 catalysts was reported to be 20 wt.% [22] while the optimum loadings of Mn and Fe in Mn-Fe/TiO2 catalysts synthesized by impregnation were both 10 wt.% [14]. In our previous article [19], we reported that it was possible to further reduce the total metal loading on Mn/TiO2 catalysts from 20 wt.% to 5 wt.% with a change in the method of synthesis from conventional impregnation to deposition, while the total metal loading of the Mn-Fe/TiO2 catalysts could be reduced from 35 wt.% to 25 wt.%. Catalysts based on Mn-Fe/TiO2 contain high amounts of active metals (about 20–25 wt.%) compared to the traditional V2O5-WO3/TiO2 system (about 7–10 wt.%). Additionally, unsupported manganese oxide in hollandite form [23] and MnOx-CeO2 [16] exhibited high NH3-SCR activity at low temperatures.
The present article deals with the preparation of Mn/TiO2, Mn-Fe/TiO2, and Mn-Fe-Ce/TiO2 using several metal precursors. Various methods of characterization were employed to understand the differences in catalyst properties and activities.

2. Results and Discussion

The SCR NO conversion profiles of the 5MnNit and 5MnAce supported catalysts are shown in Figure 1. Among the catalysts studied, Mn deposited on TiO2 showed superior catalytic activity followed by ZrO2 and Al2O3. In particular, MnAce/Ti was more active compared to the MnNit/Ti. At 250 °C, the MnAce/Ti and MnNit/Ti catalysts displayed a NO conversion of 81 and 65%, respectively. The low temperature activity of the Mn/TiO2 catalysts were compared with silica and alumina by Simirniotis et al. [24] and they concluded that Lewis acid sites, a high surface concentration of MnO2, and good redox properties were important in achieving low temperature SCR activity. For further experiments, TiO2 was chosen as the unique support.
Mn/TiO2 doped with transition metals (e.g., Ni, Cu, and Fe) had high resistance to sintering and more favorable Mn dispersion [12]. It is also reported that Mn/TiO2 catalysts promoted with transition metals showed activity for NO oxidation to NO2 [15]. In our previous publication, we reported the promotional effect of Fe on Mn/TiO2 catalysts and the optimum formulation was found to be 25 Mn0.75Fe0.25/Ti using deposition-precipitation [19]. Figure 2a shows the NO conversion profiles of the 25 wt.% Mn0.75Fe0.25/Ti catalysts with Mn and Fe precursor combinations as a function of reaction temperature. All the catalysts, except for the 25Mn0.75NitFe0.25Ace/Ti catalyst, showed full conversion above 225 °C. The 25Mn0.75AceFe0.25Nit/Ti, 25Mn0.75NitFe0.25Nit/Ti, 25Mn0.75AceFe0.25Ace/Ti, and 25Mn0.75NitFe0.25Ace/Ti catalysts exhibited NO conversion of 69.6, 55.6, 47.0, and 43.1% at 150 °C, respectively, illustrating the importance of precursors on catalyst activity.
It is also well known that the presence of Ce can enhance the SCR performance and selectivity to N2 [17]. It is also known that the presence of Ce further overcomes the SO2 deactivation and water inhibition effects [18]. Figure 2b shows the effect of Ce precursor on the optimum 25Mn0.75AceFe0.25Nit/Ti catalyst. The presence of Ce in the 25Mn0.75AceFe0.20NitCe0.05Ace/Ti catalyst showed slightly better performance while 25Mn0.75AceFe0.20NitCe0.05Nit/Ti showed less performance than the previously optimized 25Mn0.75AceFe0.25Nit/Ti catalyst. This further confirms the sensitivity of SCR catalysts to the choice of precursors. The 25Mn0.75AceFe0.25Nit/Ti, 25Mn0.75AceFe0.20NitCe0.05Ace/Ti, and 25Mn0.75AceFe0.20NitCe0.05Nit/Ti catalysts displayed NO conversions of 69.6, 73.0, and 54.4% at 150 °C, respectively. Table 1 summarizes the N2O formation data obtained at 150 °C over the 25Mn0.75AceFe0.25Nit/Ti, 25Mn0.75AceFe0.20NitCe0.05Ace/Ti, and 25Mn0.75AceFe0.20NitCe0.05Nit/Ti catalysts, which respectively produced 35, 15, and 20 ppm of N2O under wet conditions (2.3 vol% H2O). Thus, the presence of Ce can increase the selectivity to N2. A moderate Ce content in Mn-Fe/TiO2 catalysts contributed to decreased N2O formation by hindering the over oxidation of NH3, the dominant step in N2O formation. N2O formation is controlled by selective reaction of NO with NH3, limiting the oxidation of NH3. However, at a higher concentration of water (≈10 vol%), no N2O was formed for all catalysts.
Figure 3a shows the NO conversion profiles of 20–30 wt.% Mn0.75AceFe0.20NitCe0.05Ace/Ti catalysts vs. the reaction temperature. The NO conversion was in the following order: 25 Mn0.75AceFe0.20NitCe0.05Ace/Ti > 20 Mn0.75AceFe0.20NitCe0.05Ace/Ti > 30 Mn0.75AceFe0.20NitCe0.05Ace/Ti between 150–200 °C. Above 200 °C, the catalysts showed almost 100% NO conversion and it was not possible to discriminate between them. The 20, 25, and 30 wt.% Mn0.75AceFe0.20NitCe0.05Ace/Ti catalysts displayed a NO conversion of 59.8, 73.0, and 55.8% at 150 °C, respectively.
Figure 3b shows the SCR activity of the 25 wt.% MnAceFeNitCeAce/Ti catalyst with different Mn-Fe-Ce mole fractions. The highest NO conversion was attained at a Mn mole fraction of 0.75 and the lowest activity at a mole fraction of 0.725, indicating that the minimum Mn content should be 0.75. Maximum NO conversion was obtained at a Fe mole fraction of 0.20 followed by the mole fractions 0.175 and 0.225. Maximum NO conversion was obtained at a Ce mole fraction of 0.05 followed by 0.075 and 0.025. The 25Mn0.75AceFe0.20NitCe0.05Ace/Ti, 25Mn0.75AceFe0.175NitCe0.075Ace/Ti, 25Mn0.75AceFe0.225NitCe0.025Ace/Ti, and 25Mn0.725AceFe0.225NitCe0.05Ace/Ti catalysts displayed NO conversions of 73.0, 65.2, 48.4, and 41.4 at 150 °C, respectively.
The effect of space velocity (mLg−1h−1) on the most active 25Mn0.75AceFe0.20NitCe0.05Ace/Ti catalyst is shown in Figure 4. Space velocity is an important factor to be considered in the catalyst design as well as to compare to catalysts in the open literature. The catalyst displayed a NO conversion of 98, 88 and 79%, respectively, at space velocities of 360,000, 450,000, and 600,000 mL g−1 h−1 at 200 °C. The fact that at the lowest space velocity, NO conversions of above 90% can be attained at temperatures above 200 C indicates that unselective oxidation of NH3 is not a major side reaction. The temperature window between 125 and 175 °C was used to determine an apparent activation energy of 38.6 ± 4.0 kJ/mol. This value clearly indicates that the present activity measurements on powdered samples were not strongly influenced by transport limitations in the pursued temperature window of the tail-end operation (125–175 °C).
The BET (Brunauer–Emmett–Teller) surface area, H2-TPR, and NO oxidation results of the MnFe/Ti and MnFeCe/Ti catalysts are summarized in Table 2. The BET surface area of DT51-TiO2 was 83 m2/g, while those of the MnFe/Ti and MnFeCe/Ti catalysts showed an increased surface area even with 25 wt.% active metal content. Thus, pore blockage of TiO2 is unlikely and the active metal oxides are probably highly dispersed on the TiO2 support. The increased surface area was due to increased microporosity compared to TiO2 (see Figures S1 and S2, Supplementary Materials).
Ease of reduction of metals oxides is known to be an indicator for favorable low temperature SCR activity [19]. The H2 consumption profiles of the 25MnFe/Ti and 25MnFeCe/Ti catalysts are shown in Figure 5 and the integrated values (µmol/g) are summarized in Table 2. All 25MnFe/Ti catalysts showed almost similar reduction patterns. To distinguish the bimetallic reduction patterns of the MnFe/Ti catalysts, the reduction patterns of Fe/TiO2 and Mn/TiO2 were reported [19]. The Fe/TiO2 catalyst reduced from Fe2O3 to Fe at 338 °C. The Mn/TiO2 catalyst showed three peaks corresponding to step-wise reduction of MnO2 to Mn2O3, Mn2O3 to Mn3O4, and Mn3O4 to MnO [19]. The 25MnFe/Ti materials exhibited only two peaks with the first (maximum at ≈255–270 °C) corresponding to the MnO2 reduction, and the second one (maximum at ≈350–390 °C) could be due to the reduction of subsequent manganese oxide phases mixed with iron oxide. The 25Mn0.75AceFe0.25Ace/Ti and 25Mn0.75NitFe0.25Ace/Ti catalysts showed visible shoulder peaks at around 230 and 340 °C, and that of the 25Mn0.75AceFe0.25Nit/Ti and 25Mn0.75NitFe0.25Nit/Ti catalysts did not display visible shoulder peaks because of the broad nature of the reduction profiles. The origin of the shoulder peak toward lower temperature is unclear, but might be because of the presence of smaller, more easily reducible manganese oxide particles.
The 25Mn0.75AceFe0.25Nit/Ti, 25Mn0.75NitFe0.25Nit/Ti, 25Mn0.75AceFe0.25Ace/Ti, and 25Mn0.75NitFe0.25Ace/Ti catalysts consumed 4120, 4004, 3996, and 3972 µmol/g of H2, respectively. The difference in H2 consumption between the catalysts was small, but the 25Mn0.75AceFe0.25Nit/Ti catalyst was reduced at a relatively low temperature. Thus, the ease of reduction pattern and the dominating MnO2 phase (first peak) for the 25Mn0.75AceFe0.25Nit/Ti catalyst seem to be the main contributors to the superior low temperature SCR activity.
The 25MnFeCe/Ti catalysts exhibited three reduction peaks, where the first two peaks can be assigned as similar to those of the MnFe/Ti catalysts and then the third reduction peak about 550 °C is due to the reduction of CeO2 [25]. Most importantly, the 25Mn0.75AceFe0.20NitCe0.05Ace/Ti and 25Mn0.75AceFe0.20NitCe0.05Nit/Ti catalysts displayed a H2 consumption of 5040 and 4907 µmol/g, respectively. This H2 consumption, which is higher than for the 25MnFe/Ti catalysts, could be due to better dispersion of Mn, Fe, and Ce. The 25Mn0.75AceFe0.20NitCe0.05Ace/Ti catalyst was reduced at lower temperatures (≈10–15 °C) compared to the 25Mn0.75AceFe0.20NitCe0.05Nit/Ti catalyst. Thus, also in this case, the ease of reduction and the dominating MnO2 phase (first reduction peak) in the 25Mn0.75AceFe0.20NitCe0.05Ace/Ti catalyst were the main reasons for the superior SCR activity at low temperature.
The X-ray powder diffraction (XRPD) patterns of the 25MnFe/Ti and 25MnFeCe/Ti catalysts are shown in Figure 6. The TiO2 anatase phase was dominant in all catalysts and manganese oxide, iron oxide, and cerium oxides or other mixed phases of Mn, Fe, or Ce were not observed. This is a clear indication that active metal oxides are highly dispersed and/or in an amorphous state. To understand the amorphous state of the active metal oxides on the surface of the catalysts, thermal treatment at 400, 500, and 600 °C for 2 h was performed. It is anticipated that amorphous to crystalline phase transformation can happen by thermal treatment.
The XRPD patterns of the 5MnAce/Ti, 25Mn0.75AceFe0.25Nit/Ti, and 25Mn0.75AceFe0.20NitCe0.05Ace/Ti samples calcined at several temperatures are shown in Figure 7. The 5MnAce/Ti catalyst showed similar diffraction patterns of anatase-TiO2 at 400 °C and there was a small deviation compared to the TiO2 patterns at 500 °C. With a further increase in temperature to 600 °C, MnO2, Mn2O3, and anatase phases were observed. Similarly, the 25Mn0.75AceFe0.25Nit/Ti catalyst showed only anatase phases of TiO2 at 400 and 500 °C, and crystalline Mn2O3 and anatase TiO2 phases were observed at 600 °C. No diffraction patterns of iron oxide were observed. Thus, we can clearly see that the presence of Fe can cause the transformation of the mixed manganese oxide phase into particles big enough for XRD detection to consist only of the Mn2O3 phase. The 25Mn0.75AceFe0.20NitCe0.05Ace/Ti catalyst showed only anatase TiO2 phases even at 600 °C, thus the combined presence of Fe and Ce on Mn/Ti can hamper the transformation of the amorphous to the crystalline phase.
Figure 8 shows the NO conversion profiles of 5MnAce/Ti, 25Mn0.75AceFe0.25Nit/Ti, and 25Mn0.75AceFe0.20NitCe0.05Ace/Ti catalysts calcined at 400, 500, and 600 °C. The catalysts were most active when calcined at 400 °C where the catalysts calcined at 500 and 600 °C showed lower activity. When calcined at 400 °C, the catalysts were rich in the amorphous metal oxide phases (manganese, iron or cerium oxides), which are known to be SCR active. Further increase in calcination temperature resulted in the partial transformation of amorphous manganese oxide to crystalline manganese oxides (MnO2 or Mn2O3). Overall, the SCR activity of the catalysts was in parallel to the amorphous to crystalline transformation of the catalysts as also reported by Kang and Tang et al. [26,27].
The impact of the calcination temperature and transformation of active oxides can also be studied in combination with H2-TPR. The redox properties of the 5MnAce/Ti, 25Mn0.75AceFe0.25Nit/Ti, and 25Mn0.75AceFe0.20NitCe0.05Ace/Ti catalysts calcined at 400, 500, and 600 °C are shown in Figure 9. The 5MnAce/Ti catalyst showed three different reduction peaks, which correspond to stepwise reduction of MnO2 to Mn2O3 (≈260 °C), Mn2O3 to Mn3O4 (≈360 °C), and Mn3O4 to MnO (≈460 °C) [19]. Increasing the calcination temperature from 400 to 600 °C, the 5MnAce/Ti catalyst shifted the first reduction peak to higher temperatures due to strong metal–support interactions [22], and the intensity of the second reduction peak was increased, which further indicates that the SCR active MnO2 phase decreases and the Mn2O3 phase increases. The shifting of both reduction peaks to higher temperatures might be due to particle growth (sintering).
The 25Mn0.75AceFe0.25Nit/Ti and 25Mn0.75AceFe0.20NitCe0.05Ace/Ti catalysts showed almost similar reduction patterns at 400 and 500 °C of calcination temperature, further indicating that Fe and Ce are inhibiting the phase transformation of MnO2 to Mn2O3 and possibly particle growth (sintering). At 600 °C, the catalyst displayed a shift in the MnO2 reduction peak to high temperatures and the intensity of the second reduction peak was increased. Thus, the combined presence of Fe and Ce on Mn/TiO2 can increase the thermal stability.
The observed low temperature activity of Mn catalysts can also be explained by the NO to NO2 oxidation ability as reported previously [15,18]. Table 2 shows the oxidation of NO to NO2 on the 25Mn0.75Fe0.25Ti-DP and 25Mn0.75Fe0.25Ce0.05Ti-DP catalysts at 300 °C under wet conditions. All the catalysts displayed high NO to NO2 conversion of 41.6 to 66%. The 25Mn0.75AceFe0.20NitCe0.05Ace/Ti and 25Mn0.75AceFe0.25Nit/Ti catalysts displayed the highest NO to NO2 conversion. The observed NO oxidation is consistent with the increased SCR activity of the catalysts, since partial conversion of NO into NO2 is helpful to promote the fast SCR reaction, which is also known to go on at very low temperatures [28].
The surface composition as obtained by XPS (X-ray photoelectron spectroscopy) characterization is shown in Table 3. The 25Mn0.75AceFe0.25Nit/Ti and 25Mn0.75AceFe0.20NitCe0.05Ace/Ti showed a surface Mn/Fe molar ratio of 2.08 and 1.30, respectively. Thus, it appears that the precursor/promoter combination has an influence on forming MnFe oxide species on the surface of the support. Importantly, the 25Mn0.75AceFe0.25Nit/Ti catalyst showed an Oα concentration of 50.1% of the total oxygen while the Ce promoted 25Mn0.75AceFe0.20NitCe0.05Ace/Ti catalyst yielded 83.8%, respectively. High concentrations of chemisorbed oxygen have been reported to have a beneficial influence on the low-temperature SCR reaction [12,29] and is explained by an increased rate of NO to NO2 oxidation [29]. Our XPS results showed a significantly higher concentration of more reactive surface oxygen (Oα) in the 25Mn0.75AceFe0.20NitCe0.05Ace/Ti catalyst than in the 25Mn0.75AceFe0.25Nit/Ti catalyst. This is reflected in the higher NO to NO2 oxidation activity (66% vs. 60.4% conversion).

3. Experimental

3.1. Catalyst Synthesis

TiO2 in anatase form (DT-51 from Crystal Global with a S content of ≈1.25 wt%, SA = 87 m2g−1), γ-Al2O3 (Saint-Gobain, surface area of 256 m2/g), and ZrO2 (Saint-Gobain, surface area of 95 m2/g) were used as support materials. Manganese(II) nitrate tetrahydrate (Mn(NO3)2· 4H2O, Aldrich), manganese (II) acetate tetrahydrate (CH3COO)2Mn· 4H2O, Aldrich), iron(III) nitrate nonahydrate (Fe(NO3)3· 9H2O, Aldrich), iron(II) acetate (Fe(CO2CH3)2, Aldrich), cerium (III) nitrate hexahydrate (Ce(NO3)3· 6H2O, Aldrich), and Cerium (III) acetate hydrate (Ce(CH3CO2)3· xH2O, Aldrich) were used as precursors. The Mn/TiO2, Mn-Fe/TiO2, and Mn-Fe-Ce/TiO2 catalysts were synthesized by deposition-precipitation. In deposition-precipitation (DP), the required amount of metal precursors (Mn, Fe, or Ce) and 1 g of support (TiO2 or ZrO2 or γ-Al2O3) was added to 10 mL demineralized water and mixed followed by the slow addition of ammonium carbamate solution (1 mol/L, Aldrich). The resulting slurry’s aqueous phase was slowly removed by evaporation with continuous stirring followed by 12 h of oven treatment at 120 °C and finally calcined for 2 h in air at 400 °C. The catalysts were designated with total metal loading, metal composition, and metal precursor as 20–30 wt.% MnX-AceFeY-NitCeZ-Ace/Ti, respectively. Here X, Y, and Z represent molar fractions of Mn, Fe, and Ce respectively (e.g., 25Mn0.75-AceFe0.20-NitCe0.05-Ace/Ti).

3.2. Catalyst Characterization

3.2.1. X-ray Powder Diffraction

X-ray powder diffraction (XRPD) was conducted with a Huber G670 instrument. CuKα radiation in steps of 0.02° was employed with a 2θ range of 2–60°. The Debye–Scherrer equation was used to calculate the crystallite sizes.

3.2.2. Nitrogen Physisorption

BET surface areas were determined from N2 physisorption measurements on about 100 mg at 77 K with a Micromeritics ASAP 2010 apparatus. The samples were pretreated at 200 °C for 1 h before the measurement.

3.2.3. Chemisorption

H2-TPR experiments were performed on a Micromeritics Autochem-II instrument with a reducing mixture (50 mL/min) consisting of 5% H2 and balance Ar (Air Liquide) from 60 to 550 °C (10 °C/min). The H2 concentration in the effluent stream was monitored with a thermal conductivity detector (TCD).

3.2.4. X-ray Photoemission Spectroscopy

XPS measurements at room temperature were carried out with a Thermo scientific instrument. Al Kα radiation (1484.6 eV) was used and Au served as the standard for the instrument calibration. To minimize surface contamination, samples were outgassed in vacuum for 1 h in vacuum prior to the data acquisition. Deconvolution of spectra was performed using the Thermo Scientific Avantage Data system software.

3.3. Catalytic Activity Measurements

The SCR activity measurements were conducted in a fixed-bed reactor loaded with 50 mg of the fractionized (180–300 µm) catalyst at a flow rate of 300 NmL min−1 (at room temperature) at atmospheric pressure. The concentrations at inlet were: NO = 1000 ppm, NH3 = 1000 ppm, O2 = 4%, and H2O = 2.3% with He as the make up gas. The temperature was increased from 125 to 300 °C in steps of 25 °C while the NO and NH3 concentrations were measured continuously with a Thermo Electron Model 17C chemiluminescense NH3-NOx analyzer. The N2O concentration was measured by GC (Shimadzu 14 B GC, poraplot column, TCD detection). The concentrations were measured after reaching steady state conversion (approximately 45 min at each temperature).
The NO oxidation to NO2 measurements were performed in the same set up loaded with 200 mg of the fractionized (180–300 µm) catalyst at a flow rate of 300 NmL min−1 (at room temperature). The inlet concentrations were: NO = 500 ppm, O2 = 4.5% and H2O = 2.3% with He as balance gas. During the experiments the temperature was increased in steps of 50 °C from 150 to 350 °C while the NO and NO2 concentrations were measured with a Thermo scientific UV–Vis spectrophotometer (Evolution 220).

4. Conclusions

A range of precursor combinations in deposition-precipitation synthesis were tested on highly active low temperature SCR of NO with NH3 catalysts. Among the three supports (TiO2, ZrO2, and Al2O3) and two precursor combinations (nitrate vs. acetate), a monometallic 5MnAce/Ti catalyst showed good low temperature SCR activity. Among the bimetallic catalysts, the 25Mn0.75AceFe0.25Nit/Ti catalysts showed better low temperature SCR activity, and the trimetallic 25Mn0.75AceFe0.20NitCe0.05Ace/Ti catalyst showed the best low temperature SCR activity, respectively. The addition of Fe, and especially Ce, not only enhances the activity, but also the thermal stability by hindering the transformation of finely dispersed, easily reducible amorphous manganese oxide phases into larger, less easily reducible, and more crystalline particles as evidenced by H2-TPR and XRD. Fe and Ce promoted catalysts were shown by XPS to contain large amounts of surface active oxygen, further corroborating the H2-TPR and XRD results. This form of oxygen can enhance the oxidation of NO to NO2, as shown by NO oxidation measurements, and can promote fast-SCR and hence the overall activity. Furthermore, Ce lowers the selectivity toward N2O, possibly by enhancing the rate of reaction of activated NH3 with NO to N2, thus making it unavailable for over-oxidation, which can lead to N2O formation.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/11/2/259/s1, Figure S1: Adsorption/desorption isotherms of 25MnFe/Ti and 25MnFeCe/Ti catalysts with TiO2 support. Figure S2: Cumulative surface area as a function of the BJH (Barrett- Joyner-Halenda) pore width (from adsorption branch) of the 25MnFe/Ti, 25MnFeCe/Ti catalysts and TiO2 support.

Author Contributions

Conceptualization: S.S.R.P. and L.S.; Methodology: S.S.R.P., L.S.; R.F., and A.D.J.; Experimental: S.S.R.P. and L.S.; Writing of original draft: S.S.R.P.; Editing and reviewing of original draft: S.S.R.P., L.S., R.F., and A.D.J.; Revised draft: S.S.R.P. and L.S.; Project administration: R.F. and A.D.J.; Funding acquisition: R.F., B.S., and F.T.; Consulting from industrial point of view: B.S. and F.T. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by Energinet.dk through the PSO project 10521. LAB S.A, France and DONG Energy, Denmark are also thanked for their financial support.

Data Availability Statement

The data generated in this study are fully disclosed in this manuscript.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Bosch, H.; Janssen, F. Formation and control of nitrogen oxides. Catal. Today 1988, 2, 369–379. [Google Scholar]
  2. Parvulescu, V.I.; Grange, P.; Delmon, B. Catalytic removal of NO. Catal. Today 1998, 46, 233–316. [Google Scholar] [CrossRef]
  3. 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 1998, 18, 1–36. [Google Scholar] [CrossRef]
  4. Alemany, L.J.; Lietti, L.; Ferlazzo, N.; Forzatti, P.; Busca, G.; Giamello, E.; Bregani, F. Reactivity and physicochemical characterization of V2O5-WO3/TiO2 De-NOx catalysts. J. Catal. 1995, 155, 117–130. [Google Scholar]
  5. Heck, R.M. Catalytic abatement of nitrogen oxides-stationary applications. Catal. Today 1999, 53, 519–523. [Google Scholar] [CrossRef]
  6. Due-Hansen, J.; Boghosian, S.; Kustov, A.; Fristrup, P.; Tsilomelekis, G.; Ståhl, K.; Christensen, C.H.; Fehrmann, R. Vanadia-based SCR catalysts supported on tungstated and sulfated zirconia: Influence of doping with potassium. J. Catal. 2007, 251, 459–473. [Google Scholar] [CrossRef]
  7. Putluru, S.S.R.; Jensen, A.D.; Riisager, A.; Fehrmann, R. Heteropoly acid promoted V2O5/TiO2 catalysts for NO abatement with ammonia in alkali containing flue gases. Catal. Sci. Technol 2011, 1, 631–637. [Google Scholar] [CrossRef]
  8. Putluru, S.S.R.; Kristensen, S.B.; Due-Hansen, J.; Riisager, A.; Fehrmann, R. Alternative alkali resistant deNOx catalysts. Catal. Today 2012, 184, 192–196. [Google Scholar] [CrossRef]
  9. Singoredjo, L.; Korver, R.; Kapteijn, F.; Moulijn, J. Alumina Supported Manganese Oxides for the Low-Temperature Selective Catalytic Reduction of Nitric-Oxide with Ammonia. Appl. Catal. B Environ. 1992, 1, 297–316. [Google Scholar] [CrossRef]
  10. Smirniotis, P.G.; Peńa, D.A.; Uphade, B.S. Low-temperature selective catalytic reduction (SCR) of NO with NH3 by using Mn, Cr, and Cu oxides supported on Hombikat TiO2. Angew. Chem. Int. Ed. 2001, 40, 2479–2482. [Google Scholar] [CrossRef]
  11. Long, R.Q.; Yang, R.T.; Chang, R. Low temperature selective catalytic reduction (SCR) of NO with NH3 over Fe-Mn based catalysts. Chem. Commun. 2002, 5, 452–453. [Google Scholar] [CrossRef]
  12. Wu, Z.; Jiang, B.; Liu, Y. Effect of transition metals addition on the catalyst for manganese/titania for low-temperature selective catalytic reduction of nitric oxide with ammonia. Appl. Catal. B 2008, 79, 347–355. [Google Scholar] [CrossRef]
  13. Roy, S.; Viswanath, B.; Hegde, M.S.; Madras, G. Low-temperature selective catalytic reduction of NO with NH3 over Ti0.9M0.1O2-δ (M= Cr, Mn, Fe, Co, Cu). J. Phys. Chem. C 2008, 112, 6002–6012. [Google Scholar] [CrossRef]
  14. Qi, G.; Yang, R.T. Low-temperature selective catalytic reduction of NO with NH3 over iron and manganese oxides supported on titania. Appl. Catal. B Environ. 2003, 44, 217–225. [Google Scholar] [CrossRef]
  15. Wu, S.; Yao, X.; Zhang, L.; Cao, Y.; Zou, W.; Li, L.; Ma, K.; Tang, C.; Gao, F.; Dong, L. Improved low temperature NH3-SCR performance of FeMnTiOx mixed oxide with CTAB-assisted synthesis. Chem. Commun. 2015, 51, 3470–3473. [Google Scholar] [CrossRef]
  16. Qi, G.; Yang, R.T.; Chang, R. MnOx-CeO2 mixed oxides prepared by co-precipitation for selective catalytic reduction of NO with NH3 at low temperatures. Appl. Catal. B Environ. 2004, 51, 93–106. [Google Scholar] [CrossRef]
  17. Shang, T.; Hui, S.; Niu, Y.; Liang, L.; Liu, C.; Wang, D. Effect of the addition of Ce to MnOx/Ti catalyst on reduction of N2O in low-temperature SCR. Asia Pac. J. Chem. Eng. 2014, 9, 810–817. [Google Scholar] [CrossRef]
  18. Jin, R.; Liu, Y.; Wang, Y.; Cen, W.; Wu, Z.; Wang, H.; Weng, X. The role of cerium in the improved SO2 tolerance for NO reduction with NH3 over Mn-Ce/TiO2 catalyst at low temperature. Appl. Catal. B Environ. 2014, 148–149, 582–588. [Google Scholar] [CrossRef]
  19. Putluru, S.S.R.; Schill, L.; Jensen, A.D.; Siret, B.; Tabaries, F.; Fehrmann, R. Mn/TiO2 and Mn-Fe/TiO2 catalysts synthesized by deposition precipitation-promising for selective catalytic reduction of NO with NH3 at low temperatures. Appl. Catal. B Environ. 2015, 165, 628–635. [Google Scholar] [CrossRef]
  20. Kapteijn, F.; Vanlangeveld, A.D.; Moulijn, J.A.; Andreini, A.; Vuurman, M.A.; Turek, A.M.; Jehng, J.M.; Wachs, I.E. Alumina-Supported Manganese Oxide Catalysts. J. Catal. 1994, 150, 94–104. [Google Scholar] [CrossRef]
  21. Li, J.; Chen, J.; Ke, R.; Luo, C.; Hao, J. Effects of precursors on the surface Mn species and the activities for NO reduction over MnOx/TiO2 catalysts. Catal. Commun. 2007, 8, 1896–1900. [Google Scholar] [CrossRef]
  22. Peña, D.A.; Uphade, B.S.; Smirniotis, P.G. TiO2-supported metal oxide catalysts for low-temperature selective catalytic reduction of NO with NH3: I. Evaluation and characterization of first row transition metals. J. Catal. 2004, 221, 421–431. [Google Scholar] [CrossRef]
  23. Huang, Z.; Gu, X.; Wen, W.; Hu, P.; Makkee, M.; Lin, H.; Kapteijn, F.; Tang, X. A “smart” hollandite deNOx catalyst: Self-protection against alkali poisoning. Angew. Chem. Int. Ed. 2013, 52, 660–664. [Google Scholar] [CrossRef]
  24. Smirniotis, P.G.; Sreekanth, P.M.; Peña, D.A.; Jenkins, R.G. Manganese oxide catalysts supported on TiO2, Al2O3, and SiO2: A comparison for low-temperature SCR of NO with NH3. Ind. Eng. Chem. Res. 2006, 45, 6436–6443. [Google Scholar] [CrossRef]
  25. Thirupathi, B.; Smirniotis, P.G. Co-doping a metal (Cr, Fe, Co, Ni, Cu, Zn, Ce, and Zr) on Mn/TiO2 catalyst and its effect on the selective reduction of NO with NH3 at low-temperatures. Appl. Catal. B Environ. 2011, 110, 195–206. [Google Scholar] [CrossRef]
  26. Kang, M.; Park, E.D.; Kim, J.M.; Yie, J.E. Manganese oxide catalysts for NOx reduction with NH3 at low temperatures. Appl. Catal. A Gen. 2007, 327, 261–269. [Google Scholar] [CrossRef]
  27. Tang, X.; Hao, J.; Xu, W.; Li, J. Low temperature selective catalytic reduction of NOx with NH3 over amorphous MnOx catalysts prepared by three methods. Catal. Commun. 2007, 8, 329–334. [Google Scholar] [CrossRef]
  28. Iwasaki, M.; Shinjoh, H. A comparative study of “standard”, “fast” and “nO2” SCR reactions over Fe/zeolite catalyst. Appl. Catal. A Gen. 2010, 390, 71–77. [Google Scholar] [CrossRef]
  29. Shen, B.; Liu, T.; Zhao, N.; Yang, X.; Deng, L. Iron-doped Mn-Ce/TiO2 catalyst for low temperature selective catalytic reduction of NO with NH3. J. Environ. Sci. 2010, 22, 1447–1454. [Google Scholar] [CrossRef]
Figure 1. NO conversion profiles of: (a) 5MnNit and; (b) 5MnAce on various supports.
Figure 1. NO conversion profiles of: (a) 5MnNit and; (b) 5MnAce on various supports.
Catalysts 11 00259 g001
Figure 2. NO conversion profiles of catalysts prepared by different precursors: (a) 25Mn0.75Fe0.25/Ti; (b) 25Mn0.75Fe0.20Ce0.05/Ti.
Figure 2. NO conversion profiles of catalysts prepared by different precursors: (a) 25Mn0.75Fe0.25/Ti; (b) 25Mn0.75Fe0.20Ce0.05/Ti.
Catalysts 11 00259 g002
Figure 3. NO conversion profiles of: (a) 20–30 wt.% Mn0.75AceFe0.20NitCe0.05Ace/Ti catalysts; (b) 25MnAceFeNitCeAce/Ti catalysts with Mn0.75–0.725, Fe0.225–0.175, Ce0.075–0.025 mole fractions.
Figure 3. NO conversion profiles of: (a) 20–30 wt.% Mn0.75AceFe0.20NitCe0.05Ace/Ti catalysts; (b) 25MnAceFeNitCeAce/Ti catalysts with Mn0.75–0.725, Fe0.225–0.175, Ce0.075–0.025 mole fractions.
Catalysts 11 00259 g003
Figure 4. NO conversion profiles of the 25Mn0.75AceFe0.20NitCe0.05Ace/Ti catalyst at various space velocities.
Figure 4. NO conversion profiles of the 25Mn0.75AceFe0.20NitCe0.05Ace/Ti catalyst at various space velocities.
Catalysts 11 00259 g004
Figure 5. H2-TPR (temperature-programmed reduction with hydrogen) profiles of (a) 25MnFe/Ti and (b) 25MnFeCe/Ti catalysts.
Figure 5. H2-TPR (temperature-programmed reduction with hydrogen) profiles of (a) 25MnFe/Ti and (b) 25MnFeCe/Ti catalysts.
Catalysts 11 00259 g005
Figure 6. X-ray powder diffraction (XRPD) patterns of TiO2 and the (a) 25MnFe/Ti and (b) 25MnFeCe/Ti catalysts.
Figure 6. X-ray powder diffraction (XRPD) patterns of TiO2 and the (a) 25MnFe/Ti and (b) 25MnFeCe/Ti catalysts.
Catalysts 11 00259 g006
Figure 7. XRPD patterns of catalysts calcined at various temperatures: (a) 5MnAce/Ti; (b) 25Mn0.75AceFe0.25Nit/Ti; (c) 25Mn0.75AceFe0.20NitCe0.05Ace/Ti.
Figure 7. XRPD patterns of catalysts calcined at various temperatures: (a) 5MnAce/Ti; (b) 25Mn0.75AceFe0.25Nit/Ti; (c) 25Mn0.75AceFe0.20NitCe0.05Ace/Ti.
Catalysts 11 00259 g007aCatalysts 11 00259 g007b
Figure 8. NO conversion profiles of catalysts calcined at various temperatures: (a) 5MnAce/Ti; (b) 25Mn0.75AceFe0.25Nit/Ti; and (c) 25Mn0.75AceFe0.20NitCe0.05Ace/Ti.
Figure 8. NO conversion profiles of catalysts calcined at various temperatures: (a) 5MnAce/Ti; (b) 25Mn0.75AceFe0.25Nit/Ti; and (c) 25Mn0.75AceFe0.20NitCe0.05Ace/Ti.
Catalysts 11 00259 g008
Figure 9. H2-TPR (temperature-programmed reduction with hydrogen) profiles of catalysts calcined at various temperatures: (a) 5MnAce/Ti; (b) 25Mn0.75AceFe0.25Nit/Ti; and (c) 25Mn0.75AceFe0.20NitCe0.05Ace/Ti.
Figure 9. H2-TPR (temperature-programmed reduction with hydrogen) profiles of catalysts calcined at various temperatures: (a) 5MnAce/Ti; (b) 25Mn0.75AceFe0.25Nit/Ti; and (c) 25Mn0.75AceFe0.20NitCe0.05Ace/Ti.
Catalysts 11 00259 g009
Table 1. N2O formation data at 150 °C.
Table 1. N2O formation data at 150 °C.
2.3 vol.% H2O10 vol.% H2O
CatalystNO Conv.
(%)
N2O
(ppm)
Sel. N2O
(%)
N2O
(ppm)
Sel. N2O
(%)
25Mn0.75AceFe0.25Nit/Ti69.6355.000
25Mn0.75AceFe0.20NitCe0.05Ace/Ti73.0152.100
25Mn0.75AceFe0.20NitCe0.05Nit/Ti54.4203.700
Table 2. Surface area, H2-TPR (temperature-programmed reduction with hydrogen) and NO oxidation results.
Table 2. Surface area, H2-TPR (temperature-programmed reduction with hydrogen) and NO oxidation results.
CatalystSurface AreaH2 ConsumptionNO oxidation to NO2
(m2/g)(µmol/g)(%) *
25Mn0.75AceFe0.25Nit/Ti95412060
25Mn0.75NitFe0.25Nit/Ti91400458
25Mn0.75NitFe0.25Nit/Ti100399646
25Mn0.75NitFe0.25Ace/Ti100397242
25Mn0.75AceFe0.20NitCe0.05Ace/Ti102504066
25Mn0.75AceFe0.20NitCe0.05Nit/Ti96490756
* NO oxidation to NO2 at 300 °C.
Table 3. Atomic% of MnFe/Ti and MnFeCe/Ti catalysts determined by XPS (X-ray photoelectron spectroscopy).
Table 3. Atomic% of MnFe/Ti and MnFeCe/Ti catalysts determined by XPS (X-ray photoelectron spectroscopy).
CatalystTiMnFeCeMn/FeOtOα/Ot
25Mn0.75AceFe0.25Nit/Ti14.58.84.1--2.172.650.1
25Mn0.75AceFe0.20NitCe0.05Ace/Ti12.09.57.32.21.369.083.8
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Putluru, S.S.R.; Schill, L.; Jensen, A.D.; Siret, B.; Tabaries, F.; Fehrmann, R. Precursor Effect on Mn-Fe-Ce/TiO2 Catalysts for Selective Catalytic Reduction of NO with NH3 at Low Temperatures. Catalysts 2021, 11, 259. https://doi.org/10.3390/catal11020259

AMA Style

Putluru SSR, Schill L, Jensen AD, Siret B, Tabaries F, Fehrmann R. Precursor Effect on Mn-Fe-Ce/TiO2 Catalysts for Selective Catalytic Reduction of NO with NH3 at Low Temperatures. Catalysts. 2021; 11(2):259. https://doi.org/10.3390/catal11020259

Chicago/Turabian Style

Putluru, Siva Sankar Reddy, Leonhard Schill, Anker Degn Jensen, Bernard Siret, Frank Tabaries, and Rasmus Fehrmann. 2021. "Precursor Effect on Mn-Fe-Ce/TiO2 Catalysts for Selective Catalytic Reduction of NO with NH3 at Low Temperatures" Catalysts 11, no. 2: 259. https://doi.org/10.3390/catal11020259

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