Skip to main content
Erschienen in: Topics in Catalysis 13-14/2023

Open Access 06.12.2022 | Original Paper

NH3-NO SCR Catalysts for Engine Exhaust Gases Abatement: Replacement of Toxic V2O5 with MnOx to Improve the Environmental Sustainability

verfasst von: L. Consentino, Giuseppe Pantaleo, V. La Parola, C. Migliore, E. La Greca, Leonarda Francesca Liotta

Erschienen in: Topics in Catalysis | Ausgabe 13-14/2023

Aktivieren Sie unsere intelligente Suche, um passende Fachinhalte oder Patente zu finden.

search-config
loading …

Abstract

Mn-WO3/TiO2 catalysts were investigated for Selective Catalytic Reduction (SCR) of NO with NH3. The catalysts were synthesized by wetness impregnation method with different Mn loadings (1.5-3-12 wt%) on 8wt%WO3/TiO2. All three catalysts were compared with 8wt%WO3/TiO2 and bare MnOx oxide, used as references. The 1.5wt%Mn-8wt%WO3/TiO2 exhibited the highest performance in NO conversion and N2 selectivity. A commercial catalyst, based on titania supported vanadia and tungsta, (V2O5-WO3/TiO2), widely used for its high efficiency, was also investigated in the present work. The morphological, structural, redox and electronic properties of the catalysts and their thermal stability were studied by several techniques (N2 adsorption/desorption, X-ray diffraction, H2 temperature-programmed reduction, NH3 temperature programmed desorption, X-ray photoelectron spectroscopy).
The aim of this paper is to study the effect of different Mn loadings on 8wt%WO3/TiO2 with the ambition to obtain highly active and selective catalysts in a large window of temperature. The replacement of toxic vanadium used in the classic V2O5-WO3/TiO2 catalyst with MnOx in the best performing catalyst, 1.5wt%Mn-8wt%WO3/TiO2, represents an important achievement to improve the environmental sustainability.
Hinweise

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

1 Introduction

Nitrogen oxides (NOx) emitted by mobile and stationary sources are one of the major causes of air pollution and are responsible for a series of environmental problems such as acid rain, photochemical smog and the green-house effect [1]. In recent years the low temperature (100–200 °C) selective catalytic reduction (SCR) of NOx by NH3 has been investigated and metal doped titanates have attracted attention, for controlling NOx emissions from stationary sources, thanks to their elevated performances and low cost [2]. Among common catalytic devices for NO SCR in stationary plants, V2O5-WO3/TiO2 catalysts are used, however, some problems remain due to the high activity for oxidation of SO2 to SO3, the rapid decrease in activity and selectivity above 550 °C, and the toxicity of the vanadium pentoxide due to easy sublimation [3, 4]. So, NO SCR devices used in stationary sources must be installed upstream the particulate collector and flue gas desulfurization in order to meet the optimum working temperature of 350–400 °C. On the other hand, for mobile sources, this type of V2O5-WO3/TiO2 catalyst is not the best solution for diesel engine exhaust gases aftertreatment, as it cannot completely remove NOx due to the wide temperature window of exhaust gas (150–500 °C). Therefore, researchers in the field of academia and industry continue to develop new low-temperature catalysts that can work well around 250 °C.
It is well known that the nature of the catalyst is crucial for an efficient SCR of NOx by NH3 or urea. Transition metal oxides including iron (Fe), manganese (Mn), cerium (Ce), copper (Cu) and nickel (Ni) have been, recently, investigated showing extraordinary catalytic performance for NOx removal. In particular, Mn-based catalysts reported both, high NO conversion and N2 selectivity at low temperature [3, 4] since manganese oxide contains different types of labile oxygen, as described by Kapteijn et al. [5]. Nevertheless, the application of pure MnOx is not recommended due to its low surface area and poor thermal stability [6]. For such reasons, supported MnOx on Al2O3 [7], TiO2 [8], USY [9], MnOx/AC [10] and MnOx/ACF [11] have been developed. Among these, MnOx/TiO2 exhibited the best performance at low temperature [4]. It has been also considered the opportunity to dope TiO2 -supported manganese catalysts with transitional metal in order to evaluate combined effects between the two metal oxides. For example, Fe2O3, WO3, MoO3, and Cr2O3 were investigated by Wu et al. [12] finding that Mn-W/TiO2 showed the higher NOx conversion followed by Mn-Fe/TiO2, Mn-Cr/TiO2 and Mn-Mo/TiO2.
Tungsten oxide is well-known for promoting the activity and stability of the commercial vanadium-based catalyst (V2O5-WO3/TiO2). WO3 forms W-Ti-O chemical bonds over TiO2 surface [13] and stabilize anatase phase of TiO2 giving high thermal stability to the catalyst [1416]. In addition, as Shan et al. reported, tungsten increases both the surface Brӧnsted and Lewis acidity and the operation temperature window of SCR [17]. Therefore, binary metal oxide-based catalysts, as MnOx-WO3/TiO2, are noteworthy for low temperature NO SCR.
In this work we have investigated Mn-WO3/TiO2 catalysts synthesized by wetness impregnation to evaluate the effect of Mn loading on the activity and selectivity in NO conversion with the aim to extend the window of operating temperature. Moreover, we have compared the best catalyst with a commercial SCR device, based on V2O5-WO3/TiO2, in order to replace the harmful vanadium pentoxide with Mn, preserving high efficiency in NOx removal. The physical-chemical, structural and surface properties of the catalysts have been investigated by XRD, BET, H2-TPR, NH3-TPD, XRD and XPS techniques.

2 Experimental

2.1 Catalyst Preparation

All reagents were analytical grade and used as received without further purification. Firstly, we prepared the 8wt% WO3/TiO2 oxide (WTi) by wetness impregnation method of a commercial TiO2 oxide, like P25 (provided by Sigma-Aldrich). A proper amount of (NH4)10H2(W2O7)6 was mixed with oxalic acid solution with a molar ratio of the acid to paratungstate equal to 8:1. This solution was added dropwise to the TiO2 support. Subsequently, the obtained paste was dried overnight at 120 °C and calcined at 400 °C for 2 h in static air. A certain amount of TiO2 (Sigma-Aldrich) was calcined in the same conditions and used as reference. The WTi oxide was impregnated, by wetness impregnation method, with few mL of an aqueous solution of Mn(NO3)2.xH2O, in appropriate concentration in order to prepare catalysts with three Mn loadings, corresponding to 1.5, 3.0 and 12.0 wt% as Mn. After drying overnight at 120 °C, calcination was carried out at 400 °C for 2 h. The so prepared materials were labelled as xMnWTi (x = 1.5, 3, 12). Bare MnOx oxide was prepared starting from 0.5 M Mn(NO3)2.xH2O solution by precipitation with 0.5 M Na2CO3 solution. After drying at 120 °C, calcination at 400 °C for 4 h was performed.
A commercial catalyst (VWTi), provided by Hug Engineering (ElringKlinger Group), with composition 3wt% V2O5-7wt% WO3/75wt% TiO2-6wt% bentonite- 9wt% glass fiber was used as reference. Details about chemical physical properties were reported elsewhere by some of us [18].
All the samples and their relative labels are listed in Table 1.
Table 1
Chemical composition of the prepared samples and relative labels
Sample
Label
Bare MnOx
MnOx
TiO2 calcined
TiO2 calc.
8wt% WO3/TiO2
WTi
1.5wt% Mn-8wt% WO3/TiO2
1.5MnWTi
3wt% Mn-8wt% WO3/TiO2
3MnWTi
12wt% MnO2-8wt% WO3/TiO2
12MnWTi
3wt%V2O5-7wt%WO3/75wt%TiO2-6wt% bentonite- 9wt% glass fiber
VWTi

2.2 Catalysts Characterizations

Elemental analysis of the catalysts was carried out by Atomic Emission Spectroscopy (MP-AES 4200 Agilent technologies), after treatment in acidic solution with H2SO4 and HNO3 at 250 °C for complete dissolution. The real chemical composition corresponded well to the nominal one within ± 5%.
Specific surface areas (SSA) and pore volumes were determined by N2 adsorption –desorption isotherms at -196 °C using a Micromeritics ASAP 2020 equipment, through Brunauer –Emmett-Teller (BET) method in the standard pressure range 0.05–0.3 P/P0. Before the measurements, the samples were degassed at 250 °C for 2 h. By analysis of the desorption curves, using the Barrett–Joyner–Halenda (BJH) method the pore volume and pore size distribution were obtained.
X-ray diffraction (XRD) patterns were acquired with a Bruker goniometer employing Ni-filtered Cu Kα radiation. A proportional counter and 0.05° step size in 2θ were used. The assignment of the crystalline phases was based on the JPDS powder diffraction file cards [19].
Temperature programmed desorption of ammonia (NH3-TPD) was performed using the Micromeritics Autochem 2910 equipped with an ultraviolet gas analyzer (ABB, Limas 11). Before starting the ammonia adsorption experiment, the sample was pre-treated in He flow at 200 °C for 30 min. After cooling to room temperature, a stream of 5% NH3/He (30 mL/min) was flowed over the sample for 1 h. To remove the physically adsorbed ammonia, the sample was purged in flowing He (100 mL/min) at 100 °C for 1 h. Then, cooling down to room temperature, ammonia desorption was monitored with the UV gas analyser, flowing He (30 mL/min) and heating up to 600 °C (rate of 10 °C/min), holding time at 600 °C for 30 min. Before the gas detection system, a cold trap was used to condense any water desorbed from the sample. The ammonia concentration, in terms of ppm NH3 desorbed/gcat under He flow (30 mL/min) was plotted versus time and temperature.
The X-ray photoelectron spectroscopy (XPS) analyses were performed with a VG Microtech ESCA 3000 Multilab, equipped with a dual Mg/Al anode. As excitation source was used the Al Kα radiation (1486.6 eV). The sample powders were mounted on a double-sided adhesive tape. The pressure in the analysis chamber was in the range of 10− 8 Torr during data collection. The constant charging of the samples was removed by referencing all the energies to the C 1s binding energy set at 285.1 eV. Analyses of the peaks were performed with the CasaXPS software.
Hydrogen temperature programmed reduction (H2-TPR) measurements were carried out with a Micromeritics AutoChem 2910 HP Automated Catalyst Characterization System, equipped with a thermal conductivity detector (TCD). An amount of 0.1 g of sample was used for each analysis. The samples were pre-treated with a mixture of 5%vol. O2 in He, at a flow rate of 50 mL/min, heating up (10 °C/min) up to 400 °C and holding at this temperature for 30 min. After cooling down to room temperature under He flow, the gas mixture with composition 5%vol. H2 in Ar was flowed at 30 mL/min into the sample tube. During the analysis, the temperature was increased up to 1000 °C at a rate of 10 °C/min.

2.3 Catalysts Activity test

NO SCR by NH3 tests were performed in a fixed-bed U quartz reactor with an inner diameter of 12 mm. The feed gas, consisting of 1000 ppm NO, 1000 ppm NH3, 10%vol. O2 and He as balance, was led over the catalyst (60 mg) at a flow rate of 50 mL/min (STP), equivalent to a gas hourly space velocity (GHSV) of 50,000 h− 1. The activities were measured as a function of temperature heating from 25 to 400 °C with steps of 25–50 °C and the data were collected after about 40 min at each temperature, in order to achieve steady-state conditions. The inlet and outlet gas compositions were analyzed by ABB detectors, infrared (Limas 11) for NO, N2O, NO2, NH3, paramagnetic (Magnos 206) for O2. The NO conversion and N2 selectivity [20] were calculated by the following equations:
$$NO\,\,Conversion\,\% = \frac{{{{[NO]}_{in}} - {{[NO]}_{out}}}}{{{{[NO]}_{in}}}} \times 100$$
$${N_2}\,\,Selectivity\,(\% ) = \left[ {1 - \frac{{{{[N{O_2}]}_{out}} + 2{{[{N_2}O]}_{out}}}}{{{{[NO]}_{in}} + {{[N{H_3}]}_{in}} - {{[N{H_3}]}_{out}} - {{[NO]}_{out}}}}} \right] \times 100$$

3 Results and Discussion

3.1 Catalysts Characterization

3.1.1 BET and XRD Analyses

In Table 2 the morphological properties of the different catalysts are listed. As shown, the MnOx exhibited the highest surface area (119.1 m2/g), pore volume (0.36 cm3/g) and mean pore size around 11 nm. The surface area of as received TiO2 was 56 m2/g, slightly decreasing to 52 m2/g after calcination. No significant changes occurred in the pore volume and mean pore size that remained around 0.27 cm3/g and 22.0 nm, respectively.
The WO3 loading on TiO2 produced some changes in the morphology by increasing its porosity (pore volume 0.47 cm3/g, pore size 29.2 nm) and surface area (61.7 m2/g). A similar trend occurred after the impregnation of manganese on the WTi oxide, for which SSA increased until 3wt% of manganese loading, while almost stable pore volume and mean pore size values were registered in comparison with the parent WTi. Such findings suggest high dispersion of amorphous MnOx up to a loading of 3wt% in good accordance with the absence of crystalline phases for all xMnWTi catalysts, as reported in the XRD results. As far as 12MnWTi the SSA decreased in comparison with 1.5 MnWTi and as well the pore volume and mean pore size decreased, implying that some agglomeration of manganese oxide on the surface of the catalyst may occur (see also XPS characterization).
The reference commercial catalyst, VWTi, has SSA 55.6 m2/g, pore volume 0.21 cm3/g and mean pore size 15 nm [18].
Table 2
Morphological parameters of the prepared catalysts
Samples
Surface area
(m2 g-1)
BJH Pore Volume
(cm3 /g)
Mean Pore Diameter (nm) by BJH Method
MnOx
119.1
0.36
10.9
TiO2
56.0
0.28
21.1
TiO2 calc.
52.0
0.27
22.1
WTi
61.7
0.47
29.2
1.5MnWTi
74.4
0.50
26.3
3MnWTi
75.9
0.48
25.3
12MnWTi
72.0
0.40
22.1
VWTi
55.6
0.21
15.0
Figure 1 displays the XRD patterns of 12MnWTi, WTi, pure MnOx and the corresponding ICSD references. Regarding the bare MnOx, the intense peaks at 2θ = 37.1° – 42.4° – 56.1° – 66.8° were ascribed to MnO2 mainly as hexagonal phase (ICSD #29,561), though the coexistence of tetragonal phase (ICSD #759,758) cannot be excluded due to quite broad and weak peak around 2θ = 28.5° − 71.8°. In the case of WTi the sharp peaks were attributed to the anatase (ICSD #9854) and rutile (ICSD #9161) phases of titania, while no peaks corresponding tungsten oxide were observed, indicating the highly dispersion of the WO3 on the surface. Compared with the pattern of WTi, the deposition of Mn did not produce any new crystalline phase, no characteristic peaks of MnO2 were detected even at high loading (12MnWTi). As far as the VWTi reference, only peaks of the anatase phase were identified [18], being a commercial catalyst likely deposited over a less common titania.

3.1.2 H2-TPR

The reducibility of xMnWTi samples was investigated by H2-TPR experiments. Corresponding analyses were also performed on MnOx and WTi used as reference materials. The H2-TPR profiles of MnOx, WTi, and xMnWTi catalysts are shown in Fig. 2 and the theoretical and experimental H2 consumption values (mL/g) are listed in Table 3.
As reported in the literature, the MnOx reduction can occur through two-step process MnO2→Mn2O3→MnO [21, 22] according to the shape of the curve. In fact, two strong peaks at 323 and 462 °C with a slight shoulder at around 170 °C were observed for MnOx, corresponding to the stepwise reduction above mentioned. Moreover, according with the literature, the shoulder reduction peak can be attributed to reducible small clusters of surface manganese oxide species [21]. Assuming that all manganese is present as MnO2, the theoretical hydrogen consumption requested for the reduction of Mn4+ to Mn2+ is equal to 281 mL/g (Table 3). Being the experimental consumption equal to 210 mL/g it can be deduced that in the bulk MnOx the manganese species are present mainly as Mn4+ along with Mn3+, as confirmed by XPS results. In accordance with this result, it was proposed that MnO2 and Mn2O3 are the primary active phases for NO reduction in MnOx catalysts for SCR reaction, especially MnO2 [23].
The TiO2 hydrogen consumption reported in Table 3 refers to the titania oxide that was calcined under the same conditions as for MnWTi catalysts. H2-TPR of WTi catalyst (Fig. 2) showed an intense peak at 756 °C related to the reduction of W6+ to W4+ [24]. The reduction peak at 410 °C corresponds to the surface reduction Ti4+ → Ti3+ while the reduction of bulk titanium occurs at high temperature (over 800 °C) [2527].
After the deposition of bare MnOx on WTi, the reduction peaks associated to MnOx in 1.5-3-12MnWTi catalysts were shifted progressively at higher temperature indicating that MnOx strongly interacts with the support and is less easy to reduce. In good accordance with the theoretical and experimental total hydrogen consumption values (see Table 3), manganese was stabilized as Mn3+ over WTi oxide, according to XPS data, and completely reduced to Mn2+.
Table 3
Theoretical and experimental H2 consumption (mL/g) of catalysts calculated by TPR analyses
Catalyst
Reduction steps
Theoretical H2 Volume (mL/g)
Experimental H2 Volume (mL/g)
MnOx
Mn4+ → Mn2+
281
210
TiO2
Ti4+ → Ti3+
-
9
WTi
W6+ → W4+
Ti4+ → Ti3+
9
8
 
Total
17
17
1.5MnWTi
Mn3+ → Mn2+
W6+ → W4+
Ti4+ → Ti3+
3
8
8
 
Total
19
19
3MnWTi
Mn3+ → Mn2+
W6+ → W4+
Ti4+ → Ti3+
6
8
8
 
Total
22
22
12MnWTi
Mn3+ → Mn2+
W6+ → W4+
Ti4+ → Ti3+
25
7
7
 
Total
39
39

3.1.3 XPS Analyses

X ray Photoelectron Spectroscopy has been used to evaluate the changes occurring upon deposition of manganese and tungsten oxides on the titania surface. Figure 3 shows the survey spectra of MnOx, WTi and 12MnWTi. Besides the peaks due to C1s, arising by atmospheric contamination, only peaks due to Mn, Ti, W and O are visible on the surface, indicating that no precursor residues are present in the samples. Figure 4 shows the high-resolution region of Mn 2p for MnOx, 1.5MnWTi and 12MnWTi with the two components of Mn 2p3/2 and Mn 2p1/2 separated by 11.5 eV. In Table 4 the binding energies corresponding to Mn 2p3/2 are reported. Mn 2p3/2 component for MnOx can be deconvoluted with two peaks at 641.9 eV and 643.4 eV which can be attributed to Mn (III) and Mn (IV) respectively [28, 29]. The deposition of MnOx on WTi stabilizes Mn (III) oxidation state as evidenced by the position of the peaks (see Fig. 4; Table 4) in both 1.5 and 12MnWTi. Tungsten position of W 4f7/2 is, for all samples, indicative of W (VI) [3032]. Table 4 shows also the surface atomic ratios. Generally, the ratio measured is higher than the theoretical one, in accord with the fact that both W and Mn have been introduced by impregnation and are therefore localized on the surface. As far as 12MnWTi the surface Mn/Ti ratio is lower than the theoretical one and this is probably due to the high amount of Mn leading to the formation of some agglomerated particles. By looking at W/Ti surface atomic ratio it seems that the deposition of manganese occurs preferentially on titanium, as evidenced by the relative increase of the W/Ti with the increase of MnOx on the sample.
Table 4
Mn 2p3/2, W 4f7/2 binding energy (eV). W/Ti and Mn/Ti atomic ratio
Samples
Mn 2p3/2a
W 4f7/2
W/Ti (0.03)b
Mn/Tib
MnOx
641.9 (62%)
643.4 (38%)
-
-
-
WTi
-
35.7
0.10
-
1.5MnWTi
641.7 (100%)
35.8
0.11
0.04 (0.02)
12MnWTi
642.4 (100%)
35.9
0.15
0.19 (0.24)
a in parenthesis the relative percentage of the different components is reported
b in parenthesis the theoretical atomic ratio is listed

3.1.4 NH3 – TPD Analyses

The surface acidity properties of 1.5-3-12MnWTi and those of the bare MnOx and WTi oxides were investigated by NH3-TPD analyses. The ammonia desorption curves of all catalysts are shown in Fig. 5 and ammonia uptakes are reported in Table 5. All the catalysts exhibited ammonia desorption within the temperature range of 175–600°C. The temperature of the desorption peak maxima gives indications about the nature and the relative strength of acidic sites, while the area under the ammonia desorption represents the number of surface acidic sites. The acid site density was also calculated (see Table 5) by dividing the ammonia uptake by the catalysts surface area. Among the analysed samples, the WTi was characterized by the highest acidity, conversely the bare MnOx showed very low acidic behavior.
The so far reported data are in good agreement with our review article [33] where we reported that tungsten plays a considerable role increasing Brӧnsted and Lewis acidic sites. The number of acidic sites and acidic density exhibited the following order MnOx < 12MnWTi < 3MnWTi < 1.5MnWTi < WTi. As discussed in NH3-SCR performance section, by increasing manganese loading over WTi, the surface acidity promoted by tungsten oxide decreased and consequently the catalysts exhibited high activity in the low temperature range. Conversely, at low manganese loading the acidity of WTi was scarcely influenced resulting in high NO conversion of catalysts at high temperature. Summarizing, the manganese loading deposited over the surface of WTi played the key role of acidity moderator for the tungsten-based catalyst, influencing the catalytic properties of the MnWTi samples that exhibit tailored NO SCR activity at low, medium or high, temperatures.
Table 5
Acidic properties of the prepared catalysts determined by NH3-TPD.
Samples
Temperature range
(°C)
NH3 uptake
(µmol g-1)
Surface acidity density
(µmol NH3 g-1 m-2)
MnOx
215–545
18
0.15
WTi
175–600
89
1.44
1.5MnWTi
175–600
79
1.06
3MnWTi
175–600
52
0.68
12MnWTi
175–545
38
0.53

3.1.5 NO SCR by NH3 Catalytic Performances

Figures 6 and 7 show the NO conversion and N2 selectivity curves for the prepared catalysts and the commercial reference, VWTi. Pure MnOx was able to achieve considerable conversion values, above 80–90% at low temperature, in the range 100–200 °C. Then, the oxidation of ammonia was promoted at T > 200 °C and the NO SCR collapsed towards almost no conversion at 300 °C.
WTi exhibited a completely opposite trend showing extremely poor activity at low temperature and NO conversion between ~ 70–95% in a restricted range, 350–400 °C. This finding confirmed previous investigations about the relatively low efficiency of only WO3 based catalysts [34, 35]. Concerning the xMnWTi samples, it was found that the MnOx loading strongly influenced their catalytic activity. As shown in Figs. 3 and 6MnWTi was the best performing sample with NO conversion values over 80% up to ~ 98% in a wide temperature range, from 150 to 300 °C. In the case of the1.5MnWTi, the NO conversion started to increase at T > 150 °C achieving 92% at 300–350 °C, while the 12MnWTi catalyst reached almost full conversion at 200 °C, with slightly lower values near 90% at 150 and 250 °C, respectively. Based on the so far reported data it results that, as a function of the manganese content, it is possible to modulate the catalytic performances of the MnWTi catalysts in the range of low, medium or high temperature.
As previously discussed, and in agreement with our review [33], tungsten oxide plays a considerable role in NH3-NO SCR by increasing Brӧnsted and Lewis acidity and stabilizing against sintering and transition to rutile the anatase phase of TiO2 that synergically catalyzes the reaction. As reported by Zhu et al. [36] strong Lewis sites are active in the NO SCR at high temperatures (> 350 °C), weak Lewis acid sites contribute at low temperatures (< 200 °C) and Brӧnsted acid sites promote the conversion at medium temperature (250–300 °C). Therefore, it results that the acidity of the MnWTi catalysts increased or decreased as a function of a different Mn loading that affected the NO conversion efficiency in all range of temperature investigated, between 100 and 400 ° C.
Among all the xMnWTi catalysts, 3MnWTi displayed the highest activity in NOx conversion, but an ideal catalyst is expected to have high N2 selectivity in order to avoid emission of harmful nitrogen oxides as NO2 and N2O [37, 38]. As shown in Fig. 7, the selectivity of MnOx rapidly declined already at very low temperature, being less than 10% at 250 °C, while the WTi showed almost constant values not higher than 80–85% in the range 250–400 °C.
In the between of the above-mentioned catalysts, there are 12MnWTi and 3MnWTi showing a declining selectivity trend in the whole temperature window (100–400 °C). The 1.5MnWTi exhibited high selectivity, more than 90% up to 250 ° C with values between 80 − 70% at 350–400 °C. Therefore, such catalyst represents the best compromise between good catalytic performance in NOx conversion at high temperature window (250–400 °C) and high selectivity.
The drop in selectivity observed for MnOx, 3MnWTi and 12MnWTi catalysts was ascribed to the competitive oxidation reaction of ammonia catalysed by manganese, which produces NO, NO2 and N2O and consequently decreased the selectivity [5]. In order to confirm the occurrence of such reaction, in Fig. 8 the NH3, NO, NO2 and N2O concentration (ppm) trends as a function of temperature were compared for two catalysts, the worse selective and the best one, respectively, MnOx and 1.5MnWTi.
In the case of MnOx, simultaneously with the decrease in the concentration of NO and NH3 the signal of N2O appears, already at 50 °C, and increased with the temperature up to a maximum of ~ 600 ppm at 200–250 °C; then above 250 °C the N2O values decreased together with the increase of NO2.
entration (ppm) as a function of temperature for MnOx (a) and 1.5MnWTi (b) catalysts.
As for the 1.5MnWTi, that among the investigated catalyst is the most selective towards N2 formation, the NO and NH3 conversion was accompanied by low N2O formation and almost no NO2 was detected.
After identifying the best compromise between catalytic activity and N2 selectivity in the 1.5MnWTi, we focused on comparing NO conversion and N2 selectivity curves of such catalyst with those of the commercial reference (VWTi) (see Figs. 6 and 7).
It results that although 1.5MnWTi is less selective than the commercial catalyst, especially above 250 °C, it displayed similar or even enhanced (150–200 °C) NO conversion.
Considering the toxicity and low thermal stability of V2O5, that is the active phase in the commercial catalyst along with WO3, the good performances achieved by the 1.5MnWTi represent an important result for the improvement of the environmental sustainability of NH3-NO SCR devices.

4 Conclusion

Different loadings of Mn were successfully deposited by wetness impregnation over WO3/TiO2 and the resulting series of xMnWTi catalysts was investigated for SCR of NO by NH3.
The 3MnWTi catalyst showed the best catalytic activity with NO conversion values over 80% up to ~ 98% in a wide temperature range, from 150 to 300 °C, although the N2 selectivity quickly declined to ~ 60% at 200 °C and even lower at higher temperature. Conversely in the case of the 1.5MnWTi the NO conversion was higher than 90% at 300–350 °C and the selectivity to N2 varied between 90 − 80% in the range of temperature 100–350 °C. Such sample represents the best compromise between NO conversion and N2 selectivity.
BET and XRD characterizations revealed that 1.5MnWTi sample had large surface area and well dispersion of tungsten and manganese oxides. The good accordance between H2 consumption in TPR and XPS data confirmed the presence in the xMnWTi of MnO2 and Mn2O3 that are known as the primary active phases for NO SCR reaction. Surface acidity was influenced by the synergistic effect of tungsten and manganese. The former was able to increase the number of Brӧnsted and Lewis acidic sites, while the latter modulated the acidity of tungsten-based catalyst and consequently influenced the NO conversion and the selectivity to N2 of the MnWTi series making them more or less efficient at high, medium or low temperatures. Among the investigated samples, the 1.5MnWTi represents the best composition. Although with a slight decay in N2 selectivity at temperature above 250 °C, it performed almost as good as the reference commercial catalyst with the great advantage of replacing the toxic V2O5 with MnOx.

Acknowledgements

The projects TECBIA “Tecnologie a Basso Impatto Ambientale per la Produzione di Energia sui Mezzi Navali” (Progetto n. F.090041/01/X36 –CUP B98I17000680008)” and the project NAUSICA (PON “R&S 2014–2020”, grant n. ARS01_00334); project leader NAVTEC cluster are acknowledged for financial support. The authors are grateful to Ing. L. Pinauda (Hug Engineering AG) for providing the commercial catalyst, to G. Napoli (ISMN-CNR, Italy) for the economic report of the above projects, to F. Giordano (ISMN-CNR, Italy) for XRD and to N. Gallì (ISMN-CNR, Italy) for BET analyses.
Open AccessThis article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​.

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Literatur
7.
Zurück zum Zitat Guo J, Li CT, Lu P et al (2011) Research on SCR denitrification of MnOx/Al2O3 modified by CeO2 and its mechanism at low temperature. Huanjing Kexue/Environmental Sci 32:2240–2246 Guo J, Li CT, Lu P et al (2011) Research on SCR denitrification of MnOx/Al2O3 modified by CeO2 and its mechanism at low temperature. Huanjing Kexue/Environmental Sci 32:2240–2246
10.
Zurück zum Zitat Tang XL, Hao JM, Yi HH et al (2007) Selective catalytic reduction of NOx with NH3 by monolithic catalyst MnOxAC/C under low temperature in the presence of excess O2. Zhongguo Huanjing Kexue/China Environ Sci 27:845–850 Tang XL, Hao JM, Yi HH et al (2007) Selective catalytic reduction of NOx with NH3 by monolithic catalyst MnOxAC/C under low temperature in the presence of excess O2. Zhongguo Huanjing Kexue/China Environ Sci 27:845–850
19.
Zurück zum Zitat Inorganic Crystal Structure Database (ICSD) (2014) ; FIZ Karlsruhe, GmbH: Eggenstein-Leopoldshafen, Germany, 2014 Inorganic Crystal Structure Database (ICSD) (2014) ; FIZ Karlsruhe, GmbH: Eggenstein-Leopoldshafen, Germany, 2014
30.
Zurück zum Zitat Komornicki S, Radecka M, Sobas P (2004) Structural properties of TiO2 ± WO3 thin ® lms prepared by r. f. sputtering. Mater Sci 5:527–531 Komornicki S, Radecka M, Sobas P (2004) Structural properties of TiO2 ± WO3 thin ® lms prepared by r. f. sputtering. Mater Sci 5:527–531
Metadaten
Titel
NH3-NO SCR Catalysts for Engine Exhaust Gases Abatement: Replacement of Toxic V2O5 with MnOx to Improve the Environmental Sustainability
verfasst von
L. Consentino
Giuseppe Pantaleo
V. La Parola
C. Migliore
E. La Greca
Leonarda Francesca Liotta
Publikationsdatum
06.12.2022
Verlag
Springer US
Erschienen in
Topics in Catalysis / Ausgabe 13-14/2023
Print ISSN: 1022-5528
Elektronische ISSN: 1572-9028
DOI
https://doi.org/10.1007/s11244-022-01758-4

Weitere Artikel der Ausgabe 13-14/2023

Topics in Catalysis 13-14/2023 Zur Ausgabe

    Marktübersichten

    Die im Laufe eines Jahres in der „adhäsion“ veröffentlichten Marktübersichten helfen Anwendern verschiedenster Branchen, sich einen gezielten Überblick über Lieferantenangebote zu verschaffen.