Next Article in Journal
Significantly Improved Catalytic Performance of Ni-Based MgO Catalyst in Steam Reforming of Phenol by Inducing Mesostructure
Next Article in Special Issue
Part I: A Comparative Thermal Aging Study on the Regenerability of Rh/Al2O3 and Rh/CexOy-ZrO2 as Model Catalysts for Automotive Three Way Catalysts
Previous Article in Journal
Catalytic Cracking of Triglyceride-Rich Biomass toward Lower Olefins over a Nano-ZSM-5/SBA-15 Analog Composite
Previous Article in Special Issue
Use of a µ-Scale Synthetic Gas Bench for Direct Comparison of Urea-SCR and NH3-SCR Reactions over an Oxide Based Powdered Catalyst
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

VOx Surface Coverage Optimization of V2O5/WO3-TiO2 SCR Catalysts by Variation of the V Loading and by Aging

1
Paul Scherrer Institut, CH-5232 Villigen, Switzerland
2
Ecole polytechnique fédérale de Lausanne (EPFL), CH-1015 Lausanne, Switzerland
*
Author to whom correspondence should be addressed.
Catalysts 2015, 5(4), 1704-1720; https://doi.org/10.3390/catal5041704
Submission received: 24 August 2015 / Revised: 25 September 2015 / Accepted: 1 October 2015 / Published: 14 October 2015
(This article belongs to the Special Issue Automotive Emission Control Catalysts)

Abstract

:
V2O5/WO3-TiO2 selective catalytic reduction (SCR) catalysts with a V2O5 loading of 1.7, 2.0, 2.3, 2.6, 2.9, 3.2 and 3.5 wt. % were investigated in the fresh state and after hydrothermal aging at 600 °C for 16 h. The catalysts were characterized by means of nitrogen physisorption, X-ray diffraction and X-ray absorption spectroscopy. In the fresh state, the SCR activity increased with increasing V loading. Upon aging, the catalysts with up to 2.3 wt. % V2O5 exhibited higher NOx reduction activity than in the fresh state, while the catalysts with more than 2.6 wt. % V2O5 showed increasing deactivation tendencies. The observed activation and deactivation were correlated with the change of the VOx and WOx surface coverages. Only catalysts with a VOx coverage below 50% in the aged state did not show deactivation tendencies. With respect to tungsten, above one monolayer of WOx, WO3 particles were formed leading to loss of surface acidity, sintering, catalyst deactivation and early NH3 slip. An optimal compromise between activity and hydrothermal aging resistance could be obtained only with V2O5 between 2.0 and 2.6 wt. %.

Graphical Abstract

1. Introduction

Vanadium containing catalysts are used worldwide as efficient post-treatment catalysts for reducing nitrogen oxides emissions in stationary applications. In this process, NOx reacts with injected NH3 (selective catalytic reduction (SCR)) according to Equation (1) [1].
4 NO + 4 NH 3 + O 2 = 4 N 2 + 6 H 2 O
Since 2005, V-based SCR catalysts have also been applied in mobile sources such as heavy-duty Diesel vehicles, where a urea solution is used as a non-poisonous NH3 source [2]. With the most recent NOx emission regulations such as the Euro 6 standards, NH3-SCR becomes a promising technology also for light-duty Diesel engines.
The SCR catalyst typically consists of anatase TiO2 as support material, WO3 as a promoter for activity and stability and around 2 wt. % V2O5 as the active redox species [3,4,5]. Despite the conventional V2O5 notation, vanadium is actually dispersed as VOx species over the high surface area support [6,7,8]. WO3 is proposed to have multiple promoting effects including the prevention of V island formation [7], increase the number of NH3 adsorption sites [9] or improvement of thermal stability of the thermodynamically unfavorable anatase phase [4]. Extensive research has been undertaken to understand the different aspects of standard V-based SCR catalysts. These efforts were aimed at revealing mechanistic details such as the nature of the active centers and the rate determining steps [10,11,12,13,14,15,16,17], the origin of catalyst aging [18] and the electronic interactions between the various catalyst components and the poisoning phenomena [19,20,21]. For this purpose, parameters such as the synthesis of the TiO2 support material [22,23], the synthesis method, the loading and the nature of WO3 and V2O5 and the structural change upon thermal/hydrothermal treatment have been addressed.
The V content on a WO3-TiO2 (WT) support is a crucial aspect for the activity and stability of the catalyst. A low V2O5 content of 0.5–1.5 wt. % is often chosen for stationary applications, where the volume of the catalytic converter and the activity are not key properties. Instead, sulfur resistance, longevity and production costs are of higher significance [24]. In mobile sources, V-based catalysts need to be highly active due to the limited available space for the exhaust gas treatment system. Hence, the V2O5 content is increased to 2–3 wt. %. Moreover, these catalysts have to operate in a broader temperature window ranging from cold start conditions to full engine load [25]. Finally, the SCR catalyst may experience severe temperature surges because of the regeneration of the upstream diesel particulate filter.
The V loading has been investigated on various support materials such as Al2O3, ZrO2, SiO2 or TiO2 [3,7,26,27,28]. Among these, WT has often been used for V-based SCR catalyst preparation. Madia et al. examined the thermal stability of 1, 2 and 3 wt. % V2O5/WT [18]. The most active and thermally stable catalyst at 600 °C was found at 2 wt. % V2O5. The structural investigation revealed that the enhanced SCR performance is related to the amount of polymeric vanadyl surface species generated by the thermal aging. The decrease in the SCR performance for the high V-loaded catalyst upon aging was related to the loss of surface area and to the growth of three-dimensional vanadia species. Went et al. [26] investigated the different V species by varying the V2O5 loading on TiO2 from 1.3 wt. % to, 2.5, 3.0, 6.1 and 9.8 wt. %. It was shown that both monomeric and polymeric VOx species are present up to 3 wt. % V2O5, while crystalline V2O5 was only detected above 6 wt. % V2O5. Lee et al. [29] investigated V2O5/WT with a V2O5 loading of 1, 3, 5, 7 and 10 wt. % and different preparation methods. The 3 wt. % V2O5/WT exhibited the highest NOx reduction activity in the fresh state at 450 °C and after aging at 650 °C. The V loading of V2O5/WT has been the object of attention of other authors as well, e.g. Putluru et al. [30] (1.5 and 3.0 wt. % V2O5/WT), Amiridis et al. [31] (1.0, 2.0, 3.5, 3.9, 6.6, 8.5, 11.1, and 15.9 wt. % V2O5/WT), Kompio et al. [7] (0.5, 1.5, 3.0, 5.0 wt. % V2O5/WT) and Djerad et al. [3] (3 and 8 wt. % V2O5/WT). All these studies emphasize the importance of an optimal V loading on a TiO2-based SCR catalyst. However, the adjustment of the loading only slightly around 2–3 wt. % V2O5, the concentration range relevant for real-world applications, is missing.
The calcination and aging of V2O5/WT catalysts cause a change of the surface area that is directly linked to the surface density of vanadyl species and as a consequence also the catalytic performance. As Kwon et al. [32] pointed out, one can determine the optimal surface density of the vanadyl species that is crucial for a high catalytic activity. By variation of the V loading (0, 0.5, 1.0, 1.5 and 2.0 wt. % V2O5) in V2O5/TiO2 catalysts, a VOx surface density of 4.5 VOx nm−2 was found to be optimal for high NOx reduction activity. This corresponds to a surface coverage of 55–60%, based on the theoretical maximum VOx surface density of 7.9 VOx nm−2 [33].
In this study, the V2O5 loading of a V2O5/WT catalyst was fine-tuned and the importance of this parameter for the optimum performance of the catalyst is demonstrated. The catalyst composition was systematically altered by loading a WT support with 1.7, 2.0, 2.3, 2.6, 2.9, 3.2 and 3.5 wt. % V2O5. Additionally, the catalysts were hydrothermally aged in order to mimic long-term use. The aging influenced the surface area of the catalyst, which consequently altered the VOx surface coverage and the catalytic activity. It is shown that a subtle variation of the V content around the optimum value causes severe changes in the aging characteristics of V2O5/WT catalysts.

2. Results and Discussion

2.1. Catalytic Activity

The NOx reduction activity of fresh and the hydrothermally aged V2O5/WO3-TiO2 (V2O5/WT) catalysts with increasing V2O5 is reported in Figure 1a and Figure 1b, respectively. Since hydrothermal aging is more practice-relevant compared to thermal aging, the washcoated monoliths were aged at 600 °C for 16 h (GHSV = 10,000 h−1) under a continuous flow of 5 vol. % O2 and 10 vol. % H2O in a flow reactor. All curves in Figure 1a are characterized by a steep increase in activity between 200 °C and 300 °C. In the 350–450 °C temperature regime, the catalysts exhibited an efficiency that was often higher than 95%. Between 500 °C and 550 °C, the NOx reduction activity always decreased due to a selectivity loss (NH3 oxidation to N2 or NO).
Below 300 °C, the fresh catalysts showed increasing NOx reduction activity under NH3 excess (maximum DeNOx) with increasing V2O5 loading from 1.7 wt. % to 3.5 wt. %. The NOx reduction activity values measured at 250 °C are reported in Table 1 for comparison. As an example, the maximum NOx reduction activity for 2.0 wt. % V2O5 at 250 °C and 300 °C was 37% and 78%, respectively, while it was 75% (250 °C) and 96% (300 °C) for the highest loading of 3.5 wt. %. At 550 °C, the trend was reversed and high loading was no longer beneficial for a high NOx reduction activity. The catalyst with 3.5 wt. % V2O5 showed severe selectivity losses so that the originally broad operation window shrank significantly. In the ideal temperature regime of 350–450 °C, all catalysts had NOx reduction activities higher than 95 %.
Figure 1. NOx reduction activity under NH3 excess (maximum DeNOx) of 1.7–3.5 wt. % V2O5/WT (a) in the fresh state and (b) in the aged state (hydrothermal aging at 600 °C for 16 h).
Figure 1. NOx reduction activity under NH3 excess (maximum DeNOx) of 1.7–3.5 wt. % V2O5/WT (a) in the fresh state and (b) in the aged state (hydrothermal aging at 600 °C for 16 h).
Catalysts 05 01704 g001
The aged catalysts (Figure 1b) showed deactivation tendencies for high V2O5 loading. Below 300 °C, the catalysts with a loading up to 2.3 wt. % V2O5 benefited from the aging and became more active than in the fresh state. This was especially evident for the samples with 2.0 and 2.3 wt. % V2O5, whose NOx reduction activities were higher than 50% at 250 °C and 90% at 300 °C, respectively (Table 1). While for 2.6 wt. % V2O5, the NOx reduction activity was comparable to that in the fresh state, the deactivation was more pronounced at higher loadings.
Table 1. NOx reduction activity (%) at 250 °C for the fresh and the aged catalysts as function of V2O5 loading. Data are from Figure 1.
Table 1. NOx reduction activity (%) at 250 °C for the fresh and the aged catalysts as function of V2O5 loading. Data are from Figure 1.
V2O5 (wt. %)1.72.02.32.62.93.23.5
fresh 29.437.245.057.267.669.075.0
aged 45.851.850.855.048.441.235.6
In the medium temperature regime, the catalysts with a loading higher than 2.6 wt. % V2O5 suffered from deactivation. In the high temperature regime, the NOx reduction activity was below 60% for catalysts with more than 2.6 wt. % V2O5. It is important to mention that the entire temperature range of 200–550 °C is required for a complete performance test. If the catalysts are tested only up to 450 or 500 °C, the negative effect of a high V content cannot be perceived. The hydrothermal aging at 600 °C is not very severe but high enough to reveal that the V2O5 loading on WT should not exceed 2.6 wt. % to guarantee sufficient catalyst stability. Below 2.0 wt. % V2O5/WT, the aging did not affect the catalyst negatively but on the other side the activity in the low temperature regime was too low for these catalysts. A good compromise between low temperature activity, stability and high temperature selectivity was found for a loading between 2.0 and 2.6 wt. % V2O5 on WT.
Figure 2. NOx reduction activity (DeNOx) at 10 ppm NH3 slip of 1.7–3.5 wt. % V2O5/WT (a) in the fresh state and (b) in the aged state (hydrothermal aging at 600 °C for 16 h).
Figure 2. NOx reduction activity (DeNOx) at 10 ppm NH3 slip of 1.7–3.5 wt. % V2O5/WT (a) in the fresh state and (b) in the aged state (hydrothermal aging at 600 °C for 16 h).
Catalysts 05 01704 g002
Similar trends to the maximum DeNOx of Figure 1 are observed for the DeNOx at 10 ppm NH3 slip, which is shown in Figure 2. The main difference is that the effect of loading and aging were more pronounced. With a maximum permissible NH3 slip of 10 ppm, which is relevant for practical applications, often lower NOx conversions are obtained than for equimolar dosing conditions or for excess NH3 dosage [1,25,34,35].
While a high V loading was advantageous for high NOx reduction activity in the fresh state (Figure 2a), a low loading was beneficial after aging of the catalyst (Figure 2b). The catalysts with low V loading exhibited an activation, i.e. an improved NOx reduction activity compared to the fresh state. An additional feature of the highly V-loaded catalysts in the fresh state is the late NH3 slip during DeNOx at 10 ppm NH3 slip. This is observable by comparing, e.g., 3.2 wt. % V2O5 in Figure 2a and Figure 1a. The catalyst exhibited similar NOx reduction activities, evidencing that the NH3 slip is not critical; in the aged state, however, the NH3 slip was more pronounced for higher V2O5 loadings. Hence, the difference between maximum DeNOx and DeNOx at 10 ppm NH3 slip became larger. Therefore, the impact of aging can be better evaluated when the DeNOx at 10 ppm NH3 slip is considered. The reason for the enhanced significance of the deactivation is likely related to the fact that the surface acidity is indirectly included in the information delivered by the NH3 slip measurement. If a catalyst loses surface acidity, e.g. by sintering of WO3, the NH3 uptake decreases while the NH3 slip increases. Therefore, the deactivation of high V-loaded catalysts in the aged state (Figure 2b) demonstrates that these catalysts possess lower surface acidity compared to the fresh catalysts.
In the low temperature regime of 200–300 °C, where NOx conversion is determined by the catalyst activity and mass transfer limitations are negligible [25,36,37], the reaction rate constant kmass (Figure 3) was calculated according to Equation (2). With this value, the catalysts can be compared irrespective of the small loading deviations of the washcoat.
Figure 3. Reaction rate constant kmass determined at (a) 200 °C, (b) 250 °C and (c) 300 °C as a function of V2O5 loading in the fresh and the aged state.
Figure 3. Reaction rate constant kmass determined at (a) 200 °C, (b) 250 °C and (c) 300 °C as a function of V2O5 loading in the fresh and the aged state.
Catalysts 05 01704 g003
It is evident that kmass of the fresh catalysts increased from 1.7 to 2.9 wt. % V2O5, whereas it only slightly increased from 2.9 to 3.5 wt. % V2O5. This behaviour is common to all temperatures in Figure 3a–c. A higher loading than 2.9 wt. % V2O5 was therefore no longer beneficial for catalysts in the fresh state. In the aged state, the catalysts with 2.0–2.6 wt. % V2O5 exhibited a kmass of around 30, 150 and 550 cm3g−1s−1 at 200, 250 and 300 °C, respectively. At the lowest loading of 1.7 wt. % V2O5, kmass was lower compared to the loadings around 2.3 wt. % V2O5. At higher loading than 2.6 wt. % V2O5, kmass linearly decreased to about half of the maximum rate constant for 2.6 wt. % V2O5/WT. By comparing the fresh and the aged catalysts, it is apparent that they were only deactivated above 2.6 wt. % V2O5. For 1.7–2.0 wt. % V2O5, kmass increased upon aging, while no effect of aging was observed for loadings between 2.3 and 2.6 wt. % V2O5.
The performance tests in the fresh and the aged state clearly indicate that the V2O5 loading should be adjusted between 2.0 and 2.6 wt. % for a WO3/TiO2 support material. Only in this range, a compromise between good activity and hydrothermal aging resistance was obtained. The determination of the DeNOx at 10 ppm NH3 slip is an important performance test, which reveals the deactivation effect more markedly than the measurement of the maximum DeNOx.

2.2. Characterization

The washcoat material remaining from the slurries was dried and calcined at 550, 600, 650 and 700 °C for 10 h and analyzed using nitrogen physisorption (BET method), X-ray diffraction (XRD) and X-ray absorption near edge spectroscopy (XANES).
The change of the BET specific surface area (SSA) of 2.0, 2.9 and 3.5 wt. % V2O5/WT upon calcination is displayed in Figure 4a. These three samples were chosen to adequately represent the V loading range. After calcination at 550 °C, the SSA of all V-loaded catalysts was lower (by ca. 15 m2/g) than that of the WT support material irrespective of V2O5 loading. In contrast, by increasing the calcination temperature to 600 °C, it becomes evident that a high V loading negatively affected the SSA. For 2.0 wt. % V2O5/WT, the SSA decreased only by 5 m2/g (−7%), while it decreased from 64 m2/g to 38 m2/g (−41%) for 3.5 wt. % V2O5/WT. The increase in calcination temperature from 600 °C to 650 °C had a similar effect on all catalysts, i.e. the SSA decreased by further 25 m2/g. While 2.0 wt. % V2O5/WT still had a residual SSA of 35 m2/g, the SSA of 3.5 wt. % V2O5/WT decreased to 13 m2/g. At 700 °C, the loading did not play a role anymore and the SSA of all V2O5/WT was reduced to below 10 m2/g. For a loading of 2.9 wt. % V2O5, the SSA linearly decreased by 15–20 m2/g with each 50 °C increase in temperature. In contrast to the V-loaded samples, the support material WT exhibited a SSA of 44 m2/g after calcination at 700 °C. This confirms the observation that vanadium assists the sintering of the support material [18]. It is also evident that a high vanadium content accelerates sintering, i.e. sintering can start already at lower calcination temperatures. The low melting point of V2O5 is typically considered responsible for the undesired effect of V on the dispersion of the W-containing phase [38].
Figure 4. Calcination temperature dependence of (a) specific surface area (lines) and crystallite size (dashed lines) of WT, 2.0, 2.9 and 3.5 wt. % V2O5/WT. The crystallite size was obtained from the X-ray diffraction (XRD) reflections of anatase at 2Θ 25.4° and 48.0° using the Scherrer equation. (b) Specific surface area of all catalysts calcined at 600 °C.
Figure 4. Calcination temperature dependence of (a) specific surface area (lines) and crystallite size (dashed lines) of WT, 2.0, 2.9 and 3.5 wt. % V2O5/WT. The crystallite size was obtained from the X-ray diffraction (XRD) reflections of anatase at 2Θ 25.4° and 48.0° using the Scherrer equation. (b) Specific surface area of all catalysts calcined at 600 °C.
Catalysts 05 01704 g004
The impact of loading on SSA is further demonstrated in Figure 4b, where the SSA of 1.7–3.5 wt. % V2O5/WT catalysts calcined at 600 °C is shown. The change of SSA between the support material WT and up to 2.0 wt. % V2O5 was below 10 m2/g but decreased by about 5 m2/g with every additional 0.3 wt. % V2O5. It should be noticed that by comparing maximum DeNOx and SSA at 600 °C, only loadings of 2.0 and 2.3 wt. % V2O5/WT seem to be optimal. Both samples have a SSA above 50 m2/g and their activity/selectivity did not decrease upon aging.
The XRD patterns of 2.0, 3.5 wt. % V2O5/WT and WT are shown for various calcination temperatures in Figure 5a–c. Figure 6 further displays the patterns of all catalysts calcined at 600 °C with an additional inset for the WO3 reflection at 2Θ 23.5°. In all diffractograms, the anatase peaks are visible at 2Θ 25.4° and between 37 and 40°.
For 2.0 wt. % V2O5 (Figure 5a), calcination up to 600 °C did not cause formation of any other XRD visible phase. Hence, we can assume that the vanadium and tungsten species are well dispersed or that their crystallite size is below the detection limit of XRD. At 650 °C, WO3 was detected, which became more prominent at 700 °C together with the beginning of the anatase to rutile phase transformation. The crystallinity of the anatase phase visibly increased (2Θ 37.8°) with increasing calcination temperature.
Figure 5. XRD patterns of (a) 2.0 wt. % V2O5/WT, (b) 3.5 wt. % V2O5/WT, and (c) WT calcined at the indicated temperatures. All diffractograms are normalized using the anatase peak at 2Θ 25.4°.
Figure 5. XRD patterns of (a) 2.0 wt. % V2O5/WT, (b) 3.5 wt. % V2O5/WT, and (c) WT calcined at the indicated temperatures. All diffractograms are normalized using the anatase peak at 2Θ 25.4°.
Catalysts 05 01704 g005
The XRD patterns of 3.5 wt. % V2O5/WT in Figure 5b revealed the same phase evolution as in Figure 5a. However, phase changes occurred already at lower calcination temperatures than in the case of 2.0 wt. % V2O5/WT. The WO3 crystallites were already detected at 600 °C and the anatase to rutile phase transformation started at 650 °C. Despite structural changes, no V-containing phase was detected and the VOx species seemed to remain well dispersed. The changes in the speciation of the WO3 phase are associated with the presence of V in agreement with the BET observations (Figure 4). At identical calcination temperatures, WT did not present any evidence of WO3 sintering.
The crystallite size of anatase TiO2 was determined using the Scherrer equation from the XRD patterns of Figure 5 and the values are reported in Figure 4a. At 550 °C, the crystallite size was around 20 nm irrespective of the V loading. It increased already at 600 °C for catalysts with high V loading. At 650 °C, the deviation between low and high V loading was more pronounced and varied from 33 (±1) to 63 (±3) nm, respectively. At 700 °C, no significant size difference between low and high V loading was found (around 80 nm) in agreement with the similar SSA values. Both, the SSA and the crystallite size showed that the catalysts sintered with increasing V loading and increasing calcination temperature.
All catalysts were also analyzed by XRD after calcination at 600 °C (Figure 6). WO3 was not visible below 2.3 wt. % V2O5 and the diffractograms approximately matched that of the support material, WT. The contribution of WO3 appeared first for 2.6 wt. % V2O5/WT and became more prominent with increasing V loading.
Figure 6. XRD patterns of 1.7–3.5 wt. % V2O5/WT calcined at 600 °C.The inset shows a magnification around the WO3 peak at 2Θ 23.5°. All diffractograms are normalized using the anatase peak at 2Θ 25.4°.
Figure 6. XRD patterns of 1.7–3.5 wt. % V2O5/WT calcined at 600 °C.The inset shows a magnification around the WO3 peak at 2Θ 23.5°. All diffractograms are normalized using the anatase peak at 2Θ 25.4°.
Catalysts 05 01704 g006
Since XRD was not able to provide information on V because of its low loading and low aggregation state, two catalysts were selected for an element specific characterization. The local environment of the VOx species supported on WT was studied using X-ray absorption near edge structure spectroscopy (XANES).
The normalized XANES spectra of 2.0 and 3.5 wt. % V2O5 /WT calcined at 550 and 650 °C recorded at the V K-edge are shown in Figure 7. Without attempting a quantitative assessment of both oxidation state and coordination of the VOx species represented by these spectra, a qualitative discussion is sufficient to support our interpretation of the loading and sintering effects. The spectra of 2.0 and 3.5 wt. % V2O5/WT calcined at 550 °C are comparable, suggesting a similar local environment of V. This is in agreement with the XRD data of Figure 5, where no formation of any VOx phase was observed because the agglomerates were not crystalline enough or below the detection limit. Upon calcination at 650 °C, an obvious change in the XANES region is observed for 3.5 wt. % V2O5/WT. The whiteline region (up to 5.52 keV) develops three features characteristic of V2O5, suggesting that the VOx species evolved to form V2O5-like agglomerates [2,25]. This change is correlated with the increase of VOx surface coverage from 46% in 2.0 wt. % V2O5/WT to 232% in 3.5 wt. % V2O5/WT that is discussed in the next section. Hence, VOx started to adopt a different local environment when the VOx surface coverage exceeded one theoretical monolayer.
Figure 7. Normalized V K-edge X-ray absorption near edge structure spectroscopy (XANES) spectra of selected catalysts in the fresh and the aged state.
Figure 7. Normalized V K-edge X-ray absorption near edge structure spectroscopy (XANES) spectra of selected catalysts in the fresh and the aged state.
Catalysts 05 01704 g007

2.2.1. VOx Surface Coverage

The surface coverage of VOx and WOx provides additional insight for structure–activity relationships. The VOx and WOx surface coverages were calculated from the SSA considering that the theoretical saturation values for the monolayer coverage are 7.9 VOx nm−2 and 4.2 WOx nm−2, respectively [33]. As a result, Figure 8 shows the surface coverage of a low, medium and high loaded V2O5/WT catalyst at various calcination temperatures.
For both high SCR activity and selectivity, the VOx coverage should be below the monolayer level [30]. The VOx coverage in Figure 8a was below 50% for a calcination at 550 °C and below 100% for all catalysts at 600 °C. At 650 °C, 3.5 wt. % V2O5/WT exceeds the monolayer coverage and at 700 °C, all catalysts theoretically exhibit a multilayer of VOx species. This was not observable by XRD (Figure 5) but only in the XANES spectra (Figure 7), which revealed polymeric VOx with a vanadium environment similar to that of V2O5.
Figure 8c further displays the VOx coverage of all catalysts calcined at 600 °C, which varies from 24% to 78%. This corresponds well with the XRD observation and the absence of V2O5. A vanadium coverage of around 27, 36 and 46% was calculated for 2.0, 2.3 and 2.6 wt. % V2O5/WT calcined at 600 °C, respectively. By comparing the VOx coverage with the rate constants in Figure 3a–c, it can be concluded that an optimized VOx coverage should be between 25–50%. Below this value, the catalysts seem to have not enough active sites for the SCR reaction, while above 50%, the catalysts showed strong selectivity issues. This is evident from Figure 8a. The fresh 3.5 wt. % V2O5/WT catalyst had a surface coverage of 46% and was very active. Upon aging at 600 °C, the coverage increased to 78% and the catalyst experienced strong deactivation. Similar considerations apply for the activation of catalysts with a low V loading. After increasing the calcination from 550 to 650 °C, the VOx coverage of, e.g., 2.0 wt. % V2O5/WT increased from below 25% to roughly 50% (Figure 8a), which is a plausible explanation for the activation.
A study by Amiridis et al. [31] estimated an optimum V2O5 coverage of around 2 µmol/m2 (2.4 VOx/nm2 = 30% surface coverage) for V2O5/TiO2 doped with around 0.8 wt. % WO3. The optimum catalyst was chosen from the activation energy of the catalysts, which only slightly decreased with further increasing the V content. Kwon et al. [32] have shown that VOx coverage of 55–60% is optimal for the catalytic activity of V2O5/TiO2. In the present system, the optimum coverage seems to be lower, which can be explained by the enhanced surface acidity induced by WO3 [4]. Adsorbed NH3 becomes more readily available for the SCR reaction; hence a lower VOx coverage is sufficient for optimizing the catalyst performance in the presence of WO3.
Figure 8. Surface coverage of (a) VOx and (b) WOx for the support material WT and three different catalysts as function of calcination temperature. (c) Surface coverage of VOx and WOx as function of V2O5 loading at 600 °C.
Figure 8. Surface coverage of (a) VOx and (b) WOx for the support material WT and three different catalysts as function of calcination temperature. (c) Surface coverage of VOx and WOx as function of V2O5 loading at 600 °C.
Catalysts 05 01704 g008

2.2.2. WOx Surface Coverage

In the case of WOx (Figure 8b), only calcination at 550 °C produced a sub-monolayer coverage for all catalysts. Already at 600 °C, the catalysts with high V content exceeded the theoretical WOx monolayer as shown in greater detail in Figure 8c. Up to 2.0 wt. % V2O5, the WOx coverage remained below one monolayer, while it was slightly above for 2.3 wt. % V2O5 (115%). For higher V contents, the WOx surface coverage increased in agreement with the XRD observation of Figure 6. At 2.6 wt. % V2O5 or higher, WO3 crystallites were detected and the WOx surface coverage was noticeably above one monolayer.
Comparison of the surface coverage with the NOx reduction performance in Figure 1b and Figure 2b suggests that the aging at 600 °C and catalyst activation/deactivation can be correlated with the increase in surface coverage of WOx. Tungsten oxide provides acid sites to the catalyst where the NH3 is supposed to adsorb and be readily available for the SCR reaction [4]. Upon calcination at increasing temperature, the fraction of available WOx species and hence the surface acidity decrease and the catalyst increasingly loses activity. This is clearly visible from the DeNOx at 10 ppm NH3 slip (Figure 2). The decrease in acidity caused by crystallization of WO3 reduces the potential adsorption sites for NH3 thus enhancing the 10 ppm NH3 slip.

3. Experimental Section

3.1. Materials

All samples were prepared by wet impregnation of a commercial WO3/TiO2 (WT) support with NH4VO3. NH4VO3 (equivalent (eq.) to 1.7–3.5 wt. % V2O5, Sigma Aldrich, Buchs, Switzerland,) was dissolved in H2O (10 mL) and added to a 30 mL aqueous slurry of WO3/TiO2 (WT, 12 g, “Tiona DT-52”, 10 wt. % WO3 and 90 wt. % TiO2, Cristal Global, Thann, France). After the slurry was sonicated for 10 min in an ultrasonic bath, homogenized with a disperser (Miccra D-8, Schmizo AG, Zofingen, Switzerland, 20,000 rpm, 5 min) and stirred for 60 min, water was evaporated under reduced pressure and the sample was dried at 120 °C for 12 h. Finally, the sample was grinded thoroughly and was calcined at 550 °C in a muffle oven for 5 h (fresh catalyst).
For washcoating, the powders were suspended in a mixture of water (2 eq. of sample) and colloidal silicate (Ludox, 40 wt. % in H2O, Sigma Aldrich, 0.1 eq. of TiO2). After sonication of the slurry for 10 min in an ultrasonic bath, the honeycomb monoliths (cordierite, 400 cpsi, ca. 12 × 17 × 50 mm) were dip coated. The monoliths were repeatedly immersed in the slurry and dried with an air blower to reach a loading of the active material of around 1.25–1.35 g. The monoliths and the remaining slurry were dried overnight at 120 °C and calcined at 450 °C for 10 h in a muffle oven. The hydrothermal aging of the washcoated monoliths (600 °C for 16 h) was performed in a lab scale flow reactor in 20 vol% O2 and 10 vol% H2O with balance N2 at GHSV = 10,000 h−1.

3.2. Catalytic Measurements

The washcoated monoliths were tested in a laboratory test reactor described elsewhere [34,39] under a feed of 10 vol% O2, 5 vol% H2O, 500 ppm NO, 0–600 ppm NH3 with balance N2, in order to mimic realistic exhaust gas composition. The gas hourly space velocity (GHSV = volumetric gas flow/coated monolith volume) was 50,000 h−1, which is typical of SCR converters of diesel vehicles [39]. The maximum NOx reduction activity was measured by dosing NH3 in excess, i.e. at NH3/NOx = 1.2. The excess of NH3 is exploited to achieve a maximum DeNOx that is not affected by possible side reactions of NH3. Since equation 1 is in zeroth order with respect to NH3, the NH3 does not influence the equilibrium. In order to obtain a more practice-oriented value for SCR systems, the NOx reduction activity at 10 ppm NH3 slip was measured as well, as described earlier [1,25,34,35]. The mass specific rate constant (kmass) for the maximum DeNOx was calculated according to Equation (2), under the assumption of a pseudo-first order of the SCR reaction with respect to NO and zeroth order with respect to NH3 [40,41],
k m a s s = V * W · ln ( 1 X N O x )
where V* is the total flow rate at reaction condition, W the loading of the active component and XNOx the fractional NOx conversion. Although adsorption of both NH3 and NO occurs at low temperature [1], NH3 adsorption dominates on acidic SCR catalysts so that the first order SCR reaction with respect to NO is justified. The rate constant is independent of the active component loading, which is particularly important for coated monoliths where small loading deviations are unavoidable. The NOx reduction efficiency (DeNOx) was estimated according to Equation (3), [3,41],
DeNO x = C NO in  C NO x out C NO in   ·   100 %
where CinNO is the NO concentration upstream of the catalyst and CoutNOx the NO and NO2 concentrations downstream of the catalyst. Online gas analysis of the exhaust gas was performed with a Fourier transform infrared spectrometer (Nexus 670 ThermoNicolet, ThermoFisher, Schwerte, Germany) equipped with a heated gas cell.

3.3. Characterization Methods

The BET specific surface area (SSA) was measured by N2 adsorption at −196 °C on a Quantachrome Autosorb I instrument (Quantachrome Instruments, Boynton Beach, FL, USA). Prior to the measurement, the samples were outgassed at 350 °C for 10 h. Powder X-ray diffraction (PXRD) patterns were collected on a D8 ADVANCE (Bruker AXS GmbH, Karlsruhe, Germany) diffractometer using Cu Kα1 radiation (λ = 1.5406 Å). Data were recorded from 10 to 65° 2Θ using a step size of 0.03°/1 s acquisition time. The phases were identified with the X'Pert HighScore Plus software (2.0a, PANalytical B.V., Almelo, Netherlands, 2004). The crystallite size of TiO2 was determined by the Scherrer equation using the peaks at 25.4° and at 48.0°. X-ray absorption near edge structure (XANES) spectra were recorded at beamline SuperXAS of the Swiss Light Source (SLS, Villigen, Switzerland). The spectra were collected around the V K-edge (E0 = 5.465 keV) in fluorescence mode using double Si(111) crystal monochromator and a 10 µm V foil to calibrate the monochromator position. Samples were diluted with cellulose and pressed into pellets. The data were aligned, background corrected, and normalized using Athena (IFFEFIT software package, 1.2.11d, Free Software Foundation, Boston, MA, USA, 2013) [42].

4. Conclusions

We studied the activity and stability of V2O5/WO3-TiO2 SCR catalysts by systematically altering the vanadium content. Only V2O5 loadings between 2.0 and 2.6 wt. % withstand a moderate hydrothermal aging at 600 °C for 16 h and are in fact activated upon aging. Below 2.0 wt. % V2O5, sufficient activity cannot be guaranteed, while above 2.6 wt. % V2O5 deactivation upon aging increasingly occurs. On the base of NOx reduction activity at 10 ppm NH3 slip, we could show that the aging at high V2O5 loading is in fact more severe than one could anticipate from a standard measurement of the SCR activity with equimolar dosage of NH3 and NOx. The early NH3 slip is a practical and direct indicator of the loss of surface acidity caused by the sintering of WO3, which is confirmed by the XRD data. In the case of VOx species at the loadings studied in this work, XRD is not able to yield any information. The evolution from highly dispersed VOx species to a polymeric VOx environment similar to that of V2O5 upon aging of the catalyst with a high V loading was qualitatively captured by V K-edge XANES.
The catalytic performance was further correlated with the surface coverage of WOx and VOx. An optimum surface VOx coverage between 25–50 % was estimated, which can be adjusted by either the V loading or the specific surface area (calcination temperature). For the WOx surface coverage, it was shown that above one WOx monolayer, WO3 crystallites are formed, thus diminishing the NH3 uptake and hence the activity of the catalyst. It can be concluded that a V2O5 loading not higher than 2.6 wt. % should be used in SCR catalysts composed of V2O5, 10 wt. % WO3 and TiO2 in order to maintain the activity and stability of the catalyst.

Acknowledgments

The authors gratefully acknowledge financial support by Treibacher Industrie AG and the SLS for beamtime allocation at beamline SuperXAS.

Author Contributions

The experimental work was conceived and designed by A.M., M.E. and D.F.; M.E. and A.M. performed the experiments; A.M., and O.K, analyzed the data; M.E. contributed reagents/materials/analysis tools; A.M., D.F and O.K. drafted the paper. The manuscript was amended through the comments of all authors. All authors have given approval for the final version of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Koebel, M.; Elsener, M.; Kleemann, M. Urea-SCR: A promising technique to reduce NOx emissions from automotive diesel engines. Catal. Today 2000, 59, 335–345. [Google Scholar] [CrossRef]
  2. Beale, A.M.; Lezcano-Gonzalez, I.; Maunula, T.; Palgrave, R.G. Development and characterization of thermally stable supported V–W–TiO2 catalysts for mobile NH3-SCR applications. Catal. Struct. React. 2015, 1, 25–34. [Google Scholar] [CrossRef]
  3. Djerad, S.; Tifouti, L.; Crocoll, M.; Weisweiler, W. Effect of vanadia and tungsten loadings on the physical and chemical characteristics of V2O5-WO3/TiO2 catalysts. J. Mol. Catal. A 2004, 208, 257–265. [Google Scholar] [CrossRef]
  4. Forzatti, P. Present status and perspectives in de-NOx SCR catalysis. Appl. Catal. A 2001, 222, 221–236. [Google Scholar] [CrossRef]
  5. Nova, I.; Tronconi, E. Urea-SCR Technology for DeNOx After Treatment of Diesel Exhausts; Springer: New York, NY, USA, 2014; pp. 6, 25 and 68. [Google Scholar]
  6. Fierro, J.L.G. Metal Oxides: Chemistry and Applications; CRC Press: Florida, FI, USA, 2005; pp. 8–20. [Google Scholar]
  7. Kompio, P.G.W.A.; Brückner, A.; Hipler, F.; Auer, G.; Löffler, E.; Grünert, W. A new view on the relations between tungsten and vanadium in V2O5WO3/TiO2 catalysts for the selective reduction of NO with NH3. J. Catal. 2012, 286, 237–247. [Google Scholar] [CrossRef]
  8. Wachs, I.E. Recent conceptual advances in the catalysis science of mixed metal oxide catalytic materials. Catal. Today 2005, 100, 79–94. [Google Scholar] [CrossRef]
  9. Amiridis, M.D.; Duevel, R.V.; Wachs, I.E. The effect of metal oxide additives on the activity of V2O5/TiO2 catalysts for the selective catalytic reduction of nitric oxide by ammonia. Appl. Catal. B 1999, 20, 111–122. [Google Scholar] [CrossRef]
  10. Anstrom, M.; Topsøe, N.-Y.; Dumesic, J.A. Density functional theory studies of mechanistic aspects of the SCR reaction on vanadium oxide catalysts. J. Catal. 2003, 213, 115–125. [Google Scholar] [CrossRef]
  11. Gruber, M.; Hermann, K. Elementary steps of the catalytic NOx reduction with NH3: Cluster studies on reaction paths and energetics at vanadium oxide substrate. J. Chem. Phys. 2013, 139, 244701–244708. [Google Scholar] [CrossRef] [PubMed]
  12. 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]
  13. Topsøe, N.-Y. Mechanism of the Selective Catalytic Reduction of Nitric Oxide by Ammonia Elucidated by in Situ On-Line Fourier Transform Infrared Spectroscopy. Science 1994, 265, 1217–1219. [Google Scholar] [CrossRef] [PubMed]
  14. Topsøe, N.Y.; Dumesic, J.A.; Topsøe, H. Vanadia-Titania Catalysts for Selective Catalytic Reduction of Nitric-Oxide by Ammonia: I.I. Studies of Active Sites and Formulation of Catalytic Cycles. J. Catal. 1995, 151, 241–252. [Google Scholar] [CrossRef]
  15. Tronconi, E.; Nova, I.; Ciardelli, C.; Chatterjee, D.; Weibel, M. Redox features in the catalytic mechanism of the “standard” and “fast” NH3-SCR of NOx over a V-based catalyst investigated by dynamic methods. J. Catal. 2007, 245, 1–10. [Google Scholar] [CrossRef]
  16. Vittadini, A.; Casarin, M.; Selloni, A. First Principles Studies of Vanadia-Titania Monolayer Catalysts:  Mechanisms of NO Selective Reduction. J. Phys. Chem. B 2005, 109, 1652–1655. [Google Scholar] [CrossRef] [PubMed]
  17. Zheng, Y.; Jensen, A.D.; Johnsson, J.E.; Thøgersen, J.R. Deactivation of V2O5-WO3-TiO2 SCR catalyst at biomass fired power plants: Elucidation of mechanisms by lab- and pilot-scale experiments. Appl. Catal. B 2008, 83, 186–194. [Google Scholar] [CrossRef]
  18. Madia, G.; Elsener, M.; Koebel, M.; Raimondi, F.; Wokaun, A. Thermal stability of vanadia-tungsta-titania catalysts in the SCR process. Appl. Catal. B 2002, 39, 181–190. [Google Scholar] [CrossRef]
  19. Nicosia, D.; Elsener, M.; Kröcher, O.; Jansohn, P. Basic investigation of the chemical deactivation of V2O5/WO3-TiO2 SCR catalysts by potassium, calcium, and phosphate. Top. Catal. 2007, 42–43, 333–336. [Google Scholar] [CrossRef]
  20. Xie, X.; Lu, J.; Hums, E.; Huang, Q.; Lu, Z. Study on the Deactivation of V2O5-WO3/TiO2 Selective Catalytic Reduction Catalysts through Transient Kinetics. Energy Fuels 2015, 29, 3890–3896. [Google Scholar] [CrossRef]
  21. Kamata, H.; Takahashi, K.; Odenbrand, C.U.I. The role of K2O in the selective reduction of NO with NH3 over a V2O5(WO3)/TiO2 commercial selective catalytic reduction catalyst. J. Mol. Catal. A 1999, 139, 189–198. [Google Scholar] [CrossRef]
  22. Zhang, X.; Li, X.; Wu, J.; Yang, R.; Zhang, Z. Selective Catalytic Reduction of NO by Ammonia on V2O5/TiO2 Catalyst Prepared by Sol-Gel Method. Catal. Lett. 2009, 130, 235–238. [Google Scholar]
  23. Georgiadou, I.; Papadopoulou, C.; Matralis, H.K.; Voyiatzis, G.A.; Lycourghiotis, A.; Kordulis, C. Preparation, Characterization, and Catalytic Properties for the SCR of NO by NH3 of V2O5/TiO2 Catalysts Prepared by Equilibrium Deposition Filtration. J. Phys. Chem. B 1998, 102, 8459–8468. [Google Scholar] [CrossRef]
  24. 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] [CrossRef]
  25. Marberger, A.; Elsener, M.; Ferri, D.; Sagar, A.; Schermanz, K.; Kröcher, O. Generation of NH3 Selective Catalytic Reduction Active Catalysts from Decomposition of Supported FeVO4. ACS Catal. 2015, 4180–4188. [Google Scholar] [CrossRef]
  26. Went, G.T.; Leu, L.-j.; Bell, A.T. Quantitative structural analysis of dispersed vanadia species in TiO2(anatase)-supported V2O5. J. Catal. 1992, 134, 479–491. [Google Scholar] [CrossRef]
  27. Burkardt, A.; Weisweiler, W.; van den Tillaart, J.A.A.; Schäfer-Sindlinger, A.; Lox, E.S. Influence of the V2O5 Loading on the Structure and Activity of V2O5/TiO2 SCR Catalysts for Vehicle Application. Top. Catal. 2001, 16–17, 369–375. [Google Scholar] [CrossRef]
  28. Briand, L.E.; Tkachenko, O.P.; Guraya, M.; Gao, X.; Wachs, I.E.; Grünert, W. Surface-Analytical Studies of Supported Vanadium Oxide Monolayer Catalysts. J. Phys. Chem. B 2004, 108, 4823–4830. [Google Scholar] [CrossRef]
  29. Lee, B.W.; Hyun, C.; Shin, D.W. Characterization and De-NOx activity of binary V2O5/TiO2 and WO3/TiO2, and ternary V2O5-WO3/TiO2 SCR catalysts. J. Ceram. Process. Res. 2007, 8, 203–207. [Google Scholar]
  30. Putluru, S.; Schill, L.; Gardini, D.; Mossin, S.; Wagner, J.; Jensen, A.; Fehrmann, R. Superior DeNOx activity of V2O5-WO3/TiO2 catalysts prepared by deposition–precipitation method. J. Mater. Sci. 2014, 49, 2705–2713. [Google Scholar] [CrossRef]
  31. Amiridis, M.D.; Solar, J.P. Selective Catalytic Reduction of Nitric Oxide by Ammonia over V2O5/TiO2, V2O5/TiO2/SiO2, and V2O5-WO3/TiO2 Catalysts:  Effect of Vanadia Content on the Activation Energy. Ind. Eng. Chem. Res. 1996, 35, 978–981. [Google Scholar] [CrossRef]
  32. Kwon, D.W.; Park, K.H.; Hong, S.C. Influence of VOx surface density and vanadyl species on the selective catalytic reduction of NO by NH3 over VOx/TiO2 for superior catalytic activity. Appl. Catal. A 2015, 499, 1–12. [Google Scholar]
  33. Wachs, I.E. Raman and IR studies of surface metal oxide species on oxide supports: Supported metal oxide catalysts. Catal. Today 1996, 27, 437–455. [Google Scholar] [CrossRef]
  34. Kleemann, M.; Elsener, M.; Koebel, M.; Wokaun, A. Investigation of the ammonia adsorption on monolithic SCR catalysts by transient response analysis. Appl. Catal. B 2000, 27, 231–242. [Google Scholar] [CrossRef]
  35. Zhao, Y.; Hu, J.; Hua, L.; Shuai, S.; Wang, J. Ammonia Storage and Slip in a Urea Selective Catalytic Reduction Catalyst under Steady and Transient Conditions. Ind. Eng. Chem. Res. 2011, 50, 11863–11871. [Google Scholar] [CrossRef]
  36. Metkar, P.S.; Balakotaiah, V.; Harold, M.P. Experimental and kinetic modeling study of NO oxidation: Comparison of Fe and Cu-zeolite catalysts. Catal. Today 2012, 184, 115–128. [Google Scholar] [CrossRef]
  37. Koebel, M.; Madia, G.; Elsener, M. Selective catalytic reduction of NO and NO2 at low temperatures. Catal. Today 2002, 73, 239–247. [Google Scholar]
  38. Carreon, M.A.; Guliants, V.V. Ordered Meso- and Macroporous Binary and Mixed Metal Oxides. Eur. J. Inorg. Chem. 2005, 2005, 27–43. [Google Scholar] [CrossRef]
  39. Kröcher, O.; Devadas, M.; Elsener, M.; Wokaun, A.; Söger, N.; Pfeifer, M.; Demel, Y.; Mussmann, L. Investigation of the selective catalytic reduction of NO by NH3 on Fe-ZSM5 monolith catalysts. Appl. Catal. B 2006, 66, 208–216. [Google Scholar] [CrossRef]
  40. Koebel, M.; Elsener, M. Selective catalytic reduction of NO over commercial DeNOx-catalysts: experimental determination of kinetic and thermodynamic parameters. Chem. Eng. Sci. 1998, 53, 657–669. [Google Scholar] [CrossRef]
  41. Casapu, M.; Kröcher, O.; Elsener, M. Screening of doped MnOx-CeO2 catalysts for low-temperature NO-SCR. Appl. Catal. B 2009, 88, 413–419. [Google Scholar] [CrossRef]
  42. Ravel, B.; Newville, M. ATHENA, ARTEMIS, HEPHAESTUS: data analysis for X-ray absorption spectroscopy using IFEFFIT. J. Synchrotron Radiat. 2005, 12, 537–541. [Google Scholar] [CrossRef] [PubMed]

Share and Cite

MDPI and ACS Style

Marberger, A.; Elsener, M.; Ferri, D.; Kröcher, O. VOx Surface Coverage Optimization of V2O5/WO3-TiO2 SCR Catalysts by Variation of the V Loading and by Aging. Catalysts 2015, 5, 1704-1720. https://doi.org/10.3390/catal5041704

AMA Style

Marberger A, Elsener M, Ferri D, Kröcher O. VOx Surface Coverage Optimization of V2O5/WO3-TiO2 SCR Catalysts by Variation of the V Loading and by Aging. Catalysts. 2015; 5(4):1704-1720. https://doi.org/10.3390/catal5041704

Chicago/Turabian Style

Marberger, Adrian, Martin Elsener, Davide Ferri, and Oliver Kröcher. 2015. "VOx Surface Coverage Optimization of V2O5/WO3-TiO2 SCR Catalysts by Variation of the V Loading and by Aging" Catalysts 5, no. 4: 1704-1720. https://doi.org/10.3390/catal5041704

Article Metrics

Back to TopTop