Next Article in Journal
Benchmarking Perovskite Electrocatalysts’ OER Activity as Candidate Materials for Industrial Alkaline Water Electrolysis
Next Article in Special Issue
Characterization of the Evolution of Noble Metal Particles in a Commercial Three-Way Catalyst: Correlation between Real and Simulated Ageing
Previous Article in Journal
Catalytic Cracking of n-Hexadecane Using Carbon Nanostructures/Nano-Zeolite-Y Composite Catalyst
Previous Article in Special Issue
Self-Regeneration Effect of Three-Way Catalysts during Thermal Aging Procedure
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

High Surface Area VOx/TiO2/SBA-15 Model Catalysts for Ammonia SCR Prepared by Atomic Layer Deposition

Eduard-Zintl-Institut für Anorganische und Physikalische Chemie, Technical University of Darmstadt, Alarich-Weiss-Str. 8, 64287 Darmstadt, Germany
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(12), 1386; https://doi.org/10.3390/catal10121386
Submission received: 30 October 2020 / Revised: 20 November 2020 / Accepted: 25 November 2020 / Published: 28 November 2020

Abstract

:
The mode of operation of titania-supported vanadia (VOx) catalysts for NOx abatement using ammonia selective catalytic reduction (NH3-SCR) is still vigorously debated. We introduce a new high surface area VOx/TiO2/SBA-15 model catalyst system based on mesoporous silica SBA-15 making use of atomic layer deposition (ALD) for controlled synthesis of titania and vanadia multilayers. The bulk and surface structure is characterized by X-ray diffraction (XRD), UV-vis and Raman spectroscopy, as well as X-ray photoelectron spectroscopy (XPS), revealing the presence of dispersed surface VOx species on amorphous TiO2 domains on SBA-15, forming hybrid Si–O–V and Ti–O–V linkages. Temperature-dependent analysis of the ammonia SCR catalytic activity reveals NOx conversion levels of up to ~60%. In situ and operando diffuse reflection IR Fourier transform (DRIFT) spectroscopy shows N–Hstretching modes, representing adsorbed ammonia and -NH2 and -NH intermediate structures on Bronsted and Lewis acid sites. Partial Lewis acid sites with adjacent redox sites are proposed as the active sites and desorption of product molecules as the rate-determining step at low temperature. The high NOx conversion is attributed to the presence of highly dispersed VOx species and the moderate acidity of VOx supported on TiO2/SBA-15.

Graphical Abstract

1. Introduction

The emission of nitrogen oxides (NOx) into the atmosphere, mainly from the electric-power industry and daily traffic, brings about a tremendous threat to the environment as well as human health. As one of the major atmospheric pollutants, NOx has attracted increasing attention in recent years. Selective catalytic reduction (SCR) of NOx with NH3 is proved to be the most effective technology for the removal of NOx from stationary and mobile sources [1]. Commercial vanadia-based ammonia SCR catalysts are typically based on vanadia and tungstia (or molybdena) on TiO2 anatase, but vanadia supported on TiO2 has been extensively studied as a model catalyst system [2,3,4].
It is known from the literature, that SCR activities depend on the vanadia structure and the support material [5,6,7]. Mesoporous materials such as silica SBA-15 provide higher surface area than conventional powder or planar supports [8,9], and the controlled synthesis of vanadia supported on SBA-15 was reported [10,11]. However, SiO2 supported VOx-based catalysts were reported to show only poor SCR performance compared to titania-based systems [5,12]. On the other hand, Segura et al. studied vanadia/titania supported on SBA-15 and attributed the good catalytic NH3-SCR reactivity to isolated VOx or TiO2 species [13]. Reiche et al. prepared high surface area TiO2, SiO2, and TiO2-SiO2 supported vanadia systems using sol-gel methods and selective vanadia grafting and found TiO2-based catalysts to show the highest SCR activity. In mixed TiO2-SiO2 aerogels, vanadia was reported to be preferentially grafted to Ti sites [5]. Compared to these conventional synthesis techniques, atomic layer deposition (ALD) enables accurate control at the atomic scale without damaging the structure of the original matrix [9,14,15,16]. Although ALD has been widely applied to metal-oxide deposition, such as SiO2 [17,18,19], TiO2 [20,21,22,23,24], and VOx [9,25,26,27,28,29], there are only very few studies on the deposition of vanadia and titania on mesoporous SBA-15 [30], and its application to NH3-SCR [29].
A limitation of VOx-based catalysts for NOx removal by NH3-SCR is the high and narrow temperature window of acceptable deNOx efficiencies. While VOx/SiO2 catalysts are characterized by a wide temperature window but extremely low NOx conversion, attributed to the inert support and low acid site adsorption, VOx/TiO2 catalysts show higher NOx conversions within a narrow temperature window (300–400 °C), resulting from the temperature dependence of the adsorption and redox reaction behavior. One may envision that combining both support materials in a controlled manner may improve the overall catalytic performance.
While there is an ongoing debate regarding various aspects of the mechanism of VOx-based catalysts for NH3-SCR reaction, it is now widely accepted that the majority route for the catalytic reaction is an Eley-Rideal (E-R) rather than a Langmuir-Hinshelwood (L-H) mechanism [31,32,33], with the overall mechanism including the following four steps: (i) NH3 adsorption on V5+=O (Lewis acid sites) or V5+–OH (Bronsted acid sites), (ii) NO reacting with adsorbed NH3 to form intermediate NH2NO at Lewis acid sites (L acid sites) or NH3NO at Bronsted acid sites (B acid sites), (iii) rearrangement of surface species releasing the intermediate by breaking N-H bonds and coordination bonds between the nitrogenous intermediate and its adsorption site, and by reducing the surface vanadia species from V5+ to V4+ with -OH groups or water being adsorbed, (iv) reoxidation of surface V4+ (V3+ at oxygen vacancies) by O2 accompanied by desorption of surface H2O [34,35,36], Research interests have been focused on exploring the key step(s) of the whole process on a microscopic scale and explaining the temperature-dependent relation between structure and activity. To this end, literature work may be divided into studies on the nature of the acid sites for reactants adsorption, the redox reaction step, and the desorption of products, respectively.
The first aspect, i.e., the sites for reactants adsorption (L vs B acid sites), has been a matter of debate from the beginning. In brief, L acid sites were related to V=O and B acid sites to V–OH species. Earlier, Topsoe et al. had proposed B acid sites to be the catalytically active sites as they detected a dominant amount of NH4+ adsorbed onto V–OH surface sites [34,37,38,39,40], On the other hand, Ramis et al. found NH3 adsorbed on V=O sites to be thermally more stable than NH4+, implying that L acid sites were the dominant adsorption sites under reaction conditions [41], More recently, Marberger et al. supported the idea of Lewis acid sites being the active intermediate by observing a faster consumption of NH3 adsorbed at L acid sites compared to NH4+ adsorbed at B acid sites [42,43], Zhu et al. objected to the standpoint of Lewis acid sites as active sites by concluding that minority L sites (V5+=O) exhibited higher activity (TOF), while the more abundant B sites (V5+–OH) dominated the overall reaction [44], Thus, still no consensus has been reached regarding the role of L and B acid sites.
The redox reaction step has been considered to be related to the reactivity and connected to the VOx surface structures. The use of in situ electron paramagnetic resonance (EPR) and X-ray absorption spectroscopy (XAS) under reaction conditions has revealed the reducing/reoxidizing processes to be accompanied by changes of the vanadium valence and VOx surface structure [45,46,47,48]. Feng et al. reported VOx with polymeric structure to be redox-active and provided an atomic view of the change of the VOx structure with redox state [49,50,51]. Consistently, density functional theory (DFT) results showed that formation of both oxidized and reduced VOx, i.e., V5+ and V4+, kept the vanadyl bond intact for isolated VOx on TiO2, as verified by in situ EPR data for VOx/TiO2 exposed to ammonia [52]. All of these results have revealed that the redox reactivity of VOx species is structure sensitive.
Finally, we turn to the interaction between H2O and surface VOx sites. According to DFT calculations by Avdeev et al., adsorbed water changes the VOx molecular configuration by spontaneous dissociation and formation of surface V–OH groups [53,54,55]. Importantly, lattice oxygen of surface VOx species was shown to originate from adsorbed water rather than gas-phase oxygen [55]. Oxygen-18 isotope labeling studies have revealed that both terminal V=O and bridging V–O–V bonds readily exchange oxygen with water vapor [56].
To date, a detailed understanding of the role of these factors and the VOx surface structure is still missing. In our previous work, we have developed a hierarchical NH3-SCR model catalyst using low surface area silica particles as a platform [12]. Here, we report on a new model catalyst system based on high surface area mesoporous silica SBA-15. In particular, we employ ALD for controlled synthesis of titania and vanadia layers within the pores of the silica matrix and demonstrate its catalytic activity in NH3-SCR. Using UV-Vis and Raman spectroscopy and, in particular, in situ DRIFTS, we explore the nature of the surface sites upon exposure to different gas atmospheres and temperatures, including NH3-SCR reaction conditions.

2. Results

2.1. Catalyst Characterization

Table 1 gives a summary of the prepared samples and their surface area and porosity characteristics. Upon ALD deposition of TiO2 and VOx on mesoporous silica SBA-15, surface area and pore volume drop by roughly half as reflected more precisely in Table 1, and are accompanied by a reduction in pore diameter. The observed decrease in specific surface area is attributed to a coating of the (rough) mesoporous channels but also to closure of SBA-15 micropores, as discussed previously [57]. Increasing VOx deposition leads to a further (but much smaller) overall decrease of the specific surface area and pore volume (see Table 1). As shown in Figure S1, after TiO2 and VOx deposition, the nitrogen adsorption-desorption isotherms show variations of the hysteresis loop compared to bare SBA-15. In particular, the desorption branches exhibit a bulge towards lower p/p° values, indicating the presence of narrowed mesopores [58]. Thus, despite the use of ALD the pores of the silica matrix are not evenly coated, resulting in pore narrowing and possibly pore blocking, consistent with the observed overall decrease in surface area.
Figure 1a depicts wide-angle XRD patterns of TiO2/SBA-15+nxVOx (n = 1, 3, and 5) and TiO2/SBA-15. None of the diffractograms shows any peaks, thus indicating the absence of crystalline titania and vanadia phases and confirming the amorphous nature of the SBA-15 silica matrix. The broad feature at around 23° has been attributed to the walls of mesoporous silica [59]. To check for microcrystalline domains the sample TiO2/SBA-15+3xVOx was examined in more detail using UV Raman spectroscopy (see Figure S2), which was shown to be a sensitive indicator for small (micro)crystalline titania contributions in our previous work [12]. According to Figure S2, there is no indication for the presence of titania microcrystals due to the absence of characteristic anatase (394, 514, and 634 cm−1) and rutile (443 and 610 cm−1) bands [60,61], or V2O5 microcrystals. In the context of Raman detection of V2O5, it should be mentioned, however, that visible Raman spectroscopy is typically more sensitive to V2O5 microcrystals than UV Raman spectroscopy (see discussion below).
Figure 1b shows the corresponding Raman spectra at 532 nm excitation. The spectrum of TiO2/SBA-15 is characterized by broad features typical for amorphous silica dominating the weaker (surface) titania contributions [62]. Major features include the band at around 450–550 cm−1 associated with four-membered rings (D1), and the broad band at around 600–900 cm−1 assigned to symmetrical Si–O–Si stretching and three-membered rings (D2) [63]. The band located at 353 cm−1 is assigned to bending in Si–O–Si bridges [64], whereas the band at 286 cm−1 has been attributed to the A1 mode of TiO6 octahedra [65,66]. The small feature at 1064 cm−1 is characteristic of silica TO phonons [67], whereas the weak Raman peak observed at 1122 cm−1 for TiO2/SBA-15 is attributed to framework Ti–O–Si species [68]. Importantly, the addition of increasing amounts of vanadia leads to the appearance of new Raman signals at around 267, 516, 699, and 1020 cm−1. At first sight, the signals at around 267, 516, and 699 cm−1 show some similarity with those observed for crystalline V2O5, however, closer inspection reveals significant differences. The observed signals rather indicate the presence of hydrated vanadia forming xerogels V2O5·nH2O, resembling those discussed previously in the context of silica-supported vanadia, and the peak at 1020 cm−1 is attributed to the V=O stretching vibration mode of tetrahedral VOx species [69]. Based on the above results from XRD and Raman spectroscopic characterization, we can conclude that the ALD-prepared samples contain amorphous titania and vanadia on silica SBA-15 and that titania at the loading studied here (2 Ti/nm−2) does not form a conformal layer on the silica surface allowing vanadia to interact with titania and silica. According to a previous ALD study [29], deposition of 3 Ti nm−2 and 7 Ti nm−2 onto SiO2 leads to the formation of sub-monolayer titania, in agreement with the results obtained here.
In the following, we will focus on the structural and catalytic properties of sample TiO2/SBA-15+3xVOx with a vanadium loading density of 1.6 V/nm2. Table S1 and Figure 2 summarize the results of the XPS analysis. As shown in Table S1, Si, O, Ti, and V are detected at the catalyst surface, besides C. The Ti/Si and V/Si ratios correspond to 0.10 and 0.12, respectively, suggesting the presence of similar amounts of Ti and V on the surface. The results are consistent with the above picture of a mixed layer and rule out the exclusive formation of large titania or vanadia (3D) aggregates, which would strongly reduce the visibility of Ti and V in XPS. The left of Figure 2 depicts the O 1s, V 2p1/2, and V 2p3/2 photoemission located at 532.9 eV, 524.9 eV, and 517.4 eV, respectively. As expected, the former is dominated by the O 1s contribution from SiO2, by comparison with the literature [69]. Detailed analysis of the V 2p3/2 region based on literature data for binary vanadia compounds (V2O5, V2O4, V2O3) reveals the presence of V5+ at 517.6 eV (60%) and V4+ at 516.3 eV (40%), while a small contribution from V3+ at 515.7 eV cannot be ruled out [70,71]. The Ti 2p3/2 peak is composed of (at least) two contributions located at 458.7 and 459.9 eV, which can be attributed to TiO2 domains and to Ti–O–Si bonds, respectively [5,72].
Figure 3a depicts diffuse reflectance (DR) UV-Vis spectra of TiO2/SBA-15+3xVOx in comparison to TiO2/SBA-15. To clarify the influence of water from ambient on the catalyst structure, spectra were recorded under ambient (hydrated) and dehydrated conditions. Please note that bare silica SBA-15 does not show any significant UV-Vis absorption. The UV-Vis spectrum of Ti/SBA-15 is characterized by strong absorption at around 280 nm, which has been attributed to charge transfer between an oxygen ligand and the central Ti4+ ion [73,74], whereas the position of the absorption band is consistent with higher coordinated or oligomeric titania species [75]. The presence of oligomeric species would be in agreement with the literature on SiO2-supported titania, reporting two-dimensional oligomers of TiO5 domain at a loading density of about 4 Ti nm−2, which is higher than the density used in this work (2 Ti nm−2) [76].
Regarding the vanadia structure, the observed UV-Vis absorption was analyzed based on the ligand-to-metal charge transfer (LMCT) transitions observed for bulk vanadia reference compounds [77]. To this end, bands at around 240 and 290 nm were reported to originate from isolated tetrahedrally coordinated mono-vanadate ions [78], while the bands at about 270, 340, and 412 nm were assigned to poly-vanadate ions [79]. In the literature, the bands at 308, 371, and 406 nm were assigned to LMCT transitions of monomeric tetrahedral VO43−, oligomeric tetrahedral VO3, and polymeric distorted tetrahedral VO3 ions, respectively [80]. It was also reported that the bands at around 440–510 nm were due to LMCT transitions of V5+ species with a square pyramidal structure and those at 545–650 nm were due to d-d transitions of V4+ species with a square pyramidal structure. For crystalline V2O5, a broad absorption band with maxima ranging from 440 to 490 nm was observed [77]. The presence of V3+ and/or V4+ species was shown to lead to d-d transitions ranging from 600 to 800 nm [81]. A comparison of the hydrated and dehydrated state of the sample reveals the presence of a broad band at around 380–450 nm, which has been attributed to the coordination of water to V sites leading to major structural changes [74].
After dehydration, the UV-Vis absorption behavior is characterized by contributions at around 280, 320 and 400 nm. Based on the above literature results, these are attributed to monomeric/dimeric, oligomeric, and polymeric VOx species, respectively. The intensity increase at around 260 nm is assigned to Ti–O–Ti from Ti–OH or possibly V–O–Ti from V–OH. Because of the overlap of charge transfer bands of titania and vanadia [80,82], removal of the titania contribution facilitates the analysis of the surface vanadia structure (see Figure S3) [83]. As a result, Figure S3 shows the presence of an asymmetric broad absorption band located between 250–550 nm for the hydrated and between 250–450 nm for the dehydrated sample, together with the result of a fit analysis based on the above reference data [83,84]. Besides, the ratio of monomeric/dimeric/oligomeric VOx with tetrahedral coordination to polymeric VOx with (pseudo-)octahedral coordination was estimated to increase from 0.31 to 12.5 upon dehydration, underlining the structural changes associated with the removal of coordinated water.
Based on the UV-Vis spectra the bandgap energy Eg was determined using the general power-law expression suggested by Davis and Mott based on the absorbance α [85].
The value of the parameter n in Equation (1) was set to 2 following the literature [82,86], showing a linear relationship when plotting (αhν)1/2 vs (Tauc’s method) [87]. Extrapolation of the linear region to (αhν)1/2 = 0 yields Eg values of 2.99 and 3.30 eV for the hydrated and dehydrated state of the catalyst, respectively. As the Eg value is negatively correlated to the number of covalent V–O–V bonds (CVB) via the relation CVB = 14.03 − 3.95 Eg [82], an increase in Eg implies a decrease in the degree of VOx polymerization upon dehydration:
α h ν ( h ν E g ) n
Figure 3b depicts Raman spectra of TiO2/SBA-15+3xVOx under hydrated and dehydrated conditions in comparison to the spectrum of TiO2/SBA-15. As discussed above VOx-related features are observed at 268, 513, 682, and 1025 cm−1. Upon dehydration, the vanadyl (V=O) stretching mode shows an intensity increase and blueshift from ~1025 to 1038 cm−1, consistent with previous work on VOx/SiO2 and VOx/(TiO2 + SiO2) [29,69]. The position of the V=O band at 1038 cm−1 in the visible Raman spectrum is characteristic for tetrahedrally coordinated vanadium ions, and has been associated with monomeric and/or small oligomeric VOx species, consistent with the results from UV-Vis spectroscopy (see above). The presence of (micro)crystalline V2O5 in the dehydrated state of the catalyst can be excluded due to the absence of the characteristic vanadyl feature of V2O5 at 994 cm−1 [69]. The shoulder observed at around 1062 cm−1 has been attributed to V–O–Si bridging bonds [88]. The Raman bands observed at 268, 513, and 682 cm−1 for the hydrated state are consistent with the presence of hydrated vanadia forming xerogels V2O5·nH2O, as discussed above. Upon dehydration, these bands disappear due to the structural transformation of surface vanadia, leading to an increased visibility of the support-related bands at 258, 481, and 660 cm−1 discussed below.

2.2. Catalytic Performance

The NH3-SCR activity was examined within 100–450 °C using a mixture of 500 ppm NH3, 500 ppm NO, 5% O2, and N2 (balance). The temperature was increased from 100 °C to 450 °C in 50 °C steps keeping each temperature for 20 min for equilibration. Analysis of the gas phase at the reactor outlet was performed by quantitative FT-IR spectroscopy. Figure 4 depicts the catalytic performance for NOx conversion over TiO2/SBA-15+3xVOx as a function of temperature. Table 2 summarizes the NOx conversion and N2 selectivity values. The NOx conversion follows a volcano shape with an increase from 100 to 250 °C, a maximum conversion level at 250–350 °C, and a decrease above 350 °C. The large error bar in the low temperature range is related to influences of the adsorption equilibrium process on the reaction step. At high temperatures a decrease of both NOx conversion and N2 selectivity is observed, which may be attributed to increased NH3 oxidation leading to N2O generation.
Despite the similar shape of the conversion, the TiO2/SBA-15+3xVOx sample shows significantly higher catalytic activity as VOx/(SiO2 + TiO2) samples prepared by sol-gel methods [5]. In this context, it was reported that doping TiO2 with Si may result in an increase in NOx conversion and a decrease of the lower temperature limit from 360 °C for pure TiO2 to 300 °C or lower for Si-doped samples [72]. Interestingly, the catalyst exhibits an excellent N2 selectivity with values higher than 90% even at 450 °C. This behavior may be attributed to the (micro)porous structure and large surface area of the SBA-15 support, inhibiting the growth of bulk-like V2O5 species, which were suggested as main cause for N2O formation [8]. In contrast to the results shown in Figure 4, VOx/SiO2 samples (prepared by ALD) showed only low NO conversions. In fact, the presence of titania was reported to enable V–O–Ti bond formation, which was related to the number and strength of surface acid-base sites [29]. Thus, we can conclude that besides the influence of (micro)structural effects, the improved catalytic performance may be attributed to the controlled preparation of dispersed active sites and a submonolayer of TiO2 on SiO2.

2.3. Operando Characterization of TiO2/SBA-15+3xVOx

To gain insight into the NH3-SCR mechanism we applied DRIFTS in the presence of different gas environments and at different temperatures including reaction conditions (operando). Figure 5 shows DRIFT spectra of TiO2/SBA-15+3xVOx exposed to a gas mixture of 500 ppm NH3, 500 ppm NO, and 5% O2 (balanced with N2) at different temperatures from 100 °C to 450 °C at a flow rate of 50 NmL/min. In Figure 5a, the black curve (labeled ‘N2, 300 °C’) corresponds to the catalyst in a nitrogen flow at 300 °C before exposure to the reaction gas mixture. Features located at around 919 and 1036 cm−1 and at around 1849 and 2002 cm−1 have previously been assigned to V–O–V and V=O stretching modes of surface VOx species and their first overtones, respectively [2,50,89,90]. The feature at 1203 cm−1 has been ascribed to L acid sites on TiO2 [91]. The shoulder at around 1640 cm−1 and the broad band within the range 3500–3800 cm−1 can be attributed to OH bending and stretching of water, respectively [2,72]. In the OH stretching region, the sharp peak at 3745 cm−1 is characteristic of isolated Si–OH groups, while the small feature at 3656 cm−1 can be associated with V–OH [74]. The Si–OH feature shows an asymmetric broad tail at around 3650 cm−1, which has been assigned to Ti–OH [92].
Upon exposure to the reaction gas mixture, two new features appeared at 1418 and 1455 cm−1, which can be attributed to the asymmetric N–H bending vibration of NH4+ linked to B acid sites, while the shoulder and small feature at 1176 and 1606 cm−1 have been associated with the symmetric and symmetric N–H bending vibrations of NH3 linked to L acid sites, respectively [93,94]. According to Figure 5a, the observed N–H modes show a different temperature-dependent behavior. Whereas the feature at 1606 cm−1 decreased with temperature and disappeared at about 250 °C, the band at about 1418 cm−1 kept intact.
In Figure 5b, upon exposure to the reaction gas mixture, new IR bands are observed at 2803 and 3050 cm−1, which can be attributed to N–H stretching of NH4+ linked to B acid sites, while the bands located at 3159, 3278, and 3370 cm−1 are assigned to N–H stretching modes of NH3 coordinated to L acid sites [41,94,95,96]. As the temperature increases, the bands at 3278 and 3370 cm−1 show an intensity decrease and completely disappear at temperatures above 300 °C, while the feature at 3159 cm−1 is detectable until 450 °C. Similarly, the bands at 2803 and 3050 cm−1 related to NH4+ on B acid sites can be observed within the whole temperature range.
Figure 6 depicts DRIFT spectra of TiO2/SBA-15+3xVOx exposed to different gas environments at 300 °C at a total flow rate of 50 NmL/min within (a) 2600–3500 cm−1 and (b) 3500–4000 cm−1. Following the assignments discussed in the context of Figure 5, the features at 2808 and 3046 cm−1 are assigned to NH4+ coordinated to B acid sites, and those at 3150, 3275, and 3356 cm−1 to NH3 bound to L acid sites (see Figure 6a).
Similarly to the effect of temperature, the bands linked to B acid sites (2808 and 3046 cm−1) and the band linked to L acid sites at 3150 cm−1 are observed independent of the composition of the gas atmosphere, while those at 3275 and 3356 cm−1 can only be detected upon exposure to gas mixtures containing NH3. That is to say, that NH4+ on V–OH sites is stable towards reaction no matter which redox conditions, while NH3 adsorbed on specific L acid sites was proved to be reactive. In the corresponding high-wavenumber region (see Figure 6b) IR bands are observed at 3655 and 3737 cm−1 in all gas mixtures and are attributed to V–OH and Si–OH stretching vibrations, respectively, as discussed above, while the broad absorption band extending to ~3600 cm−1 originates from Ti–OH (see above) and the background from the fine structure of adsorbed water.
Upon closer inspection, the IR spectra in Figure 6 show gas phase-dependent variations. Upon the first exposure to 20 vol% O2, no acid site-related bands are observed. However, water-related background peaks are detected indicating the rearrangement of surface hydroxyl groups into water under oxidizing conditions [55]. Switching to the reductive atmosphere, i.e., NH3/NO, results in the appearance of the five N–H stretching modes related to L and B acid sites (see above), while during the following O2 treatment the L site related features at 3275 and 3356 cm−1 disappear. On the other hand, the 3150 cm−1 band assigned to NH3 coordinated to V=O sites, is still detected under oxidizing conditions. In the high-wavenumber region, the Si–OH, Ti–OH, and V–OH signals first show a decrease and then a partial recovery in oxygen. Next, the catalyst was exposed separately to NH3 and NO, and then to a NO/O2 mixture oxidation. In the presence of NH3 reduction, the L site related features completely reappear and the water-related peaks show a weak positive increase indicating that NH3 itself could be reduced to NHn thereby transferring hydrogen atoms to adjacent hydroxyl groups to form water. Upon oxidation by NO or NO + O2 the features at 3275 and 3356 cm−1 disappeared, while the water-related peaks significantly increased, suggesting the promotional effect of NO regarding NH3 oxidation, leading to intermediate formation (e.g., NH2–NO or NH–NO), consistent with results from DFT [97], and mass spectroscopy [98].
Finally, the catalyst was exposed to reaction conditions, i.e., an NH3/NO/O2 mixture, leading to the re-appearance of all five acid site-related features, while the peaks for Si–OH, Ti–OH, and V–OH decreased in intensity. It is worth mentioning that there is a distinct difference in the IR spectra for exposure to the NH3/NO mixture and the subsequent exposure to NH3 and NO. In the former case, the presence of NH3 was sufficient to competitively adsorb on the surface resulting in H2O desorption, while in the latter case, NH3 initially adsorbed on the surface was oxidized by NO resulting in NH2–NO/N–NO and water formation with reduced competitive adsorption [98]. These above results are consistent with the strength of adsorption following the order NH3 > H2O > NH2–NO intermediate, in agreement with the literature [99,100]. Interestingly, the addition of oxygen (NH3/NO/O2) leads to a different, distinct state of the catalyst regarding the presence of surface species, which falls between the behavior observed for NH3 exposure and subsequent exposure to NH3 and NO, and is also different to that obtained by subsequent exposure to NH3/NO and oxygen.
In summary, the DRIFTS results shown in Figure 5 and Figure 6 reveal important differences in the adsorption/reactivity behavior of NH3 species attached to acid sites. The N–H stretching linked to B acid sites at 2803 and 3050 cm−1, which have been assigned to coordinate and suspended N–H bonds [41], respectively, did not respond to the different atmospheres and different temperatures, suggesting that the B acid sites were active for adsorption but not for reaction. On the other hand, the L acid site related N-H stretching features at 3275 and 3356 cm−1, assigned to -NH2 and -NH species, showed a more distinct response to the nature of the gas environment and the temperature, while the 3150 cm−1 feature, assigned to adsorbed NH3 species, remained passive towards reaction. Regarding the reaction mechanism, we draw the following conclusions: (a) NO reacts with NH3 connected to the inert site (3150 cm−1). (b) NH3 adsorbs to active L acid site forming NH2 by H-abstraction (3275 and 3356 cm−1), followed by its interaction with NO leading to NH2NO intermediate formation and further reaction to N2 and H2O. (c) In case of N2O formation, NH2NO may be expected to lead to NHNO intermediates and finally N–NO.
Figure 7 depicts the results of the in situ detection of the exhaust gas during different temperature SCR experiments from 100 to 450 °C. Each temperature was kept for 20 min to reach an equilibrium state. When increasing the temperature, desorption peaks were observed indicating changes in the surface concentration as a function of temperature. At lower temperatures, the presence of NH3 peaks but absence of NO desorption peaks (not shown) suggests a surface reaction following the E-R rather than the L-H route, in agreement with the literature [32,33]. In addition, insight into the reaction mechanism can be obtained. To this end, three steps have been distinguished, i.e., adsorption of reactants (mainly NH3), reaction, and desorption of products (such as H2O, NO2, and N2O). According to Figure 7, there are low temperature desorption peaks of NH3 at 25/100 °C and 100/150 °C, which are attributed to NH3 desorption from the B acid site, while high temperature NH3 peaks above 150 °C are assigned to desorption from L acid site [5,101,102]. These peaks disappeared at about 300/350 °C, consistent with the disappearance of L acid sites in DRIFTS (see Figure 5). The strong water desorption peak at 25/100 °C originates from condensation of OH groups, e.g., after desorption of NH3 bound to B acid sites [5]. The first desorption peak of NO2 at 150/200 °C can be attributed to the distortion and release of NO2 from nitrite/nitrate species by NO molecules connected to surface oxygen sites [103,104]. The second peak at 200/250 °C shows that NO oxidation is faster than NO2 desorption at 200 °C. Similarly, the first desorption peak for N2O detected at 200/250 °C is indicative of its formation on active sites and subsequent chemisorption at temperatures <250 °C [105].
All in all, catalyst characterization of TiO2/SBA-15+3xVOx reveals a relatively high surface area (>350 m2/g) and the presence of surface hybrid V–O–M bonds connecting the vanadia aggregates to the substrate. The NH3-SCR catalytic performance follows a volcano shape with temperature, with an optimal deNOx value of about 60% at 300 °C. Finally, the operando DRIFTS results reveal N–H stretching features related to B and L acid sites.

3. Discussion

3.1. Surface Vanadia Structure

Previously, highly dispersed vanadia and titania on SBA-15 have been studied in detail under hydrated and dehydrated conditions [74]. Despite the lower loading ranges discussed (V: 0.00001–0.7 V/nm2; Ti: 0.001–0.7 Ti/nm2) valuable structural information for the present sample (1.6 V/nm2; 2 Ti/nm2) was obtained regarding the nature of the surface species, in particular, as all samples are well below monolayer coverage [26,106,107]. This is supported by the absence of Raman features due to crystalline V2O5 and TiO2, implying the presence of amorphous domains, consisting of dispersed titania and vanadia species on the silica surface. On the other hand, the detailed surface composition after successive deposition of titania and vanadia needs closer examination, especially as it was reported that vanadium is preferentially deposited on titanium sites [26,108].
For TiO2 deposited on SBA-15, the Ti loading was quantified as 2 Ti/nm2, which is significantly below monolayer coverage (6–8 Ti/nm2) [106]. DR UV-Vis spectra show an absorbance band at about 260 nm, attributed to titanium in 5-fold coordination (see Figure 3) [76]. The UV-Vis edge energies were determined to be 3.2 and 3.0 for the hydrated and dehydrated state, respectively. Comparison with edge energies of titania structures ranging from isolated tetrahedrally coordinated TiO4 (4.3 eV) to oligomeric TiO5 species (3.4 eV), suggests the presence of oligomeric TiO5 species in the VOx/TiO2–SiO2 sample. Regarding the effect of water on the titania structure, no difference was found between the dehydrated and hydrated state at extremely low Ti loading (<0.05 Ti/nm2), in contrast to higher loadings (> 0.1 Ti/nm2) [74]. In fact, adsorbed water was reported to interact with adjacent titanium atoms (Ti–O–Ti) in oligomeric titania, leading to an edge energy 0.2 eV higher than in the dehydrated state, consistent with the behavior observed here [74]. High temperature calcination is expected to lead to an increase in Ti–O–Ti linkages as part of domain TiO2 [26] without formation of crystalline TiO2, as discussed above (see above). While anatase and rutile related Raman bands are absent, the feature at 286 cm−1 is attributed to TiO6 octahedra as part of amorphous TiO2 domains on the silica surface.
After vanadia deposition, adsorption-desorption isotherms are characterized by a steep hysteresis, attributed to cavitation-controlled evaporation in a narrow range of the pore neck, while some pore blockage or seepage may not be excluded. The surface area shows only a small decrease with increasing VOx deposition, while no significant differences in the NLDFT pore-size distribution were observed. This behavior implies that the ALD deposited VOx did not block the pore openings, but was rather selectively deposited inside the matrix on TiO2 domain sites, forming ink-bottle pore [109].
Although the ALD method permitted the vanadium precursor to react with all surface adsorption sites, including Ti–OH and Si–OH hydroxyls, preferential deposition seems to be inevitable [29]. According to Table 1, the sample with 1xVOx deposition shows a lower V:Ti ratio of 0.24, at which vanadium is expected to be mainly loaded on TiO2 domains [26]. For the 3xVOx and 5xVOx samples close to unit V:Ti ratios of 0.85 and 1.08 are observed, which can be interpreted as sub- and over-saturated coverage of VOx on TiO2, respectively. It should be mentioned that the VOx deposition rate showed differences among the three samples. In fact, based on the growth rate for 1xVOx, leading to a V/Ti ratio of 0.24, the extrapolated V/Ti ratio for 3xVOx should be 0.72, which is slightly smaller than the actual value of 0.85, while the extrapolated V/Ti ratio for 5xVOx should be 1.20, a value slightly larger than the observed value of 1.08. The increase in VOx growth rate is attributed to a change in deposition site from titania to silica when approaching TiO2 saturation coverage. On the other, the mobility of surface VOx species was reported to reach its lowest surface free energy (SFE) at elevated temperature under reaction conditions, resulting in a two-dimensional spreading of VOx species on silica and titania [110,111]. To this end, it is worth mentioning that the IR spectra recorded at elevated temperature under reaction conditions reveal a hydroxyl signal at 3740 cm−1 due to surface Si–OH groups (see Figure 5), indicating that the deposition of TiO2 and VOx did not lead to complete coverage of the silica surface.
The UV-Vis DRS and Raman spectra provide structural information on the VOx surface species (see Figs. 3 and S3). It is well known that the structure of surface VOx depends sensitively on the gas environment [112]. UV-Vis spectra of the hydrated state are characterized by a 250–550 nm band, consistent with polymeric VOx with pseudo-octahedral coordination, and by a band at 250–450 nm for the dehydrated catalyst, reflecting the presence of tetrahedrally coordinated oligomeric VOx species, overlapping with the signal of the TiO2/SiO2 substrate in the dehydrated state. The presence of H2O can hydrate the surface VOx species leading to the formation of V2O5·nH2O-like gels, as evidenced by Raman features at 267, 516, 699, and ~1020 cm−1 under ambient conditions (see Figure 1 and Figure 3). Regarding the position of the V=O stretching bands (~1020 and 1038 cm−1) also the effect of moisture needs to be taken into account [69]. Upon dehydration, a 13 cm−1 blueshift of the V=O feature is observed (see Figure 3), consistent with the literature [69], leading to the formation of tetrahedrally coordinated vanadium ions. In the dehydrated state, the position of the V=O band at 1038 cm−1 is located between the values 1031 cm−1 and 1042 cm−1 reported for bare TiO2 and SiO2, respectively [29,113,114], suggesting the simultaneous existence of vanadia species on both oxides attached via V–O–Ti and V–O–Si bands.
Summarizing, regarding the structure of the VOx surface species we propose based on our findings that (i) titania is dispersed on the silica surface in domains containing low oligomerized TiO5, (ii) VOx is preferentially deposited on titania sites at low VOx loadings (1xVOx), inducing the formation of dispersed oligomeric species, (iii) increasing VOx loading results in the deposition of V on both Ti and Si sites (3xVOx, 5xVOx).

3.2. Rate-Determining Step

In agreement with the literature, the observed temperature-dependent efficiency for NO removal follows a volcano shape, with a maximum NO conversion at 250–350 °C. N2 selectivities are near 100% within 100–300 °C and decrease to 90% as temperatures rise to 450 °C (see Figure 4).
The reason for the reactivity changing with temperature was related to the different processes determining the rate. Generally, the SCR reaction process on the VOx catalyst could be divided into three steps, i.e., adsorption of reactants, redox reaction, and desorption of products. On the one hand, it was stated that the redox reaction was rate-determining at low temperature, while adsorption of NH3 on surface acid sites was the rate-determining step (RDS) at high temperature [5,115]. On the other hand, desorption of generated H2O was considered the low-temperature RDS, and the reductive reaction of NH3 the high-temperature RDS [116]. To explore possible structural factors related to catalytic reactivity, operando DRIFT was applied.
Based on the DRIFT spectra in Figure 5 and Figure 6 part of the L acid sites are proposed as active sites, due to their gas phase and temperature dependent behavior. These sites can be associated with the N-H stretching of coordinated amide, while the stable L acid sites can be related to the N–H vibration of coordinated NH3. Thus, the detailed H-abstraction ability is important even for the same type of adsorption sites. According to the literature, dispersed oligomeric VOx species are believed to be the main active species, providing oxygen (V=O or V–O–V) [7,56,117,118], for H-abstraction of adsorbed NH3 to form an amide intermediate [98]. To this end, the samples TiO2/SBA-15-3xVOx and TiO2/SBA-15-5xVOx with a larger fraction of oligomeric VOx show a better NOx conversion than TiO2/SBA-15-1xVOx (not shown). Also, the changes of the acid site-related signals with increasing temperature demonstrate that at temperatures < 300 °C the H-abstraction step is slower than the adsorption of NH3 (see Figure 5). The presence of adsorbed NH3 at low temperatures was evidenced by NH3 desorption peaks. On the other hand, desorption of water was observed at very low temperatures, indicating that its desorption was slower than its formation by reaction. Thus, we propose the RDS at low temperatures to be related to the desorption of products, especially water. To this end, additional water vapor was reported to suppress the reactivity irreversibly, indicative of the occupation of active sites by water [72]. Our conclusions are also supported by the fact that N2O formed at temperatures below 200 °C was adsorbed on the surface rather than being released into the gas phase.
Except for the redox properties surrounding the acid site, the acid strength is another important factor concerning the SCR reaction, relevant for both the adsorption of reactants (mainly NH3 in E-R route) and the desorption of products (such as H2O, NO2, and N2). The acidity of binary oxides and ternary oxides VOx/TiO2/SiO2 has been the subject of previous studies, both on solid solutions and supported systems [5,29,30,108]. For binary systems, the surface acidity was found to increase as VOx/SiO2 < TiO2/SiO2 < VOx/TiO2 [30]. Regarding VOx/TiO2/SiO2 systems prepared by ALD [29], samples with lower VOx loading (1 V/nm2, 3 Ti/nm2) and higher VOx loading (2 V/nm2, 3 Ti/nm2 ) were compared towards their acidity behavior, showing a decrease (increase) in weak/medium (strong) strength acid sites with loading. Furthermore, strong acid sites were assigned to VOx species on titania and weak acid sites to VOx species on silica [29]. In our study, NH3 desorption peaks for TiO2/SBA-15+1xVOx were detected at higher temperature as compared to 3xVOx and 5xVOx (see Figure S4), consistent with the formation of stronger acid sites at lower V loading on titania. The N2O desorption signal also follows the trend of the VOx acid sites as 1xVOx > 3xVOx > 5xVOx. For the 1xVOx sample, even at 300/350 °C, a weak N2O desorption peak was detected. The 3xVOx and 5xVOx samples showed a very similar desorption behavior of reactants and products. At temperatures > 300 °C, no more desorption of NH3 but an increase in the N2O concentration was detected, reflecting an increase in the H-abstraction ability. Thus the RDS at high temperatures may be related to the adsorption of NH3.

3.3. Structure-Activity Relationship

In the sections above, we discussed the structure of VOx loaded on TiO2/SBA-15 and several aspects affecting the NH3-SCR activity. Compared to previous VOx-based model systems with similar V loading on SiO2, TiO2, or SiO2 + TiO2 prepared by other techniques, our VOx/TiO2/SBA-15 catalyst shows an improved NOx conversion and operation window [5,119,120,121]. For NH3-SCR over supported VOx catalysts, high reactivity has been related to crystalline (anatase) TiO2 [5,122,123]. Mixing SiO2 with TiO2, to be used as support for vanadia, introduces extra acid sites, and is accompanied by the transformation of crystalline into amorphous titania, resulting in a decrease in the catalytic performance compared to bare titania [5,124]. However, by adjusting the Si/Ti ratio, a volcano-shape of the SCR reactivity with Si content was obtained [124], indicating synergetic effects regarding the reactivity, possibly related to the presence of Ti–O–Si surface sites. Kobayashi et al. studied the acidity of TiO2–SiO2 mixed systems, also showing a volcano-shaped relationship as a function of composition, with a maximum at a Si/Ti ratio of 1 [124]. On the other hand, according to other studies, using the same sol-gel method, the best SCR activity was observed for lower Si/Ti ratios such as 2/8 [72,124]. The mismatch between the acidity and reactivity may be explained by the RDS discussion above, as within the low and medium temperature range, the desorption step was rate determining for the strong acid sites. In present work, ALD deposition of TiO2 onto SiO2 results in the formation of oligomeric TiO2 domains, with a limited boundary between TiO2 and SiO2, in contrast to the more uniform mixture of Ti and Si obtained by the sol-gel method. The improved catalytic performance of the TiO2/SBA-15+3xVOx compared to other SiO2/TiO2-based VOx catalysts can therefore be explained by the comparable amounts of V and Ti, allowing V to be deposited on TiO2 and boundary sites (Ti–O–Si). This leads to the formation of VOx species supported by hybrid sites (V–O–Ti, V–O–Si), thus increasing the total number of acid sites. Consistent with this scenario, for lower V loading (1xVOx), significantly lower NOx conversion is observed (not shown), due to the preferential deposition of V on (amorphous) TiO2, while for higher V loading (5xVOx), a similar behavior as for the 3xVOx catalyst was obtained, due accessibility of both TiO2 and Ti–O–Si boundary sites.
Meanwhile, it is well-known that the SCR reactivity of VOx/TiO2 catalysts shows a narrow temperature window within the high temperature range, strongly decreasing below 300 °C [125], VOx/SiO2 catalysts are characterized by low SCR reactivity but a wider temperature adaptability even at low temperatures [126]. Besides, according to the literature, produced H2O was adsorbed on surface active sites because the activation barriers of the NH2NO decomposition on the catalyst surface were much lower than those calculated for the gas-phase reaction [127]. Thus, in the low temperature range, the desorption of water and other product molecules was considered rate-determining. Keranen et al. reported that VOx supported on pure TiO2 led to a larger number of strong acid sites but less weak/medium ones than on a mixture of TiO2 and SiO2 [29], implying strong adsorption of NH3 and H2O. In fact, Figure 7 reveals the appearance of desorption peaks of water during heating steps. Thus, a limited mixture of TiO2 and SiO2 decreases the surface acid strength, facilitates desorption of H2O and other products and is therefore expected to improve the SCR activity at low temperatures.
Another aspect that is proposed to influence the low-temperature SCR performance is the preparation of the TiO2 domains on the SiO2 support [26]. In contrast to other methods used for the preparation of mixed SiO2 + TiO2 supports, such as sol-gel based synthesis [5] and co-precipitation [124], the ALD approach enables controlled deposition of sub-monolayer titania and vanadia, resulting in micro-domain areas of TiO2 and well dispersed VOx species. The deposition of VOx on small domains of an oxide less reducible than vanadia but more reducible than the support is expected to improve the dispersion of surface VOx species [128]. Furthermore, the high surface area of the mesoporous matrix can be largely preserved by the low-temperature ALD approach and allows to further increase the VOx loading for higher SCR catalytic performance.

4. Experimental

4.1. Chemicals

The chemicals used for the preparation of SBA-15, including triblock copolymer Pluronic P123 (molecular weight = 5800, EO20PO70EO20) and tetraethylorthosilicate (>99%, TEOS), were purchased from Sigma-Aldrich Chemie GmbH (Munich, Germany). As ALD precursors for TiO2 and VOx, titanium tetrachloride (>99%, TiCl4) and vanadium oxytrichloride (>99%, VOCl3) were employed, respectively, which were purchased from Sigma-Aldrich Chemie GmbH (Munich, Germany). As oxygen source in the ALD process demineralized water was used.

4.2. Synthesis

For the preparation of SBA-15 [129], 4.0 g P123 was completely dispersed in a mixture of 120 mL HCl solution (2 M) and 30 mL water in a PP bottle at 35 °C with 250 r/min stirring for 2 h. Then 10 mL TEOS was added at 35 °C with 400 r/min stirring for 20 h. After that, the bottle was put in an oven at 85 °C with hydrothermal crystallization treatment for 24 h. Using vacuum filtration, the as-made SBA-15 was collected, followed by calcination to remove the template P123 in a muffle furnace heated to 550 °C for 12 h (heating rate: 1.5 °C/min).
VOx/TiO2/SBA-15 samples were prepared by ALD using a deposition system described in detail elsewhere [12]. Briefly, three ALD cycles of titania were deposited onto silica SBA-15. Finally, VOx was deposited with 1, 3, and 5 cycles leading to different V loadings. For the ALD of TiO2, the TiCl4 precursor was evaporated in a feeding bottle at 40 °C and reacted in the reaction chamber at 120 °C for 60 s. Similarly, for VOx ALD, the VOCl3 precursor was evaporated at 40 °C and reacted at 60 °C for 60 s. In both processes, water at room temperature acted as oxidant. Between reaction steps the system was purged (N2, 60 s).
The sample with 3 cycles of TiO2 on SBA-15 is labeled as ‘TiO2/SBA-15′, and samples with different cycles of VOx on TiO2/SBA-15 as ‘TiO2/SBA-15+nxVOx‘ (n = 1, 3, 5). Hydrated samples correspond to samples exposed to ambient conditions for at least one week after synthesis. Samples labeled as ‘dehydrated’ were heated at 500–600 °C for 1 h under a controlled atmosphere (20% O2/80% N2) and then cooled to room temperature.

4.3. Catalytic Performance

The SCR performance was measured in a commercial reactor (CCR1000, Linkam Scientific Instruments, Tadworth, UK) in combination with quantitative gas-phase FT-IR spectroscopy, employing a Tensor 27 instrument equipped with a deuterated and L-alanine-doped triglycine sulfate (DLaTGS) detector (Bruker, Mannheim, Germany). Using mass-flow controllers, the inlet gas concentration was adjusted to 500 ppm NO, 500 ppm NH3, and 5% O2 (balanced with N2), by mixing 2000 ppm NO/N2 (±0.25% abs.), 2000 ppm NH3/N2 (±0.25% abs.), O2 (≥99.999%), and N2 (≥99.999%). During the NH3-SCR process, the contents of NO, NH3, and O2 in the feed were set to 500 ppm, 500 ppm, and 5%, respectively, balanced with N2. The weight of the catalyst sample in the reactor bed was 15 mg and the gas mixture passed through the catalyst bed at a flow rate of 50 NmL·min−1, resulting in a gas hourly space velocity (GHSV) of 40,000 h−1. The temperature was increased from 100 °C to 450 °C at 50 °C steps. Each temperature was kept for 20 min to allow for equilibration. The outlet gas mixture was analyzed by an FT-IR spectrometer equipped with a low volume gas cell (25 mL, LFT, Axiom Analytical, Inc, Tustin, CA, USA). The gas cell was heated to 120 °C to avoid the condensation of water produced by the catalytic reaction. IR spectra were continuously recorded with a resolution of 4 cm−1. For quantitative analysis of the IR active gas-phase components (NO, NO2, NH3, N2O) calibration curves were recorded.
By combining the outlet concentrations of NO, NO2, N2O, and NH3 with the inlet concentrations, the N2 concentration was calculated based on nitrogen mass conservation, assuming that nitrogen was present only as gas-phase molecular species. The NOx conversion and N2 selectivity were calculated using the following Equations (2) and (3):
NO x   conversion   =   [ NO x ] in [ NO x ] out [ NO x ] in 100 %
N 2   selectivity   =   2 [ N 2 ] out [ NO x ] in [ NO x ] out + [ NH 3 ] in [ NH 3 ] out 100 %
where the subscripts ‘in’ and ‘out’ represent the inlet and outlet flow, respectively.

4.4. Characterization

N2 adsorption-desorption measurements were carried out on a NOVA 3000e system (Quantachrome, Boynton Beach, FL, USA) to determine the specific surface area using the Brunauer-Emmett-Teller (BET) method and the porosity characteristics using non-local density functional theory (NLDFT). Quantification of the vanadium content of the samples was done by inductively coupled plasma optical emission spectroscopy (ICP-OES).
XRD experiments were carried out in transmission geometry on an X-ray powder diffractometer (StadiP, Stoe & Cie GmbH, Darmstadt, Germany) with a Mythen 1K (Dectris, Baden-Daettwil, Switzerland) detector. For the measurements, a Cu Kα1 radiation (λ = 1.540598 Å) and a Ge [110] monochromator was employed.
XP spectra were recorded on an SSX 100 ECSA spectrometer (Surface Science Laboratories Inc., Minneapolis, MN, USA), equipped with a monochromatic Al-Kα X-ray source (1486.6 eV), in a constant analyzer energy (CAE) mode at a 36° detection angle with 0.1 eV resolution. Data analysis included subtraction of a Shirley background and a peak-fit analysis using Gaussian−Lorentzian product functions with 45% Lorentzian share. Atomic concentrations were calculated using the following relative sensitivity factors (RSFs): 0.537 (C 1s), 2.930 (O 1s), 0.817 (Si 2p), 7.810 (Ti 2p), and 9.660 (V 2p).
UV-Vis diffuse reflectance spectroscopy (UV-Vis DRS) was carried out on a V-770 spectrometer (Jasco, Pfungstadt, Germany) equipped with a praying mantis mirror cell and a high-temperature reaction chamber (Harrick Scientific Products Inc., Pleasantville, NY). Halogen and deuterium lamps were used as light sources in the Vis and UV, respectively. A Peltier-cooled PbS detector was employed for detection. Spectra were recorded from 800 to 200 nm with a spectral resolution of 0.5 nm. MgO was used as a white standard.
Visible Raman spectra (532 nm excitation) were recorded on a Holo Spec f/1.8i spectrometer (Holographic Imaging Spectrograph, Kaiser Optical Systems, Los Altos, CA, USA) equipped with a transmission grating and a charge-coupled device (CCD) detector. 532 nm radiation obtained by frequency doubling from a Nd:YAG laser (Cobolt Inc., Kassel, Germany) was employed for excitation. The spectrometer was calibrated by the emission lines of a standard argon lamp. The laser power was set to 1.6 mW at the position of the sample as measured by a power meter (Ophir, North Logan, UT, USA).
UV Raman spectra (256.7 nm excitation) were performed with a tunable Ti:sapphire solid-state laser (Indigo-S, Coherent, Santa Clara, CA, USA). The Raman scattered light was dispersed in a three-stage spectrometer (TriVista 555, Princeton Instruments, Krailing, Germany) used in subtractive mode and detected by a charged coupled device (CCD) camera (Spec10:2KBUV, Princeton Instruments, Krailing, Germany), cooled to −120 °C with liquid nitrogen. The spectral resolution was 1 cm−1. All spectra were recorded at room temperature by irradiating the samples for 900 s with a laser power of 2.6 mW at a repetition rate of 5 kHz.
DRIFTS was carried on a Vertex 70 spectrometer (Bruker, Mannheim, Germany) equipped with a liquid N2 cooled mercury-cadmium-telluride (MCT) detector and a commercial Harrick cell (Harrick Scientific Products Inc., Pleasantville, NY, USA). As a background standard, KBr powder was used. The spectral resolution of DRIFTS was set to 1 cm−1. Blank spectra were recorded by passing pure N2 through the loaded sample, and the actual gas-phase spectra were obtained by subtracting the blank spectra. For in situ/operando experiments, the gas phase at the outlet of the cell was continuously analyzed by quantitative FT-IR spectroscopy as described above.

5. Conclusions

An NH3-SCR model catalyst was prepared by controlled deposition of titania and vanadia onto a mesoporous high surface area silica support by use of ALD. The final VOx/TiO2/SBA-15 catalyst retained a large surface area, characterized by domains of amorphous titania on silica and the formation of dispersed oligomeric VOx surface species. The presence of similar amounts of deposited Ti and V allows vanadium to be anchored to both TiO2 and Ti–O–Si boundary sites, thus increasing the total number of acid sites.
Compared to other catalysts with similar composition, the SBA-15-based catalyst used in the present study shows a better NH3-SCR performance and a wider temperature window for operation. The superior catalytic behavior is attributed to several factors, including the controlled ALD synthesis, which leads to a high dispersion of VOx oligomeric species due to the high surface area of the mesoporous matrix. Considering the analysis of the RDS at low temperature, desorption of products is of great importance due to their strong interaction with acid sites. The structure of VOx species with amorphous TiO2 domains on the SiO2 substrate, forming hybrid supports of V–O–Ti and V–O–Si, provided more acid sites, especially weak acid sites, thus leading to a better catalytic SCR performance than VOx/TiO2 and also VOx on atomically mixed SiO2 + TiO2 supports prepared by sol-gel methods.
Our findings demonstrate the potential of using high surface area VOx/TiO2/SBA-15 model catalysts for gaining new insight into the factors determining the mode of operation of supported vanadia catalysts used for NH3-SCR.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/10/12/1386/s1, Figure S1. N2 adsorption-desorption isotherms of TiO2/SBA-15+nVOx (n = 1, 3, and 5) materials and SBA-15, Figure S2. UV-Raman (256.7 nm) of the TiO2/SBA-15+3xVOx sample. Prior to VOx deposition the TiO2/SBA-15 sample was calcined at 500 °C for 2 h, Table S1. Results of the XPS analysis of TiO2/SBA-15+3xVOx, Figure S3. UV-Vis DRS of TiO2/SBA-15+3xVOx sample using TiO2/SBA-15 as the reference background. (a) hydrated, (b) dehydrated, Figure S4. In situ detection of the exhaust gas during NH3-SCR reaction over(a) TiO2/SBA-15+1xVOx and (b) TiO2/SBA-15+5xVOx catalysts at 100–450 °C. The gas feed consisted of 500 ppm NH3, 500 ppm NO, and 5% O2 (balanced with N2). The total flow rate was 50 NmL/min (GHSV = 40,000 h−1)

Author Contributions

Original idea, C.H.; Conceptualization, J.S. and C.H.; methodology, J.S.; formal analysis, J.S.; writing—original draft preparation, J.S.; writing—review and editing, C.H.; supervision, C.H.; project administration, C.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Deutsche Forschungsgemeinschaft (DFG-FOR1583) And the APC was funded by the German Research Foundation and the Open Access Publishing Fund of the Technical University of Darmstadt.

Acknowledgments

The authors thank Karl Kopp for performing the XPS analysis, Martin Brodrecht for N2 adsorption-desorption measurements, Kathrin Hofmann for XRD measurements, as well as Anastasia Filtschew for assistance with the DRIFT spectrometer. Financial support by the Deutsche Forschungsgemeinschaft (DFG-FOR1583) is gratefully acknowledged. We furthermore acknowledge support by the German Research Foundation and the Open Access Publishing Fund of the Technical University of Darmstadt.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lai, J.K.; Wachs, I.E. A Perspective on the Selective Catalytic Reduction (SCR) of NO with NH3 by Supported V2O5–WO3/TiO2 Catalysts. ACS Catal. 2018, 8, 6537–6551. [Google Scholar] [CrossRef]
  2. Putluru, S.S.R.; Schill, L.; Godiksen, A.; Poreddy, R.; Mossin, S.; Jensen, A.D.; Fehrmann, R. Promoted V2O5/TiO2 catalysts for selective catalytic reduction of NO with NH3 at low temperatures. Appl. Catal. B Environ. 2016, 183, 282–290. [Google Scholar] [CrossRef] [Green Version]
  3. Zhang, S.; Zhong, Q.; Wang, Y. Effect of rutile phase on V2O5 supported over TiO2 mixed phase for the selective catalytic reduction of NO with NH3. Appl. Surf. Sci. 2014, 314, 112–118. [Google Scholar] [CrossRef]
  4. Zhu, M.; Lai, J.K.; Tumuluri, U.; Ford, M.E.; Wu, Z.; Wachs, I.E. Reaction Pathways and Kinetics for Selective Catalytic Reduction (SCR) of Acidic NOx Emissions from Power Plants with NH3. ACS Catal. 2017, 7, 8358–8361. [Google Scholar] [CrossRef]
  5. Reiche, M. Vanadia grafted on TiO2–SiO2, TiO2 and SiO2 aerogels Structural properties and catalytic behaviour in selective reduction of NO by NH3. Appl. Catal. B Environ. 1999, 23, 187–203. [Google Scholar] [CrossRef]
  6. Engweiler, J.; Baiker, A. Vanadia supported on titania aerogels morphological properties and catalytic behaviour in the selective reduction of nitric oxide by ammonia. Appl. Catal. A Gen. 1994, 120, 187–205. [Google Scholar] [CrossRef]
  7. He, G.; Lian, Z.; Yu, Y.; Yang, Y.; Liu, K.; Shi, X.; Yan, Z.; Shan, W.; He, H. Polymeric vanadyl species determine the low-temperature activity of V-based catalysts for the SCR of NOx with NH3. Sci. Adv. 2018, 4, eaau4637. [Google Scholar] [CrossRef] [Green Version]
  8. Song, I.; Youn, S.; Lee, H.; Lee, S.G.; Cho, S.J.; Kim, D.H. Effects of microporous TiO2 support on the catalytic and structural properties of V2O5/microporous TiO2 for the selective catalytic reduction of NO by NH3. Appl. Catal. B Environ. 2017, 210, 421–431. [Google Scholar] [CrossRef]
  9. Ruff, P.; Schumacher, L.; Rogg, S.; Hess, C. Atomic Layer Deposition-Assisted Synthesis of Embedded Vanadia Catalysts. ACS Catal. 2019, 9, 6349–6361. [Google Scholar] [CrossRef]
  10. Hess, C. Direct correlation of the dispersion and structure in vanadium oxide supported on silica SBA-15. J. Catal. 2007, 248, 120–123. [Google Scholar] [CrossRef]
  11. Hess, C. Characterization of the synthesis and reactivity behavior of nanostructured vanadia model catalysts using XPS and vibrational spectroscopy. Surf. Sci. 2006, 600, 3695–3701. [Google Scholar] [CrossRef] [Green Version]
  12. Hess, C.; Waleska, P.; Ratzka, M.; Janssens, T.V.W.; Rasmussen, S.B.; Beato, P. Hierarchical Vanadia Model Catalysts for Ammonia Selective Catalytic Reduction. Top. Catal. 2017, 60, 1631–1640. [Google Scholar] [CrossRef]
  13. Segura, Y.; Chmielarz, L.; Kuśtrowski, P.; Cool, P.; Dziembaj, R.; VanSant, E.F. Characterisation and reactivity of vanadia–titania supported SBA-15 in the SCR of NO with ammonia. Appl. Catal. B Environ. 2005, 61, 69–78. [Google Scholar] [CrossRef]
  14. Lu, J.L.; Elam, J.W.; Stair, P.C. Atomic layer deposition—Sequential self-limiting surface reactions for advanced catalyst “bottom-up” synthesis. Surf. Sci. Rep. 2016, 71, 410–472. [Google Scholar] [CrossRef] [Green Version]
  15. Johnson, R.W.; Hultqvist, A.; Bent, S.F. A brief review of atomic layer deposition: From fundamentals to applications. Mater. Today 2014, 17, 236–246. [Google Scholar] [CrossRef]
  16. Miikkulainen, V.; Leskelä, M.; Ritala, M.; Puurunen, R.L. Crystallinity of inorganic films grown by atomic layer deposition: Overview and general trends. J. Appl. Phys. 2013, 113, 021301. [Google Scholar] [CrossRef]
  17. Li, S.; Xu, J.; Wang, L.; Yang, N.; Ye, X.; Yuan, X.; Xiang, H.; Liu, C.; Li, H. Effect of post-deposition annealing on atomic layer deposited SiO2 film for silicon surface passivation. Mater. Sci. Semicond. Process. 2020, 106. [Google Scholar] [CrossRef]
  18. Fares, C.; Islam, Z.; Haque, A.; Kneiß, M.; Von Wenckstern, H.; Grundmann, M.; Tadjer, M.; Ren, F.; Pearton, S.J. Effect of Annealing on the Band Alignment of ALD SiO2 on (AlxGa1−x)2O3 for x = 0.2–0.65. ECS J. Solid State Sci. Technol. 2019, 8, P751–P756. [Google Scholar] [CrossRef]
  19. Ding, Y.; Zhang, Y.; Ren, Y.M.; Orkoulas, G.; Christofides, P.D. Machine learning-based modeling and operation for ALD of SiO2 thin-films using data from a multiscale CFD simulation. Chem. Eng. Res. Des. 2019, 151, 131–145. [Google Scholar] [CrossRef]
  20. Fengler, S.; Kriegel, H.; Schieda, M.; Gutzmann, H.; Klassen, T.; Wollgarten, M.; Dittrich, T. Charge Transfer in c-Si(n++)/TiO2(ALD) at the Amorphous/Anatase Transition: A Transient Surface Photovoltage Spectroscopy Study. ACS Appl. Mater. Interfaces 2020, 12, 3140–3149. [Google Scholar] [CrossRef]
  21. Yang, Y.; Wei, W.; Wang, S.; Huang, T.; Yuan, M.; Zhang, R.; Yang, W.; Zhang, T.; Sun, Y.; Yuan, Y.J.; et al. Suppression of Photoinduced Surface Oxidation of Vanadium Dioxide Nanostructures by Blocking Oxygen Adsorption. ACS Omega 2019, 4, 17735–17740. [Google Scholar] [CrossRef] [PubMed]
  22. Hirschberg, C.; Jensen, N.S.; Boetker, J.; Madsen, A.; Kääriäinen, T.O.; Kaariainen, M.-L.; Hoppu, P.; George, S.M.; Murtomaa, M.; Sun, C.C.; et al. Improving Powder Characteristics by Surface Modification Using Atomic Layer Deposition. Org. Process. Res. Dev. 2019, 23, 2362–2368. [Google Scholar] [CrossRef]
  23. Kim, T.W.; Uchida, S.; Kondo, T.; Segawa, H. Optimization of TiO2 compact layer formed by atomic layer deposition for efficient perovskite solar cells. Appl. Phys. Lett. 2019, 115, 203902. [Google Scholar] [CrossRef]
  24. Jo, W.J.; Katsoukis, G.; Frei, H. Ultrathin Amorphous Silica Membrane Enhances Proton Transfer across Solid-to-Solid Interfaces of Stacked Metal Oxide Nanolayers while Blocking Oxygen. Adv. Funct. Mater. 2020, 30. [Google Scholar] [CrossRef] [Green Version]
  25. Yang, G.Q.; Wang, H.; Gong, T.; Song, Y.H.; Feng, H.; Ge, H.Q.; Ge, H.B.; Liu, Z.T.; Liu, Z.W. Understanding the active-site nature of vanadia-based catalysts for oxidative dehydrogenation of ethylbenzene with CO2 via atomic layer deposited VOx on γ-Al2O3. J. Catal. 2019, 380, 195–203. [Google Scholar] [CrossRef]
  26. Samek, I.A.; Bobbitt, N.S.; Snurr, R.Q.; Stair, P.C. Interactions of VOx Species with Amorphous TiO2 Domains on ALD-Derived Alumina-Supported Materials. J. Phys. Chem. C 2018, 123, 7988–7999. [Google Scholar] [CrossRef]
  27. Song, G.Y.; Oh, C.; Sinha, S.; Son, J.; Heo, J. Facile Phase Control of Multivalent Vanadium Oxide Thin Films (V2O5 and VO2) by Atomic Layer Deposition and Postdeposition Annealing. ACS Appl. Mater. Interfaces 2017, 9, 23909–23917. [Google Scholar] [CrossRef]
  28. Held, A.; Kowalska-Kuś, J.; Millot, Y.; Averseng, F.; Calers, C.; Valentin, L.; Dzwigaj, S. Influence of the Preparation Procedure of Vanadium-Containing SiBEA Zeolites on Their Catalytic Activity in Propene Epoxidation. J. Phys. Chem. C 2018, 122, 18570–18582. [Google Scholar] [CrossRef]
  29. Keränen, J.; Guimon, C.; Auroux, A.; Iiskola, E.I.; Niinistö, L. Gas-phase synthesis, structure and surface acid–base properties of highly dispersed vanadia/titania/silica catalysts. Phys. Chem. Chem. Phys. 2003, 5, 5333–5342. [Google Scholar] [CrossRef]
  30. Keränen, J.; Guimon, C.; Iiskola, E.; Auroux, A.; Niinistö, L. Atomic layer deposition and surface characterization of highly dispersed titania/silica-supported vanadia catalysts. Catal. Today 2003, 78, 149–157. [Google Scholar] [CrossRef]
  31. Lei, Z.; Long, A.; Wen, C.; Zhang, J.; Chen, B. Experimental and Kinetic Study of Low Temperature Selective Catalytic Reduction of NO with NH3 over the V2O5/AC Catalyst. Ind. Eng. Chem. Res. 2011, 50, 5360–5368. [Google Scholar] [CrossRef]
  32. Jiang, Y.; Gao, X.; Zhang, Y.; Wu, W.; Luo, Z.; Cen, K. PbCl2-poisoning kinetics of V2O5/TiO2 catalysts for the selective catalytic reduction of NO with NH3. Environ. Prog. Sustain. Energy 2015, 34, 1085–1091. [Google Scholar] [CrossRef]
  33. Yang, S.; Wang, C.; Ma, L.; Peng, Y.; Qu, Z.; Yan, N.; Chen, J.; Chang, H.; Li, J. Substitution of WO3in V2O5/WO3–TiO2 by Fe2O3 for selective catalytic reduction of NO with NH3. Catal. Sci. Technol. 2013, 3, 161–168. [Google Scholar] [CrossRef]
  34. Topsøe, N.Y.; Anstrom, M.; Dumesic, J. Raman, FTIR and Theoretical Evidence for Dynamic Structural Rearrangements of Vanadia/Titania DeNOx Catalysts. Catal. Lett. 2001, 76, 11–20. [Google Scholar] [CrossRef]
  35. Damma, D.; Ettireddy, P.R.; Reddy, B.M.; Smirniotis, P.G. A review of low temperature NH3-SCR for removal of NOx. Catalysts 2019, 9, 349. [Google Scholar] [CrossRef] [Green Version]
  36. Han, L.; Cai, S.; Gao, M.; Hasegawa, J.Y.; Wang, P.; Zhang, J.; Shi, L.; Zhang, D. Selective Catalytic Reduction of NOx with NH3 by Using Novel Catalysts: State of the Art and Future Prospects. Chem. Rev. 2019, 119, 10916–10976. [Google Scholar] [CrossRef]
  37. Topsoe, N. Characterization of the nature of surface sites on vanadia-titania catalysts by FTIR. J. Catal. 1991, 128, 499–511. [Google Scholar] [CrossRef]
  38. Topsoe, N.Y.; Dumesic, J.A.; Topsoe, H. Vanadia-titania catalysts for selective catalytic reduction of nitric-oxide by ammonia: II Studies of active sites and formulation of catalytic cycles. J. Catal. 1995, 151, 241–252. [Google Scholar] [CrossRef]
  39. Topsøe, N.Y.; Slabiak, T.; Clausen, B.S.; Srnak, T.Z.; Dumesic, J.A. Influence of water on the reactivity of vanadia/titania for catalytic reduction of NOx. J. Catal. 1992, 134, 742–746. [Google Scholar] [CrossRef]
  40. Topsøe, N.Y.; Topsøe, H. Combined in-situ FTIR and on-line activity studies: Applications to vanadia-titania DeNOx catalyst. Catal. Today 1991, 9, 77–82. [Google Scholar] [CrossRef]
  41. 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]
  42. Marberger, A.; Ferri, D.; Elsener, M.; Kröcher, O. The Significance of Lewis Acid Sites for the Selective Catalytic Reduction of Nitric Oxide on Vanadium-Based Catalysts. Angew. Chem. Int. Ed. 2016, 55, 11989–11994. [Google Scholar] [CrossRef] [PubMed]
  43. Marberger, A.; Ferri, D.; Elsener, M.; Sagar, A.; Artner, C.; Schermanz, K.; Kröcher, O. Relationship between structures and activities of supported metal vanadates for the selective catalytic reduction of NO by NH3. Appl. Catal. B Environ. 2017, 218, 731–742. [Google Scholar] [CrossRef]
  44. Zhu, M.; Lai, J.K.; Tumuluri, U.; Wu, Z.; Wachs, I.E. Nature of Active Sites and Surface Intermediates during SCR of NO with NH3 by Supported V2O5–WO3/TiO2 Catalysts. J. Am. Chem. Soc. 2017, 139, 15624–15627. [Google Scholar] [CrossRef]
  45. Risse, T.; Hollmann, D.; Brückner, A. Chapter 1. In situ electron paramagnetic resonance (EPR)—A unique tool for analysing structure and reaction behaviour of paramagnetic sites in model and real catalysts. Catalysists 2015, 27, 1–32. [Google Scholar] [CrossRef]
  46. Silversmit, G.; Poelman, H.; Sack, I.; Buyle, G.; Marin, G.B.; De Gryse, R. An in-situ Reduction/Oxidation XAS Study on the EL10V8 VOx/TiO2(Anatase) Powder Catalyst. Catal. Lett. 2006, 107, 61–71. [Google Scholar] [CrossRef]
  47. Silversmit, G.; Van Bokhoven, J.A.; Poelman, H.; Van Dereerden, A.; Marin, G.B.; Reyniers, M.; De Gryse, R. The Structure of a VOxTiO2anatase Powder Catalyst under Reduction and Oxidation at 623K. Phys. Scr. 2005, 115, 798. [Google Scholar] [CrossRef] [Green Version]
  48. Silversmit, G.; Van Bokhoven, J.A.; Poelman, H.; Van Der Eerden, A.M.; Marin, G.B.; Reyniers, M.F.; De Gryse, R. The structure of supported and unsupported vanadium oxide under calcination, reduction and oxidation determined with XAS. Appl. Catal. A Gen. 2005, 285, 151–162. [Google Scholar] [CrossRef]
  49. Feng, H.; Elam, J.W.; Libera, J.A.; Pellin, M.J.; Stair, P.C. Oxidative dehydrogenation of cyclohexane over alumina-supported vanadium oxide nanoliths. J. Catal. 2010, 269, 421–431. [Google Scholar] [CrossRef]
  50. Feng, Z.; Lu, J.; Feng, H.; Stair, P.C.; Elam, J.W.; Bedzyk, M.J. Catalysts Transform While Molecules React: An Atomic-Scale View. J. Phys. Chem. Lett. 2013, 4, 285–291. [Google Scholar] [CrossRef]
  51. Feng, Z.; Cheng, L.; Kim, C.Y.; Elam, J.W.; Zhang, Z.; Curtiss, L.A.; Zapol, P.; Bedzyk, M.J. Atomic-Scale Study of Ambient-Pressure Redox-Induced Changes for an Oxide-Supported Submonolayer Catalyst: VOx/α-TiO2(110). J. Phys. Chem. Lett. 2012, 3, 2845–2850. [Google Scholar] [CrossRef]
  52. Arnarson, L.; Rasmussen, S.B.; Falsig, H.; Lauritsen, J.V.; Moses, P.G. Coexistence of Square Pyramidal Structures of Oxo Vanadium (+5) and (+4) Species Over Low-Coverage VOX/TiO2 (101) and (001) Anatase Catalysts. J. Phys. Chem. C 2015, 119, 23445–23452. [Google Scholar] [CrossRef]
  53. Avdeev, V.I.; Bedilo, A.F. Molecular mechanism of propane oxidative dehydrogenation on surface oxygen radical sites of VOx/TiO2 catalysts. Res. Chem. Intermed. 2015, 42, 5237–5252. [Google Scholar] [CrossRef]
  54. Avdeev, V.I.; Bedilo, A.F. Molecular Mechanism of Oxygen Isotopic Exchange over Supported Vanadium Oxide Catalyst VOx/TiO2. J. Phys. Chem. C 2013, 117, 2879–2887. [Google Scholar] [CrossRef]
  55. Avdeev, V.I.; Tapilin, V.M. Water Effect on the Electronic Structure of Active Sites of Supported Vanadium Oxide Catalyst VOx/TiO2(001). J. Phys. Chem. C 2010, 114, 3609–3613. [Google Scholar] [CrossRef]
  56. Jehng, J.-M.; Deo, G.; Weckhuysen, B.M.; Wachs, I.E. Effect of water vapor on the molecular structures of supported vanadium oxide catalysts at elevated temperatures. J. Mol. Catal. A Chem. 1996, 110, 41–54. [Google Scholar] [CrossRef]
  57. Ruff, P.; Lauterbach, S.; Kleebe, H.J.; Hess, C. Surface structuring of mesoporous materials by controlled synthesis of nanocavities. Microporous Mesoporous Mater. 2016, 235, 160–169. [Google Scholar] [CrossRef]
  58. Thielemann, J.P.; Girgsdies, F.; Schlögl, R.; Hess, C. Pore structure and surface area of silica SBA-15: Influence of washing and scale-up. Beilstein J. Nanotechnol. 2011, 2, 110–118. [Google Scholar] [CrossRef] [Green Version]
  59. Huang, Y.; Gao, D.; Tong, Z.; Zhang, J.; Luo, H. Oxidation of NO over cobalt oxide supported on mesoporous silica. J. Nat. Gas Chem. 2009, 18, 421–428. [Google Scholar] [CrossRef]
  60. Lubas, M.; Jasinski, J.; Sitarz, M.; Kurpaska, L.; Podsiad, P. Raman spectroscopy of TiO2 thin films formed by hybrid treatment for biomedical applications. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2014, 133, 867–871. [Google Scholar] [CrossRef]
  61. Zhang, Q.; Ma, L.; Shao, M.; Huang, J.; Ding, M.; Deng, X.; Wei, X.; Xu, X. Anodic Oxidation Synthesis of One-Dimensional TiO2Nanostructures for Photocatalytic and Field Emission Properties. J. Nanomater. 2014, 2014, 1–14. [Google Scholar] [CrossRef] [Green Version]
  62. Waleska, P.S.; Rupp, S.; Hess, C. Operando Multiwavelength and Time-Resolved Raman Spectroscopy: Structural Dynamics of a Supported Vanadia Catalyst at Work. J. Phys. Chem. C 2018, 122, 3386–3400. [Google Scholar] [CrossRef]
  63. Tallant, D.R.; Bunker, B.C.; Balfe, C.A.; Brinker, C.J. Raman Spectra of Rings in Silicate Materials. MRS Proc. 1986, 73, 261. [Google Scholar] [CrossRef]
  64. Chligui, M.; Guimbretiere, G.; Canizares, A.; Matzen, G.; Vaills, Y.; Simon, P. New Features in the Raman Spectrum of Silica: Key-Points in the Improvement on Structure Knowledge. Available online: https://hal.archives-ouvertes.fr/hal-00520823/document (accessed on 27 November 2020).
  65. Huang, T.; Hu, Z.; Xu, G.S.; Zhang, X.L.; Zhang, J.Z.; Chu, J.H. Inherent optical behavior and structural variation in Na0.5Bi0.5TiO3-6%BaTiO3 revealed by temperature dependent Raman scattering and ultraviolet-visible transmittance. Appl. Phys. Lett. 2014, 104, 111908. [Google Scholar] [CrossRef]
  66. Kreisel, J.; Glazer, A.M.; Jones, G.; Thomas, P.A.; Abello, L.; Lucazeau, G. An x-ray diffraction and Raman spectroscopy investigation of A-site substituted perovskite compounds: The (Na1−xKx)0.5Bi0.5TiO3(0lexle1) solid solution. J. Phys. Condens. Matter 2000, 12, 3267–3280. [Google Scholar] [CrossRef]
  67. Walrafen, G.E.; Chu, Y.C.; Hokmabadi, M.S. Raman spectroscopic investigation of irreversibly compacted vitreous silica. J. Chem. Phys. 1990, 92, 6987–7002. [Google Scholar] [CrossRef]
  68. Li, C.; Xiong, G.; Liu, J.; Ying, P.; Xin, A.Q.; Feng, Z. Identifying Framework Titanium in TS-1 Zeolite by UV Resonance Raman Spectroscopy. J. Phys. Chem. B 2001, 105, 2993–2997. [Google Scholar] [CrossRef]
  69. Hess, C.; Tzolova-Müller, G.; Herbert, R. The Influence of Water on the Dispersion of Vanadia Supported on Silica SBA-15: A Combined XPS and Raman Study. J. Phys. Chem. C 2007, 111, 9471–9479. [Google Scholar] [CrossRef]
  70. Eberhardt, M.A.; Proctor, A.; Houalla, M.; Hercules, D.M. Investigation of V Oxidation States in Reduced V/Al2O3Catalysts by XPS. J. Catal. 1996, 160, 27–34. [Google Scholar] [CrossRef]
  71. Sawatzky, G.A.; Post, D. X-ray photoelectron and Auger spectroscopy study of some vanadium oxides. Phys. Rev. B 1979, 20, 1546–1555. [Google Scholar] [CrossRef]
  72. Pan, Y.; Zhao, W.; Zhong, Q.; Cai, W.; Li, H. Promotional effect of Si-doped V2O5/TiO2 for selective catalytic reduction of NOx by NH3. J. Environ. Sci. 2013, 25, 1703–1711. [Google Scholar] [CrossRef]
  73. Gao, X.; Bare, S.R.; Fierro, J.L.G.; Wachs, I.E. Structural Characteristics and Reactivity/Reducibility Properties of Dispersed and Bilayered V2O5/TiO2/SiO2 Catalysts. J. Phys. Chem. B 1999, 103, 618–629. [Google Scholar] [CrossRef]
  74. Nitsche, D.; Hess, C. Structure of Isolated Vanadia and Titania: A Deep UV Raman, UV–Vis, and IR Spectroscopic Study. J. Phys. Chem. C 2016, 120, 1025–1037. [Google Scholar] [CrossRef]
  75. Waleska, P.; Hess, C. Structural Dynamics of Dispersed Titania During Dehydration and Oxidative Dehydrogenation Studied by In Situ UV Raman Spectroscopy. Catal. Lett. 2018, 148, 2537–2547. [Google Scholar] [CrossRef]
  76. Gao, X.; Bare, S.R.; Fierro, J.L.G.; Banares, M.A.; Wachs, I.E. Preparation and in-Situ Spectroscopic Characterization of Molecularly Dispersed Titanium Oxide on Silica. J. Phys. Chem. B 1998, 102, 5653–5666. [Google Scholar] [CrossRef]
  77. Beck, B.; Harth, M.; Hamilton, N.G.; Carrero, C.A.; Uhlrich, J.J.; Trunschkec, A.; Shaikhutdinov, S.; Schubert, H.; Freund, H.; Schlögl, R.; et al. Partial oxidation of ethanol on vanadia catalysts on supporting oxides with different redox properties compared to propane. J. Catal. 2012, 296, 120–131. [Google Scholar] [CrossRef] [Green Version]
  78. Schraml-Marth, M.; Pohl, M.; Krauss, H.-L.; Wokaun, A. Spectroscopic investigation of the structure of silica-supported vanadium oxide catalysts at submonolayer coverages. J. Chem. Soc. Faraday Trans. 1991, 87, 2635–2646. [Google Scholar] [CrossRef]
  79. Arena, F.; Frusteri, F.; Martra, G.; Coluccia, S.; Parmaliana, A. Surface structures, reduction pattern and oxygen chemisorption of V2O5/SiO2 catalysts. J. Chem. Soc. Faraday Trans. 1997, 93, 3849–3854. [Google Scholar] [CrossRef]
  80. 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 Gen. 2015, 499, 1–12. [Google Scholar] [CrossRef]
  81. Burcham, L.J.; Deo, G.; Gao, X.; Wachs, I.E. In situ IR, Raman, and UV-Vis DRS spectroscopy of supported vanadium oxide catalysts during methanol oxidation. Top. Catal. 2000, 11, 85–100. [Google Scholar] [CrossRef]
  82. Gao, X.; Wachs, I.E. Investigation of Surface Structures of Supported Vanadium Oxide Catalysts by UV−vis−NIR Diffuse Reflectance Spectroscopy. J. Phys. Chem. B 2000, 104, 1261–1268. [Google Scholar] [CrossRef]
  83. Srinivas, D.; Hölderich, W.; Kujath, S.; Valkenberg, M.; Raja, T.; Saikia, L.; Hinze, R.; Ramaswamy, V. Active sites in vanadia/titania catalysts for selective aerial oxidation of β-picoline to nicotinic acid. J. Catal. 2008, 259, 165–173. [Google Scholar] [CrossRef]
  84. Due-Hansen, J.; Rasmussen, S.B.; Mikolajska, E.; Bañares, M.A.; Avila, P.C.; Fehrmann, R. Redox behaviour of vanadium during hydrogen–oxygen exposure of the V2O5-WO3/TiO2 SCR catalyst at 250 °C. Appl. Catal. B Environ. 2011, 107, 340–346. [Google Scholar] [CrossRef]
  85. Davis, E.A.; Mott, N.F. Conduction in non-crystalline systems V. Conductivity, optical absorption and photoconductivity in amorphous semiconductors. Philos. Mag. 1970, 22, 0903–0922. [Google Scholar] [CrossRef]
  86. Ghosh, A.; Bhattacharya, S. Optical and other physical properties of semiconducting cadmium vanadate glasses. J. Appl. Phys. 2007, 101, 83511. [Google Scholar] [CrossRef]
  87. Tauc, J. Absorption edge and internal electric fields in amorphous semiconductors. Mater. Res. Bull. 1970, 5, 721–729. [Google Scholar] [CrossRef]
  88. Waleska, P.S.; Hess, C. Oligomerization of Supported Vanadia: Structural Insight Using Surface-Science Models with Chemical Complexity. J. Phys. Chem. C 2016, 120, 18510–18519. [Google Scholar] [CrossRef]
  89. Borovkov, V.Y.; Mikheeva, E.P.; Zhidomirov, G.M.; Lapina, O.B. Theoretical and Experimental Studies of the Nature of the Catalytic Activity of VOx/TiO2 Systems. Kinet. Catal. 2003, 44, 710–717. [Google Scholar] [CrossRef]
  90. Bulushev, D.A.; Reshetnikov, S.I.; Kiwi-Minsker, L.; Renken, A. Deactivation kinetics of V/Ti-oxide in toluene partial oxidation. Appl. Catal. A Gen. 2001, 220, 31–39. [Google Scholar] [CrossRef] [Green Version]
  91. Kamata, H. 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 Chem. 1999, 139, 189–198. [Google Scholar] [CrossRef]
  92. Lu, J.; Kosuda, K.M.; Van Duyne, R.P.; Stair, P.C. Surface Acidity and Properties of TiO2/SiO2 Catalysts Prepared by Atomic Layer Deposition: UV−visible Diffuse Reflectance, DRIFTS, and Visible Raman Spectroscopy Studies. J. Phys. Chem. C 2009, 113, 12412–12418. [Google Scholar] [CrossRef] [Green Version]
  93. Mu, J.C.; Li, X.Y.; Sun, W.B.; Fan, S.Y.; Wang, X.Y.; Wang, L.; Qin, M.C.; Gan, G.Q.; Yin, Z.F.; Zhang, D.K. Inductive Effect Boosting Catalytic Performance of Advanced Fe1−xVxO delta Catalysts in Low-Temperature NH3 Selective Catalytic Reduction: Insight into the Structure, Interaction, and Mechanisms. ACS Catal. 2018, 8, 6760–6774. [Google Scholar] [CrossRef]
  94. Cheng, K.; Liu, J.; Zhao, Z.; Wei, Y.; Jiang, G.; Duan, A. Direct synthesis of V–W–Ti nanoparticle catalysts for selective catalytic reduction of NO with NH3. RSC Adv. 2015, 5, 45172–45183. [Google Scholar] [CrossRef]
  95. Zhu, L.; Zhong, Z.; Yang, H.; Wang, C. Effect of MoO3 on vanadium based catalysts for the selective catalytic reduction of NOx with NH3 at low temperature. J. Environ. Sci. 2017, 56, 169–179. [Google Scholar] [CrossRef]
  96. Topsoe, N.; Topsoe, H.; Dumesic, J.A. Vanadia/Titania Catalysts for Selective Catalytic Reduction (SCR) of Nitric-Oxide by Ammonia. J. Catal. 1995, 151, 226–240. [Google Scholar] [CrossRef]
  97. Anstrom, M. Density functional theory studies of mechanistic aspects of the SCR reaction on vanadium oxide catalysts. J. Catal. 2003, 213, 115–125. [Google Scholar] [CrossRef]
  98. Farber, M.; Harris, S.P. Kinetics of ammonia-nitric oxide reactions on vanadium oxide catalysts. J. Phys. Chem. 1984, 88, 680–682. [Google Scholar] [CrossRef]
  99. Bredow, T.; Homann, T.; Jug, K. Adsorption of NO, NH3 and H2O on V2O5/TiO2 catalysts. Res. Chem. Intermed. 2004, 30, 65–73. [Google Scholar] [CrossRef]
  100. Yao, H.; Chen, Y.; Zhao, Z.; Wei, Y.; Liu, Z.; Zhai, D.; Liu, B.; Xu, C. Periodic DFT study on mechanism of selective catalytic reduction of NO via NH3 and O2 over the V2O5 (001) surface: Competitive sites and pathways. J. Catal. 2013, 305, 67–75. [Google Scholar] [CrossRef]
  101. Nickl, J.; Dutoit, D.; Baiker, A.; Scharf, U.; Wokaun, A. Vanadia catalysts for the selective catalytic reduction of NO by NH3 prepared by vapour deposition of vanadyl alkoxide onto various supports. Ber. Bunsenges. Phys. Chem. 1993, 97, 217–228. [Google Scholar]
  102. Pinaeva, L.; Suknev, A.; Budneva, A.; Paukshtis, E.; Bal’Zhinimaev, B. On the role of oxygen in the reaction of NO reduction by NH3 over monolayer V2O5 TiO2 catalyst. J. Mol. Catal. A Chem. 1996, 112, 115–124. [Google Scholar] [CrossRef]
  103. Mihai, O.; Creaser, D.; Olsson, L. Adsorption and Oxidation Investigations over Pt/Al2O3 Catalyst: A Microcalorimetric Study. Catalists 2016, 6, 73. [Google Scholar] [CrossRef] [Green Version]
  104. Zhang, S.; Zhang, B.; Liu, B.; Sun, S. A review of Mn-containing oxide catalysts for low temperature selective catalytic reduction of NOx with NH3: Reaction mechanism and catalyst deactivation. RSC Adv. 2017, 7, 26226–26242. [Google Scholar] [CrossRef] [Green Version]
  105. Klyushina, A.; Pacultová, K.; Karásková, K.; Jirátová, K.; Ritz, M.; Fridrichová, D.; Volodarskaja, A.; Obalová, L. Effect of preparation method on catalytic properties of Co-Mn-Al mixed oxides for N2O decomposition. J. Mol. Catal. A Chem. 2016, 425, 237–247. [Google Scholar] [CrossRef]
  106. Hamilton, N.; Wolfram, T.; Müller, G.T.; Hävecker, M.; Kröhnert, J.; Carrero, C.; Schomäcker, R.; Trunschkec, A.; Schlögl, R. Topology of silica supported vanadium–titanium oxide catalysts for oxidative dehydrogenation of propane. Catal. Sci. Technol. 2012, 2, 1346–1359. [Google Scholar] [CrossRef] [Green Version]
  107. Deo, G.; Wachs, I. Effect of Additives on the Structure and Reactivity of the Surface Vanadium Oxide Phase in V2O5/TiO2 Catalysts. J. Catal. 1994, 146, 335–345. [Google Scholar] [CrossRef]
  108. Jehng, J.M.; Wachs, I.E. The molecular structures and reactivity of V2O5/TiO2/SiO2 catalysts. Catal. Lett. 1992, 13, 9–19. [Google Scholar] [CrossRef]
  109. Rouquerol, F.O.; Rouquerol, J.; Sing, K.S.W.; Llewellyn, P.L.; Maurin, G. Adsorption by Powders and Porous Solids: Principles, Methodology and Applications, 2nd ed.; Elsevier/AP: Amsterdam, The Netherlands, 2014. [Google Scholar]
  110. Haber, J.; Machej, T.; Czeppe, T. The phenomenon of wetting at solid/solid interface. Surf. Sci. 1985, 151, 301–310. [Google Scholar] [CrossRef]
  111. Wang, C.B.; Cai, Y.; Wachs, I.E. Reaction-Induced Spreading of Metal Oxides onto Surfaces of Oxide Supports during Alcohol Oxidation: Phenomenon, Nature, and Mechanisms. Langmuir 1999, 15, 1223–1235. [Google Scholar] [CrossRef]
  112. Zhu, H.; Ould-Chikh, S.; Dong, H.; Llorens, I.; Saih, Y.; Anjum, D.H.; Hazemann, J.L.; Basset, J.M. VOx/SiO2Catalyst Prepared by Grafting VOCl3 on Silica for Oxidative Dehydrogenation of Propane. Chem. Cat. Chem. 2015, 7, 3332–3339. [Google Scholar] [CrossRef]
  113. Moisii, C.; Curran, M.D.; Van De Burgt, L.J.; Stiegman, A.E. Raman spectroscopy of discrete silica supported vanadium oxide: Assignment of fundamental stretching modes. J. Mater. Chem. 2005, 15, 3519–3524. [Google Scholar] [CrossRef]
  114. Nuguid, R.J.G.; Ferri, D.; Marberger, A.; Nachtegaal, M.; Kröcher, O. Modulated Excitation Raman Spectroscopy of V2O5/TiO2: Mechanistic Insights into the Selective Catalytic Reduction of NO with NH3. ACS Catal. 2019, 9, 6814–6820. [Google Scholar] [CrossRef]
  115. Lietti, L.; Alemany, J.L.; Forzatti, P.; Busca, G.; Ramis, G.; Giamello, E.; Bregani, F. Reactivity of V2O5-WO3/TiO2 catalysts in the selective catalytic reduction of nitric oxide by ammonia. Catal. Today 1996, 29, 143–148. [Google Scholar] [CrossRef]
  116. Arnarson, L.; Falsig, H.; Rasmussen, S.B.; Lauritsen, J.V.; Moses, P.G. A complete reaction mechanism for standard and fast selective catalytic reduction of nitrogen oxides on low coverage VO /TiO2(0 0 1) catalysts. J. Catal. 2017, 346, 188–197. [Google Scholar] [CrossRef]
  117. Giakoumelou, I.; Fountzoula, C.; Kordulis, C.; Boghosian, S. Molecular structure and catalytic activity of V2O5/TiO2 catalysts for the SCR of NO by NH3: In situ Raman spectra in the presence of O2, NH3, NO, H2, H2O, and SO2. J. Catal. 2006, 239, 1–12. [Google Scholar] [CrossRef]
  118. Dong, L.; Sun, C.; Tang, C.; Zhang, B.; Zhu, J.; Liu, B.; Gao, F.; Hu, Y.; Dong, L.; Chen, Y. Investigations of surface VOx species and their contributions to activities of VOx/Ti0.5Sn0.5O2 catalysts toward selective catalytic reduction of NO by NH3. Appl. Catal. A Gen. 2012, 431, 126–136. [Google Scholar] [CrossRef]
  119. Zhang, Q.M.; Song, C.; Lyu, G.; Bin, F.; Pang, H.T.; Song, J. Effect of metal oxide partial substitution of V2O5 in V2O5–WO3/TiO2 on selective catalytic reduction of NO with NH3. J. Ind. Eng. Chem. 2015, 24, 79–86. [Google Scholar] [CrossRef]
  120. Zhang, S.; Zhong, Q. Promotional effect of WO3 on O2 over V2O5/TiO2 catalyst for selective catalytic reduction of NO with NH3. J. Mol. Catal. A Chem. 2013, 373, 108–113. [Google Scholar] [CrossRef]
  121. Wang, P.; Guo, R.T. The promotion effect of copper doping on the potassium resistance of V/TiO2 catalyst for selective catalytic reduction of NO with NH3. Chem. Pap. 2017, 71, 2253–2259. [Google Scholar] [CrossRef]
  122. Handy, B.E.; Baiker, A.; Schraml-Marth, M.; Wokaun, A. Vanadia supported on TiO2-SiO2 mixed oxide gels: Structure of the dispersed phase and activity for the selective catalytic reduction of NO with NH3. J. Catal. 1992, 133, 1–20. [Google Scholar] [CrossRef]
  123. Handy, B.E.; Maciejewski, M.; Baiker, A. Vanadia, vanadia-titania, and vanadia-titania-silica gels: Structural genesis and catalytic behavior in the reduction of nitric oxide with ammonia. J. Catal. 1992, 134, 75–86. [Google Scholar] [CrossRef]
  124. Kobayashi, M.; Kuma, R.; Masaki, S.; Sugishima, N. TiO2-SiO2 and V2O5/TiO2-SiO2 catalyst: Physico-chemical characteristics and catalytic behavior in selective catalytic reduction of NO by NH3. Appl. Catal. B Environ. 2005, 60, 173–179. [Google Scholar] [CrossRef]
  125. Won, J.M.; Kim, M.S.; Hong, S.C. Effect of vanadium surface density and structure in VOx/TiO2 on selective catalytic reduction by NH3. Korean J. Chem. Eng. 2018, 35, 2365–2378. [Google Scholar] [CrossRef]
  126. Dutoit, D.C.M.; Reiche, M.A.; Baiker, A. Vanadia-silica aerogels—Structure and catalytic properties in selective reduction of NO by NH3. Appl. Catal. B-Environ. 1997, 13, 275–288. [Google Scholar] [CrossRef]
  127. Soyer, S.; Uzun, A.; Senkan, S.; Onal, I. A quantum chemical study of nitric oxide reduction by ammonia (SCR reaction) on V2O5 catalyst surface. Catal. Today 2006, 118, 268–278. [Google Scholar] [CrossRef]
  128. Dai, H.; Bell, A.T.; Iglesia, E. Effects of molybdena on the catalytic properties of vanadia domains supported on alumina for oxidative dehydrogenation of propane. J. Catal. 2004, 221, 491–499. [Google Scholar] [CrossRef]
  129. Zhao, D.; Sun, J.; Li, Q.; Stucky, G.D. Morphological Control of Highly Ordered Mesoporous Silica SBA-15. Chem. Mater. 2000, 12, 275–279. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns and (b) Raman spectra of the samples TiO2/SBA-15 + nxVOx (n = 1, 3 and 5) compared to TiO2/SBA-15 under ambient conditions. (i) TiO2/SBA-15, (ii) TiO2/SBA-15+1xVOx, (iii) TiO2/SBA-15+3xVOx, (iv) TiO2/SBA-15+5xVOx.
Figure 1. (a) XRD patterns and (b) Raman spectra of the samples TiO2/SBA-15 + nxVOx (n = 1, 3 and 5) compared to TiO2/SBA-15 under ambient conditions. (i) TiO2/SBA-15, (ii) TiO2/SBA-15+1xVOx, (iii) TiO2/SBA-15+3xVOx, (iv) TiO2/SBA-15+5xVOx.
Catalysts 10 01386 g001
Figure 2. (a) O 1s, V 2p, and Ti 2p photoemission of TiO2/SBA-15+3xVOx. Detailed view of (b) the V 2p3/2 region (c) the Ti 2p3/2 region. The colored lines represent the result of a least-square fit analysis.
Figure 2. (a) O 1s, V 2p, and Ti 2p photoemission of TiO2/SBA-15+3xVOx. Detailed view of (b) the V 2p3/2 region (c) the Ti 2p3/2 region. The colored lines represent the result of a least-square fit analysis.
Catalysts 10 01386 g002
Figure 3. (a) DR UV-is and (b) Raman spectra of TiO2/SBA-15+3xVOx under ambient (hydrated) and dehydrated conditions in comparison to dehydrated TiO2/SBA-15.
Figure 3. (a) DR UV-is and (b) Raman spectra of TiO2/SBA-15+3xVOx under ambient (hydrated) and dehydrated conditions in comparison to dehydrated TiO2/SBA-15.
Catalysts 10 01386 g003
Figure 4. Catalytic performance of TiO2/SBA-15+3xVOx in NH3-SCR. The temperature was increased from 100 °C to 450 °C. The feed gas consisted of 500 ppm NH3, 500 ppm NO, and 5% O2 balanced with N2. The total flow rate was 50 NmL min−1 (GHSV = 40,000 h−1).
Figure 4. Catalytic performance of TiO2/SBA-15+3xVOx in NH3-SCR. The temperature was increased from 100 °C to 450 °C. The feed gas consisted of 500 ppm NH3, 500 ppm NO, and 5% O2 balanced with N2. The total flow rate was 50 NmL min−1 (GHSV = 40,000 h−1).
Catalysts 10 01386 g004
Figure 5. DRIFT spectra of TiO2/SBA-15+3xVOx exposed to a gas mixture of 500 ppm NH3, 500 ppm NO, and 5% O2 (balanced with N2) at temperatures from 100 °C to 450 °C within (a) 850–3800 cm−1, (b) 2600–3600 cm−1. The total flow rate was 50 NmL/min (GHSV = 40,000 h−1).
Figure 5. DRIFT spectra of TiO2/SBA-15+3xVOx exposed to a gas mixture of 500 ppm NH3, 500 ppm NO, and 5% O2 (balanced with N2) at temperatures from 100 °C to 450 °C within (a) 850–3800 cm−1, (b) 2600–3600 cm−1. The total flow rate was 50 NmL/min (GHSV = 40,000 h−1).
Catalysts 10 01386 g005
Figure 6. DRIFT spectra of TiO2/SBA-15+3xVOx exposed to different gas environments at 300 °C and at a total gas flow rate of 50 NmL/min (GHSV = 40,000 h−1) within (a) 2600–3500 cm−1, (b) 3500–4000 cm−1.
Figure 6. DRIFT spectra of TiO2/SBA-15+3xVOx exposed to different gas environments at 300 °C and at a total gas flow rate of 50 NmL/min (GHSV = 40,000 h−1) within (a) 2600–3500 cm−1, (b) 3500–4000 cm−1.
Catalysts 10 01386 g006
Figure 7. In situ detection of the exhaust gas during NH3-SCR reaction of TiO2/SBA-15+3xVOx at 100–450 °C. The feed consisted of 500 ppm NH3, 500 ppm NO, and 5% O2 (balanced with N2) at a total flow rate of 50 NmL/min (GHSV = 40,000 h−1).
Figure 7. In situ detection of the exhaust gas during NH3-SCR reaction of TiO2/SBA-15+3xVOx at 100–450 °C. The feed consisted of 500 ppm NH3, 500 ppm NO, and 5% O2 (balanced with N2) at a total flow rate of 50 NmL/min (GHSV = 40,000 h−1).
Catalysts 10 01386 g007
Table 1. Surface, porosity, and V loading characteristics of the prepared samples.
Table 1. Surface, porosity, and V loading characteristics of the prepared samples.
SamplesStotal
[m2/g]
Dp
[nm]
Vtotal
[cm3/g]
LV
[V/nm2]
V:Ti
(wt%) a
SBA-159526.960.82----
TiO2/SBA-15+1xVOx4106.330.490.41.4:5.8
TiO2/SBA-15+3xVOx3666.330.431.65.1:5.9
TiO2/SBA-15+5xVOx3556.330.432.06.0:5.6
a From ICP-OES.
Table 2. Catalytic performance of TiO2/SBA-15+3xVOx in NH3-SCR as a function of temperature. The feed gas consisted of 500 ppm NH3, 500 ppm NO, and 5% O2 balanced with N2. The flow rate was 50 NmL min−1 (GHSV = 40,000 h−1).
Table 2. Catalytic performance of TiO2/SBA-15+3xVOx in NH3-SCR as a function of temperature. The feed gas consisted of 500 ppm NH3, 500 ppm NO, and 5% O2 balanced with N2. The flow rate was 50 NmL min−1 (GHSV = 40,000 h−1).
Temperature, °C100150200250300350400450
NOx conversion, %33.433.151.361.159.960.553.241.9
N2 selectivity, %10099.499.398.197.095.192.890.0
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shen, J.; Hess, C. High Surface Area VOx/TiO2/SBA-15 Model Catalysts for Ammonia SCR Prepared by Atomic Layer Deposition. Catalysts 2020, 10, 1386. https://doi.org/10.3390/catal10121386

AMA Style

Shen J, Hess C. High Surface Area VOx/TiO2/SBA-15 Model Catalysts for Ammonia SCR Prepared by Atomic Layer Deposition. Catalysts. 2020; 10(12):1386. https://doi.org/10.3390/catal10121386

Chicago/Turabian Style

Shen, Jun, and Christian Hess. 2020. "High Surface Area VOx/TiO2/SBA-15 Model Catalysts for Ammonia SCR Prepared by Atomic Layer Deposition" Catalysts 10, no. 12: 1386. https://doi.org/10.3390/catal10121386

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