Next Article in Journal
Converting Hybrid Mechanisms to Electron Transfer Mechanism by Increasing Biochar Pyrolysis Temperature for the Degradation of Sulfamethoxazole in a Sludge Biochar/Periodate System
Next Article in Special Issue
Catalytic Degradation of Toluene over MnO2/LaMnO3: Effect of Phase Type of MnO2 on Activity
Previous Article in Journal
Study on Optimum Preparation Conditions of ZnIn2S4 to Effectively Reduce Cr(VI) under Visible Light Radiation
Previous Article in Special Issue
Abundant Oxygen Vacancies Induced by the Mechanochemical Process Boost the Low-Temperature Catalytic Performance of MnO2 in NH3-SCR
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Protection Effect of Ammonia on CeNbTi NH3-SCR Catalyst from SO2 Poisoning

1
The Key Laboratory of Advanced Materials of Ministry of Education, School of Materials Science and Engineering, Tsinghua University, Beijing 100084, China
2
National Institute of Clean-and-Low-Carbon Energy (NICE), Beijing 102209, China
3
Advanced Materials Institute, Graduate School at Shenzhen, Tsinghua University, Shenzhen 518055, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Catalysts 2022, 12(11), 1430; https://doi.org/10.3390/catal12111430
Submission received: 18 October 2022 / Revised: 8 November 2022 / Accepted: 11 November 2022 / Published: 14 November 2022

Abstract

:
CeNbTi catalyst was poisoned in different sulfur poisoning atmospheres at 300 °C for 6 h and then was evaluated for selective catalytic reduction (SCR) of NOx with NH3. The catalyst deactivation upon SO2 exposure was effectively inhibited in the presence of NH3. Temperature-programmed decomposition (TPD) analyses were applied to identify deposit species on the poisoned catalysts by comparison with several groups of reference samples. Diffuses reflectance infrared Fourier transform spectroscopy (DRIFTS) over CeNbTi catalysts with different poisoning pretreatments and gas purging sequences were designed to investigate the roles of NH3 in the removal of surface sulfites and sulfates. More ammonium sulfates including ammonium bisulfate and ammonium cerium sulfate were generated instead of inert cerium sulfate in these conditions. The mechanisms about the formation and transformation of surface deposits upon sulfur poisoning w/wo NH3 were explored, which provided a basis for developing Ce-based mixed oxides as SCR catalysts for stationary sources.
Keywords:
ceria; SCR; sulfur; NH3; TPD; DRIFTS; deposit

1. Introduction

Over the past decades nitrogen oxides (NOx) that are produced from mobile and stationary sources have led to many environment problems [1]. The selective catalytic reduction (SCR) of NOx using ammonia as the reductant is one of the most promising deNOx technologies [2]. Ceria-based catalysts have attracted plenty of interest due to some advantageous features including high NOx conversion, nontoxicity, and low cost [3,4,5,6]. Recently, the activity and hydrothermal stability of ceria-based catalysts have been improved by inducing WO3 and SnO2 as a promoter, realizing over 90% NOx conversion at 300−550 °C after hydrothermal-aging at 1000 °C, as reported by He et al. [7,8]. This means the application scenarios of ceria-based catalysts are broadened, such as used downstream of the diesel particulate filter (DPF). However, in typical application scenarios, especially coal-fired power plants or the aftertreatment system of marine diesel engines, the effluent gases always contain a certain concentration of SO2. Thus, SO2 resistance is one of the most important characteristics of SCR catalysts. The deposition of cerium sulfates and ammonium bisulfate has been recognized as the main factor of deactivation of ceria-based catalysts. The deactivation of these catalysts in the presence of SO2 is sensitive to reaction temperatures. Zhang et al. [9] reported that with the increase of reaction temperature from 180 to 300 °C, the content of cerium sulfates increased significantly, accompanied with the decreased deposition of ammonium sulfates/bisulfates. Zhang et al. [10] found that the sulfation process was gradually worsened with raising the treating temperature, and the sulfate species over CeO2 changed from surface sulfates to bulk-like ones, and then to bulk ones. These studies indicate that different sulfur species are formed on ceria-based catalysts at different temperature ranges, which affect NH3-SCR activity significantly.
In addition to reaction temperatures, other gases, especially ammonia, influence the type of sulfur species that is formed during SO2 poisoning. Research on the effects of ammonia on SO2 poisoning have been focused on Cu-CHA catalysts [11,12,13]. For example, as investigated by Wijayanti et al. [11,12], ammonium-sulfur species formed over Cu-SSZ-13 during sulfur poisoning under SCR conditions at 300 °C, reducing the availability of copper sites for the redox SCR cycle due to the site-blocking effect of the deposits [12]. Wang et al. [13] found that copper sulfate was present in all the Cu/SAPO-34 samples that were sulfated at various temperatures, whereas ammonia sulfate was only found over the sample at 250 °C. They suggested that both copper sulfate and ammonia sulfate species decrease the SCR reaction rate of the Cu/SAPO-34 catalyst by reducing the number of Cu2+ sites. Moreover, Jangjou et al. [14] found that ammonia can drive sulfur from a more thermodynamically copper sulfate to less stable ammonium sulfate over Cu/SAPO-34 catalysts. The results showed that ammonium affected the types of sulfur-containing species. This appears favorable, as NH3 can react with pre-adsorbed sulfur on the catalyst to form ammonium-sulfur species, which decomposes at lower temperatures in comparison to the other sulfate forms.
Besides, similar phenomena have also been observed over Fe-W and Nb-V/Ce catalysts in recent studies [15,16], that the introduction of NH3 in the sulfation process at 300 °C dramatically reduced the formation of surface metal sulfate species. Fe2(SO4)3 can decompose in NH3 atmosphere at approximately 220 °C, helping reduce sulfate deposit on the catalyst [15]. The NH3-SCR performances of metal oxide catalysts (such as γ-Fe2O3, CeO2, and γ-MnO2) in a sulfur-containing atmosphere were closely related to the difference in the deposition/decomposition ability of the formed sulfate species as reported by An et al. [17]. Adsorbed NH3 and H2O can promote the decomposition of metal sulfate Me(SO4)y, proven by temperature-programmed surface reaction (TPSR) of NH3.
Despite the above studies, the SO2 poisoning over ceria-based catalysts in an SCR environment, to our knowledge, has not been comprehensively studied. Different from Cu-CHA, ceria-based catalysts tend to deactivate more quickly in sulfur poisoning due to the formation of cerium sulfate. This may have a positive effect on the activity if the formation of cerium sulfate can be inhibited in the NH3-containing atmospheres. Therefore, sulfur poisoning in different atmospheres is worth studying for ceria-based catalysts. In our previous work [18,19], the addition of niobium on CeWTi catalyst resulted in high NOx conversion in a wide temperature window and significantly improved sulfur resistance at low and medium temperatures. In this work, CeNbTi was chosen as a typical ceria-based oxides catalyst, and 300 °C was chosen as a typical temperature for sulfur poisoning of SCR catalysts. The sulfur poisoning mechanism of CeNbTi catalyst in an SCR environment and specifically the role of NH3 were explored.

2. Experimental

2.1. Catalyst Preparation

CeNbTi catalyst was prepared by a co-precipitation method. C12H7NbO24 (Aladdin, China) and Ce(NO3)3·6H2O (Aladdin) were dissolved in deionized water with a Nb2O5:CeO2 mass ratio of 2:1. The precursors were mixed with commercial TiO2 powders (DT51, Cristal, Saudi Arabia) in deionized water according to a (Nb2O5 + CeO2):TiO2 molar ratio of 3:7. Diluted ammonium hydroxide (Beijing Chem., Beijing, China) was then added in the mixed solution as a precipitating agent under vigorous stirring. The obtained precipitate was dried at 110 °C overnight and calcined in a muffle furnace at 600 °C for 5 h. The loading of CeO2 and Nb2O5 are 12 wt.% and 22 wt.%, respectively.
The poisoned catalysts were obtained by treating the as-received CeNbTi in a gas flow consisting of 500 ppm NO (when used), 500 ppm NH3 (when used), 200 ppm SO2 (when used), 5% H2O (when used), 5% O2, and balanced N2 at 300 °C for 6 h.
For reference, the deposited catalysts were prepared by an incipient wetness impregnation method. Ammonium sulfite (Aladdin), ammonium hydrogen sulfate (Aladdin), cerium sulfate (Aladdin), and ammonium cerium sulfate (Aladdin) were impregnated on the CeNbTi catalyst with a nominal loading of 2 wt.%. The aqueous solution was added dropwise to the catalyst powders. The mixture was dried at 80 °C for 12 h, and the obtained samples were denoted as (NH4)2SO3/CeNbTi, NH4HSO4/CeNbTi, Ce(SO4)2/CeNbTi, and (NH4)4Ce(SO4)4/CeNbTi, respectively.

2.2. Activity Measurement

The NH3-SCR activity measurement was carried out in a fixed bed reactor with powder catalyst (200 mg, 40–60 mesh). The reaction gas mixture consisted of 500 ppm NO, 500 ppm NH3, 5% O2, 5% H2O, and N2 in balance. The gas hourly space velocity (GHSV) was 80,000 h−1 and the reaction temperature was set at 300 °C. The NOx (NO, NO2, and N2O) concentrations in the outlet gases were measured by a Nicolet 380 infrared (IR) spectrometer (Thermo Fisher, Waltham, MA, USA). The NOx conversion was calculated as follows:
NO x   conversion   ( % ) = [ NO ] in + [ NO 2 ] in [ NO ] out [ NO 2 ] out [ NO ] in + [ NO 2 ] in × 100

2.3. Catalyst Characterization

Nitrogen adsorption isotherms were measured on a JW-BK200(JWGB, China) instrument. All the samples were degassed at 220 °C for 1 h before the nitrogen adsorption measurements. The Brunner–Emmett–Teller (BET) surface area was calculated by a multipoint BET method.
Temperature-programmed decomposition (TPD) analyses were performed on an MKS 2030 Fourier transform infrared (FT-IR) gas analyzer (MKS Instruments, USA). A total of 100 mg sample was pretreated in N2 at 100 °C for 30 min. Then, the sample was ramped to 1000 °C at a rate of 10 °C/min in N2.
In situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) was performed on a Nicolet 6700 IR spectrometer (Thermo Fisher). The spectra were calculated by the Kubelka–Munk function. The reaction atmosphere consisted of 200/500 ppm NH3 (when used), 100/200 ppm SO2 (when used), 5% O2, and N2 in balance. The catalyst in a diffuse reflectance IR cell was pretreated by N2 (100 mL/min) at 500 °C for 30 min to remove traces of organic residues. Then, the sample was cooled down to 100 °C. Afterward, the samples were purged with the reaction atmosphere and the spectra were collected as a function of time.

3. Results

3.1. NH3-SCR Activity

The NOx conversions of the used and poisoned catalysts were obtained from the isothermal activity test at 300 °C and the results are shown in Figure 1. After running in a standard NH3-SCR condition for 6 h, the used CeNbTi catalyst still exhibited high SCR activity, with a temperature window of 195–420 °C in which the NOx conversion exceeded 80% (Figure S1). When only SO2 and O2 are applied for a sulfur poisoning test, the catalyst is deactivated severely with the NOx conversion declining sharply to 62%. Such a poisoning effect is maintained in the presence of water. Nevertheless, the runnings in NH3-containing atmospheres prevent the catalyst deactivation to a great extent, and these poisoned catalysts show ca. 90% NOx conversion. Again, the influence of water can be omitted. To judge the role of NO, it was removed from the treating atmosphere and the NOx conversion of the obtained catalyst increased slightly from 88% to 91%. A similar negligible role of H2O can also be observed. All these demonstrate that ammonia is responsible for the greatly improved sulfur resistance of the ceria-based catalyst.

3.2. TPD Analyses

3.2.1. Identification of Surface Deposits

To identify the deposits on the poisoned catalysts, the TPD experiments were performed in N2. Almost no NOx (including NO, NO2, and N2O) were observed within the whole temperature range (not shown), indicating their poor adsorption when competing with sulfur oxides. Figure 2 shows the evolution of SO2 and NH3 signals during the TPD processes. The NH3 and SO2 desorption were estimated by peak fitting and the results are summarized in Table 1. Since the sulfur poisoning tests were performed at 300 °C, the NH3 signal also starts at this temperature in Figure 2a. The NH3 release (16 μmol/g) for the catalyst that was treated in a simple SCR condition (without SO2) is derived from desorption of ammonia that is adsorbed at surface acid sites. The ammonia desorption is doubled in the SO2-containing treating atmospheres. According to the reports of Xu et al. [9] and Song et al. [20], the increased ammonia can be related to the formation of ammonium sulfates including ammonium bisulfate, ammonium sulfate, and ammonium cerium sulfate. The presence of H2O or NO does not affect the ammonium adsorption on the catalyst upon the sulfur poisoning treatments. Since the specific surface areas of catalysts almost have no changes with or without NH3-containing SO2 poisoning atmosphere and the maximal amount of NH4HSO4 (about 17 μmol/g) is much smaller than that of metal sulfates and ammonium metal sulfates, the physical poisoning (covering effect) of the deposited NH4HSO4 should not be the main factor to deactivate the catalyst.
Compared with NH3, the SO2 signal appears at much higher temperatures (450–1000 °C) in Figure 2b. Interestingly, there are two SO2 peaks over the catalyst that was treated in the simple SCR condition without SO2. The distinct SO2 peak centered at 730 °C with a shoulder at 900 °C is associated with the decomposition of titanium sulfates in the original material [18,21,22]. After sulfur poisoning, the major peak shifts towards a lower temperature (685 °C) and another small peak appears at 565 °C. According to the previous studies [23,24], the low-temperature peak is ascribed to the decomposition of metal sulfites. The overlapped peak at 685 °C is attributed to the decomposition of ammonium sulfates and cerium sulfates [25,26]. When ammonia is added into the poisoning atmosphere, both the SO2 peaks at 565 and 685 °C decrease in intensity significantly. As listed in Table 1, about two thirds of sulfites and about one third of sulfate deposits are reduced on those catalysts.
To further verify the deposits on the treated catalysts, one group of reference samples was prepared and their TPD curves are shown in Figure 2c,d. Among these samples, (NH4)2SO3, NH4HSO4, Ce(SO4)2, and (NH4)4Ce(SO4)4 were chosen as the representatives of ammonium sulfites, ammonium bisulfates, cerium sulfates, and ammonium ceria sulfates, respectively. The NH3 desorption occurs at a much lower temperature (around 250 °C) over (NH4)2SO3/CeNbTi. Comparatively, the NH3 desorption peak appears at 340 and 380 °C for the NH4HSO4 and (NH4)4Ce(SO4)4 impregnated catalysts, respectively, which are close to those (400 °C) over the catalysts that were poisoned in the NH3-containing atmospheres Figure 2a.
As shown in Figure 2d, a small SO2 desorption peak at 525 °C is only found over (NH4)2SO3/CeNbTi, indicating the origin of the 565 °C peak in Figure 2b. It is interesting that a distinct peak at 700 °C is also observed for this sample. One possible explanation is the re-adsorption of SO2 from sulfite decomposition with ceria acting as strong basic sites, which releases at higher temperatures. Another possibility that cannot be excluded is that sulfites may be readily oxidized to sulfates by active oxygen in ceria during the impregnation and TPD processes. Similar phenomena about delayed release of SO2 have been observed over many catalysts including VTi, VWTi, and Cu-SSZ-13 [26,27,28]. All the impregnated sulfates show a similar SO2 desorption peak at 600–750 °C. Therefore, it is difficult to identify specific sulfates from the 685 °C desorption peaks in Figure 2b in this way. Nevertheless, some important information can still be deduced. Considering the fact that almost only cerium sulfates (Ce(SO4)2 and Ce2(SO4)3) are generated in the poisoning atmospheres without NH3 and the total SO2 desorption declines over the SO2 + NH3 co-treated catalysts (Figure 2b), the additional NH3 release over the latter catalysts (Figure 2a) should be ascribed to the decomposition of ammonium sulfates such as NH4HSO4 and (NH4)4Ce(SO4)4. All these indicate that the formation of cerium sulfates is inhibited to a large extent in the presence of NH3.

3.2.2. Effect of Ammonia on Surface Deposits

To investigate the mechanism of NH3 addition on the sulfur resistance of CeNbTi catalyst, several experiments were performed to characterize structural changes after the durability tests. No obvious differences were observed in the XRD patterns (Figure S2) of the fresh and poisoned catalysts. Only the characteristic peaks of anatase TiO2 are observed, and those of ceria, niobia, or sulfates do not appear due to their low crystallinity. In order to further identify the deposits on the poisoned catalyst, another group of reference samples was prepared by treating CeNbTi catalyst with SO2 + O2 and NH3 in different sequences before TPD experiments. As shown in Figure 3, a short-time (30 min) pretreatment in SO2 + O2 resulted in similar TPD curves of NH3 and SO2 as those for a much longer period (6 h). The sulfite-related SO2 desorption peak of the former catalyst is even stronger than that in Figure 3b, indicating the transformation of sulfites to sulfates proceeded with time-on-stream. After that, NH3 alone was introduced to react with the sulfates/sulfites on the SO2 + O2 pretreated catalyst and then the TPD measurement was performed. Here, the sulfite-related peak almost diminishes due to the formation and following decomposition of ammonium sulfite. As shown by the TG/DSC curves in Figure S3, ammonium sulfite decomposes easily even at temperatures below 120 °C. It implies that once it was formed by a reaction of metal sulfites with the introduced ammonia, ammonia sulfite would decompose subsequently during the pretreatment. As a result, few metal sulfites exist on the catalyst after NH3 purging (Figure 3b). Meanwhile, a small NH3 desorption peak occurs at 400 °C (Figure 3a), assigned to the decomposition of ammonium sulfates and chemisorbed ammonia on the surface acid sites.
Then, a reverse sequence was adopted for purging gases. The catalyst was first pretreated in NH3 for 30 min, aiming to saturate NH3 adsorption, and then the feeding gas was switched to SO2 + O2. It is interesting to find that the sulfite-related SO2 peak significantly decreases in intensity and almost no NH3 desorbs. In this case, the adsorbed NH3 is firstly consumed by gaseous SO2 via forming ammonium sulfite and subsequent decomposition, as well as driven by competitive adsorption of SOx. Then, the continuous passing with SO2 + O2 produces ceria-based sulfite and sulfate species. It is noted that the sulfate-related SO2 desorption peak decreases slightly in intensity after purging with NH3 (Figure 3b). It demonstrates that NH3 can remove surface sulfites but release the deposited sulfates to a small extent within the limited time. Thus, the reduced sulfate deposits in the presence of NH3 (Figure 3b) should be ascribed to a great extent to other factors such as competitive adsorption between NH3 and SO2.

3.3. Infrared Studies

3.3.1. Surface Groups Identification

Ex situ IR spectra of the treated CeNbTi catalysts were compared with reference samples to determine surface deposit species. As shown in Figure 4a, four main bands are found in the spectrum of ammonium sulfite. The bands at 1183, 1027, and 930 cm−1 are assigned to sulfite species [29], and the one at 1502 cm−1 is assigned to NH4+ on Brønsted acid sites [30]. In the case of ammonium hydrogen sulfate (Figure 4b), the band at 1478 cm−1 is attributed to asymmetric bending vibrations of NH4+ and the bands at 1326 and 1260 cm−1 are assigned to the S=O vibration absorption singles of HSO4 [31]. For cerium sulfate (Figure 4c), the bands at 1367 and 1300 cm−1 are related to the asymmetric stretching frequencies of O=S=O species [32]. The band at 1660 cm−1 is assigned to O-H stretching modes from the reaction between SO2 and surface hydroxyl groups [33,34]. As for ammonium cerium sulfate, five characteristic bands are observed in Figure 4d. The distinct band at 1275 cm−1 with a shoulder at 1200 cm−1 is ascribed to bidentate sulfate species [35]. The band at 1067 cm−1 is attributed to the S-O v3 vibrations of Ce-bonded bidentate SO42− in C2v symmetry [36]. The bands at 1606 and 1468 cm−1 are assigned to the NH4+ on Brønsted acid sites [37].
The IR spectrum of the CeNbTi that was treated in a simple SCR condition is shown in Figure 4e as the background. The band at 1617 cm−1 is assigned to adsorb H2O from the reaction between SO2 and surface hydroxyl groups [33]. The band at 965 cm−1 is attributed to characteristic vibrations of Nb=O bond in surface NbOx species [23] and the one at 1128 cm−1 is due to modes of Nb-O-Nb stretching vibration [38,39]. The stretching modes of Ce-O-Ce bands generally results in a band at 729 cm−1 [34], which shifts to 782 cm−1 due to interacting with Nb2O5 [38]. As for the spectrum that was collected over the catalyst upon exposure to SO2 + O2 (Figure 4f), the sulfate-related bands can be observed. There are two broad bands ranging at 1370–1300 and 1275–1180 cm−1 that are ascribed to the asymmetric stretching frequencies of the O=S=O bond and S-O bond of bidentate sulfate species, respectively. SO2 + O2 + NH3 +NO was chosen as a representative of three NH3-containing poisoning atmospheres. As shown in Figure 4g, the spectrum of the obtained catalyst exhibits a characteristic band at 1446 cm−1. It is assigned to the asymmetric bending vibrations of ionic NH4+ that is bound to Brønsted acid sites that are provided by surface sulfates (i.e., NH4HSO4 and (NH4)4Ce(SO4)2). Considering the similar intensities of the 1478 and 1326 cm−1 bands for NH4HSO4 (Figure 4b), the extra intensified band at 1446 cm−1 may mean the formation of ammonium cerium sulfate to a great extent. Another evidence is the band at 1067 cm−1 that is attributed to Ce-bonded bidentate sulfates. The broad band at 1275–1180 cm−1 may come from HSO4−1, CeSO4, and (NH4)4Ce(SO4)4. Other adsorption bands of sulfates at 1381 and 1354 cm−1 are attributed to asymmetric stretching frequencies of O=S=O species that are related to the formation of Ce(SO4)2.

3.3.2. In Situ SO2 + O2 Adsorption

Figure 5a shows the DRIFT spectra of adsorbed species over CeNbTi catalyst in a SO2 + O2 flow at 100 °C as a function of time. As the signals that were obtained by performing the DRIFTS experiments at 300 °C are much weaker, the adsorption of reactants and products becomes more significant by reducing the operating temperature to 100 °C. Hereby, the sulfur poisoning and ammonium protecting effects are amplified, although the specific mechanisms may be somewhat different at two temperatures. The band at 1268 cm−1 is assigned to bidentate sulfates [35], while several small features at 1180, 1024, and 925 cm−1 are attributed to sulfite species that are associated with different adsorption sites [33]. The band at 1086 cm−1 could be assigned to v3 vibrations of bidentate SO42− in C2v symmetry [36], while those at 1383 and 1300 cm−1 could be attributed to the asymmetric stretching frequencies of O=S=O species for sulfates that were adsorbed on metal oxides [32]. The band at 1430 cm−1 is attributed to the formed S−OH from the reaction between SO2 and metal cations [18]. The band at 1630 cm−1 is assigned to O-H stretching modes from the reaction between SO2 and surface hydroxyl groups [33,34]. All these bands increase sharply in intensity with time, indicating the deposition of a large amount of sulfite/sulfate species on CeNbTi catalyst upon sulfur poisoning.

3.3.3. In Situ Surface Reaction between Ad-Species and NH3/SO2

As mentioned above, fewer sulfates and sulfites formed when SO2 poisoning tests were performed in the NH3-containing atmospheres. In order to understand how NH3 affects the adsorbed sulfate/sulfite species, DRIFTS was performed to characterize the evolution of surface species on the SO2 + O2 pretreated catalyst upon exposure to NH3 and the results are shown in Figure 5b. The sulfate bands (1428, 1383, 1300, 1268, and 1086 cm−1) and sulfite bands (1024 and 925 cm−1) were observed for the pretreated catalyst. After purging with NH3 for 30 min, a new band appears at 1676 cm−1, assigned to NH4+ species that are bound to Bronsted acid sites [38,40]. Meanwhile, one band that was assigned to sulfate species shifts from 1430 to 1452 cm−1. The regular IR spectrum of ammonium bisulfate and ammonium cerium sulfate shows a NH4+ characteristic band at 1478 (Figure 4b) and 1468 cm−1 (Figure 4c), respectively. Both the asymmetric bending vibrations of NH4+ and the formed S−OH from metal sulfates contribute to this band. Thus, it implies the generation of ammonium cerium sulfate and ammonium bisulfate after the introduction of NH3. Meanwhile, the metal sulfate bands (1383 and 1300 cm−1) disappear finally, suggesting the transformation of Ce(SO4)2 to (NH4)4Ce(SO4)4 since NH4+ bonding with SO42− consumes O=S=O in Ce(SO4)2 [41], and the latter species decomposes subsequently (Figure S4).
A reverse DRIFT spectroscopy was also performed to explore the evolution of surface species on the NH3 pretreated catalyst upon exposure to SO2 + O2. As shown in Figure 4c, the Brønsted acid site-related bands (1660 and 1451 cm−1) and Lewis acid site-related bands (1600, 1210, and 1024 cm−1) are observed initially for the pretreated catalyst [10,34,42,43]. After the introduction of SO2, the sulfate-related bands (1434, 1266, and 1068 cm−1) form with time-on-stream. Importantly, the bands at 1383 and 1300 cm−1 that were attributed to CeSO4 can hardly be observed, illustrating that introduced SOx prefers to react with adsorbed NH3 to form NH4HSO4 and (NH4)4Ce(SO4)4. Meanwhile, the sulfite-related bands at 1022 and 925 cm−1 appear slowly. By comparison with the case of the as-received catalyst (Figure 5c), it indicates that the adsorbed ammonia reduces the formation of metal sulfates effectively. That is, the competitive adsorption between NH3 and SO2 would improve the sulfur resistance of the catalyst greatly.
Furthermore, in order to judge the discrepancy of NH3 and SO2 adsorption capacity over the CeNbTi catalyst, a competitive adsorption test was performed in a gas stream containing 200 ppm NH3, 200 ppm SO2, and 5% O2. At the initial stage (1 min) in Figure 5d, NH3 is preferentially adsorbed on Brønsted acid sites with the symmetric bending vibrations of NH4+ at 1618 cm−1, and then (2 min) the signal of S-O bond in SO42− appears at 1300 cm−1. After that (3 min), overlapped signals of the asymmetric bending vibrations of adsorbed NH4+ and the S−OH vibrations of SO42− appear at 1430 cm−1. The vibration signals of NH4+ and SO42− show a blue shift to a higher wavenumber and a red shift to lower wavenumber, respectively. This was derived from more NH3 adsorption forming NH4+[NH3]n structure [44], O-H from the reaction between SO2 and surface hydroxyl groups, and SO42− influenced by ammonia. These results demonstrate that the ammonia co-adsorbs on the CeNbTi catalyst with SO2 and could promote metal sulfates species transforming to ammonium sulfate-type and ammonium metal sulfate-type species.

4. Discussion

Intrinsically, NH3 as an alkaline gas can facilitate SO2 adsorption on ceria-based catalysts by the formation of ammonium sulfate-type species. Figure 6 shows the evolution of the typical sulfate bands (1300–1256 cm−1) and sulfite band (925 cm−1) over the CeNbTi catalyst that was treated by different poisoning procedures that were derived from Figure 5a,c,d, respectively. The peak areas of these bands were integrated and drawn as a function of time. In a simplified atmosphere (SO2 + O2), the sulfate bands increase in intensity continuously (Figure 6a), with cerium sulfate as the main product (Equation (2)), proven by the infrared spectra in Figure 4f and Figure 5a. It also correlates with the TG results (Table 1) that after the treatment in SO2+O2 for 6 h, about 13% of CeO2 transforms to cerium sulfate by assuming cerium sulfate as the dominating metal sulfate. Similarly, the sulfite band also obviously increases in intensity (Figure 6b), with the transformation of about 3% of CeO2 to cerium sulfites.
Significantly, the formation of sulfates/sulfites is inhibited on the NH3 pretreated catalyst and reaches a quasi-equilibrium quickly. One possible elimination route of sulfites by ammonia is the reaction between the pre-absorbed NH3 and sulfur dioxide to form ammonia sulfite (Equation (3)) which can decompose feasibly at much lower temperatures (Figure S3). When NH3, SO2, and O2 co-purge over the catalyst, almost no sulfite band is observed.
Meanwhile, the deposition amount of the sulfates decreases sharply on the NH3 pretreated catalyst. Ammonium bisulfate is more likely to form on niobium-modified catalysts [18], which can protect ceria from sulfur poisoning. As shown by the DRIFTS result of SO2 + O2 adsorption (Figure 5c), typical signals of HSO42− (1268 cm−1) and NH4+ that are adsorbed on Brønsted acid sites that are provided by surface sulfates (1434 cm−1) are observed. Therefore, it is reasonable that NH4HSO4 deposits on the pretreated catalyst according to Equation (4).
Previous studies [21,45] discovered that the decomposition of NH4HSO4 occurs at the temperature range of 320–450 °C and NH4HSO4 can react with ceria to form ammonium cerium sulfate at lower temperatures according to Equation (5). Jangjou [14] found that the ammonia could cause the transformation from CuxSOy to (NH4)xCuySOz at 210 °C on the Cu-SSZ-13 catalyst. Similarly, it can be seen in Figure 5b that (NH4)4Ce(SO4)4 species forms on the sulfated CeNbTi catalyst via Equation (6) upon NH3 purging. Compared with Ce(SO4)2, this material can serve as an active species to a certain extent via the redox cycle between (NH4)4Ce(SO4)4 and NH4Ce(SO4)2 (Figure S5). The competitive adsorption of NH3 and SO2 on the catalyst further reduces the surface sulfate deposition. NH3 adsorbs preferentially on the catalyst (Figure 5d), and this preferential occupation of surface active sites by NH3 reduces SOx adsorption. As listed in Table 1, it can be calculated that about 4–5% of CeO2 is deactivated by assuming (NH4)4Ce(SO4)4 as the distinct deposit when treating the catalyst in the SO2 + O2 + NH3 + NO atmosphere.
CeO2 + 2SO2 + O2 → Ce(SO4)2
2 NH 3 +   SO 2 + H 2 O   ( NH 4 ) 2 SO 3
2NH3 + 2SO2 + O2 + 2H2O → 2NH4HSO4
4NH4HSO4 + CeO2 → (NH4)4Ce(SO4)4 + 2H2O
Ce(SO4)2 + 4NH3 + 2SO2 + 2H2O + O2 → (NH4)4Ce(SO4)4
The degradation degree of CeNbTi catalyst by SO2 poisoning under SCR conditions (NH3 + NO + O2) is much slighter than SO2 + O2 exposure alone. Based on the above results and analysis, ammonia inhibits sulfur poisoning in NH3-SCR to a great extent via two main ways. On the one hand, ammonia would react with sulfites and form ammonium-sulfite which can subsequently decompose. Consequently, there is much less sulfite residue on the catalyst in the presence of ammonia. On the other hand, ammonia can pre-occupy surface adsorptions sites separating SO2 and active sites to reduce the sulfation of active metal. Besides, ammonia can transform cerium sulfate to ammonium cerium sulfate with relatively higher reactivity than cerium sulfate. The corresponding SO2 deactivation mechanisms in the presence and absence of ammonia are presented primarily in Figure 7. The present work discloses the inhibition effect of NH3 on sulfur poisoning of CeNbTi catalyst, which provides a comprehensive mechanistic insight and enables the development of ceria-based oxide catalysts for nitrogen oxide emission control in stationary sources. It should be noted that the above findings are based on the TPD and DRIFTS results, in which the suggestion of surface species still suffers from some uncertainty and remains to be further validated.

5. Conclusions

The deactivation of CeNbTi catalysts by exposure to SO2 in different atmospheres was investigated. The TPD and DRIFTS techniques were employed to characterize different species that were deposited on the catalyst and investigate their formation/decomposition/transformation mechanisms. Several major conclusions were drawn as follows:
(1)
After exposure to SO2 + O2 for 6 h, the NOx conversion at 300 °C of the catalyst decreases from almost 100% to ca. 60%, while the presence of NH3 in the poisoning atmospheres can reserve the catalyst with about 90% NH3-SCR activity. Additionally, humid conditions do not result in any obvious changes in the deactivation degree of sulfur poisoning with or without NH3.
(2)
The types and amounts of sulfates/sulfites depend importantly on the sulfur poisoning atmospheres. When the poisoning is performed in the presence of NH3, the total amount of sulfites and sulfates that are deposited on the catalyst is reduced by 44% compared with that in absence of ammonia. Competitive adsorption between NH3 and SO2 is suggested to be one of the dominating factors for the decreased surface deposits. The pre-occupying NH3 is confirmed to protect ceria active sites from reacting with SO2.
(3)
With the introduction of ammonia, not only the amounts of sulfates/sulfites decrease significantly, but also the types of sulfates change a lot from metal sulfates to ammonium sulfates. The ready decomposition behaviors of NH4HSO4 (and (NH4)2SO4), as well as the transformation of cerium sulfates with more sulfate radicals that are bonded per cerium atom, can also transform to cerium ammonium sulfates which facilitate the decrease in sulfate deposits. Compared with inert metal sulfates, cerium ammonium sulfates even contribute to NH3-SCR reaction with the redox cycle between Ce3+ and Ce4+ in the ammonium sulfates. Additionally, metal sulfites can be partially converted into ammonium sulfite in a NH3-containing atmosphere which is easy to decompose even at low temperatures, resulting in fewer sulfite deposits.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12111430/s1, Figure S1: (a) NOx conversions and (b) N2 selectivities of CeNbTi catalysts after the treatments in different atmospheres at 300 ºC for 6 h. Reaction conditions: 500 ppm NO, 500 ppm NH3, 5% O2, 5% H2O, N2 in balance, and GHSV = 80,000 h−1; Figure S2: XRD patterns of (a) CeNbTi catalyst and those after poisoning in (b) SO2 + O2, (c) SO2 + O2 + H2O and (d) SO2 + O2 + NH3 + NO + H2O; Figure S3: TG and DSC curves of (NH4)2SO3 in N2; Figure S4: The outlet gas concentrations after introducing NH3 to Ce(SO4)2 at 300 °C; Figure S5: NOx conversion over NbTi mixed oxides and those impregnated with (NH4)4Ce(SO4)4, Ce(SO4)2 and (NH4)2SO3. Reaction conditions: 500 ppm NO, 500 ppm NH3, 5% O2, 5% H2O, N2 in balance, and GHSV = 80,000 h−1. References [46,47,48] are cited in supplementary materials.

Author Contributions

Conceptualization, Y.G., L.C. and X.W.; Data curation, Y.G., L.C. and Z.M.; Formal analysis, Y.G. and L.C.; Funding acquisition, X.W. and D.W.; Investigation, Y.G. and L.C.; Methodology, L.C. and X.W.; Project administration, D.W.; Resources, Z.M., D.W. and B.W.; Supervision, D.W. and B.W.; Validation, X.Z.; Writing—original draft, Y.G. and L.C.; Writing—review and editing, Y.G., L.C., X.W., R.R. and Z.S. All authors have read and agreed to the published version of the manuscript.

Funding

The National Key R&D Program of China (No. 2017YFC0211202) and the Key Laboratory of Advanced Materials of Ministry of Education (No. 2016AML01).

Data Availability Statement

Data available upon request from the corresponding author.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work that was reported in this paper.

References

  1. Kwon, D.W.; Park, K.H.; Ha, H.P.; Hong, S.C. The role of molybdenum on the enhanced performance and SO2 resistance of V/Mo-Ti catalysts for NH3-SCR. Appl. Surf. Sci. 2019, 481, 1167–1177. [Google Scholar] [CrossRef]
  2. Han, L.P.; Cai, S.X.; Gao, M.; Hasegawa, J.Y.; Wang, P.L.; Zhang, J.P.; Shi, L.Y.; Zhang, D.S. 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]
  3. Chen, L.; Si, Z.C.; Wu, X.D.; Weng, D.; Ran, R.; Yu, J. Rare earth containing catalysts for selective catalytic reduction of NOx with ammonia: A Review. J. Rare Earth 2014, 32, 907–917. [Google Scholar] [CrossRef]
  4. Mosrati, J.; Atia, H.; Eckelt, R.; Lund, H.; Agostini, G.; Bentrup, U.; Rockstroh, N.; Keller, S.; Armbruster, U.; Mhamdi, M. Nb-modified Ce/Ti oxide catalyst for the selective catalytic reduction of NO with NH3 at low temperature. Catalysts 2018, 8, 175. [Google Scholar] [CrossRef] [Green Version]
  5. Jiang, Y.; Bao, C.; Liu, S.; Liang, G.; Lu, M.; Lai, C.; Shi, W.; Ma, S. Enhanced activity of Nb-modified CeO2/TiO2 catalyst for the selective catalytic reduction of NO with NH3. Aerosol Air Qual. Res. 2018, 18, 2121–2130. [Google Scholar] [CrossRef] [Green Version]
  6. Liu, K.; He, H.; Yu, Y.; Yan, Z.; Yang, W.; Shan, W. Quantitative study of the NH3-SCR pathway and the active site distribution over CeWOx at low temperatures. J. Catal. 2019, 369, 372–381. [Google Scholar] [CrossRef]
  7. Liu, J.J.; He, G.Z.; Shan, W.P.; Yu, Y.B.; Huo, Y.L.; Zhang, Y.; Wang, M.; Yu, R.; Liu, S.S.; He, H. Introducing tin to develop ternary metal oxides with excellent hydrothermal stability for NH3 selective catalytic reduction of NOx. Appl. Catal. B Environ. 2021, 291, 120125. [Google Scholar] [CrossRef]
  8. Shan, W.P.; Yu, Y.B.; Zhang, Y.; He, G.Z.; Peng, Y.; Li, J.H.; He, H. Theory and practice of metal oxide catalyst design for the selective catalytic reduction of NOx with NH3. Catal. Today 2021, 375, 292–301. [Google Scholar] [CrossRef]
  9. Zhang, W.; Liu, G.; Jiang, J.; Tan, Y.; Wang, Q.; Gong, C.; Shen, D.; Wu, C. Temperature sensitivity of the selective catalytic reduction (SCR) performance of Ce–TiO2 in the presence of SO2. Chemosphere 2020, 243, 125419. [Google Scholar] [CrossRef] [PubMed]
  10. Zhang, L.; Zou, W.; Ma, K.; Cao, Y.; Xiong, Y.; Wu, S.; Tang, C.; Gao, F.; Dong, L. Sulfated temperature effects on the catalytic activity of CeO2 in NH3-selective catalytic reduction conditions. J. Phys. Chem. C 2015, 119, 1155–1163. [Google Scholar] [CrossRef]
  11. Wijayanti, K.; Xie, K.; Kumar, A.; Kamasamudram, K.; Olsson, L. Effect of gas compositions on SO2 poisoning over Cu/SSZ-13 used for NH3-SCR. Appl. Catal. B Environ. 2017, 219, 142–154. [Google Scholar] [CrossRef]
  12. Wijayanti, K.; Leistner, K.; Chand, S.; Kumar, A.; Kamasamudram, K.; Currier, N.W.; Yezerets, A.; Olsson, L. Deactivation of Cu-SSZ-13 by SO2 exposure under SCR conditions. Catal. Sci. Technol. 2016, 6, 2565–2579. [Google Scholar] [CrossRef]
  13. Wang, C.; Wang, J.; Wang, J.; Yu, T.; Shen, M.; Wang, W.; Li, W. The effect of sulfate species on the activity of NH3-SCR over Cu/SAPO-34. Appl. Catal. B Environ. 2017, 204, 239–249. [Google Scholar] [CrossRef]
  14. Jangjou, Y.; Wang, D.; Kumar, A.; Li, J.; Epling, W.S. SO2 poisoning of the NH3-SCR reaction over Cu-SAPO-34: Effect of ammonium sulfate versus other S-containing species. ACS Catal. 2016, 6, 6612–6622. [Google Scholar] [CrossRef]
  15. Wang, Y.; Yi, W.; Yu, J.; Zeng, J.; Chang, H. Novel methods for assessing the so2 poisoning effect and thermal regeneration possibility of MOx–WO3/TiO2 (M = Fe, Mn, Cu, and V) Catalysts for NH3-SCR. Environ. Sci. Technol. 2020, 54, 12612–12620. [Google Scholar] [CrossRef] [PubMed]
  16. Lian, Z.; Liu, F.; Shan, W.; He, H. Improvement of Nb doping on SO2 resistance of VOx/CeO2 catalyst for the selective catalytic reduction of NOx with NH3. J. Phys. Chem. C 2017, 121, 7803–7809. [Google Scholar] [CrossRef]
  17. An, D.; Yang, S.; Zou, W.; Sun, J.; Tan, W.; Ji, J.; Tong, Q.; Sun, C.; Li, D.; Dong, L. Unraveling the so2 poisoning effect over the lifetime of MeOx (Me = Ce, Fe, Mn) catalysts in low-temperature NH3-SCR: Interaction of reaction atmosphere with surface species. J. Phys. Chem. C 2022, 126, 12168–12177. [Google Scholar] [CrossRef]
  18. Ma, Y.; Cheng, S.; Wu, X.; Shi, Y.; Cao, L.; Liu, L.; Ran, R.; Si, Z.; Liu, J.; Weng, D. Low-temperature solid-state ion-exchange method for preparing Cu-SSZ-13 selective catalytic reduction catalyst. ACS Catal. 2019, 8, 6962–6973. [Google Scholar] [CrossRef]
  19. Ma, Z.; Weng, D.; Wu, X.; Si, Z.; Wang, B. A novel Nb–Ce/WOx–TiO2 catalyst with high NH3-SCR activity and stability. Catal. Commun. 2012, 27, 97–100. [Google Scholar] [CrossRef]
  20. Song, L.; Chao, J.; Fang, Y.; He, H.; Li, J.; Qiu, W.; Zhang, G. Promotion of ceria for decomposition of ammonia bisulfate over V2O5-MoO3/TiO2 catalyst for selective catalytic reduction. Chem. Eng. J. 2016, 303, 275–281. [Google Scholar] [CrossRef]
  21. Ye, D.; Qu, R.; Zheng, C.; Cen, K.; Gao, X. Mechanistic investigation of enhanced reactivity of NH4HSO4 and NO on Nb-and Sb-doped VW/Ti SCR catalysts. Appl. Catal. A Gen. 2018, 549, 310–319. [Google Scholar] [CrossRef]
  22. Ye, D.; Qu, R.; Song, H.; Zheng, C.; Gao, X.; Luo, Z.; Ma, N.; Cen, K. Investigation of the promotion effect of WO₃ on the decomposition and reactivity of NH₄HSO₄ with NO on V₂O₅–WO₃/TiO₂ SCR catalysts. RSC Adv. 2016, 6, 55584–55592. [Google Scholar] [CrossRef]
  23. Leistner, K.; Mihai, O.; Wijayanti, K.; Kumar, A.; Kamasamudram, K.; Currier, N.W.; Yezerets, A.L.; Olsson, L. Comparison of Cu/BEA, Cu/SSZ-13 and Cu/SAPO-34 for ammonia-SCR reactions. Catal. Today 2015, 258, 49–55. [Google Scholar] [CrossRef]
  24. Udupa, M.R. Thermal decomposition of cerium (IV), cerium (III), chromium (III) and titanium (IV) sulphates. Thermochim. Acta 1982, 57, 377–381. [Google Scholar] [CrossRef]
  25. Youn, S.; Song, I.; Lee, H.; Cho, S.J.; Kim, D.H. Effect of pore structure of TiO2 on the SO2 poisoning over V2O5/TiO2 catalysts for selective catalytic reduction of NOx with NH3. Catal. Today 2018, 303, 19–24. [Google Scholar] [CrossRef]
  26. Ma, Z.; Wu, X.; Feng, Y.; Si, Z.; Weng, D.; Shi, L. Low-temperature SCR activity and SO2 deactivation mechanism of Ce-modified V2O5–WO3/TiO2 catalyst. Prog. Nat. Sci. Mater. 2015, 25, 342–352. [Google Scholar] [CrossRef] [Green Version]
  27. Li, Y.; Xiong, J.; Lin, Y.; Guo, J.; Zhu, T. Distribution of SO2 oxidation products in the SCR of NO over V2O5/TiO2 catalysts at different temperatures. Ind. Eng. Chem. Res. 2020, 59, 5177–5185. [Google Scholar] [CrossRef]
  28. Jangjou, Y.; Do, Q.; Gu, Y.; Lim, L.G.; Sun, H.; Wang, D.; Kumar, A.; Li, J.H.; Grabow, L.C.; Epling, W.S. Nature of Cu active centers in Cu-SSZ-13 and their responses to SO2 exposure. ACS Catal. 2018, 8, 1325–1337. [Google Scholar] [CrossRef]
  29. Shen, B.X.; Liu, T. Deactivation of MnOx-CeOx/ACF catalysts for low-temperature NH3-SCR in the presence of SO2. Acta Phys. Chim. Sin. 2010, 26, 3009–3016. [Google Scholar]
  30. Kwon, D.W.; Nam, K.B.; Hong, S.C. The role of ceria on the activity and SO2 resistance of catalysts for the selective catalytic reduction of NOx by NH3. Appl. Catal. B Environ. 2015, 166, 37–44. [Google Scholar] [CrossRef]
  31. Qing, M.; Lei, S.; Kong, F.; Liu, L.; Zhang, W.; Wang, L.; Guo, T.; Su, S.; Hu, S.; Wang, Y.; et al. Analysis of ammonium bisulfate/sulfate generation and deposition characteristics as the by-product of SCR in coal-fired flue gas. Fuel 2022, 313, 122790. [Google Scholar] [CrossRef]
  32. Liu, X.; Wang, P.; Shen, Y.; Zheng, L.; Han, L.; Deng, J.; Zhang, J.P.; Wang, A.Y.; Ren, W.; Gao, F.; et al. Boosting SO2-Resistant NOx Reduction by modulating electronic interaction of short-range Fe–O coordination over Fe2O3/TiO2 catalysts. Environ. Sci. Technol. 2022, 56, 11646–11656. [Google Scholar] [CrossRef]
  33. Xu, Z.; Impeng, S.; Jia, X.; Wang, F.; Shen, Y.; Wang, P.; Zhang, D. SO2-Tolerant catalytic reduction of NOx by confining active species in TiO2 nanotubes. Environ. Sci. Nano 2022, 9, 2121. [Google Scholar] [CrossRef]
  34. Jin, R.; Liu, Y.; Wang, Y.; Cen, W.; Wu, Z.; Wang, H.; Weng, X. The role of cerium in the improved SO2 tolerance for NO reduction with NH3 over Mn-Ce/TiO2 catalyst at low temperature. Appl. Catal. B Environ. 2014, 148, 582–588. [Google Scholar] [CrossRef]
  35. Chang, H.; Chen, X.; Li, J.; Ma, L.; Wang, C.; Liu, C.; Schwank, J.; Hao, J. Improvement of activity and SO2 tolerance of Sn-modified MnOx–CeO2 catalysts for NH3-SCR at low temperatures. Environ. Sci. Technol. 2013, 47, 5294–5301. [Google Scholar] [CrossRef]
  36. Ye, D.; Qu, R.; Song, H.; Gao, X.; Luo, Z.; Ni, M.; Cen, K. New insights into the various decomposition and reactivity behaviors of NH4HSO4 with NO on V2O5/TiO2 catalyst surfaces. Chem. Eng. J. 2016, 283, 846–854. [Google Scholar] [CrossRef]
  37. Gao, F.; Tang, X.; Yi, H.; Li, J.; Zhao, S.; Wang, J.; Chu, C.; Li, C. Promotional mechanisms of activity and SO2 tolerance of Co-or Ni-doped MnOx-CeO2 catalysts for SCR of NOx with NH3 at low temperature. Chem. Eng. J. 2017, 317, 20–31. [Google Scholar] [CrossRef]
  38. Ma, Z.; Wu, X.; Si, Z.; Weng, D.; Ma, J.; Xu, T. Impacts of niobia loading on active sites and surface acidity in NbOx/CeO2–ZrO2 NH3–SCR catalysts. Appl. Catal. B Environ. 2015, 179, 380–394. [Google Scholar] [CrossRef] [Green Version]
  39. Burcham, L.J.; Datka, J.; Wachs, I.E. In situ vibrational spectroscopy studies of supported niobium oxide catalysts. J. Phys. Chem. B. 1999, 103, 6015–6024. [Google Scholar] [CrossRef]
  40. Cao, L.; Chen, L.; Wu, X.; Ran, R.; Xu, T.; Chen, Z.; Weng, D. TRA and DRIFTS studies of the fast SCR reaction over CeO2/TiO2 catalyst at low temperatures. Appl. Catal. A Gen. 2018, 557, 46–54. [Google Scholar] [CrossRef]
  41. Zhang, W.; Liu, G.; Jiang, J.; Tan, Y.; Wang, Q.; Gong, C.; Shen, D.; Wu, C. Sulfation effect of Ce/TiO2 catalyst for the selective catalytic reduction of NOx with NH3: Mechanism and kinetic studies. RSC Adv. 2019, 9, 32110. [Google Scholar] [CrossRef] [PubMed]
  42. Shan, W.; Liu, F.; He, H.; Shi, X.; Zhang, C. A superior Ce-W-Ti mixed oxide catalyst for the selective catalytic reduction of NOx with NH3. Appl. Catal. B Environ. 2012, 115, 100–106. [Google Scholar] [CrossRef]
  43. Chen, L.; Si, Z.; Wu, X.; Weng, D. DRIFT study of CuO–CeO2–TiO2 mixed oxides for NOx reduction with NH3 at low temperatures. ACS Appl. Mater. Interfaces 2014, 6, 8134–8145. [Google Scholar] [CrossRef] [PubMed]
  44. Lonyi, F.; Valyon, J. A TPD and IR study of the surface species formed from ammonia on zeolite H-ZSM-5, H-mordenite and H-beta. Thermochim. Acta 2001, 373, 53–57. [Google Scholar] [CrossRef]
  45. Tong, T.; Chen, J.; Xiong, S.; Yang, W.; Yang, Q.; Yang, L.; Li, J. Vanadium-density-dependent thermal decomposition of NH4HSO4 on V2O5/TiO2 SCR catalysts. Catal. Sci. Technol. 2019, 9, 3779–3787. [Google Scholar] [CrossRef]
  46. Casari, B.M.; Langer, V. Two Ce(SO4)2·4H2O polymorphs: Crystal structure and thermal behavior. J. Solid State Chem. 2007, 180, 1616–1622. [Google Scholar] [CrossRef]
  47. Yang, Y.; Yang, R. Study on the thermal decomposition of tetrahydrated cerie sulphate. Thermochim. Acta 1992, 202, 301–306. [Google Scholar] [CrossRef]
  48. Singh Mudher, K.D.; Keskar, M.; Venugopal, V. Solid state reactions of CeO2, ThO2 and PuO2 with ammonium sulphate. J. Nucl. Mater. 1999, 265, 146–153. [Google Scholar] [CrossRef]
Figure 1. SCR activity of the treated CeNbTi catalysts. Reaction conditions: 500 ppm NO, 500 ppm NH3, 5% O2, 5% H2O, N2 in balance, 300 °C, and GHSV = 80,000 h−1.
Figure 1. SCR activity of the treated CeNbTi catalysts. Reaction conditions: 500 ppm NO, 500 ppm NH3, 5% O2, 5% H2O, N2 in balance, 300 °C, and GHSV = 80,000 h−1.
Catalysts 12 01430 g001
Figure 2. TPD curves of (a,c) NH3 and (b,d) SO2 over the treated (a,b) and deposited (c,d) catalysts.
Figure 2. TPD curves of (a,c) NH3 and (b,d) SO2 over the treated (a,b) and deposited (c,d) catalysts.
Catalysts 12 01430 g002
Figure 3. TPD curves of (a) NH3 and (b) SO2 over the catalyst that was pretreated in SO2 + O2 and NH3 in different sequences.
Figure 3. TPD curves of (a) NH3 and (b) SO2 over the catalyst that was pretreated in SO2 + O2 and NH3 in different sequences.
Catalysts 12 01430 g003
Figure 4. IR spectra of (ad) reference samples and (eg) treated CeNbTi catalysts.
Figure 4. IR spectra of (ad) reference samples and (eg) treated CeNbTi catalysts.
Catalysts 12 01430 g004
Figure 5. DRIFT spectra of (a) the catalyst exposed to SO2 + O2 as a function of time, (b) the SO2 + O2 pretreated catalyst exposed to NH3, (c) the NH3 pretreated catalyst exposed to SO2 + O2, and (d) the catalyst exposed to NH3 + SO2 +O2 at 100 °C as a function of time.
Figure 5. DRIFT spectra of (a) the catalyst exposed to SO2 + O2 as a function of time, (b) the SO2 + O2 pretreated catalyst exposed to NH3, (c) the NH3 pretreated catalyst exposed to SO2 + O2, and (d) the catalyst exposed to NH3 + SO2 +O2 at 100 °C as a function of time.
Catalysts 12 01430 g005
Figure 6. Evolution of (a) typical sulfate bands (1300–1256 cm−1) and (b) sulfite band (925 cm−1) as a function of time over CeNbTi catalysts that were treated with different poisoning procedures.
Figure 6. Evolution of (a) typical sulfate bands (1300–1256 cm−1) and (b) sulfite band (925 cm−1) as a function of time over CeNbTi catalysts that were treated with different poisoning procedures.
Catalysts 12 01430 g006
Figure 7. Schematic SO2 deactivation mechanisms over CeNbTi in the presence and absence of NH3.
Figure 7. Schematic SO2 deactivation mechanisms over CeNbTi in the presence and absence of NH3.
Catalysts 12 01430 g007
Table 1. Amounts of NH3 and SO2 that were released from the treated catalysts during the TPD tests.
Table 1. Amounts of NH3 and SO2 that were released from the treated catalysts during the TPD tests.
SampleSBET
(m2/g)
NH3
(μmol/g)
SO2 (μmol/g)
Peak Temperature (°C)Deposit *
565685730900
NH3 + NO + O211316009834-
SO2 + O2850351859834220
SO2 + O2 + H2O870352129838247
SO2 + O2 + NH3 + NO9335171319837148
SO2 + O2 +NH3 + NO + H2O8633151299832144
SO2 + O2 + NH38934171259832142
* Estimated by the sum of peaks at 565 and 685 °C.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gao, Y.; Cao, L.; Wu, X.; Zhang, X.; Ma, Z.; Ran, R.; Si, Z.; Weng, D.; Wang, B. Protection Effect of Ammonia on CeNbTi NH3-SCR Catalyst from SO2 Poisoning. Catalysts 2022, 12, 1430. https://doi.org/10.3390/catal12111430

AMA Style

Gao Y, Cao L, Wu X, Zhang X, Ma Z, Ran R, Si Z, Weng D, Wang B. Protection Effect of Ammonia on CeNbTi NH3-SCR Catalyst from SO2 Poisoning. Catalysts. 2022; 12(11):1430. https://doi.org/10.3390/catal12111430

Chicago/Turabian Style

Gao, Yang, Li Cao, Xiaodong Wu, Xu Zhang, Ziran Ma, Rui Ran, Zhichun Si, Duan Weng, and Baodong Wang. 2022. "Protection Effect of Ammonia on CeNbTi NH3-SCR Catalyst from SO2 Poisoning" Catalysts 12, no. 11: 1430. https://doi.org/10.3390/catal12111430

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