Next Article in Journal
Progress in Nanoporous Templates: Beyond Anodic Aluminum Oxide and Towards Functional Complex Materials
Next Article in Special Issue
Mesoporous Palladium N,N’-Bis(3-Allylsalicylidene)o-Phenylenediamine-Methyl Acrylate Resins as Heterogeneous Catalysts for the Heck Coupling Reaction
Previous Article in Journal
Effects of Printing Parameters on the Fit of Implant-Supported 3D Printing Resin Prosthetics
Previous Article in Special Issue
Synthesis of Petal-Like MnO2 Nanosheets on Hollow Fe3O4 Nanospheres for Heterogeneous Photocatalysis of Biotreated Papermaking Effluent
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of SO2 on the Selective Catalytic Reduction of NOx over V2O5-CeO2/TiO2-ZrO2 Catalysts

1
Key Laboratory of Energy Thermal Conversion and Control of Ministry of Education, School of Energy and Environment, Southeast University, Nanjing 210096, China
2
State Power Environmental Protection Research Institute, Nanjing 210031, China
*
Authors to whom correspondence should be addressed.
Materials 2019, 12(16), 2534; https://doi.org/10.3390/ma12162534
Submission received: 23 June 2019 / Revised: 2 August 2019 / Accepted: 5 August 2019 / Published: 9 August 2019

Abstract

:
The effect of SO2 on the selective catalytic reduction of NOx by NH3 over V2O5-0.2CeO2/TiO2-ZrO2 catalysts was studied through catalytic activity tests and various characterization methods, like Brunner−Emmet−Teller (BET) surface measurement, X-ray diffraction (XRD), transmission electron microscopy (TEM), X-ray fluorescence (XRF), hydrogen temperature-programmed desorption (H2-TPR), X-ray photoelectron spectroscopy (XPS) and in situ diffused reflectance infrared Fourier transform spectroscopy (DRIFTS). The results showed that the catalyst exhibited superior SO2 resistance when the volume fraction of SO2 was below 0.02%. As the SO2 concentration further increased, the NOx conversion exhibited some degree of decline but could restore to the original level when stopping feeding SO2. The deactivation of the catalyst caused by water in the flue gas was reversible. However, when 10% H2O was introduced together with 0.06% SO2, the NOx conversion was rapidly reduced and became unrecoverable. Characterizations indicated that the specific surface area of the deactivated catalyst was significantly reduced and the redox ability was weakened, which was highly responsible for the decrease of the catalytic activity. XPS results showed that more Ce3+ was generated in the case of reacting with SO2. In situ DRIFTS results confirmed that the adsorption capacity of SO2 was enhanced obviously in the presence of O2, while the SO2 considerably refrained the adsorption of NH3. The adsorption of NOx was strengthened by SO2 to some extent. In addition, NH3 adsorption was improved after pre-adsorbed by SO2 + O2, indicating that the Ce3+ and more oxygen vacancy were produced.

1. Introduction

Selective catalytic reduction (SCR) catalysts are commonly severely deactivated by SO2, which is abundantly present in flue gas. There are two mechanisms that can explain the sulfur poisoning of catalysts. Firstly, SO2 reacts with NH3 and vapor in the oxygen atmosphere, producing sulfate species including ammonium sulfate and ammonium bisulfate. These sulfate substances can deposit on the catalytic surface and cause pore plugging. As a result, the specific surface area and pore volume decrease observably, ending up with the catalyst deactivation. Studies have shown that the thermal decomposition temperature range of ammonium sulfate and ammonium bisulfate are 213–308 °C and 308–419 °C [1], respectively. It is still hard for these sulfate species to decompose over a traditional V/TiO2 catalyst. In the second, SO2 can react with the active center atoms and produce metal sulfates, which will decrease catalyst activity [2,3]. The former deactivation is reversible and the catalysts can be reactivated by washing and high-temperature processing. However, the later deactivation is irreversible.
The mechanism of catalyst sulfur poisoning has been extensively studied. Wei et al. [4] claimed that SO2 significantly reduced the adsorption of NH3 on the Lewis acid sites. At the same time, SO2 would react with NH4+ to form NH4HSO3, thereby reducing NOx conversion. However, Jiang et al. [5] proposed that SO2 had little effect on the adsorption of NH3, and conversely promoted the formation of new Bronsted acid sites. The weakening of NO adsorption on the catalyst surface was the main cause of deactivation. With the deposition of sulfate on the catalyst surface, less NO participated in the SCR reaction, resulting in a decrease in NOx conversion. Similarly, Liu et al. [6] claimed that SO2 had a significant inhibitory effect on the reduction of NOx due to the deposition of sulfate substances. It hindered the adsorption of NO and the production of the intermediate ammonium nitrate, resulting in a decrease in catalyst activity. Moreover, Pan et al. [7] found that the main reason for MnOx/TiO2 catalysts deactivation was that the active center atom manganese was sulfated. Besides, the presence of SO2 caused ammonium sulfate to deposit on the catalyst, decreasing the adsorption of NO. Gu et al. [8] reported that the main reason for the deactivation of Ce/TiO2 catalyst was that SO2 could react with the catalyst to form high thermally stable Ce(SO4)2 and Ce2(SO4)3, and Xu et al. obtained similar results [9].
On the basis of the above literatures, it is found that there are still many controversies about the mechanism of catalyst sulfur poisoning. Moreover, there is no report found currently on the sulfur poisoning mechanism of V2O5-CeO2/TiO2-ZrO2 catalysts, which were reported to be an excellent SCR catalyst with a wide temperature range [10,11,12]. Our previous studies indicated that the V2O5-0.2CeO2/TiO2-ZrO2 catalyst exhibited superior catalytic performance as well as high tolerance of SO2 and H2O.In this paper, the catalytic activity of V2O5-0.2CeO2/TiO2-ZrO2 in the presence of different SO2 content was further studied. Various characterization methods such as the Brunner−Emmet−Teller (BET) surface measurement, X-ray diffraction (XRD), transmission electron microscopy (TEM), X-ray fluorescence (XRF), hydrogen temperature-programmed desorption (H2-TPR), X-ray photoelectron spectroscopy (XPS) and in situ diffused reflectance infrared Fourier transform spectroscopy (DRIFTS) were employed to study the poisoning mechanism from a microcosmic aspect.

2. Materials and Methods

2.1. Catalyst Preparation

The Ti–Zr carrier was prepared by a coprecipitation method with the molar ratio of Ti:Zr = 1:1. An equal amount of TiCl4 solution and ZrOCl2 8H2O were dissolved in deionized water and stirred constantly. With stirring, NH3·H2O was slowly added until the pH reached 10. The obtained solution was aged at room temperature for 24 h. Then, the precipitate was washed with deionized water until the supernatant contained no Cl. Finally, the sample was dried at 110 °C for 12 h and then calcined at 450 °C for 4 h in a muffle furnace.
A step-by-step impregnation method was used to prepare V2O5-0.2CeO2/TiO2–ZrO2. A certain amount of CeNO3·6H2O and Ti–Zr powder were added to deionized water. The obtained suspension was stirred at room temperature for 2 h, followed by stirred at 85 °C for 4 h. After dried at 110 °C for 12 h and calcined in a muffle furnace at 450 °C for 4 h, the Ce/Ti–Zr sample was obtained. Thereafter, the resulting Ce/Ti–Zr was impregnated with NH4VO3 solution in the same manner. The obtained sample was denoted as V-0.2Ce/Ti–Zr, where the content of V2O5 loading was 1 wt. % and the molar ratio of Ce to Ti–Zr support = 0.2.
For short, the V-0.2Ce/Ti–Zr catalyst after reaction with 0.06% SO2 for 2 h was denoted as S-V-0.2Ce/Ti–Zr, while the catalyst after reaction with 10 vol% H2O and 0.06% SO2 for 2 h was denoted as HS-V-0.2Ce/Ti–Zr, respectively.

2.2. Activity Measurements

The SCR activity measurement of catalysts was carried out in a fixed-bed reactor with the inner diameter of 7 mm. The 0.3 g catalyst (40–60 mesh) was placed in the reactor with a gas hourly space velocity (GHSV) of 20000 h−1. Typically, the total gas flow was 100 mL/ min. The simulated gas was composed of 0.08% NO, 0.08% NH3, 5% O2, 5%/10% H2O (when used), 0.02%/0.04%/0.06% SO2 (when used) and N2 as the balanced gas. The NO, NO2 and NOx concentration were persistently monitored with a flue gas analyzer (Testo350-XL).

2.3. Catalyst Characterization

The BET was measured using a specific surface area and pore size analyzer V-Sorb 2800P (Beijing Gold APP, Beijing, China). The sample was pretreated under vacuum at 250 °C for 5 h, and the adsorbate was high purity nitrogen.
XRD patterns were carried out for phase analysis via a SmartLab™ X-ray diffractometer (Rigaku, Tokyo, Japan). Cu target acted as the X-ray source.
The morphology of the catalyst was determined by TEM (Thermo Fisher Scientific, Waltham, Massachusetts, America). The catalyst sample was dispersed in an ethanol solution after thoroughly ground, followed by shaken under ultrasonic waves for 15 min.
XRF spectrometer (Thermo Fisher Scientific, Waltham, Massachusetts, MA, America) was used to analyze the content of component in the catalyst sample.
H2-TPR was carried out in a quartz U-tube reactor connected to a thermal conduction detector (TCD) using an H2–Ar mixture (10% H2 by volume) as reductant (Finetec Instruments, Hangzhou, Zhejiang, China). The temperature range during the test was for 25 °C to 800 °C with a heating rate of 10 °C /min.
XPS analysis was performed on a PHI Quantera II system (Ulvac-PHI, Chigasaki, Kanagawa Prefecture, Japan). The binding energies were referenced to the C 1 s line at 284.8 eV from adventitious carbon.
In situ DRIFTS studies were performed on a Nicolet 6700 spectrometer (Thermo Fisher Scientific, Waltham, Massachusetts, MA, USA). The scanning wave number ranged from 400 cm−1 to 4000 cm−1. Before the test, the sample was pretreated with N2 at 400 °C for 1 h to remove impurities. The background at a certain temperature was collected during the cooling process.

3. Results and Discussion

3.1. Effects of SO2 and H2O on Catalyst Activity

Figure 1 showed the catalytic activity results with different concentration SO2 over the V-0.2Ce/Ti–Zr catalyst at 250 °C. In the absence of SO2, NOx conversion of the V-0.2Ce/Ti–Zr catalyst was approximately 90%, indicating that the catalyst had high activity at the low temperature. After 0.02% SO2 was introduced, NOx conversion remained stable. However, NOx conversion decreased rapidly to 71% and 65% at 20 min respectively when 0.04% and 0.06% SO2 were added. Furthermore, it could be found that the catalytic activity could recover after stopping SO2. It was inferred that SO2 and SO3 reacted with NH3 in the early stage, leading to the decrease of NH3 as well as the catalytic performance. As the reaction went on, SO2 and SO3 reacted with CeO2 and produced Ce3+ in the presence of excess oxygen, which could strengthen the acid sites on the catalyst and increased NOx conversion [13].
The effect of H2O on the catalytic activity of the catalyst was investigated and the results were shown in Figure 2. It was found that the activity of the catalyst was rapidly decreased with the introduction of H2O. However, after the H2O was stopped, the activity gradually recovered almost to the original level, thus indicating that the effect of H2O on the catalyst was reversible.
The activity of the V-0.2Ce/Ti–Zr catalyst was tested at 250 °C in the presence of 10% H2O and different concentrations of SO2, and the results were shown in Figure 3. After the introduction of SO2 and H2O, NOx conversion decreased rapidly to less than 30%, which was much lower than that in the presence of SO2 or H2O alone. After SO2 and H2O were removed, NOx conversion increased to some extent but could not be restored to the initial level, indicating an irreversible deactivation occurred. Studies showed that SO2 would react with NH3 to form ammonium sulfates, and block active sites on the surface of the catalyst.
Furthermore, the performance of the HS-V-0.2Ce/Ti–Zr catalyst at different temperatures was also tested, and the results were shown in Figure 4. It could be found that the mid-temperature (200–300 °C) activity of the HS-V-0.2Ce/Ti–Zr catalyst was significantly decreased, while it at a higher temperature (>300 °C) was obviously enhanced. One literature indicated that the formation of Ce(SO4)2 led to a decrease in the activity of the catalyst at moderate temperatures [14]. At the same time, Xu et al. [15] pointed out that the sulphate was produced during the poisoning process, which had a certain activity at high temperatures and could improve the high temperature activity of the catalyst.

3.2. Physico-Chemical Characterization of Catalysts

3.2.1. BET Analysis

The specific surface area of the catalysts before and after being poisoned by SO2 was illustrated in Table 1. Compared with the fresh sample, the BET surface area of the S-V-0.2Ce/Ti–Zr catalyst decreased from 54.45 m2/g to 25.07 m2/g, and that of the HS-V-0.2Ce/Ti–Zr catalyst further reduced to 15.27 m2/g. While with comparison to the fresh counterpart, there was no apparent change on the pore volume. The specific surface area of the catalyst was related to the adsorption capacity of NH3 during the SCR reaction, thereby further affecting the denitrification activity.
In order to further investigate the reason for the significant decrease in the BET specific surface area, the pore size distribution of the catalyst was mapped. As shown in Figure 5, the pore size range of the fresh catalyst was mainly concentrated at 2 nm to 10 nm, indicating that the mesopores contributed the most to the specific surface area of the catalyst. For the catalyst after reaction with O2, the number of macropores and mesopores (in the range of 5.3 nm to 50 nm) increased, and the increase of pore diameter in HS- V-0.2Ce/Ti–Zr catalyst was more pronounced. It was found that the mesopores (mainly 2 nm to 10 nm) contributed the most to the specific surface area of the catalyst. Therefore, it was presumed that the number of mesopores determined the NH3-SCR activity of the V-0.2Ce/Ti–Zr catalyst. After the reaction in presence of H2O and SO2, mesopores in the range of 5–50 nm contributed greatly to the specific surface area of the catalyst, while the mesopores and micropores with pore diameters less than 5 nm gradually decreased. Based on the above analysis, it was speculated that in the presence of SO2 and H2O, (NH4)2SO4 and Ce(SO4)2 formed during the reaction were adsorbed on the catalyst, and blocked 2 to 5 nm mesopores and micropores around the active component of the catalyst. These caused a decrease in specific surface area of the catalyst, further inhibiting catalytic activity.

3.2.2. XRD Analysis

To further study the microstructure of the poisoned V-0.2Ce/Ti–Zr, an XRD analysis was performed. As shown in Figure 6, catalysts before and after being poisoned by SO2 showed similar spectra. According to our previous research [12], these diffraction peaks mainly corresponded to ZrTiO2, ZrO2, V2O5, CeO2 and TiO2. Diffraction peaks of substances such as (NH4)2SO4 or Ce(SO4)2 were not detected. The results illustrated that no sulphate with good crystal form was formed, or the amount of sulphate formed on the surface was small and highly dispersed.

3.2.3. TEM and XRF Analysis

TEM was carried out to study the morphology changes of catalysts poisoned by H2O and SO2. As shown in Figure 7, the surface of the fresh catalyst was uniform and had a good dispersion. While for the poisoned one, it was clearly observed that the surface was covered with some substances, and the particles were agglomerate.
Element content of the catalyst was tested and the results were shown in Table 2. No S element was detected in the S-V-0.2Ce/Ti–Zr catalyst, while a trace amount of the S element was detected in the HS-V-0.2Ce/Ti–Zr catalyst. The results revealed that when SO2 was separately introduced, substantially no S element was present on the catalyst surface, or its content was extremely low. However, once H2O was introduced together, SO2 was more likely to participate in the reaction, and was present on the surface in the form of ammonium sulfate or Ce(SO4)2.

3.2.4. H2-TPR Analysis

TPR profiles for catalysts before and after poisoned by SO2 were presented in Figure 8. For the fresh catalyst, five reduction peaks were observed at 343 °C, 418 °C, 507 °C, 580 °C and 732 °C, respectively. Studies have shown that the low temperature reduction of V/TiO2 catalyst is mainly related to the monomer vanadium or highly dispersed vanadium species [16]. Held et al. proposed that the reduction peak at about 730 °C was mainly related to V2O4 [17]. Based upon our previous work, peaks centered at 343 °C and 580 °C were due to the reduction of vanadium from V5+ to V4+ and V4+ to V3+, respectively. The reduction peak centered at 418 °C was attributed to the surface (α) reduction processed of CeO2, and the subsurface layers and deeper regions of the catalyst nanoparticles were reduced at 507 °C.
It could be found that the redox ability of the S-V-0.2Ce/Ti–Zr catalyst was significantly weakened, according to the fact that several peaks disappeared (343 °C, 418 °C and 580 °C). After reaction with SO2 and H2O, the redox ability of the HS-V-0.2Ce/Ti–Zr catalyst was further attenuated, and only one reduction peak appeared at 458 °C was detected. Peaks centered at 495 °C and 458 °C were due to the reduction of Ce4+, while the peak at 653 °C was attributed to the reduction of V5+ to V3+ and the reduction of bulk phase Ce4+. It could be found that the disappearance of reduction peaks related with vanadium was an important cause of catalytic deactivation. For S-V-0.2Ce/Ti–Zr sample, the α-reduction peak of Ce4+ disappeared, which should be related to the reduction of Ce4+ to Ce3+. It was reported that the concentration of Ce3+ was used to assess the amount of oxygen formed. The oxygen vacancies were driven by the transition from Ce4+ to Ce3+, accelerating the transport of active oxygen species and facilitating the reaction vacancies, which could explain the phenomenon that the activity of the catalyst decreased first and then gradually increased. After being poisoned by SO2 and H2O, the intensity of the β reduction peak (458 °C) was significantly weakened, and the reduction peak of the vanadium oxide completely disappeared. It was speculated that the formed ammonium hydrogen sulfate was deposited on the surface of the catalyst or more stable Ce2(SO4)3 was formed, which hindered the conversion of vanadium active species and led to a reduction of the redox ability.

3.2.5. XPS Analysis

Figure 9a showed the XPS results of Ce 3d. Peaks u′′′, u′′, u and v′′′, v′′, v can be corresponded to Ce4+, while peaks u′ and v′ were assigned to Ce3+, respectively. It can be seen from Figure 9a that Ce in the catalyst mainly existed in the Ce4+ valence state. As shown in Table 3, after reacting with 0.06% SO2 for 2 h, the ratio of Ce3+/ (Ce4++Ce3+) in S-V-0.2Ce/Ti–Zr catalysts was increased to 25.15% compared with that in V-0.2Ce/Ti–Zr (24.94%), indicating more Ce3+ was generated. However, in the presence of SO2 and H2O, the ratio of Ce3+/(Ce4+ + Ce3+) was reduced again to 24.93%. Hence, it was deduced that the introduction of SO2 promoted the conversion of Ce4+ and Ce3+, while the formed ammonium hydrogen sulfate on the surface of the catalyst could block the path of mutual conversion between Ce4+ and Ce3+.
As shown in Figure 9b, the three peaks in O 1 s can be assigned to oxygen vacancy, adsorbed oxygen and oxygen lattice from higher binding energy to lower binding energy. The adsorbed oxygen and oxygen lattice were marked as Oβ and Oα, respectively [18,19]. The oxygen vacancy was formed due to the evolution of the oxygen lattice [20].
As can be seen in Table 3, the ratio of Oβ/(Oβ + Oα) was reduced significantly in the S-V-0.2Ce/Ti–Zr catalysts and even worse in the HS-V-0.2Ce/Ti–Zr catalysts. The existence of Oβ could promote the oxidation of NO to NO2, which can explain the decrease in activity of the catalysts poisoned by SO2.

3.3. In Situ DRIFTS Study

3.3.1. Sulfur Dioxide Adsorption.

Figure 10 showed the DRIFTS spectra of the V-0.2Ce/Ti–Zr catalyst in the flow of 0.06% SO2 at 50 °C and then purged by N2 with increasing temperatures from 50 °C to 400 °C. The peak at 1633 cm−1 was linked to H2O vibration produced by the reaction of SO2 and hydroxyl on the catalytic surface [8]. With the addition of SO2 at 50 °C, peaks at 1376 cm−1, 1338 cm−1, 1265 cm−1, 1097 cm−1 and 1049 cm−1 were detected. Based on the research of Peak et al. [21], the triply degenerate asymmetric stretching ν3 band were accessible to FTIR investigation, and would split into three bands when the bidentate sulfate complex was formed. Therefore, it was deduced that the bands at 1265 cm−1, 1097 cm−1 and 1049 cm−1 were attributed to bidentate sulfate on V-0.2Ce/Ti–Zr. The band at 1338 cm–1 was assigned to the adsorbed SO2, which mainly exists in the form of SO32−. The band at 1376 cm−1 might be due to the asymmetric vibration of O = S = O covalent groups (SO42–). When the temperature reached 100 °C, the adsorption peak at 1376 cm−1 disappeared [22,23].
Figure 11a shows the in situ DRIFTS spectra of the V-0.2Ce/Ti–Zr catalyst in the flow of 0.06% SO2 at 250 °C and then purged with N2. After adding SO2 for 5 min, bands at 1363 cm−1, 1344 cm−1, 1295 cm−1, 1083 cm−1 and 1047 cm−1 were detected with increasing intensity. The bands at 1295 cm−1, 1107 cm−1 and 1050 cm−1 were contributed to the bidentate sulfate on V-0.2Ce/Ti–Zr while the band at 1344 cm–1 was assigned to the adsorbed SO2 (SO32–). The band at 1363 cm−1 could be attributed to the asymmetric vibration of O = S = O covalent groups (SO42–). It had been found that the O = S = O asymmetric covalent groups arose from the VOSO4 adsorption peak that appeared at 1383 cm−1 [24,25]. Hence, we deduced the reaction between the adsorbed SO2 with V2O5 in the catalyst to form the VOSO4 intermediate.
Figure 11b shows the DRIFTS spectra of the V-0.2Ce/Ti–Zr catalyst in the flow of O2 + SO2 at 250 °C and then purged with N2. In general, catalysts exhibited similar peaks, while the intensity of SO2 adsorption was enhanced significantly in the presence of O2.

3.3.2. Effect of SO2 on NH3 Adsorption

Figure 12 showed the NH3 adsorption results on V-0.2Ce/Ti–Zr in the presence of SO2. It could be found that original adsorption capacity of NH3 was very weak, and only the band at 1182 cm−1 was observed [26]. Then NH3 was switched off and SO2 + O2 was introduced. The NH3 adsorption peak was replaced by SO42− adsorption peak quickly. Bands at 1278 cm−1, 1178 cm−1 and 1050 cm−1 were assigned to SO42− three-fold degeneracy asymmetric stretching vibration v3 and the band at 1340 cm−1 was assigned to the adsorbed SO2 while the band at 1373 cm−1 could be attributed to the asymmetric vibration of O = S = O covalent groups (SO42–).
In Figure 13, the catalysts were pretreated by 0.06% SO2 + O2 for 30 min, and then exposed to 800 ppm NH3, followed by treating with SO2 and O2 again. After adding NH3, several bands at 1660 cm−1, 1600 cm−1, 1440 cm−1 and 1220 cm−1 were detected. The band at 1600 cm−1 and 1220 cm−1 was associated with the asymmetric and symmetric deformation vibration of NH3 adsorbed on Lewis acid sites, while the band at 1660 cm−1 and 1440 cm−1 were due to the asymmetric and symmetric deformation of NH4+ bound to Brönsted acid sites [26,27]. Compared with the results of Figure 9, it could be found that the intensity of bands due to adsorbed NH3 was significantly increased after being pretreated by SO2 and O2. It was deduced that the presence of SO2 and O2 could promote the transformation from Ce4+ to Ce3+ (the formation of Ce2(SO4)3), resulting in the enhancement of the NH3 adsorption ability. Furthermore, it was observed that the intensity of the NH3 adsorption peaks did not change significantly after the introduction of SO2 and O2 again.

3.3.3. Effect of SO2 on NO Adsorption

The competitive adsorption behavior of SO2 with NO2 in the presence of O2 was investigated, and in situ DRIFTS results were shown in Figure 14. As shown in Figure 15, bands appeared at 1375 cm−1 and 1315 cm−1 were due to cis-N2O22−, and it could be found that the intensity of the bands increased gradually and the purge of N2 had little influence on the adsorption bands. Band at 1190 cm−1 were attributed to the bridging nitrates [28]. After 10 min, SO42− adsorption peak appeared at 1046 cm−1. As the reaction went on, it divided into two adsorption peaks (1049 cm−1 and 1035 cm−1). Along with the purge of N2, band at 1095 cm−1 assigned to bidentate sulfate was observed. Furthermore, weak SO42− adsorption band was detected at 1135 cm−1.
In Figure 15, the catalyst was first treated by NO + O2 for 60 min until the adsorption was saturated and 0.06% SO2 was added into the cell, and then SO2 were introduced into the in-situ cell for various times, followed by N2 purging. As shown in Figure 15, bands due to the cis-N2O22− were also detected at 1375 cm−1 and 1315 cm−1. Band at 1188 cm−1 was assigned to the adsorbed NOx species. Nitrate species adsorption peak appeared at 1144 cm−1 and 1067 cm−1 after NO + O2 pre-adsorption. Once feeding SO2, the intensity of the bands at 1375 cm−1, 1315 cm−1, 1188 cm−1 and 1144 cm−1 were strengthened and the purge of N2 had little influence on the adsorption bands, while the band at 1067 cm−1 gradually disappeared. The SO42− adsorption peak appeared at 1046 cm−1.
Based on the analysis above, it could be concluded that the presence of SO2 could promote NO adsorption on the catalyst surface. In addition, our previous research has a certified catalyst that mainly follows the Eley–Rideal (E-R) mechanism, and NO adsorption on the catalyst surface restrains the SCR reaction [12]. Thus, it can explain the decrease of catalyst activity when SO2 was injected.

3.3.4. Effect of SO2 on NH3 + NO + O2 Co-Adsorption

As shown in Figure 16, in the presence of NH3 + NO + O2, only the band at 1205 cm−1 due to adsorbed NH3 was observed. After 0.06% SO2 was injected, adsorption intensity of bands due to adsorbed NH3 decreased rapidly. These results indicated that the presence of SO2 significantly weakened the NH3 adsorption, which was consistent with the result of Figure 12. After the introduction of SO2 for 20 min, bands at 1373 cm−1, 1280 cm−1, 1095 cm−1 and 1043 cm−1 were detected. The band at 1373 cm−1 could be attributed to the asymmetric vibration of O = S = O covalent groups and bands at 1280 cm−1, 1095 cm−1 and 1043 cm−1 were assigned to SO42− three-fold degeneracy asymmetric stretching vibration v3. In our previous study [12], the SCR reaction over V-0.2Ce/Ti–Zr obeyed the E-R mechanism. Hence, as an important reaction substance, the weakening of NH3 adsorption would greatly inhibit the SCR reaction.

3.4. Possible SO2 Reaction Mechanism Over The Catalyst

The adsorption ability of SO2 had something with its concentration. There was no adsorption when SO2 concentration was low, which explained the stable activity of catalysts at low SO2 concentration. With the increase of the SO2 concentration, SO2 and SO42− adsorption appeared. SO2 was adsorbed on the surface in the form of SO32. The consumption of surface OH species meant that SO2 was able to react with surface hydroxyl groups to form adsorbed H2O. It was speculated that in the presence of O2, SO2 would react with NH3 to form (NH4)2SO4, which could deposit on the surface of the catalyst and reduce the catalyst activity. The sulfate products could combine with Ce to form more stable Ce(SO4)2. The reaction might be proposed as follows:
SO2(g) + O2−(a) →SO32−(a),
SO2(g) + 2OH(a) →SO32− (a) + H2O,
O2(g) →2[O](a),
SO32− (a) + [O](a) →SO42− (a)
SO42− (a) + 2NH4+(a)→(NH4)2SO4
Ce4+ + SO42− →Ce(SO4)2
According to the in-situ DRIFTS results and other study [19], SO2(SO32–) could react with V5+–OH to form VOSO4 intermediate.
SO32− (a) + V5+-OH(a) →VOSO4
With the time passing, absorbed SO2 (SO32−) contacted with the active component Ce and caused the transformation from Ce4+ to Ce3+, increasing the oxygen vacancy and enhancing the NH3 adsorption ability, which can explain the recovery of activity. The presence of O2 promoted the redox of SO2 and formation of sulfate Ce2(SO4)3.
3SO2 + 2CeO2 + O2 →Ce2(SO4)3

4. Conclusions

Catalyst activity test results showed that low concentration SO2 (<0.02%) had little influence on catalyst activity. With increasing SO2 concentration, NOx conversion of the catalyst gradually decreased but could restore to the original level when stopping feeding SO2. A reversible deactivation would occur in the presence of H2O. However, in the presence of SO2 and H2O, the catalyst was irreversibly deactivated, which was worse than that of adding SO2 or H2O alone.
The characterization results showed that the BET specific surface area of the catalysts poisoned by SO2 were reduced, and the redox capacity were significantly weakened. The sulfuration of the active component Ce and the deposition of ammonium sulfate might be the causes of the deactivation. XPS analysis showed that the presence of SO2 promoted the generation of Ce3+, which probably promoted the SCR reaction.
In situ DRIFTS analysis indicated that the adsorption capacity of SO2 was enhanced in the presence of O2. The presence of SO2 would inhibit the adsorption of NH3. At the same time, the NO adsorption ability of the catalyst was enhanced to some extent by SO2. Moreover, the NH3 adsorption ability of the catalyst with pre-adsorption of SO2 + O2 was markedly enhanced, indicating that when SO2 + O2 existed, more Ce3+ and oxygen vacancy were produced. These conclusions might contribute to a better understanding of the SO2 poisoning mechanism over the V2O5–CeO2/TiO2–ZrO2 catalyst.

Author Contributions

Formal analysis, P.W.; S.W. and W.G.; Investigation, Y.Z.; Methodology, K.Z. and K.S.; Supervision, K.S.; Writing-original draft, Y.Z.

Acknowledgments

This work was supported by the Key Research and Development Projects of Jiangsu Province (BE2017716), National Key R&D Program of China (2017YFB0603201) and Environmental nonprofit industry research subject (2016YFC0208102).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fan, Y.Z.; Cao, F.H. Thermal decomposition kinetics of ammonium sulfate. J. Chem. Eng. Chin. Univ. 2011, 25, 341–346. [Google Scholar]
  2. Zhang, W.; Chen, H.; Xia, Y. Effects of Nd-modification on the activity and SO2 resistance of MnOx/TiO2 catalyst for low-temperature NH3-SCR. New J. Chem. 2018, 42, 12845–12852. [Google Scholar]
  3. Qi, G.; Yang, R.T. Low-temperature selective catalytic reduction of NO with NH3 over iron and manganese oxides supported on titania. Appli. Catal. B: Environ. 2003, 44, 217–225. [Google Scholar] [CrossRef]
  4. Wei, L.; Cui, S.; Guo, H. Mechanism of SO2 Influence on Mn/TiO2 for Low Temperature SCR Reaction; Han, Y., Ed.; Chinese Materials Conference; Springer: Berlin, Germany, 2017; pp. 789–796. [Google Scholar]
  5. Jiang, B.Q.; Wu, Z.B.; Liu, Y. DRIFT study of the SO2 effect on low-temperature SCR reaction over Fe-Mn/TiO2. J Phys Chem. C 2010, 114, 4961–4965. [Google Scholar] [CrossRef]
  6. Liu, F.; He, H. Selective catalytic reduction of NO with NH3 over manganese substituted iron titanate catalyst: Reaction mechanism and H2O/SO2 inhibition mechanism study. Catal. Today. 2010, 153, 70–76. [Google Scholar] [CrossRef]
  7. Pan, S.; Luo, H.; Li, L. H2O and SO2 deactivation mechanism of MnOx/MWCNTs for low-temperature SCR of NOx with NH3. J. Mol. Catal. A: Chem. 2013, 377, 154–161. [Google Scholar] [CrossRef]
  8. Gu, T.; Liu, Y.; Weng, X. The enhanced performance of ceria with surface sulfation for selective catalytic reduction of NO by NH3. Catal. Commun. 2010, 12, 310–313. [Google Scholar] [CrossRef]
  9. Xu, Q.; Yang, W.; Cui, S. Sulfur resistance of Ce-Mn/TiO2 catalysts for low-temperature NH3–SCR. R. Soc. Open Sci. 2018, 5, 171846. [Google Scholar] [CrossRef]
  10. Yamaguchi, T. Recent progress in solid superacid. Cheminform 1990, 61, 1–25. [Google Scholar] [CrossRef]
  11. 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. Appli. Catal. B Environ. 2015, 166–167, 37–44. [Google Scholar] [CrossRef]
  12. Zhang, Y.; Guo, W.; Wang, L. Characterization and activity of V2O5-CeO2/TiO2-ZrO2 catalysts for NH3-selective catalytic reduction of NOx. Chin. J. Catal. 2015, 36, 1701–1710. [Google Scholar] [CrossRef]
  13. Zhang, Y.; Zhu, X.; Shen, K. Influence of ceria modification on the properties of TiO2-ZrO2 supported V2O5 catalysts for selective catalytic reduction of NO by NH3. J. Colloid Interface Sci. 2012, 376, 233–238. [Google Scholar] [CrossRef] [PubMed]
  14. Zhang, Y.; Yue, X.; Huang, T. In Situ DRIFTS Studies of NH3-SCR Mechanism over V2O5-CeO2/TiO2-ZrO2 Catalysts for Selective Catalytic Reduction of NOx. Materials 2018, 11, 1307. [Google Scholar] [CrossRef] [PubMed]
  15. Xu, W.; He, H.; Yu, Y. Deactivation of a Ce/TiO2 Catalyst by SO2 in the Selective Catalytic Reduction of NO by NH3. J. Phys. Chem. C 2009, 113, 4426–4432. [Google Scholar] [CrossRef]
  16. Pena, M.L.; Dejoz, A.; Fornes, V. V-containing MCM-41 and MCM-48 catalysts for the selective oxidation of propane in gas phase. Appli. Catal. A Gen. 2001, 209, 155–164. [Google Scholar] [CrossRef]
  17. Held, A.; Kowalska-Kuś, J.; Nowińska, K. Epoxidation of propene on vanadium species supported on silica supports of different structure. Catal. Commun. 2012, 17, 108–113. [Google Scholar] [CrossRef]
  18. Jiang, H.; Wang, C.; Wang, H. Synthesis of highly efficient MnOx catalyst for Low-Temperature NH3-SCR prepared from Mn-MOF-74 template. Mater. Lett. 2016, 168, 17–19. [Google Scholar] [CrossRef]
  19. Liu, J.; Guo, R.; Li, M. Enhancement of the SO2 resistance of Mn/TiO2 SCR catalyst by Eu modification: A mechanism study. Fuel 2018, 223, 385–393. [Google Scholar] [CrossRef]
  20. Zhang, L.; Zhang, D.; Zhang, J. Design of meso-TiO2@MnOx-CeOx/CNTs with a core-shell structure as DeNOx catalysts: Promotion of activity, stability and SO2-tolerance. Nanoscale 2013, 5, 9821. [Google Scholar] [CrossRef] [PubMed]
  21. Peak, D.; Ford, R.G.; Sparks, D.L. An in situ ATR-FTIR investigation of sulfate bonding mechanisms on goethite. J. Colloid Inter. Sci. 1999, 218, 289–299. [Google Scholar] [CrossRef] [PubMed]
  22. Luo, T.; Gorte, R.J. Characterization of SO2-poisoned ceria-zirconia mixed oxides. Appli. Catal. B Environ. 2004, 53, 77–85. [Google Scholar] [CrossRef]
  23. Yamaguchi, T.; Jin, T.; Tanabe, K. Structure of acid sites on sulfur-promoted iron oxide. J. Phys. Chem. 1986, 90, 3148–3152. [Google Scholar] [CrossRef]
  24. Zhang, L.; Li, L.L.; Cao, Y.; Yao, X.J.; Ge, C.Y.; Gao, F.; Yu, D.; Tang, C.J.; Dong, L. Getting insight into the influence of SO2 on TiO2/CeO2 for the selective catalytic reduction of NO by NH3. Appli. Cataly. B Environ. 2015, 165, 589–598. [Google Scholar] [CrossRef]
  25. Liu, Y.; Shu, H.; Xu, Q. FT-IR study of the catalytic oxidation of SO2 during the process of selective catalytic reduction of NO with NH3 over commercial catalysts. J. Fuel Chem. Technol. 2015, 43, 1018–1024. [Google Scholar] [CrossRef]
  26. Zhang, Q.; Fan, J.; Ning, P. In situ DRIFTS investigation of NH3-SCR reaction over CeO2 /zirconium phosphate catalyst. Appli. Surf. Sci. 2018, 435, 1037–1045. [Google Scholar] [CrossRef]
  27. Gelves, J.F.; Dorkis, L.; Márquez, M.A. Activity of an iron Colombian natural zeolite as potential geo-catalyst for NH3-SCR of NOx. Cataly. Today 2019, 320, 112–122. [Google Scholar] [CrossRef]
  28. Jin, C.M.; Xin, Y.L.; Wen, B.S. Enhancement of low-temperature catalytic activity over highly dispersed Fe-Mn/Ti Catalyst for selective catalytic reduction of NOx with NH3. Ind. Eng. Chem. Res. 2018, 57, 10159–10169. [Google Scholar]
Figure 1. NOx conversion of V-0.2Ce/Ti–Zr in the presence of SO2 (250 °C). Reaction condition: NH3 = NO = 0.08%, O2 = 5%, N2 as balance, SO2 content: 0.02%, 0.04% and 0.06%, all by volume).
Figure 1. NOx conversion of V-0.2Ce/Ti–Zr in the presence of SO2 (250 °C). Reaction condition: NH3 = NO = 0.08%, O2 = 5%, N2 as balance, SO2 content: 0.02%, 0.04% and 0.06%, all by volume).
Materials 12 02534 g001
Figure 2. NOx conversion of V-0.2Ce/Ti–Zr in the presence of H2O (250 °C).
Figure 2. NOx conversion of V-0.2Ce/Ti–Zr in the presence of H2O (250 °C).
Materials 12 02534 g002
Figure 3. NOx conversion of V-0.2Ce/Ti–Zr in the presence of H2O and SO2 (250 °C; reaction condition: NH3 = NO = 0.08%, O2 = 5%, H2O = 10%, SO2 = 0.02%, 0.04% and 0.06%, all by volume.
Figure 3. NOx conversion of V-0.2Ce/Ti–Zr in the presence of H2O and SO2 (250 °C; reaction condition: NH3 = NO = 0.08%, O2 = 5%, H2O = 10%, SO2 = 0.02%, 0.04% and 0.06%, all by volume.
Materials 12 02534 g003
Figure 4. NOx conversion of the V-0.2Ce/Ti–Zr and HS-V-0.2Ce/Ti–Zr catalyst.
Figure 4. NOx conversion of the V-0.2Ce/Ti–Zr and HS-V-0.2Ce/Ti–Zr catalyst.
Materials 12 02534 g004
Figure 5. Pore size distribution of the fresh and poisoned catalysts. (a) V-0.2Ce/Ti–Zr; (b) S-V-0.2Ce/Ti–Zr and (c) HS-V-0.2Ce/Ti–Zr.
Figure 5. Pore size distribution of the fresh and poisoned catalysts. (a) V-0.2Ce/Ti–Zr; (b) S-V-0.2Ce/Ti–Zr and (c) HS-V-0.2Ce/Ti–Zr.
Materials 12 02534 g005
Figure 6. XRD patterns of the catalysts (a) V-0.2Ce/Ti–Zr, (b) S-V-0.2Ce/Ti–Zr and (c) HS-V-0.2Ce/Ti–Zr.
Figure 6. XRD patterns of the catalysts (a) V-0.2Ce/Ti–Zr, (b) S-V-0.2Ce/Ti–Zr and (c) HS-V-0.2Ce/Ti–Zr.
Materials 12 02534 g006
Figure 7. TEM pictures of the catalysts before and after poisoning by SO2. (a) V-0.2Ce/Ti–Zr and (b) HS-V-0.2Ce/Ti–Zr.
Figure 7. TEM pictures of the catalysts before and after poisoning by SO2. (a) V-0.2Ce/Ti–Zr and (b) HS-V-0.2Ce/Ti–Zr.
Materials 12 02534 g007
Figure 8. Hydrogen temperature-programmed desorption (H2-TPR) patterns of the catalysts before and after being poisoned by SO2. (a) V-0.2Ce/Ti–Zr, (b) S-V-0.2Ce/Ti–Zr and (c) HS-V-0.2Ce/Ti–Zr.
Figure 8. Hydrogen temperature-programmed desorption (H2-TPR) patterns of the catalysts before and after being poisoned by SO2. (a) V-0.2Ce/Ti–Zr, (b) S-V-0.2Ce/Ti–Zr and (c) HS-V-0.2Ce/Ti–Zr.
Materials 12 02534 g008
Figure 9. XPS spectra of Ce 3d and O 1s on catalysts before and after being poisoned by SO2. (a) V-0.2Ce/Ti–Zr; (b) S-V-0.2Ce/Ti–Zr and (c) HS-V-0.2Ce/Ti–Zr.
Figure 9. XPS spectra of Ce 3d and O 1s on catalysts before and after being poisoned by SO2. (a) V-0.2Ce/Ti–Zr; (b) S-V-0.2Ce/Ti–Zr and (c) HS-V-0.2Ce/Ti–Zr.
Materials 12 02534 g009
Figure 10. In situ DRIFTS spectra of V-0.2Ce/Ti–Zr treated in flowing 0.06% SO2 at 50 °C and then purged by N2 at a different temperature.
Figure 10. In situ DRIFTS spectra of V-0.2Ce/Ti–Zr treated in flowing 0.06% SO2 at 50 °C and then purged by N2 at a different temperature.
Materials 12 02534 g010
Figure 11. In situ DRIFTS spectra recorded at 250 °C in a flow of 0.06% SO2 and 0.06% SO2 + 5% O2 over V-0.2Ce/Ti–Zr. (a) SO2 adsorption and (b) SO2 + O2 co-adsorption.
Figure 11. In situ DRIFTS spectra recorded at 250 °C in a flow of 0.06% SO2 and 0.06% SO2 + 5% O2 over V-0.2Ce/Ti–Zr. (a) SO2 adsorption and (b) SO2 + O2 co-adsorption.
Materials 12 02534 g011
Figure 12. In situ DRIFTS spectra of V-0.2Ce/Ti–Zr pretreated by NH3 exposed to NH3 + SO2 + O2 for various times.
Figure 12. In situ DRIFTS spectra of V-0.2Ce/Ti–Zr pretreated by NH3 exposed to NH3 + SO2 + O2 for various times.
Materials 12 02534 g012
Figure 13. In situ DRIFTs spectra of V-0.2Ce/Ti–Zr pretreated by SO2 + O2 exposed to NH3 for various times and then treated by SO2 + O2 again.
Figure 13. In situ DRIFTs spectra of V-0.2Ce/Ti–Zr pretreated by SO2 + O2 exposed to NH3 for various times and then treated by SO2 + O2 again.
Materials 12 02534 g013
Figure 14. In situ DRIFTS spectra taken at 250 °C in a flow of 0.08% NO + 0.06% SO2 + 5% O2 on V-0.2Ce/Ti–Zr.
Figure 14. In situ DRIFTS spectra taken at 250 °C in a flow of 0.08% NO + 0.06% SO2 + 5% O2 on V-0.2Ce/Ti–Zr.
Materials 12 02534 g014
Figure 15. In situ DRIFTS spectra taken at 250 °C in a flow of 0.08% NO + 0.06% SO2 + 5% O2 over the NO + O2 presorbed on V-0.2Ce/Ti–Zr.
Figure 15. In situ DRIFTS spectra taken at 250 °C in a flow of 0.08% NO + 0.06% SO2 + 5% O2 over the NO + O2 presorbed on V-0.2Ce/Ti–Zr.
Materials 12 02534 g015
Figure 16. In situ DRIFTS spectra of the NH3 + NO + O2 reaction on V-0.2Ce/Ti–Zr (250 °C) with 0.06% SO2.
Figure 16. In situ DRIFTS spectra of the NH3 + NO + O2 reaction on V-0.2Ce/Ti–Zr (250 °C) with 0.06% SO2.
Materials 12 02534 g016
Table 1. BET surface area and pore volume of the fresh and poisoned catalyst.
Table 1. BET surface area and pore volume of the fresh and poisoned catalyst.
SampleBET Surface Area (m2/g)Pore Volume (mL/g)
V-0.2Ce/Ti–Zr54.450.14
S-V-0.2Ce/Ti–Zr25.070.16
HS-V-0.2Ce/Ti–Zr15.270.11
Table 2. XRF results of the catalysts before and after being poisoned by SO2.
Table 2. XRF results of the catalysts before and after being poisoned by SO2.
SampleVCeTiZrS
V-0.2Ce/Ti–Zr0.72312.5121.234.660
S-V-0.2Ce/Ti–Zr0.6312.36519.7526.660
HS-V-0.2Ce/Ti–Zr0.55812.93220.4626.41.13
Table 3. Peak area ratio of XPS.
Table 3. Peak area ratio of XPS.
Area Ratio
Sample C e 3 + C e 3 + + C e 4 + O β O β + O α
V-0.2Ce/Ti–Zr 24.9428.14
S-V-0.2Ce/Ti–Zr25.1520.71
HS-V-0.2Ce/Ti–Zr24.9318.67

Share and Cite

MDPI and ACS Style

Zhang, Y.; Wu, P.; Zhuang, K.; Shen, K.; Wang, S.; Guo, W. Effect of SO2 on the Selective Catalytic Reduction of NOx over V2O5-CeO2/TiO2-ZrO2 Catalysts. Materials 2019, 12, 2534. https://doi.org/10.3390/ma12162534

AMA Style

Zhang Y, Wu P, Zhuang K, Shen K, Wang S, Guo W. Effect of SO2 on the Selective Catalytic Reduction of NOx over V2O5-CeO2/TiO2-ZrO2 Catalysts. Materials. 2019; 12(16):2534. https://doi.org/10.3390/ma12162534

Chicago/Turabian Style

Zhang, Yaping, Peng Wu, Ke Zhuang, Kai Shen, Sheng Wang, and Wanqiu Guo. 2019. "Effect of SO2 on the Selective Catalytic Reduction of NOx over V2O5-CeO2/TiO2-ZrO2 Catalysts" Materials 12, no. 16: 2534. https://doi.org/10.3390/ma12162534

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