Next Article in Journal
Atomic Layer Deposition for Preparation of Highly Efficient Catalysts for Dry Reforming of Methane
Next Article in Special Issue
Dry Reforming of Methane over NiLa-Based Catalysts: Influence of Synthesis Method and Ba Addition on Catalytic Properties and Stability
Previous Article in Journal
High-Performing PGM-Free AEMFC Cathodes from Carbon-Supported Cobalt Ferrite Nanoparticles
Previous Article in Special Issue
Performance and Stability of Metal (Co, Mn, Cu)-Promoted La2O2SO4 Oxygen Carrier for Chemical Looping Combustion of Methane
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Byproduct Analysis of SO2 Poisoning on NH3-SCR over MnFe/TiO2 Catalysts at Medium to Low Temperatures

Institute of Environmental Engineering, National Chiao Tung University, Hsinchu 300, Taiwan
*
Author to whom correspondence should be addressed.
Catalysts 2019, 9(3), 265; https://doi.org/10.3390/catal9030265
Submission received: 18 February 2019 / Revised: 11 March 2019 / Accepted: 11 March 2019 / Published: 15 March 2019
(This article belongs to the Special Issue Catalysts Deactivation, Poisoning and Regeneration)

Abstract

:
The byproducts of ammonia-selective catalytic reduction (NH3-SCR) process over MnFe/TiO2 catalysts under the conditions of both with and without SO2 poisoning were analyzed. In addition to the NH3-SCR reaction, the NH3 oxidation and the NO oxidation reactions were also evaluated at temperatures of 100–300 °C to clarify the reactions occurred during the SCR process. The results indicated that major byproducts for the NH3 oxidation and NO oxidation tests were N2O and NO2, respectively, and their concentrations increased as the reaction temperature increased. For the NH3-SCR test without the presence of SO2, it revealed that N2O was majorly from the NH3-SCR reaction instead of from NH3 oxidation reaction. The byproducts of N2O and NO2 for the NH3-SCR reaction also increased after increasing the reaction temperature, which caused the decreasing of N2-selectivity and NO consumption. For the NH3-SCR test with SO2 at 150 °C, there were two decay stages during SO2 poisoning. The first decay was due to a certain amount of NH3 preferably reacted with SO2 instead of with NO or O2. Then the catalysts were accumulated with metal sulfates and ammonium salts, which caused the second decay of NO conversion. The effluent N2O increased as poisoning time increased, which was majorly from oxidation of unreacted NH3. On the other hand, for the NH3-SCR test with SO2 at 300 °C, the NO conversion was not decreased after increasing the poisoning time, but the N2O byproduct concentration was high. However, the SO2 led to the formation of metal sulfates, which might inhibit NO oxidation reactions and cause the concentration of N2O gradually decreased as well as the N2-selectivity increased.

1. Introduction

Nitrogen oxides (NOx, NO and NO2) produced from stationary sources are major air pollutants that lead to environmental concerns such as photochemical smog and acid rain [1]. The most effective technology for the removal of NOx emission from coal-fired power plants is ammonia-selective catalytic reduction (NH3-SCR; SCR hereafter) [2]. The traditional SCR catalysts are active within the temperature window of 300–400 °C [3,4]. Even though some of the traditional catalyst compositions such as V2O5–WO3/ TiO2 or Fe-zeolite-based catalysts can lower down their working temperature window to be as low as 250 or even 200 °C [5,6,7,8,9,10], there is still a strong demand in developing SCR catalysts to be active at less than 200 °C and placing them downstream of the electrostatic precipitator and desulfurizer [11,12,13].
Literature data showed that Mn-based catalysts have good activity for low-temperature SCR [14,15,16,17]. Moreover, in iron containing SCR catalysts, the introduction of Mn could obviously enhance the low-temperature activity, probably due to the fact that synergistic effect between iron and manganese species [18]. It was reported that the MnFe/TiO2 could improve the activity, stability and SO2 durability of the SCR catalysts using NH3 as the reducing agent [19,20,21,22,23].
In the past, there have been extensive studies usingFourier-transform infrared spectroscopy (FTIR) for understanding the mechanism of SCR reaction on the surface of MnFe catalysts [24,25,26,27,28]. In addition, several different types of reaction mechanisms have been proposed including the typical Eley–Rideal mechanism and Langmuir–Hinshelwood mechanism. For the Eley–Rideal mechanism, it is assumed that the gaseous NO directly reacts with an activated ammonia surface complex [29]. On the other hand, the Langmuir–Hinshelwood reaction mechanism involves that a surface NO complex reacts with an activated ammonia [3,30]. Moreover, Yang et al. [31] used in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS-FTIR) to reveal the mechanism of low-temperature SCR reaction over the MnFe spinel. The results indicated that the contribution of Eley–Rideal mechanism to NO conversion increased after increasing the reaction temperature.
Although many studies have been done on the SCR mechanism, to the authors’ best knowledge, there has been no work on clarifying reactions occurred in the medium to low-temperature SCR system with and without the presence of SO2. Therefore, this study employed MnFe/TiO2 catalyst to study the oxidation of NH3, the oxidation of NO and the NH3-SCR reaction with and without the presence of SO2 at 100–300 °C. The reactants and byproducts of gaseous NO, NH3, N2O, and NO2 in the effluent streams as well as solid byproducts of ammonium salts and metal sulfates on the catalysts were analyzed. The results can offer useful information to understand the reaction pathway at different operation conditions for the application of medium to low-temperature SCR catalysts.

2. Results and Discussion

2.1. Oxidation Reactions of NH3 and NO without SO2

To understand the products and byproducts of oxidation reactions, the NH3 and NO oxidation reactions over the MnFe/TiO2 catalyst were studied within the temperature range of 100–300 °C, which was the most active temperature region for MnFe/TiO2 catalyst in SCR reaction [32]. In addition, the N-balance of NH3 and NO oxidation tests were calculated by Equations (1) and (2), respectively.
N-balance  in   NH 3   oxidation   =   [ NH 3 ] out   + [ N 2 O ] out   + [ NO ] out + [ NO 2 ] out   [ NH 3 ] in   × 100 %
N-balance  in   NO   oxidation   =   [ NO ] out   + [ NO 2 ] out   [ NO ] in   × 100 %
The results of N-balance and outlet gas concentrations in the NH3 oxidation reaction without SO2 over MnFe/TiO2 catalyst at different reaction temperatures are shown in Figure 1. It can be seen that the major oxidation product of NH3 was N2O (i.e., Equation (3) shown later in the Materials and Method section) instead of NO (Equation (4)). In addition, N2O increased from 31 ppm to 219 ppm after increasing the reaction temperature from 100 °C to 300 °C. On the other hand, the effluent NO concentration was only 36 ppm at most, which occurred at reaction temperature of 300 °C; but no NO2 was found in the effluent at all tested temperatures.
The results of N-balance and outlet gas concentrations in the NO oxidation reaction over MnFe/TiO2 catalyst at different reaction temperatures are shown in Figure 2, it can be seen that NO could be oxidized to NO2, and the concentration of NO2 was significantly increased from 10 to 309 ppmv for reaction temperatures from 100 to 300 °C.
The above results indicated that NH3 and NO oxidations (Equations (3)–(5)) increased after increasing the reaction temperature, and the major products of NH3 oxidation and NO oxidation were N2O and NO2, respectively. In addition, it can be observed that both the N-balance results shown in Figure 1 and Figure 2 were very high at 97~100% in the 100–300 °C range. This indicates that we can detect almost all the reaction species.

2.2. SCR Reactions without SO2

Figure 3 shows the NO consumption (in ppmv) and outlet gas concentrations of N2O, NO2 and NH3 over the MnFe/TiO2 catalyst for reaction temperatures of 100–300 °C. It can be seen that the MnFe/TiO2 catalyst maintained high NO consumption in the SCR reaction. When raising reaction temperature from 100 to 200 °C, the NO consumption in the SCR reaction increased slightly from 450 ppm to 495 ppm, and the outlet concentration of NH3 was slightly decreased from 63 to 0 ppm.
At the reaction temperature of 200 °C, the concentrations of N2O and NO2 were 175 and 0 ppm, respectively, with 98% of NO consumption in the SCR reactions (500 ppm NO reacted with 500 ppm NH3). On the other hand, the N2O from NH3 oxidation (500 ppm NH3 reacted with O2) at the reaction temperature of 200 °C was 156 ppm (as indicated by Figure 1). Therefore, it can be seen that in the SCR system, lower N2O percentage ( 175 ppm   N 2 O 500   ppm   NO   +   500   ppm   NH 3 = 17.5%) was being produced as compared to that during only NH3 oxidation ( 156   ppm   N 2 O   500   ppm   NH 3 = 31.2%).
The “fast SCR” (Equation (8)), first proposed in 1986 [33], proceeds at a much higher reaction rate than “standard SCR” reactions (Equations (6) and (7)) was developed to improve deNOx efficiency at lower temperatures [34,35]. As indicated by Figure 2 for the NO oxidation reaction, NO could be oxidized to NO2 at temperatures higher than 150 °C, but NO2 was not detected at temperature below 200 °C in the SCR test as seen in Figure 3. This might be due to the fact that NO2 would react with NO and NH3 according to the fast SCR reaction (Equation (8)).
After further raising the reaction temperature from 200 to 300 °C, NO consumption in the SCR reaction was slightly decreased from 495 ppm to 455 ppm, and concentrations of N2O and NO2 increased from 175 to 270 ppm and 0 to 36 ppm, respectively. When reaction temperature increased, oxidation reactions occurred more quickly as indicated by Figure 1 and Figure 2. Hence the reason for slight decreases in the NO consumption at above 200 °C might be due to the fact that NO was oxidized (Equation (5)) as also demonstrated in the literature [26,36,37,38]. Moreover, the results of Figure 3 also indicated that products of the SCR reaction gradually changed from N2 to N2O when raising the temperature from 100 to 300 °C. Therefore, the N2-selectivity of SCR reaction was gradually decreased from 93% to 41% after increasing the reaction temperature from 100 °C to 300 °C.

2.3. SCR Reactions with SO2

To clarify the byproducts of SCR reactions with SO2, NO consumption in the SCR reaction and outlet gas concentrations in the NH3-SCR reaction over MnFe/TiO2 catalyst at 150 and 300 °C are shown in Figure 4a,b, respectively. It can be seen in Figure 4a that there were four stages during the SCR test period. At stage I where no SO2 was introduced, the NO consumption in the SCR reaction was very high (495 ppm out of 500 ppm). After 150 ppm SO2 was introduced, the NO consumption in the SCR reaction decreased rapidly within 60 min SO2 poisoning (stage II), and they remained roughly stable for another 60 min (stage III), then decreased gradually with time again (stage IV).
In our previous work [32], it was found that the gradual decrease (stage II) of NO consumption in the SCR reaction was probably due to the fact that SO2 competed with NO to react with NH3 and form ammonium salts. However, the MnFe/TiO2 catalyst still had good activity at stage III. Thus, during this stage, there was no sufficient NH3 to react with NO, which caused NO consumption remained relatively lower but stable at around 425 ppm from 60 to 120 min SO2 poisoning time as compared to stages I and II. At stage IV, NO consumption in the SCR reaction decreased significantly from 420 ppm to 153 ppm for poisoning time from 120 to 360 min, which was due to accumulation of metal sulfates and ammonium salts which blocked the active sites of catalyst.
On the other hand, no effluent NH3 was detected from stage I to stage III. Then NH3 slip occurred at stage IV as seen in Figure 4a, which was due to less active sites and decreasing NO consumption in the SCR reaction. One can also see that concentrations of N2O slightly decreased from 76 to 63 ppm within 120 min SO2 poisoning time (stage II and III), then increased gradually to 86 ppm after 360 min SO2 poisoning time (stage IV). This result indicated that SO2 would react with NH3 and form ammonium salts, which cause concentrations of N2O slightly decreased at stages II and III. Then NH3 slip occurred at stage IV, which caused increasing N2O concentrations.
For the SCR test with SO2 at 300 °C as shown in Figure 4b, it can be seen that the SO2 poisoning effect was negligible as compared to the poisoning test at low temperature. When the reaction temperature was at 300 °C, the NO consumption in the SCR reaction remained around 460 ppm, which was almost the same as those without SO2 poisoning. Moreover, during the SCR activity tests no SO2 concentration was detected, which indicated that all the gas phase SO2 molecules might be adsorbed and/or reacted with metal catalysts.
The main reason for the inhibition of SO2 poisoning might be attributed to different SCR reaction mechanisms at different temperatures. Literature results showed that at lower temperatures (<200 °C), SCR reactions would follow the Langmuir−Hinshelwood mechanism. In the Langmuir−Hinshelwood mechanism, SO2 would compete with NO to be adsorbed on the active sites, which cause the decreases of SCR efficiencies. However, when reaction temperature increased, the SCR reaction mechanism would transform from Langmuir−Hinshelwood mechanism to Eley–Rideal mechanism. In the high temperature range (>200 °C), the SCR reaction mainly followed the Eley–Rideal mechanism, over which the gaseous NO could directly react with an activated ammonia [30,39,40].
One can also see from Figure 4b that at reaction temperature of 300 °C, the SCR system did not have NH3 slip. The concentrations of NO2 and N2O decreased from 30 to 0 ppm and from 270 to 172 ppm, respectively; and the N2-selectivity increased from 41% to 62% for the SCR reaction from stage II to stage IV. The result indicated that by increasing the reaction temperature to 300 °C the SO2 poisoning effect could be inhibited, and the N2 selectivity can be enhanced due to the decreased N2O concentrations.
The amount of sulfate species on the catalysts was estimated by thermo-gravimetric analysis (TGA) analysis and the results are shown in Figure 5 in terms of differential thermogram (DTG) spectra. The weight loss profiles of all samples showed three distinct decomposition steps: (1). the weight loss at low temperature (<200 °C) was assigned to the water desorption on the catalyst surface. (2). the weight loss at 200–400 °C could be attributed to decomposition of ammonium salts [41,42,43]. (3). the weight loss at high temperature (>670 °C) was originated from metal sulfates [44,45].
It can be seen from Figure 5 that the fresh catalyst had only one major weight loss peak, which appeared at 50–150 °C and corresponded to H2O desorption on the catalyst surface. On the other hand, decomposition peaks of both ammonium salts (200–400 °C) and metal sulfates (670–900 °C) were observed on all MnFe/TiO2 catalysts poisoned at temperature ranges of 100–200 °C. When reaction temperatures were 250 °C and 300 °C, there were no decomposition peak of ammonium salts. This reveals that ammonium salts were not formed on the catalyst surface at temperatures above 250 °C. As also noted in Figure 5, there was another peak in the range of temperature of 400–600 °C for the fresh catalyst and poisoned catalyst at 100 °C. However, since we cannot find related literature discussed on this, so the reason for this peak is not clear.
The amounts of sulfate species of fresh and poisoned catalysts are listed in Table 1. It is observed that the amounts of ammonium salts on the catalysts decreased from 2.3 wt.% to negligible amounts by increasing the SCR reaction temperature from 100 to 300 °C. This indicated that the reaction temperature would directly affect the formation of ammonium salts. It is noted that rigid quantification of metal sulfates accumulated on the poisoned catalyst was not possible via the DTG data because some ammonium salts could also be transformed into metal sulfates during the continuous TGA heating process [46]. Hence we can only confirm that all the poisoned catalysts had roughly similar amounts of metal sulfates during the temperature from 100 °C to 250 °C (within experimental error of TGA instrument, ±0.3 wt.%), except for the case at 300 °C.
At high temperature of 300 °C, the NO consumption was not affected by SO2. Thus, it is easy to predict that adding an excessive amount of NH3 tends to be oxidized and forming N2O and NO as observed in Figure 1. However, at low temperature of 150 °C, the NO consumption was significantly affected by SO2. Thus, to ensure the gradual decrease of NO consumption at stage II of Figure 4a was due to the competition between SO2 and NO to react with NH3, different inlet amounts of NH3 were tested to study the SO2 poisoning mechanism at temperature of 150 °C. The results are shown in Figure 6. One can see that adding different amounts of NH3 had a similar effect on NO consumption at stages I and IV; but it had different NO consumptions at stages II and III. On the other hand, the NH3 outlet concentrations were different at all stages. At stage I, NH3 only reacted with a certain amount of NO and thus extra NH3 slip was detected in the outlet gas. When SO2 was introduced, NH3 not only reacted with NO but also reacted with SO2. Therefore, both the outlet concentrations of NH3 and NO consumption decreased at stages II and III. Besides, when increasing NH3 amount to above 550 ppm (i.e., NH3/NO molar ratio of 1.1), the first decay of NO consumption at stage II could be inhibited. This indicated that NH3 concentration was sufficient to react with NO and SO2. The result shown in Figure 6 indicated that if a sufficient amount of NH3 can be provided to react with both SO2 and NO, then the first decay of NO consumption could be inhibited at early time of stage II. However, after 120 min of SO2 poisoning (stage IV), the NO consumption could not be affected by different amounts of NH3. Thus, the more injection amount of NH3 led to eventually the more amount of NH3 slip to the atmosphere.

2.4. Product and Byproduct Analysis

Since different reactions occurred during the SCR process, i.e., Equations (3)–(12), were considered both with and without SO2, thus the percentages of product and byproducts can be calculated. Table 2 lists formulas for calculating the percentages of N-containing product (N2) as well as gaseous and solid byproducts (NO, N2O and salts) formed during the SCR process. The byproducts could be formed by the SCR reaction as well as by the NH3 oxidation or the NO oxidation reactions. Because it is difficult to clarify that the outlet NO was from NH3 oxidation or from the unreacted NO, therefore the NO formation from NH3 oxidation (Equation (4)) is neglected and all the effluent NO was assumed to be only from the unreacted NO. This is an acceptable assumption since the formation of NO from NH3 oxidation was very minor (0~7%) at reaction temperatures of 100~300 °C as observed from Figure 1. Besides, it was assumed that the fast SCR reaction (Equation (8)) only served as the intermediate reaction at the low temperature SCR process [35]. Thus, the fast SCR reaction was not considered in the calculation of the percentages of all N-containing product and byproducts.
Based on formulas shown in Table 3, the results on percentages of all N-containing species during the NH3-SCR process tested at 150 °C are shown in Figure 7a; and those tested at 300 °C are shown in Figure 7b. It can be seen from Figure 7a that at SCR operation temperature of 150 °C and without the presence of SO2, the major product of SCR process appeared to be N2. This indicated that MnFe/TiO2 catalyst can serve as a good catalyst and achieve high N2 selectivity at low temperature of 150 °C when SO2 was not presented in the system. The minor presence of N2O byproduct in the exhaust was majorly from the SCR reaction rather than from the NH3 oxidation reaction.
On the other hand, when SO2 was introduced at 150 °C, the percentages of N2 and N2O decreased after increasing the poisoning time as observed in Figure 7a. Because NH3 could not be reacted with NO due to decreased availability of active sites, it tends to increase the NO and NH3 slip to the atmosphere. In addition, the portion of N2O from NH3 oxidation was also gradually increased as increasing the SO2 poisoning time. After 360 min of SO2 poisoning, the exhaust N2O from NH3 oxidation was more than from SCR reaction. One can also see that percentages of ammonium salts were similar at different SO2 poisoning times. This indicated that certain amounts of NH3 would preferentially be reacted with SO2. In addition, the remaining NH3 would then be reacted with NO or O2. Moreover, it can be seen that percentages of NO slip were higher than percentages of NH3 slip at different SO2 poisoning times, which is due to the fact that NH3 not only reacted with NO but also reacted with SO2 and O2.
For results at 300 °C as seen in Figure 7b, it is observed that NO2 from NO oxidation disappeared during the SO2 poisoning. However, the unreacted NO (NO slip) appeared during the SO2 poisoning. This indicated that the formation of metal sulfates at 300 °C might inhibit the NO oxidation reaction. At high temperature, the SCR reaction mainly followed the Eley–Rideal mechanism, so the continuous decreasing of N2O concentration as poisoning time increases and no SO2 concentration in the outlet gases might be related to the reaction between SO2 and activated NH3 instead of the NH3 oxidation reaction at 300 °C. Moreover, the total percentages of N2 and N2O from SCR reactions remained almost the same no matter SO2 was presented in the system or not. The low N2 selectivity revealed that MnFe/TiO2 catalyst may not be a good candidate for SCR process at 300 °C unless the space velocity can be reduced to further enhance the more complete reduction of NO to N2 instead of forming the N2O byproduct. However, it is interested to note that the N2 selectivity gradually increased and the percentages of N2O from SCR reaction gradually decreased after increasing SO2 poisoning time. This indicated that SO2 promotion phenomenon might exist at 300 °C, which was attributed to the formation of SO42− on the catalyst surface. This increased NH3 adsorption and promoted NH3 reaction with NO via Eley−Rideal mechanism.

2.5. Reaction Pathways

From the above results, one could surmise the reaction pathway in the NH3-SCR system with/without SO2 at 150 °C and 300 °C as shown in Figure 8. When the SCR system was at 150 °C without SO2, it can be seen from the top left plot that a fraction of NH3 and NO would be oxidized to N2O and NO2, respectively. In addition, the major reaction product of SCR reaction was N2 instead of N2O. On the other hand, when increasing temperature to 300 °C, it can be seen from the bottom left plot that NH3 would be oxidized to both N2O and NO. Moreover, the major reaction product of SCR reaction was N2O instead of N2, which was revealed by the low N2-selectivity as seen in Figure 7b.
When SO2 was introduced at 150 °C, it can be seen from the top right plot of Figure 8 that SO2 would be reacted with both NH3 and metal catalyst, which resulted in the formation of ammonium salts and metal sulfates, respectively. In addition, because NO could not be adsorbed on metal sulfates, therefore it could not be reacted with NH3 at low temperature. As a result, unreacted NH3 (NH3 slip) turned out to be the major N-containing species in addition to the unreacted NO.
When increasing temperature to 300 °C, it can be seen from the bottom right plot that ammonium salts would not be formed in the presence of SO2, but SO2 would react with metal catalyst to form metal sulfates. In addition, gaseous NO could directly react with adsorbed ammonia via Eley−Rideal mechanism [29]. The major reaction product of SCR reaction gradually changed from N2O to N2 after increasing poisoning time.

3. Materials and Method

3.1. Reactions in SCR System

In the SCR system, it may contain oxidation reactions, SCR reactions and SO2 poisoning reactions [46,47,48,49,50,51,52].
NH3 oxidation:
2NH3 + 2O2 → N2O + 3H2O
4NH3 + 5O2 → 4NO + 6H2O
NO oxidation:
2NO + O2 → 2NO2
SCR reactions:
4NO + 4NH3 + O2 → 4N2 + 6H2O
4NO + 4NH3 + 3O2 → 4N2O + 6H2O
NO + NO2+ 2NH3 → 2N2 + 3H2O (fast SCR)
SO2 oxidation and poisoning:
SO2 + ½ O2 → SO3
SO3 + 2NH3 + H2O → (NH4)2SO4
SO3 + NH3 + H2O → NH4HSO4
SO2 + metal → metal sulfates (e.g., MnSO4 and Ti(SO4)2)
In this study, experimental tests were designed to clarify the products and byproducts of the above reactions.

3.2. Synthesis of MnFe/TiO2 Catalysts

Mn and Fe metal oxides were supported on TiO2 (in the form of TiO(OH)2) by the co-precipitation method. In a typical procedure, 8 g of TiO2 (China Steel Corp., Kaohsiung, Taiwan), 11.57 g of ferric nitrate 9-hydrate (99%,J.T. Baker, Radnor, PA, USA), 7.13 g of manganese (II) acetate tetrahydrate (99%, Merck, Kenilworth, NJ, USA) and D.I. water (76 g) were mixed then adjusted to pH = 10 with 25 wt.% ammonia solution to form a precipitate. It was filtered and washed thoroughly with D.I. water, then dried at 120 °C for 12 h. Finally, the material was calcined at 350 °C for 6 h in air.

3.3. Catalyst Reaction

The NH3 oxidation, NO oxidation and SCR activity tests were carried out at atmospheric pressure in a fixed-bed reactor loaded with sieved pelletized (16–30 mesh) catalysts. The operation conditions of inlet gas concentrations, reaction temperatures, and gas hourly space velocity for different tests are shown in Table 3. Under typical NH3 oxidation, NO oxidation, and SCR reaction tests, the concentrations of NH3 and/or NO were the same at 500 ppmv, and the SO2 concentration was 150 ppmv if SO2 poisoning effect was considered. The feed gases were mixed in a gas mixer. Then the catalysts were preheated in the reactor for 30 min to ensure that an isothermal reaction temperature was reached. During the oxidation and SCR test, the NO and SO2 concentrations at the inlet and outlet of the reactor were monitored by a NO/SO2 analyzer (Ultramat 23, SIEMENS, Munich, Germany). In addition, the concentrations of NH3, N2O, and NO2 at the inlet and outlet of the reactor were monitored by a FTIR Spectrophotometer (Bomem MB 104,San Jose, CA, USA and ITRI, Hsinchu, Taiwan).
The NO consumption due to SCR reactions (Equations (6)–(8)) must be subtracted by the NO oxidation to NO2 (Equation (5)). Thus, the NO consumption of SCR, [NO]SCR is defined by:
[NO]SCR = [NO]in − ([NO]out − [NO2]out)
Since the N-containing product and byproduct of SCR reaction are N2 and N2O from Equations (6)–(8), thus the N2-selectivity of SCR reactions was calculated by
N 2  selectivity of SCR reaction  = [ 1 [ NO ] out +   [ NH 3 ] out + [ N 2 O ] out   + [ NO 2 ] out   [ NO ] in + [ NH 3 ] in     ] × 100 %
The TGA was conducted to determine the sulfates species forming on the surface of the catalysts with a NETZSCH TG 209 F1 apparatus. The heating program was carried out under airflow of 10 mL/min with a heating rate of 10 °C/min from room temperature to 900 °C.

4. Conclusions

This study employed MnFe/TiO2 catalyst to study the product/byproducts for the oxidation of NH3, the oxidation of NO and the NH3-SCR reaction with/without SO2 to understand the reaction pathway of medium to low-temperature SCR process. For SCR operation temperature of 150 °C without the presence of SO2, the major product of SCR process appeared to be N2. The minor presence of N2O byproduct was majorly from the SCR reaction instead of from the NH3 oxidation reaction. Moreover, the result indicated that products of the SCR reaction gradually changed from N2 to N2O when raising the temperature from 100 to 300 °C. Therefore, the N2-selectivity of SCR reaction was gradually decreased. One the other hand when SO2 was introduced at 150 °C, the percentages of N2 and N2O decreased after increased poisoning time. However, when increasing temperature to 300 °C, the percentages of N2 increased while that of N2O decreased after increasing the poisoning time. This indicated the existence of SO2 promotion effect on the NH3-SCR at 300 °C.

Author Contributions

Conceptualization, T.L. and H.B.; methodology, T.L. and H.B.; validation, H.B.; formal analysis, T.L.; investigation, T.L.; resources, H.B.; data curation, T.L. and H.B.; writing—original draft preparation, T.L.; writing—review and editing, H.B.; visualization, T.L. and H.B.; supervision, H.B.; project administration, H.B.; funding acquisition, H.B.

Funding

This research received no external funding.

Acknowledgments

The authors gratefully acknowledge the financial support from the Ministry of Science and Technology, Taiwan through grant No: MOST 105-3113-E-009-003.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Busca, G.; Lietti, L.; Ramis, G.; Berti, F. Chemical and mechanistic aspects of the selective catalytic reduction of NOx by ammonia over oxide catalysts: A review. Appl. Catal. B Environ. 1998, 18, 1–36. [Google Scholar] [CrossRef]
  2. Chang, Y.Y.; Yan, Y.L.; Tseng, C.H.; Syu, J.Y.; Lin, W.Y.; Yuan, Y.C. Development of an Innovative Circulating Fluidized-Bed with Microwave System for Controlling NOx. Aerosol Air Qual. Res. 2012, 12, 379–386. [Google Scholar] [CrossRef]
  3. Garcia-Bordeje, E.; Pinilla, J.L.; Lazaro, M.J.; Moliner, R.; Fierro, J.L.G. Role of sulphates on the mechanism of NH3-SCR of NO at low temperatures over presulphated vanadium supported on carbon-coated monoliths. J. Catal. 2005, 233, 166–175. [Google Scholar] [CrossRef]
  4. Gan, L.N.; Guo, F.; Yu, J.; Xu, G.W. Improved Low-Temperature Activity of V2O5-WO3/TiO2 for Denitration Using Different Vanadium Precursors. Catalysts 2016, 6, 25. [Google Scholar] [CrossRef]
  5. Kompio, P.G.W.A.; Bruckner, A.; Hipler, F.; Auer, G.; Loffler, E.; Grunert, W. A new view on the relations between tungsten and vanadium in V2O5-WO3/TiO2 catalysts for the selective reduction of NO with NH3. J. Catal. 2012, 286, 237–247. [Google Scholar] [CrossRef]
  6. Krocher, O. Selective Catalytic Reduction of NOx. Catalysts 2018, 8, 459. [Google Scholar] [CrossRef]
  7. Balle, P.; Geiger, B.; Kureti, S. Selective catalytic reduction of NOx by NH3 on Fe/HBEA zeolite catalysts in oxygen-rich exhaust. Appl. Catal. B Environ. 2009, 85, 109–119. [Google Scholar] [CrossRef]
  8. Kim, M.H.; Yang, K.H. The Role of Fe2O3 Species in Depressing the Formation of N2O in the Selective Reduction of NO by NH3 over V2O5/TiO2-Based Catalysts. Catalysts 2018, 8, 134. [Google Scholar]
  9. Lee, S.M.; Kim, S.S.; Hong, S.C. Systematic mechanism study of the high temperature SCR of NOx by NH3 over a W/TiO2 catalyst. Chem. Eng. Sci. 2012, 79, 177–185. [Google Scholar]
  10. Ning, R.; Chen, L.; Li, E.; Liu, X.; Zhu, T. Applicability of V2O5-WO3/TiO2 Catalysts for the SCR Denitrification of Alumina Calcining Flue Gas. Catalysts 2019, 9, 220. [Google Scholar] [CrossRef]
  11. Gao, R.H.; Zhang, D.S.; Liu, X.G.; Shi, L.Y.; Maitarad, P.; Li, H.R.; Zhang, J.P.; Cao, W.G. Enhanced catalytic performance of V2O5-WO3/Fe2O3/TiO2 microspheres for selective catalytic reduction of NO by NH3. Catal. Sci. Technol. 2013, 3, 191–199. [Google Scholar] [CrossRef]
  12. Xu, W.Q.; He, H.; Yu, Y.B. 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]
  13. Yang, R.; Huang, H.F.; Chen, Y.J.; Zhang, X.X.; Lu, H.F. Performance of Cr-doped vanadia/titania catalysts for low-temperature selective catalytic reduction of NOx with NH3. Chin. J. Catal. 2015, 36, 1256–1262. [Google Scholar] [CrossRef]
  14. Kong, Z.J.; Wang, C.; Ding, Z.N.; Chen, Y.F.; Zhang, Z.K. Enhanced activity of MnxW0.05Ti0.95−xO2−δ for selective catalytic reduction of NOx with ammonia by self-propagating high-temperature synthesis. Catal. Commun. 2015, 64, 27–31. [Google Scholar] [CrossRef]
  15. Lian, Z.H.; Liu, F.D.; He, H.; Shi, X.Y.; Mo, J.S.; Wu, Z.B. Manganese-niobium mixed oxide catalyst for the selective catalytic reduction of NOx with NH3 at low temperatures. Chem. Eng. J. 2014, 250, 390–398. [Google Scholar] [CrossRef]
  16. Gao, C.; Shi, J.W.; Fan, Z.Y.; Gao, G.; Niu, C.M. Sulfur and Water Resistance of Mn-Based Catalysts for Low-Temperature Selective Catalytic Reduction of NOx: A Review. Catalysts 2018, 8, 11. [Google Scholar] [CrossRef]
  17. Putluru, S.S.R.; Schill, L.; Godiksen, A.; Poreddy, R.; Mossin, S.; Jensen, A.D.; Fehrmann, R. Promoted V2O5/TiO2 catalysts for selective catalytic reduction of NO with NH3 at low temperatures. Appl. Catal. B Environ. 2016, 183, 282–290. [Google Scholar] [CrossRef]
  18. Gil, S.; Garcia-Vargas, J.M.; Liotta, L.F.; Pantaleo, G.; Ousmane, M.; Retailleau, L.; Giroir-Fendler, A. Catalytic Oxidation of Propene over Pd Catalysts Supported on CeO2, TiO2, Al2O3 and M/Al2O3 Oxides (M = Ce, Ti, Fe, Mn). Catalysts 2015, 5, 671–689. [Google Scholar] [CrossRef]
  19. Liu, L.; Gao, X.; Song, H.; Zheng, C.H.; Zhu, X.B.; Luo, Z.Y.; Ni, M.J.; Cen, K.F. Study of the Promotion Effect of Iron on Supported Manganese Catalysts for No Oxidation. Aerosol Air Qual. Res. 2014, 14, 1038–1046. [Google Scholar] [CrossRef]
  20. Shi, J.; Zhang, Z.H.; Chen, M.X.; Zhang, Z.X.; Shangguan, W.F. Promotion effect of tungsten and iron co-addition on the catalytic performance of MnOx/TiO2 for NH3-SCR of NOx. Fuel 2017, 210, 783–789. [Google Scholar] [CrossRef]
  21. Zhu, L.; Zhong, Z.P.; Yang, H.; Wang, C.H. NH3-SCR Performance of Mn-Fe/TiO2 Catalysts at Low Temperature in the Absence and Presence of Water Vapor. Water Air Soil Poll. 2016, 227, 476. [Google Scholar] [CrossRef]
  22. Deng, S.C.; Zhuang, K.; Xu, B.L.; Ding, Y.H.; Yu, L.; Fan, Y.N. Promotional effect of iron oxide on the catalytic properties of Fe-MnOx/TiO2 (anatase) catalysts for the SCR reaction at low temperatures. Catal. Sci. Technol. 2016, 6, 1772–1778. [Google Scholar] [CrossRef]
  23. Dong, L.F.; Fan, Y.M.; Ling, W.; Yang, C.; Huang, B.C. Effect of Ce/Y Addition on Low-Temperature SCR Activity and SO2 and H2O Resistance of MnOx/ZrO2/MWCNTs Catalysts. Catalysts 2017, 7, 181. [Google Scholar] [CrossRef]
  24. Mu, W.T.; Zhu, J.; Zhang, S.; Guo, Y.Y.; Su, L.Q.; Li, X.Y.; Li, Z. Novel proposition on mechanism aspects over Fe-Mn/ZSM-5 catalyst for NH3-SCR of NOx at low temperature: Rate and direction of multifunctional electron-transfer-bridge and in situ DRIFTs analysis. Catal. Sci. Technol. 2016, 6, 7532–7548. [Google Scholar] [CrossRef]
  25. Chen, T.; Guan, B.; Lin, H.; Zhu, L. In situ DRIFTS study of the mechanism of low temperature selective catalytic reduction over manganese-iron oxides. Chin. J. Catal. 2014, 35, 294–301. [Google Scholar] [CrossRef]
  26. Zhou, G.Y.; Zhong, B.C.; Wang, W.H.; Guan, X.J.; Huang, B.C.; Ye, D.Q.; Wu, H.J. In situ DRIFTS study of NO reduction by NH3 over Fe-Ce-Mn/ZSM-5 catalysts. Catal. Today 2011, 175, 157–163. [Google Scholar] [CrossRef]
  27. Lin, Q.C.; Li, J.H.; Ma, L.; Hao, J.M. Selective catalytic reduction of NO with NH3 over Mn-Fe/USY under lean burn conditions. Catal. Today 2010, 151, 251–256. [Google Scholar] [CrossRef]
  28. Zhang, K.; Yu, F.; Zhu, M.Y.; Dan, J.M.; Wang, X.G.; Zhang, J.L.; Dai, B. Enhanced Low Temperature NO Reduction Performance via MnOx-Fe2O3/Vermiculite Monolithic Honeycomb Catalysts. Catalysts 2018, 8, 100. [Google Scholar] [CrossRef]
  29. Liu, F.D.; He, H.; Ding, Y.; Zhang, C.B. Effect of manganese substitution on the structure and activity of iron titanate catalyst for the selective catalytic reduction of NO with NH3. Appl. Catal. B Environ. 2009, 93, 194–204. [Google Scholar] [CrossRef]
  30. Shu, Y.; Sun, H.; Quan, X.; Chen, S.O. Enhancement of Catalytic Activity Over the Iron-Modified Ce/TiO2 Catalyst for Selective Catalytic Reduction of NOx with Ammonia. J. Phys. Chem. C 2012, 116, 25319–25327. [Google Scholar] [CrossRef]
  31. Yang, S.J.; Xiong, S.C.; Liao, Y.; Xiao, X.; Qi, F.H.; Peng, Y.; Fu, Y.W.; Shan, W.P.; Li, J.H. Mechanism of N2O Formation during the Low-Temperature Selective Catalytic Reduction of NO with NH3 over Mn-Fe Spinel. Environ. Sci. Technol. 2014, 48, 10354–10362. [Google Scholar] [CrossRef] [PubMed]
  32. Lee, T.Y.; Liou, S.Y.; Bai, H.L. Comparison of titania nanotubes and titanium dioxide as supports of low-temperature selective catalytic reduction catalysts under sulfur dioxide poisoning. J. Air Waste Manag. 2017, 67, 292–305. [Google Scholar] [CrossRef] [PubMed]
  33. Tuenter, G.; Vanleeuwen, W.F.; Snepvangers, L.J.M. Kinetics and Mechanism of the NOx Reduction with NH3 on V2O5-WO3-TiO2 Catalyst. Ind. Eng. Chem. Prod. Res. Dev. 1986, 25, 633–636. [Google Scholar] [CrossRef]
  34. Kang, M.; Park, E.D.; Kim, J.M.; Yie, J.E. Manganese oxide catalysts for NOx reduction with NH3 at low temperatures. Appl. Catal. A Gen. 2007, 327, 261–269. [Google Scholar] [CrossRef]
  35. Ruggeri, M.P.; Grossale, A.; Nova, I.; Tronconi, E.; Jirglova, H.; Sobalik, Z. FTIR in situ mechanistic study of the NH3-NO/NO2 “Fast SCR” reaction over a commercial Fe-ZSM-5 catalyst. Catal. Today 2012, 184, 107–114. [Google Scholar] [CrossRef]
  36. Park, K.H.; Lee, S.M.; Kim, S.S.; Kwon, D.W.; Hong, S.C. Reversibility of Mn Valence State in MnOx/TiO2 Catalysts for Low-temperature Selective Catalytic Reduction for NO with NH3. Catal. Lett. 2013, 143, 246–253. [Google Scholar] [CrossRef]
  37. Cao, F.; Xiang, J.; Su, S.; Wang, P.Y.; Hu, S.; Sun, L.S. Ag modified Mn-Ce/gamma-Al2O3 catalyst for selective catalytic reduction of NO with NH3 at low-temperature. Fuel Process. Technol. 2015, 135, 66–72. [Google Scholar] [CrossRef]
  38. Xu, H.D.; Fang, Z.T.; Cao, Y.; Kong, S.; Lin, T.; Gong, M.C.; Chen, Y.Q. Influence of Mn/(Mn plus Ce) Ratio of MnOx-CeO2/WO3-ZrO2 Monolith Catalyst on Selective Catalytic Reduction of NOx with Ammonia. Chin. J. Catal. 2012, 33, 1927–1937. [Google Scholar] [CrossRef]
  39. Liu, F.D.; Asakura, K.; He, H.; Shan, W.P.; Shi, X.Y.; Zhang, C.B. Influence of sulfation on iron titanate catalyst for the selective catalytic reduction of NOx with NH3. Appl. Catal. B Environ. 2011, 103, 369–377. [Google Scholar] [CrossRef]
  40. Shu, Y.; Aikebaier, T.; Quan, X.; Chen, S.; Yu, H.T. Selective catalytic reaction of NOx with NH3 over Ce-Fe/TiO2-loaded wire-mesh honeycomb: Resistance to SO2 poisoning. Appl. Catal. B Environ. 2014, 150, 630–635. [Google Scholar] [CrossRef]
  41. Zhang, L.F.; Zhang, X.L.; Lv, S.S.; Wu, X.P.; Wang, P.M. Promoted performance of a MnOx/PG catalyst for low-temperature SCR against SO2 poisoning by addition of cerium oxide. RSC Adv. 2015, 5, 82952–82959. [Google Scholar] [CrossRef]
  42. Pourkhalil, M.; Moghaddam, A.Z.; Rashidi, A.; Towfighi, J.; Mortazavi, Y. Preparation of highly active manganese oxides supported on functionalized MWNTs for low temperature NOx reduction with NH3. Appl. Surf. Sci. 2013, 279, 250–259. [Google Scholar] [CrossRef]
  43. Wang, X.B.; Gui, K.T. Fe2O3 particles as superior catalysts for low temperature selective catalytic reduction of NO with NH3. J. Environ. Sci. 2013, 25, 2469–2475. [Google Scholar] [CrossRef]
  44. Jin, R.B.; Liu, Y.; Wang, Y.; Cen, W.L.; Wu, Z.B.; Wang, H.Q.; Weng, X.L. 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]
  45. Shen, B.X.; Zhang, X.P.; Ma, H.Q.; Yao, Y.; Liu, T. A comparative study of Mn/CeO2, Mn/ZrO2 and Mn/Ce-ZrO2 for low temperature selective catalytic reduction of NO with NH3 in the presence of SO2 and H2O. J. Environ. Sci. China 2013, 25, 791–800. [Google Scholar] [CrossRef]
  46. Lee, T.; Bai, H. Metal Sulfate Poisoning Effects over MnFe/TiO2 for Selective Catalytic Reduction of NO by NH3 at Low Temperature. Ind. Eng. Chem. Res. 2018, 57, 4848–4858. [Google Scholar] [CrossRef]
  47. Busca, G.; Larrubia, M.A.; Arrighi, L.; Ramis, G. Catalytic abatement of NOx: Chemical and mechanistic aspects. Catal. Today 2005, 107–108, 139–148. [Google Scholar] [CrossRef]
  48. Kapteijn, F.; Singoredjo, L.; Andreini, A.; Moulijn, J.A. Activity and Selectivity of Pure Manganese Oxides in the Selective Catalytic Reduction of Nitric-Oxide with Ammonia. Appl. Catal. B Environ. 1994, 3, 173–189. [Google Scholar] [CrossRef]
  49. Wang, Y.L.; Li, X.X.; Zhan, L.; Li, C.; Qiao, W.M.; Ling, L.C. Effect of SO2 on Activated Carbon Honeycomb Supported CeO2-MnOx Catalyst for NO Removal at Low Temperature. Ind. Eng. Chem. Res. 2015, 54, 2274–2278. [Google Scholar] [CrossRef]
  50. Yang, W.W.; Liu, F.D.; Xie, L.J.; Lan, Z.H.; He, H. Effect of V2O5 Additive on the SO2 Resistance of a Fe2O3/AC Catalyst for NH3-SCR of NOx at Low Temperatures. Ind. Eng. Chem. Res. 2016, 55, 2677–2685. [Google Scholar] [CrossRef]
  51. Qiu, L.; Wang, Y.; Pang, D.D.; Ouyang, F.; Zhang, C.L.; Cao, G. Characterization and Catalytic Activity of Mn-Co/TiO2 Catalysts for NO Oxidation to NO2 at Low Temperature. Catalysts 2016, 6, 9. [Google Scholar] [CrossRef]
  52. Lippits, M.J.; Gluhoi, A.C.; Nieuwenhuys, B.E. A comparative study of the selective oxidation of NH3 to N2 over gold, silver and copper catalysts and the effect of addition of Li2O and CeOx. Catal. Today 2008, 137, 446–452. [Google Scholar] [CrossRef]
Figure 1. N-balance and outlet gas concentrations in the NH3 oxidation reaction without SO2 over MnFe/TiO2 catalyst at different reaction temperatures. Reaction conditions: [NH3] = 500 ppm, [O2] = 10%, balanced with N2, GHSV = 50,000 h−1.
Figure 1. N-balance and outlet gas concentrations in the NH3 oxidation reaction without SO2 over MnFe/TiO2 catalyst at different reaction temperatures. Reaction conditions: [NH3] = 500 ppm, [O2] = 10%, balanced with N2, GHSV = 50,000 h−1.
Catalysts 09 00265 g001
Figure 2. N-balance and outlet gas concentrations in the NO oxidation reaction without SO2 over MnFe/TiO2 catalyst at different reaction temperatures. Reaction conditions: [NO] = 500 ppm, [O2] = 10%, balanced with N2, GHSV = 50,000 h−1.
Figure 2. N-balance and outlet gas concentrations in the NO oxidation reaction without SO2 over MnFe/TiO2 catalyst at different reaction temperatures. Reaction conditions: [NO] = 500 ppm, [O2] = 10%, balanced with N2, GHSV = 50,000 h−1.
Catalysts 09 00265 g002
Figure 3. NO consumption and outlet gas concentrations in the NH3-SCR reaction without SO2 over MnFe/TiO2 catalyst at different reaction temperatures. Reaction conditions: Reaction temperature = 100~300 °C, [NO] = 500 ppm, [NH3] = 500 ppm, [O2] = 10%, balanced with N2, and GHSV = 50,000 h−1.
Figure 3. NO consumption and outlet gas concentrations in the NH3-SCR reaction without SO2 over MnFe/TiO2 catalyst at different reaction temperatures. Reaction conditions: Reaction temperature = 100~300 °C, [NO] = 500 ppm, [NH3] = 500 ppm, [O2] = 10%, balanced with N2, and GHSV = 50,000 h−1.
Catalysts 09 00265 g003
Figure 4. NO consumption and outlet gas concentrations in the NH3-SCR reaction with SO2 over MnFe/TiO2 catalyst at (a) 150 °C and (b) 300 °C. Reaction conditions: [NO] = 500 ppm, [NH3] = 500 ppm, [SO2] = 150 ppm, [O2] = 10%, balanced with N2, and GHSV = 50,000 h−1.
Figure 4. NO consumption and outlet gas concentrations in the NH3-SCR reaction with SO2 over MnFe/TiO2 catalyst at (a) 150 °C and (b) 300 °C. Reaction conditions: [NO] = 500 ppm, [NH3] = 500 ppm, [SO2] = 150 ppm, [O2] = 10%, balanced with N2, and GHSV = 50,000 h−1.
Catalysts 09 00265 g004aCatalysts 09 00265 g004b
Figure 5. DTG spectra of fresh catalyst and catalysts poisoned at different reaction temperatures after 6 h poisoning time.
Figure 5. DTG spectra of fresh catalyst and catalysts poisoned at different reaction temperatures after 6 h poisoning time.
Catalysts 09 00265 g005
Figure 6. NO consumption and outlet NH3 concentrations in the NH3-SCR reaction with SO2 over MnFe/TiO2 catalyst at different ammonium amounts. Reaction conditions: Reaction temperature = 150 °C, [NO] = 500 ppm, [NH3] = 500~650 ppm, [SO2] = 150 ppm, [O2] = 10%, balanced with N2, and GHSV = 50,000 h−1.
Figure 6. NO consumption and outlet NH3 concentrations in the NH3-SCR reaction with SO2 over MnFe/TiO2 catalyst at different ammonium amounts. Reaction conditions: Reaction temperature = 150 °C, [NO] = 500 ppm, [NH3] = 500~650 ppm, [SO2] = 150 ppm, [O2] = 10%, balanced with N2, and GHSV = 50,000 h−1.
Catalysts 09 00265 g006
Figure 7. Product and byproduct percentages of N-containing species during the SCR process over MnFe/TiO2 catalyst with SO2 poisoning at (a) 150 °C and (b) 300 °C. The inlet NH3 and NO molar concentration ratio was 1:1; and the percentage calculation formulas were based on those listed in Table 3.
Figure 7. Product and byproduct percentages of N-containing species during the SCR process over MnFe/TiO2 catalyst with SO2 poisoning at (a) 150 °C and (b) 300 °C. The inlet NH3 and NO molar concentration ratio was 1:1; and the percentage calculation formulas were based on those listed in Table 3.
Catalysts 09 00265 g007aCatalysts 09 00265 g007b
Figure 8. Proposed reaction pathway in the NH3-SCR system with/without SO2 at 150 °C and 300 °C. The reaction pathway was based on inlet NH3 and NO molar concentration ratio of 1:1.
Figure 8. Proposed reaction pathway in the NH3-SCR system with/without SO2 at 150 °C and 300 °C. The reaction pathway was based on inlet NH3 and NO molar concentration ratio of 1:1.
Catalysts 09 00265 g008
Table 1. Amounts of ammonium salts and metal sulfates deposited on the catalyst surfaces.
Table 1. Amounts of ammonium salts and metal sulfates deposited on the catalyst surfaces.
MnFe/TiO2 Poisoned with Different Reaction TemperaturesAmmonium Salts a (by weight) %Metal Sulfates b (by Weight) %
Fresh0.00.0
100 °C2.34.21 c
150 °C1.84.11 c
200 °C1.14.03 c
250 °C0.14.08 c
300 °C0(−0.1)4.52 c
a The amount of ammonium salts was calculated by weight difference between the fresh and poisoned catalysts from the TGA spectrum of 200–400 °C. b The amount of metal sulfates was calculated by weight difference between the fresh and poisoned catalysts from the TGA spectrum of 670–900 °C. c Rigid quantification of the amounts of metal sulfates is not possible via the DTG data because it was possible that some of the ammonium salts could transform into metal sulfates during the heating process of TGA.
Table 2. Formulas for calculating the N-containing product and byproduct percentages during the SCR process.
Table 2. Formulas for calculating the N-containing product and byproduct percentages during the SCR process.
Product and ByproductReaction EquationPercentage Calculation Equation
NO2NO oxidation to NO2 (Equation (5)):
2NO + O2 → 2NO2
The percentage of NO2 from NO oxidation
[ NO 2 ] NO   ( % )   =   [   [ NO 2 ] out [ NO ] in + [ NH 3 ] in   ]   × 100 %
Salts
(NH4)2SO4
NH4HSO4
Ammonium salts
Equations (10) and (11):
SO3 + 2NH3 + H2O (NH4)2SO4
SO3 + NH3 + H2O → NH4HSO4
The percentage of salts
[ Salt ] ( % )   = [   [ NH 3 ] Salt #   [ NO ] in + [ NH 3 ] in   ]   × 100 %
#the NH3 consumption of salts reaction ( [ NH 3 ] Salt ) were indicated by result of Figure 6
N2ONH3 slip oxidation to form N2O
[N2O] slip (%)
The percentage of N2O from NH3 slip oxidation
[ N 2 O ]   slip   ( % )   =   [   [ NH 3 ] out   ×   [ N 2 O ] [ NH 3 ] *   [ NO ] in + [ NH 3 ] in   ]   × 100 %
* [ N 2 O ] [ NH 3 ] : from results of NH3 oxidation test (Figure 1)
NH3 oxidation to form N2O
Equation (3):
2NH3 + 2O2 → N2O + 3H2O
The percentage of N2O from NH3 oxidation
[ N 2 O ] NH 3   ( % )   =   1 [ 2 [ NO ] SCR [ NO ] in + [ NH 3 ] in ] × 100 % [ NO 2 ] NO ( % ) [ Salt ] ( % ) + [ N 2 O ] Slip ( % )
SCR reaction to form N2O
Equation (7):
4NO + 4NH3 + 3O2 → 4N2O + 6H2O
The percentage of N2O from SCR reaction
[ N 2 O ] SCR   ( % )   =   [   2 [ N 2 O ] out [ NO ] in + [ NH 3 ] in ]   × 100 % [ N 2 O ] NH 3 ( % )
N2SCR reaction to form N2
Equation (6):
4NO + 4NH3 + O2 → 4N2 + 6H2O
The percentage of N2 from SCR reaction
[ N 2 ] SCR   ( % )   =   [   2 [ NO ] SCR [ NO ] in + [ NH 3 ] in ]   × 100 % [ N 2 O ] SCR ( % )
Table 3. Operation conditions of experimental tests in this study.
Table 3. Operation conditions of experimental tests in this study.
NH3 (ppmv)NO (ppmv)SO2 (ppmv)O2 (%)Temperature (°C)GHSV (hr−1)
NH3 oxidation test5000010%100–30050,000
NO oxidation test0500010%100–30050,000
SCR test without SO2500500010%100–30050,000
SCR test with SO250050015010%150 & 30050,000
SCR test with SO2 at different NH3 amounts500–60050015010%15050,000

Share and Cite

MDPI and ACS Style

Lee, T.; Bai, H. Byproduct Analysis of SO2 Poisoning on NH3-SCR over MnFe/TiO2 Catalysts at Medium to Low Temperatures. Catalysts 2019, 9, 265. https://doi.org/10.3390/catal9030265

AMA Style

Lee T, Bai H. Byproduct Analysis of SO2 Poisoning on NH3-SCR over MnFe/TiO2 Catalysts at Medium to Low Temperatures. Catalysts. 2019; 9(3):265. https://doi.org/10.3390/catal9030265

Chicago/Turabian Style

Lee, Tsungyu, and Hsunling Bai. 2019. "Byproduct Analysis of SO2 Poisoning on NH3-SCR over MnFe/TiO2 Catalysts at Medium to Low Temperatures" Catalysts 9, no. 3: 265. https://doi.org/10.3390/catal9030265

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