Next Article in Journal
Controlled Synthesis of Heterostructured SnO2-CuO Composite Hollow Microspheres as Efficient Cu-Based Catalysts for the Rochow Reaction
Next Article in Special Issue
Mobility of NH3-Solvated CuII Ions in Cu-SSZ-13 and Cu-ZSM-5 NH3-SCR Catalysts: A Comparative Impedance Spectroscopy Study
Previous Article in Journal
Continuous Dimethyl Carbonate Synthesis from CO2 and Methanol Using Cu-Ni@VSiO as Catalyst Synthesized by a Novel Sulfuration Method
Previous Article in Special Issue
Recent Progress in Atomic-Level Understanding of Cu/SSZ-13 Selective Catalytic Reduction Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

DRIFT Study on Promotion Effect of the Keggin Structure over V2O5-MoO3/TiO2 Catalysts for Low Temperature NH3-SCR Reaction

1
Laboratory of Catalysis Chemistry and Nanoscience, Department of Chemistry and Chemical Engineering, College of Environmental and Energy Engineering, Beijing University of Technology, Beijing 100124, China
2
Collaborative Innovation Center of Electric Vehicles in Beijing, Beijing 100081, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally.
Catalysts 2018, 8(4), 143; https://doi.org/10.3390/catal8040143
Submission received: 10 March 2018 / Revised: 30 March 2018 / Accepted: 30 March 2018 / Published: 3 April 2018
(This article belongs to the Special Issue Selective Catalytic Reduction of NOx)

Abstract

:
Heteropoly acids (HPAs) with the Keggin structure have been widely used in NOx removal. Two kinds of catalysts (those with and without the Keggin structure) are prepared for studying the effect of the Keggin structure on the NH3-SCR reaction. A series of in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFT) analyses are conducted to investigate the surface-adsorbed species on the catalysts during the SCR reaction. The mechanism for enhancing low-temperature activity of the catalysts is proposed. Furthermore, the effect of NH4+ in the Keggin structure is also investigated. Results indicate that both the Langmuir–Hinshelwood (L-H) and Eley–Rideal (E-R) mechanisms occurred in the NH3-SCR reaction over the catalyst with the Keggin structure (Cat-A); in addition, when more acid sites are provided, NOx species activity is improved and more NH4+ ions participate in reaction over Cat-A, thus promoting SCR activity.

Graphical Abstract

1. Introduction

NOx is an air pollutant that can result in a variet of pollution, such acid rain, optical chemical smog, and the greenhouse effect. NOx is emitted when using fossil fuels in industrial facilities and automobiles [1,2]. Selective catalytic reduction (SCR) has been extensively used to remove NOx emissions from stationary sources such as coal-fired boilers [3]. The commercialized catalyst used in SCR has mainly been V2O5-WO3/TiO2 (or V2O5-MoO3/TiO2), which has high activity and good SO2 tolerance. However, it has a very narrow operating temperature (300–400 °C), high conversion of SO2 to SO3, and low N2 selectivity at high temperatures [4,5]. Therefore, developing an NH3-SCR catalyst with good catalytic performance at low temperature is desirable.
Heteropoly acids (HPAs) have unique structure and extraordinarily strong acidity, and, as such, HPAs have attracted the attention of many researchers [6,7]. The main kind of HPAs used in catalytic applications is Keggin HPAs, such as H3PW12O40, H3PMo12O40, and H3SiW12O40 [8]. It has been found that NO can be effectively absorbed on 12-tungstophosphoric acid at 150 °C and decomposed at about 450 °C upon rapid heating. It has been reported that small polar molecules (such as NO, NH3, and H2O) can access the secondary structure of Keggin HPAs and that SO2 cannot influence adsorption of NO. Supplementary experiments showed that NO can replace the structural water present between Keggin units of HPAs [9,10]. Later, some researchers used aromatic hydrocarbons loaded with Pt and Pd to improve the activity of NO reduction [11,12]. Putluru et al. reported that HPAs can promote SCR activity and shows excellent alkali deactivation resistance [13]. Weng et al. used 12-tungstaphosphoric acid loaded on CeO2 and found that the catalyst had significantly improved performance in the SCR reaction and improved SO2 poisoning resistance [14]. Ammonium salts are more available than HPAs, and using 12-tungstophosphoric acid ammonium salt for NOx removal has been reported. Moffat et al. found an interesting phenomenon; specifically, that the ammonium salt of 12-tungstophosphoric acid exhibited better NOx removal capacity in the case without NH3 because bound NH4+ on the solid can react with absorbed NO2 to produce N2, and ammonium salt can be regenerated from gaseous NH3 [9]. However, there are few reports about using ammonium salts of HPAs for SCR, and further research on the effect of the Keggin structure and that of the NH4+ of ammonium salts on SCR is lacking.
In this research, V2O5-(NH4)3PMo12O40/TiO2 and a conventional oxide catalyst are prepared for studying the effect of the Keggin structure on the NH3-SCR reaction. A series of in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFT) analyses are conducted to investigate surface-adsorbed species. Fourier Transform infrared spectroscopy (FT-IR) is used to evaluate the effect of NH4+ of ammonium salts on the catalysts.

2. Results and Discussion

2.1. Catalytic Activity

Catalytic performances of the catalysts with and without the Keggin structure at elevated temperature are shown in Figure 1. Structures of the two samples were determined using X-ray diffraction (XRD) and Raman spectroscopy (Figures S1 and S2, respectively). X-ray Fluorescence Spectrometer (XRF) results of elemental analysis and the BET specific surface area data are shown in Table S1. Results show that Cat-A, which has the Keggin structure, had the same composition and BET surface area as Cat-B, which has an oxide phase. As shown in Figure 1, Cat-A had much higher SCR activity than Cat-B. NO conversion over Cat-A reached 93% at 220 °C and showed no obvious decline until 350 °C. In contrast, NO conversion over Cat-B was only 67% at 220 °C. Clearly, the catalyst with the Keggin structure had improved NO conversion in low temperature. Both catalysts had excellent N2 selectivity of 99% below 250 °C. The time-on-stream stability was investigated (Figure S5). The NOx conversion over Cat-A decreased from 100% to 95% in 20 h, while, for Cat-B, it decreased from 99% to 92%. Therefore, Cat-A exhibited better time-on-stream stability than Cat-B.

2.2. Adsorption Behaviors of Reactants on the Surface of the Catalysts (In Situ DRIFT)

2.2.1. NH3 Adsorption on the Surface of the Catalysts

In situ DRIFT spectra of NH3 adsorption on Cat-A and Cat-B at 250 °C are shown in Figure 2. As seen in Figure 2A, there were six bands: 1211, 1391, 1454, 1596, 1685, and 1737 cm−1. The peaks centered at 1454 cm−1 and in the range of 1750–1650 cm−1 can be, respectively, attributed to asymmetric and symmetric bending (δas and δs) vibrations of NH4+ species on Brönsted acid sites [15,16,17]. The bands at 1211 and 1391 cm−1 can be assigned to N-H bonds in NH3 that is chemisorbed on Lewis acid sites [18].
As illustrated in Figure 2B, DRIFT spectra for NH3 adsorption over Cat-B were similar to those for NH3 adsorption over Cat-A. A broad absorption band at 1448 cm−1 was ascribed to δas NH4+ vibration and the bands at 1750–1650 cm−1 were due to δs NH4+ vibration [19]. Coordinated δs NH3 (1230 and 1394 cm−1) and δas NH3 species (1605 cm−1) on Lewis acid sites were also observed [20,21]. It is evident that the intensity of the band assigned to Lewis acid sites (1230 cm−1) is much weaker than that over Cat-A. It is known that HPAs have stronger (Brönsted) acidity, but it was reported that (NH4)3PMo12O40 provides Lewis acid sites that were strong at high temperature because of the oxygen-deficient Keggin structure [22]. The NH3-desorption peaks in the temperature programmed desorption of ammonia (NH3-TPD) was interconnected to the concentration of acid sites [23]. Combined with the results for NH3-TPD (Figure S3), the peak area of Cat-A was more than Cat-B, which suggested that the concentration of Lewis acid sites was more than Cat-B. As reported, the high amount of Lewis acid site could improve the low temperature activity [24]. Therefore, the Keggin structure not only provides more Brönsted acid sites, but also offered Lewis acid sites that are beneficial for low-temperature SCR catalytic activity.

2.2.2. Co-Adsorption of NO + O2 on the Surface of the Catalysts

Figure 3 shows DRIFT results of NO + O2 co-adsorption on the surface of Cat-A and Cat-B. Adsorbed species on the two catalysts are clearly very different. As seen in Figure 3A, there are several distinct bands at 1347, 1513, 1640, 1668, 1786, 1751, 1860, and 1976 cm−1. A broad band at 1347 cm−1 appeared first and can be ascribed to free nitrate ions [25]. Later, smaller bands at 1513, 1640, and 1668 cm−1 were observed, and these can be assigned to bidentate nitrate, adsorbed NO2, and adsorbed N2O4, respectively [20,26]. Bands at 1751 and 1786 cm−1 can be ascribed to adsorption of trans-(NO)2 [25,27]. A weaker band at 1860 cm−1 can be assigned to surface bound NO [28]. The band at 1976 cm−1 increased with time and can be ascribed to NO in the secondary structure of Keggin anions [28]. For Cat-B (Figure 3B), a series of peaks can be ascribed to monodentate nitrates (1317 and 1490 cm−1), free nitrate ions (1340 cm−1), and adsorbed NO2 (1647 cm−1) [29,30]. For Cat-B, there were fewer kinds of adsorbed NOx species than for Cat-A. Over Cat-B, there were more monodentate nitrates than free nitrate ions. It is reported that, for NSR, NOx was primarily trapped by the metal ions (such as K or Li) in free nitrate ions, thus leading to increased trapping capacity of NOx [31,32]. It is likely that NH4+ in Cat-A plays the same role in trapping NOx, and this indicates that the NOx storage behavior of Cat-A is better than that of Cat-B.

2.2.3. NO + O2 Adsorption on the Surface of the Catalysts after NH3 Pre-Adsorption

The reaction of NO + O2 with pre-adsorbed NH3 at 250 °C on the surface of both catalysts was investigated using DRIFT to understand the reactivity of NH3 adsorbed on the catalysts. As seen in Figure 4A, several peaks that correspond to coordinated NH3 (L) and NH4+ (B) species on the surface of Cat-A are observed. When NO + O2 was added, Lewis and Brönsted acidity decreased gradually within 10 min, and, after 20 min, several bands (1349, 1514, 1638, 1786, and 1944 cm−1) that correspond to adsorbed NOx species appeared and grew with time. The adsorption behaviors of Cat-B (Figure 4B) were similar to those of Cat-A. Over time, all of the peaks gradually disappeared, accompanied by the appearance of nitrate species. These results imply that coordinated NH3 and NH4+ species participate in the NH3-SCR reaction and that both were active species.

2.2.4. NH3 Adsorption on the Surface of the Catalysts after NO + O2 Pre-Adsorption

The catalysts were pre-treated with NO + O2 at 250 °C for 60 min, and then NH3 was introduced to investigate the reaction of NH3 with nitrate species over the catalysts. As seen in Figure 5A, several peaks that correspond to nitrate species were found in the spectra after treatment with NO + O2. With the introduction of NH3, the amounts of free nitrate ions (1347 cm−1), bidentate nitrates (1513 cm−1), NO2 (1640 cm−1), and N2O4 (1668 cm−1) decreased quickly within 1 min. Surface NO (1840 cm−1) and NOH+ (1976 cm−1) in the Keggin structure gradually decreased in 5 min, implying that NO in the secondary structure of Keggin anions participates in the reaction. However, the peak at 1786 cm−1, which corresponds to trans-(NO)2, is still present after purging with NH3 for 30 min, and this suggests that the Eley–Rideal (E-R) mechanism occurred [26]. For Cat-B (Figure 5B), a series of peaks that correspond to nitrate species were observed in the spectra. After NH3 was introduced for 1 min, NO2 (1647 cm−1) disappeared, and this can enhance the process of the “fast-SCR” reaction. Compared with Cat-A, free nitrate ions of Cat-B decreased more slowly and monodentate nitrates disappeared gradually within 5 min. It has been reported that monodentate nitrates and NO2 molecules had higher activity in the SCR reaction than free nitrate ions. In our results, free nitrate ions for Cat-A, which had the Keggin structure, also easily participated in the reaction, and this was possibly because NOx was trapped by the NH4+ of the Keggin structure in free nitrate ions, such as in the case of NOx storage/reduction (NSR) catalysts where NOx can easily be trapped.

2.3. Effect of NH4+ in (NH3)4PMo12O40 for NH3-SCR

It is known that NH3 adsorption and storage is important for NH3-SCR [33,34]. Compared with oxide SCR catalysts, Cat-A, which has the Keggin structure, has many NH4+ ions. Moffat et al. reported that NH4+ of ammonium 12-tungstophosphate can interact with NO2 and NO and can be regenerated from ammonia gas at 150 °C [9]. Hence, it is essential to know whether NH4+ of the Keggin structure participates in NH3-SCR and if NH4+ can be regenerated for Cat-A.
If NH4+ of the Keggin structure participates in the reaction, the SCR reaction on Cat-A would continue longer than that on Cat-B without NH3. To verify this assumption, NO conversion of Cat-A and Cat-B after cutting off NH3 was investigated, and the results are shown in Figure 6. As seen in Figure 6, NO conversions on the two catalysts decreased after cutting off NH3. For Cat-A, NO conversion remained 100% for 14 min, whereas NO conversion over Cat-B kept remained 100% for only 4 min. Thus, it is obvious that Cat-A, which has the Keggin structure, had higher NH3 storage than Cat-B, and this suggests that NH4+ of the Keggin structure participates in the reaction.
Change in the Keggin structure during successive adsorption of NO + O2 and NH3 was investigated using FT-IR spectra to further verify participation of NH4+ in the SCR reaction. As seen in Figure 7, vibrational bands for NH4+, P-O, and Mo-O-Mo in the Keggin structure are observed at 1402, 1062, and 906 cm−1, respectively. After adsorption of NO, the peak at 1402 cm−1 disappeared. NH3 was added after cutting off NO + O2, and the peak at 1402 cm−1 reappeared. Moreover, the peaks at 1062 and 906 cm−1 remained stable during the whole reaction process, and this suggests that HPAs did not change during the reaction. Thus, it was determined that NH4+ of the Keggin structure participates in the SCR reaction and was regenerated from ammonia gas. Meanwhile, the Keggin structure (of the HPAs) was unchanged in the reaction. In situ Raman spectra of Cat-A were recorded (Figure S4) to further study the stability of the Keggin structure in the SCR reaction. The results indicate that the peaks corresponding to the Keggin structure did not shift, and this suggests that there was no structural perturbation during the reaction.

2.4. Reaction Mechanism

From DRIFT spectra for NH3 adsorption on the two samples, it is observed that NH3 was adsorbed on Brönsted and Lewis acid sites in the forms of NH4+ and NH3, respectively. After NO + O2 was passed over the catalyst surface that was pretreated with NH3, the adsorbed NH3 species gradually vanished in 10 min, and this indicates that both Brönsted and Lewis acid sites were active centers. As shown in DRIFT spectra of NO + O2 adsorption, NO was adsorbed and oxidized to NO2 and other nitrate species on the two samples. There were more kinds of NOx species adsorbed on Cat-A than on Cat-B. Different results were observed when NH3 was passed over the surface of each of the two samples that were pretreated with NO + O2 at 250 °C. Compared with Cat-B, not all of the adsorbed NOx species participated in the reaction over Cat-A, and most of the adsorbed NOx species quickly vanished in 1 min, which indicates that both the E-R and L-H mechanisms occurred [35,36,37,38]. Noticeably, the peaks of NOH+ disappeared after 5 min, and this suggests that NO in the secondary structure of the Keggin anions (NOH+) also participated in reaction. The effect of NH4+ in the Keggin structure for the SCR reaction was investigated, and the results reveal that NH4+ of the Keggin structure participates in the SCR reaction as an NH3 pool and can be regenerated from ammonia gas.
Based on the above discussion, a mechanism for improving the low-temperature activity of Cat-A was proposed; the proposed mechanism is shown in Figure 8 and is described as follows.
Both the E-R and L-H mechanisms occurred in the NH3-SCR reaction. For the E-R mechanism, NH3 was adsorbed on Lewis acid sites and reacted directly with gaseous NO and NO2. This was followed by decomposition into N2 and H2O. For the L-H mechanism, NO was adsorbed, and then most of it was oxidized to nitrate species (NO3 and NO2-adsorbed), which reacted with NH3-adsorbed species. It is noted that the catalyst with Keggin structure had more Brönsted acid Lewis acid sites than the catalyst with oxide phase. In addition, NH4+ of the Keggin structure reacted by the L-H mechanism and was regenerated during the reaction; thus, low-temperature SCR activity was promoted via the L-H mechanism.
For Cat-B, the mechanisms were similar to Cat-A. Both the E-R and L-H mechanisms also occurred in the reaction. For the E-R mechanism, gaseous NO and NO2 could react with NH3 on Lewis acid sites. For the L-H mechanism, NO could be adsorbed and then oxidized to nitrate species. Most of the nitrate species were monodentate nitrates and NO2-adsorbed not free NO3, and the nitrate species could react with NH3-adsorbed species. The results were in agreement with many SCR catalysts with oxide phase [25,26,39].

3. Experimental

3.1. Catalyst Preparation

Two kinds of catalysts were prepared via impregnation, and the main chemical compositions were the same. Specifically, (NH4)3PMo12O40 loading was 20 wt % and V2O5 loading was 1 wt %.
The catalyst with the Keggin structure was prepared as follows. NH4H2PO4 (0.6 g, Fuchen, Tianjin, China, 99%) and (NH4)6Mo7O24 (11.5 g, Fuchen, Tianjin, China, 99%) were dissolved in distilled water (100 mL). Solution pH was adjusted to a value of about 1. TiO2 anatase powder (40.0 g, Xinhua, Chongqing, China, 99%) was added to the precursor solution and solid to liquid ratio (g/mL) was 2:5. The mixture was stirred at 80 °C for 5 h and then dried at 120 °C for 3 h. The solid was ground into a powder and calcined at 400 °C for 5 h. NH4VO3 (0.6 g, Fuchen, Tianjin, China, 99%) was then added to a solution of oxalic acid, and the weight of H2C2O4·2H2O (Fuchen, Tianjin, China, 99%) was double that of NH4VO3. Desiccation and calcination conditions were the same as above, and the obtained catalyst was denoted as Cat-A.
Catalysts with an oxide phase were prepared via multiple impregnation methods to avoid generating the Keggin structure. (NH4)6Mo7O24 (11.5 g) was dissolved in 100 mL of distilled water. TiO2 anatase powder was impregnated with the precursor solution and stirred at 80 °C for 5 h. The mixture was then dried at 120 °C and was calcined at 400 °C for 5 h. NH4H2PO4 (0.6 g) was then dissolved in 100 mL of distilled water. V2O5 was loaded using the above conditions, and this sample was denoted as Cat-B.

3.2. Catalytic Activity Test

Catalytic activities of the samples were measured in a fixed bed quartz reactor (9 mm i.d.) with 0.4 mL of catalysts (40–60 mesh). Typical reactant gas was a mixture of 1000 ppm NO, 1000 ppm NH3, 5% O2, 5% water vapor, and 350 ppm SO2 with a balance of N2 under a flow of 500 mL/min. The space velocity was 40,000 h−1. Concentrations of NO and NH3 were continuously detected using a Thermo Scientific 17i NOx chemiluminescence analyzer. N2O was monitored using a Bruker Tensor 27 FTIR spectrometer. The reaction system was maintained at each reaction temperature for 30 min before analysis. The equations used to calculate NOx conversion and N2 selectivity are as follows:
NO   conversion ( % )   =   [ NO ] in ( [ NO ] out + [ NO 2 ] out ) [ NO ] in × 100 %
N 2   selectivity ( % )   =   ( 1 [ NO 2 ] out + 2 [ N 2 O ] out [ NH 3 ] in + [ NO ] in [ NH 3 ] out [ NO ] out   ) × 100 %

4. Conclusions

Two kinds of catalysts (one with and one without the Keggin structure) are synthesized to investigate the effect of the Keggin structure on the NH3-SCR reaction. Cat-A, which had the Keggin structure, exhibits better low-temperature SCR performance. NH3 is adsorbed on both Brönsted and Lewis acid sites over the two catalysts but the catalyst with Keggin structure has more Brönsted and Lewis acid sites. Moreover, there are more kinds of adsorbed NOx species on Cat-A than on Cat-B, and the adsorbed NOx species are mainly free nitrate ions. The two catalysts follow both E-R and L-H mechanisms. Compared with Cat-B, most of the adsorbed NOx species for Cat-A react quickly with gaseous NH3. In addition, NH4+ of the Keggin structure participates in the reaction via the L-H mechanism and is recovered by ammonia gas in the flow. Thereby, more acid sites are provided, adsorbed NOx species activity is improved, and more NH4+ ions participate in the L-H mechanism over Cat-A, which had the Keggin structure, thus promoting SCR activity.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/8/4/143/s1, Characterization methods: XRD and Raman results (Figures S1 and S2); compositions and BET surface areas of Cat-A and Cat-B (Table S1); NH3-TPD results (Figure S3); in situ sequential Raman spectra results (Figure S4); the time-on-stream stability of the catalysts (Figure S5).

Acknowledgments

The work was supported by the National Key R&D Program of China (2017YFC0210303-02) and the National Natural Science Foundation of China (201577005). The authors thank Ran Liu from Shiyanjia lab for assistance of BET and NH3-TPD analysis (www.Shiyanjia.com).

Author Contributions

Rui Wu and Hong He conceived and designed the experiments; Rui Wu experiments; Rui Wu, Ningqiang Zhang and Wenge Qu analyzed the data; Liyun Song and Lingcong Li contributed analysis tools; Rui Wu wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Amiridis, M.D.; Duevel, R.V.; Wachs, I.E. The effect of metal oxide additives on the activity of V2O5/TiO2 catalysts for the selective catalytic reduction of nitric oxide by ammonia. Appl. Catal. B Environ. 1999, 20, 111–122. [Google Scholar] [CrossRef]
  2. 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]
  3. Li, P.; Liu, Q.; Liu, Z. Behaviors of NH4HSO4 in SCR of NO by NH3 over different cokes. Chem. Eng. J. 2012, 181–182, 169–173. [Google Scholar] [CrossRef]
  4. Kamata, H.; Takahashi, K.; Odenbrand, C.I. Surface acid property and its relation to SCR activity of phosphorus added to commercial V2O5 (WO3)/TiO2 catalyst. Catal. Lett. 1998, 53, 65–71. [Google Scholar] [CrossRef]
  5. Putluru, S.S.R.; Schill, L.; Godiksen, A.; Poreddy, R.; Mossin, S.; Jensen, A.D.; Fehrmann, R. Promoted V2O5/TiO2 catalysts for selective catalytic reduction of NO with NH3 at low temperatures. Appl. Catal. B Environ. 2016, 183, 282–290. [Google Scholar] [CrossRef] [Green Version]
  6. Hill, C.L. Foreword: Polyoxometalates in catalysis. J. Mol. Catal. A Chem. 1996, 114, 1. [Google Scholar] [CrossRef]
  7. Chen, Y.; Wang, Y.; Lu, J.; Wu, C. The viscosity reduction of nano-keggin-K3PMo12O40 in catalytic aquathermolysis of heavy oil. Fuel 2009, 88, 1426–1434. [Google Scholar] [CrossRef]
  8. Mizuno, N.; Misono, M. Heterogeneous catalysis. Chem. Rev. 1998, 98, 199–218. [Google Scholar] [CrossRef] [PubMed]
  9. Bélanger, R.; Moffat, J.B. The sorption and reduction of nitrogen oxides by 12-tungstophosphoric acid and its ammonium salt. Catal. Today 1998, 40, 297–306. [Google Scholar] [CrossRef]
  10. Yang, R.T.; Chen, N. A new approach to decomposition of nitric oxide using sorbent/catalyst without reducing gas: Use of heteropoly compounds. Ind. Eng. Chem. Res. 1994, 33, 825–831. [Google Scholar] [CrossRef]
  11. Vaezzadeh, K.; Petit, C.; Pitchon, V. The removal of NOx from a lean exhaust gas using storage and reduction on H3PW12O40·6H2O. Catal. Today 2002, 73, 297–305. [Google Scholar] [CrossRef]
  12. Yoshimoto, R.; Ninomiya, T.; Okumura, K.; Niwa, M. Cooperative effect induced by the mixing of Na-ZSM-5 and Pd/H3PW12O40/SiO2 in the selective catalytic reduction of NO with aromatic hydrocarbons. Appl. Catal. B Environ. 2007, 75, 175–181. [Google Scholar] [CrossRef]
  13. Putluru, S.S.R.; Jensen, A.D.; Riisager, A.; Fehrmann, R. Heteropoly acid promoted V2O5/TiO2 catalysts for NO abatement with ammonia in alkali containing flue gases. Catal. Sci. Technol. 2011, 1, 631–637. [Google Scholar] [CrossRef]
  14. Weng, X.; Dai, X.; Zeng, Q.; Liu, Y.; Wu, Z. DRIFT studies on promotion mechanism of H3PW12O40 in selective catalytic reduction of NO with NH3. J. Colloid Interface Sci. 2016, 461, 9–14. [Google Scholar] [CrossRef] [PubMed]
  15. Lietti, L.; Nova, I.; Ramis, G.; Dall’Acqua, L.; Busca, G.; Giamello, E.; Forzatti, P.; Bregani, F. Characterization and Reactivity of V2O5–MoO3/TiO2 De-NOx SCR Catalysts. J. Catal. 1999, 187, 419–435. [Google Scholar] [CrossRef]
  16. Ettireddy, P.R.; Ettireddy, N.; Boningari, T.; Pardemann, R.; Smirniotis, P.G. Investigation of the selective catalytic reduction of nitric oxide with ammonia over Mn/TiO2 catalysts through transient isotopic labeling and in situ FT-IR studies. J. Catal. 2012, 292, 53–63. [Google Scholar] [CrossRef]
  17. Song, Z.; Zhang, Q.; Ma, Y.; Liu, Q.; Ning, P.; Liu, X.; Wang, J.; Zhao, B.; Huang, J.; Huang, Z. Mechanism-dependent on the different CeO2 supports of phosphotungstic acid modification CeO2 catalysts for the selective catalytic reduction of NO with NH3. J. Taiwan Inst. Chem. E 2017, 71, 277–284. [Google Scholar] [CrossRef]
  18. Yao, X.; Zhang, L.; Li, L.; Liu, L.; Cao, Y.; Dong, X.; Gao, F.; Deng, Y.; Tang, C.; Chen, Z.; et al. Investigation of the structure, acidity, and catalytic performance of CuO/Ti0.95Ce0.05O2 catalyst for the selective catalytic reduction of NO by NH3 at low temperature. Appl. Catal. B Environ. 2014, 150–151, 315–329. [Google Scholar] [CrossRef]
  19. Thirupathi, B.; Smirniotis, P.G. Co-doping a metal (Cr, Fe, Co, Ni, Cu, Zn, Ce, and Zr) on Mn/TiO2 catalyst and its effect on the selective reduction of NO with NH3 at low-temperatures. Appl. Catal. B Environ. 2011, 110, 195–206. [Google Scholar] [CrossRef]
  20. You, Y.; Chang, H.; Zhu, T.; Zhang, T.; Li, X.; Li, J. The poisoning effects of phosphorus on CeO2-MoO3/TiO2 DeNOx catalysts: NH3-SCR activity and the formation of N2O. Mol. Catal. 2017, 439, 15–24. [Google Scholar] [CrossRef]
  21. Zeng, Y.; Zhang, S.; Wang, Y.; Zhong, Q. CeO2 supported on reduced TiO2 for selective catalytic reduction of NO by NH3. J. Colloid Interface Sci. 2017, 496, 487–495. [Google Scholar] [CrossRef] [PubMed]
  22. Li, W.; Oshihara, K.; Ueda, W. Catalytic performance for propane selective oxidation and surface properties of 12-molybdophosphoric acid treated with pyridine. Appl. Catal. A Gen. 1999, 182, 357–363. [Google Scholar] [CrossRef]
  23. Boningari, T.; Ettireddy, P.R.; Somogyvari, A.; Liu, Y.; Vorontsov, A.; McDonald, C.A.; Smirniotis, P.G. Influence of elevated surface texture hydrated titania on Ce-doped Mn/TiO2 catalysts for the low-temperature SCR of NOx under oxygen-rich conditions. J. Catal. 2015, 325, 145–155. [Google Scholar] [CrossRef]
  24. Pappas, D.K.; Boningari, T.; Boolchand, P.; Smirniotis, P.G. Novel manganese oxide confined interweaved titania nanotubes for the low-temperature Selective Catalytic Reduction (SCR) of NOx by NH3. J. Catal. 2016, 334, 1–13. [Google Scholar] [CrossRef]
  25. Liu, Z.; Zhang, S.; Li, J.; Ma, L. Promoting effect of MoO3 on the NOx reduction by NH3 over CeO2/TiO2 catalyst studied with in situ DRIFTS. Appl. Catal. B Environ. 2014, 144, 90–95. [Google Scholar] [CrossRef]
  26. Liu, J.; Li, X.; Zhao, Q.; Ke, J.; Xiao, H.; Lv, X.; Liu, S.; Tadé, M.; Wang, S. Mechanistic investigation of the enhanced NH3-SCR on cobalt-decorated Ce-Ti mixed oxide: In situ FTIR analysis for structure-activity correlation. Appl. Catal. B Environ. 2017, 200, 297–308. [Google Scholar] [CrossRef]
  27. Valyon, J.; Hall, W. Surface Species Formed from NO on Copper Zeolites. ChemInform 1993, 24. [Google Scholar] [CrossRef]
  28. Herring, A.M.; McCormick, R.L. In situ infrared study of the absorption of nitric oxide by 12-tungstophosphoric acid. J. Phys. Chem. B 1998, 102, 3175–3184. [Google Scholar] [CrossRef]
  29. Chen, L.; Li, J.; Ge, M. DRIFT Study on Cerium−Tungsten/Titiania Catalyst for Selective Catalytic Reduction of NOx with NH3. Environ. Sci. Technol. 2010, 44, 9590–9596. [Google Scholar] [CrossRef] [PubMed]
  30. Li, Z.; Li, J.; Liu, S.; Ren, X.; Ma, J.; Su, W.; Peng, Y. Ultra hydrothermal stability of CeO2-WO3/TiO2 for NH3-SCR of NO compared to traditional V2O5-WO3/TiO2 catalyst. Catal. Today 2015, 258, 11–16. [Google Scholar] [CrossRef]
  31. Toops, T.J.; Smith, D.B.; Epling, W.S.; Parks, J.E.; Partridge, W.P. Quantified NOx adsorption on Pt/K/gamma-Al2O3 and the effects of CO2 and H2O. Appl. Catal. B Environ. 2005, 58, 255–264. [Google Scholar] [CrossRef]
  32. Büchel, R.; Strobel, R.; Baiker, A.; Pratsinis, S.E. Flame-Made Pt/K/Al2O3 for NOx Storage–Reduction (NSR) Catalysts. Top. Catal. 2009, 52, 1799–1802. [Google Scholar] [CrossRef]
  33. Kamasamudram, K.; Currier, N.W.; Chen, X.; Yezerets, A. Overview of the practically important behaviors of zeolite-based urea-SCR catalysts, using compact experimental protocol. Catal. Today 2010, 151, 212–222. [Google Scholar] [CrossRef]
  34. Schmieg, S.J.; Oh, S.H.; Kim, C.H.; Brown, D.B.; Lee, J.H.; Peden, C.H.F.; Kim, D.H. Thermal durability of Cu-CHA NH3-SCR catalysts for diesel NOx reduction. Catal. Today 2012, 184, 252–261. [Google Scholar] [CrossRef]
  35. Ramis, G.; Bregani, F.; Forzatti, P. Fourier transform-infrared study of the adsorption and coadsorption of nitric oxide, nitrogen dioxide and ammonia on vanadia-titania and mechanism of selective catalytic reduction. Appl. Catal. 1990, 64, 259–278. [Google Scholar] [CrossRef]
  36. Topsøe, N.-Y. Characterization of the nature of surface sites on vanadia-titania catalysts by FTIR. J. Catal. 1991, 128, 499–511. [Google Scholar] [CrossRef]
  37. Lietti, L.; Svachula, J.; Forzatti, P.; Ramis, G.; Bregani, P. Surface and catalytic properties of Vanadia-Titania and Tungsta-Titania systems in the Selective Catalytic Reduction of nitrogen oxides. Catal. Today 1993, 17, 131–139. [Google Scholar] [CrossRef]
  38. Topsøe, N.-Y. Mechanism of the selective catalytic reduction of nitric oxide by ammonia elucidated by in situ on-line Fourier transform infrared spectroscopy. Science 1994, 265, 1217–1219. [Google Scholar] [CrossRef] [PubMed]
  39. Peng, Y.; Li, K.; Li, J. Identification of the active sites on CeO2–WO3 catalysts for SCR of NOx with NH3: An in situ IR and Raman spectroscopy study. Appl. Catal. B Environ. 2013, 140, 483–492. [Google Scholar] [CrossRef]
Figure 1. NO conversion of the two catalysts. (Reaction conditions: NO 1000 ppm, NH3/NO = 1, O2 5%, H2O 5%, SO2 350 ppm, N2 balance, and GHSV = 40,000 h−1.).
Figure 1. NO conversion of the two catalysts. (Reaction conditions: NO 1000 ppm, NH3/NO = 1, O2 5%, H2O 5%, SO2 350 ppm, N2 balance, and GHSV = 40,000 h−1.).
Catalysts 08 00143 g001
Figure 2. DRIFT spectra for NH3 adsorption at 250 °C on: (A) Cat-A; and (B) Cat-B.
Figure 2. DRIFT spectra for NH3 adsorption at 250 °C on: (A) Cat-A; and (B) Cat-B.
Catalysts 08 00143 g002
Figure 3. DRIFT spectra for NO + O2 adsorption at 250 °C on: (A) Cat-A; and (B) Cat-B.
Figure 3. DRIFT spectra for NO + O2 adsorption at 250 °C on: (A) Cat-A; and (B) Cat-B.
Catalysts 08 00143 g003
Figure 4. DRIFT spectra of NO + O2 reacted with pre-adsorbed NH3 species at 250 °C on: (A) Cat-A; and (B) Cat-B.
Figure 4. DRIFT spectra of NO + O2 reacted with pre-adsorbed NH3 species at 250 °C on: (A) Cat-A; and (B) Cat-B.
Catalysts 08 00143 g004
Figure 5. DRIFT spectra of NH3 reacted with pre-adsorbed NO + O2 species at 250 °C on: (A) Cat-A; and (B) Cat-B.
Figure 5. DRIFT spectra of NH3 reacted with pre-adsorbed NO + O2 species at 250 °C on: (A) Cat-A; and (B) Cat-B.
Catalysts 08 00143 g005
Figure 6. NO conversion of Cat-A and Cat-B over time after cutting off NH3. Reaction conditions: T = 250 °C, [NO] = 1000 ppm, [O2] = 5%, GHSV = 40,000 h−1, and N2 balance gas.
Figure 6. NO conversion of Cat-A and Cat-B over time after cutting off NH3. Reaction conditions: T = 250 °C, [NO] = 1000 ppm, [O2] = 5%, GHSV = 40,000 h−1, and N2 balance gas.
Catalysts 08 00143 g006
Figure 7. FT-IR spectra of Cat-A with: (A) no adsorption; (B) adsorption of NO + O2; and (C) after cutting off NO + O2, adsorption of NH3.
Figure 7. FT-IR spectra of Cat-A with: (A) no adsorption; (B) adsorption of NO + O2; and (C) after cutting off NO + O2, adsorption of NH3.
Catalysts 08 00143 g007
Figure 8. Schematic of the proposed NH3-SCR reaction mechanism over Cat-A.
Figure 8. Schematic of the proposed NH3-SCR reaction mechanism over Cat-A.
Catalysts 08 00143 g008

Share and Cite

MDPI and ACS Style

Wu, R.; Zhang, N.; Li, L.; He, H.; Song, L.; Qiu, W. DRIFT Study on Promotion Effect of the Keggin Structure over V2O5-MoO3/TiO2 Catalysts for Low Temperature NH3-SCR Reaction. Catalysts 2018, 8, 143. https://doi.org/10.3390/catal8040143

AMA Style

Wu R, Zhang N, Li L, He H, Song L, Qiu W. DRIFT Study on Promotion Effect of the Keggin Structure over V2O5-MoO3/TiO2 Catalysts for Low Temperature NH3-SCR Reaction. Catalysts. 2018; 8(4):143. https://doi.org/10.3390/catal8040143

Chicago/Turabian Style

Wu, Rui, Ningqiang Zhang, Lingcong Li, Hong He, Liyun Song, and Wenge Qiu. 2018. "DRIFT Study on Promotion Effect of the Keggin Structure over V2O5-MoO3/TiO2 Catalysts for Low Temperature NH3-SCR Reaction" Catalysts 8, no. 4: 143. https://doi.org/10.3390/catal8040143

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