Next Article in Journal
Metal Micro-Monoliths for the Kinetic Study and the Intensification of the Water Gas Shift Reaction
Next Article in Special Issue
Electrochemical Promotion of Nanostructured Palladium Catalyst for Complete Methane Oxidation
Previous Article in Journal
The Role of Impregnated Sodium Ions in Cu/SSZ-13 NH3-SCR Catalysts
Previous Article in Special Issue
Low-Temperature Electrocatalytic Conversion of CO2 to Liquid Fuels: Effect of the Cu Particle Size
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A CeO2/ZrO2-TiO2 Catalyst for the Selective Catalytic Reduction of NOx with NH3

1
Center for Excellence in Regional Atmospheric Environment, Institute of Urban Environment, Chinese Academy of Sciences, Xiamen 361021, China
2
Ningbo Urban Environment Observation and Research Station-NUEORS, Institute of Urban Environment, Chinese Academy of Sciences, Ningbo 315800, China
3
School of Environmental and Biological Engineering, Nanjing University of Science and Technology, Nanjing 210094, China
4
State Key Joint Laboratory of Environment Simulation and Pollution Control, Research Center for Eco-Environmental Sciences, Chinese Academy of Sciences, Beijing 100085, China
*
Author to whom correspondence should be addressed.
Catalysts 2018, 8(12), 592; https://doi.org/10.3390/catal8120592
Submission received: 30 September 2018 / Revised: 24 November 2018 / Accepted: 27 November 2018 / Published: 30 November 2018
(This article belongs to the Special Issue Emissions Control Catalysis)

Abstract

:
In this study, CeZr0.5TiaOx (with a = 0, 1, 2, 5, 10) catalysts were prepared by a stepwise precipitation approach for the selective catalytic reduction of NOx with NH3. When Ti was added, all of the Ce-Zr-Ti oxide catalysts showed much better catalytic performances than the CeZr0.5Ox. Particularly, the CeZr0.5Ti2Ox catalyst showed excellent activity for broad temperature range under high space velocity condition. Through the control of pH value and precipitation time during preparation, the function of the CeZr0.5Ti2Ox catalyst could be controlled and the structure with highly dispersed CeO2 (with redox functions) on the surface of ZrO2-TiO2 (with acidic functions) could be obtained. Characterizations revealed that the superior catalytic performance of the catalyst is associated with its outstanding redox properties and adsorption/activation functions for the reactants.

1. Introduction

NOx (mainly NO and NO2) in the atmosphere plays critical roles in the formation of severe air pollution problems, such as haze, acid rain, and photochemical smog. In the last few decades, great efforts have been devoted to the development of NOx emission control technologies [1,2,3]. Selective catalytic reduction of NOx with NH3 (NH3-SCR) has been widely applied for the removal of NOx generated from stationary sources for many years, and it has also been used for the control of NOx emission from diesel vehicles [2,4].
Catalysts play an important role in the development of NH3-SCR technology [5,6]. Vanadium-based catalyst (especially V2O5-WO3/TiO2), with excellent SO2 resistance, is the most widely used NH3-SCR catalyst for NOx emission control from power plants, and it was also applied on diesel vehicles as the first generation of SCR catalyst [4]. However, this catalyst system still has some problems, including the toxicity of active V2O5, narrow temperature window, and low thermal stability [2].
There has been strong interest in developing a vanadium-free catalyst that can be used on diesel vehicles [5,6,7,8,9,10,11]. Ce is a key component in three-way catalysts for emission control in automobiles for gasoline. CeO2 provides an oxygen storage function through redox cycling between Ce3+ and Ce4+. In recent years, Ce has also attracted great attention for applications as a support [12,13], promoter [14,15,16,17,18], or main active component [19,20,21,22,23,24,25,26] for NH3-SCR catalysts.
Pure Ce oxide is not suitable for use as an NH3-SCR catalyst [27,28]. When Zr oxide was introduced into Ce oxide, the thermal stability and the oxygen storage capacity of the oxide could be significantly improved. Therefore, Ce-Zr oxide was investigated for NH3-SCR [12,13,29,30,31,32,33,34]. In the NH3-SCR reaction, both redox functions and acidic functions of the catalyst are needed [4,35]. Therefore, a high dispersion of active sites and close coupling of redox with acid sites is the way to design a highly efficient NH3-SCR catalyst.
In this study, starting from a preparation of Ce-Zr oxide by the co-preparation method, we developed a Ce-Zr-Ti oxide catalyst using a stepwise precipitation approach, under the theoretical guidance of the close combination of the Ce-Zr oxide with strong redox functions and Ti oxide with excellent acid properties [4,5]. This obtained catalyst showed superior catalytic performance for NH3-SCR.

2. Results and Discussion

2.1. NH3-SCR Activity

Figure 1A presents the NOx conversion over the catalysts with different Ti contents under a relatively high gas hourly space velocity (GHSV) of 200,000 h−1. The CeZr0.5Ox just exhibited over 50% NOx conversion in a narrow temperature range of 350–425 °C. When Ti was introduced, all of the Ce-Zr-Ti oxide catalysts exhibited much better activities. With the increase in Ti content, the low temperature firstly increased and then decreased. As a result, the CeZr0.5Ti2Ox catalyst presented the best activity in a low temperature range, together with a high NOx conversion in a wide temperature range. On the other hand, the variation in high temperature activity with Ti content was contrary to that of low temperature activity, with the activity of CeZr0.5Ti2Ox slightly lower than those of the other Ce-Zr-Ti oxide catalysts in a high temperature range. In addition, adding Ti to the catalyst also enhanced the N2 selectivity, and the Ce-Zr-Ti oxide catalysts all presented higher N2 selectivity than CeZr0.5Ox (Figure 1B).
The influences of H2O and space velocity on the NOx conversion over CeZr0.5Ti2Ox were tested and the results are shown in Figure 2. The existence of 5% H2O in the flow gas decreased the low temperature activity, but enhanced the high temperature activity. As a result, over 80% NOx conversion could still be achieved from 250 to 450 °C. When the GHSV was decreased from 200,000 h−1 to 100,000 h−1, the activity of the catalyst at low temperatures was obviously improved.

2.2. Separated NO/NH3 Oxidation

To analyze the effects of Ti on the catalyst, separated NO oxidation and NH3 oxidation tests were carried out for the CeZr0.5Ox and CeZr0.5Ti2Ox (Figure 3). The NO2 production during NO oxidation over the CeZr0.5Ti2Ox was clearly higher than that over CeZr0.5Ox at a low temperature. Since the presentation of NO2 in the reaction gas could promote the SCR reaction at a low temperature by accelerating the fast SCR process (2 NH3 + NO + NO2 → 2N2 + 3H2O), the enhanced low-temperature activity by the introduction of Ti should be associated with the promoted oxidation of NO to NO2 over CeZr0.5Ti2Ox [10,35]. In addition, the introduction of Ti also promoted NH3 oxidation over the catalyst at a high temperature. The NH3-SCR reaction route at a high temperature mainly follows the Eley-Rideal mechanism, and the activation of NH3 to form NH2 species by oxidation plays the key role for the reaction with NO to form N2 and H2O, owing to NH2 + NO(g) → N2 + H2O. Therefore, promoted NH3 oxidation would be beneficial for the improvement of high temperature activity.

2.3. XRD

The X-ray diffraction (XRD) results of the CeZr0.5Ox and Ce-Z-Ti oxide catalysts are presented in Figure 4. Both CeO2 and ZrO2 were detected in CeZr0.5Ox. With the increase of Ti, the peaks for CeO2 and ZrO2 became more and more weak, and only anatase TiO2 was observed for CeZr0.5Ti10Ox. Only weak peaks for CeO2 with cubic fluorite structures (PDF# 43-1002) were observed in the CeZr0.5Ti2Ox, indicating that the introduction of Ti had induced the structural change of the CeZr0.5Ox, and the crystallizations of Ce, Zr and Ti oxides in CeZr0.5Ti2Ox were significantly inhibited. As a result, the CeZr0.5Ti2Ox (165.1 m2/g) showed a higher Brunauer–Emmett–Teller (BET) surface area than CeZr0.5Ox (113.5 m2/g).

2.4. H2-TPR

The H2 temperature-programmed reduction (H2-TPR) profiles of CeZr0.5Ox and CeZr0.5Ti2Ox are presented in Figure 5. The CeZr0.5Ox exhibited two peaks at 496 and 755 °C due to the surface and bulk reductions of CeO2 (as detected by XRD), respectively [31,36,37,38]. During the test, coordinatively unsaturated surface oxygen anions are easily reduced by H2 in the low temperature region, while the bulk oxygen species are reduced only after the transportation to the surface [39]. With the addition of Ti, a sharp H2 consumption peak appeared at 567 °C, which indicates that another type of Ce species might be formed. Considering the XRD results, this sharp peak might be associated with the reduction of the highly dispersed Ce species from Ce4+ to Ce3+ [22,34]. In addition, the H2 consumption of CeZr0.5Ti2Ox was much higher than that of CeZr0.5Ox at a low temperature. The H2-TPR results clearly indicated the enhancement of redox functions for CeZr0.5Ti2Ox.
Previous studies have indicated that the redox properties of NH3-SCR catalyst play a dominant role in the low temperature activity [35,40,41]. Therefore, the enhanced redox function of CeZr0.5Ti2Ox would beneficial for low temperature activity.

2.5. NOx/NH3-TPD

To investigate the NOx and NH3 adsorption/desorption properties of CeZr0.5Ox and CeZr0.5Ti2Ox, NOx temperature-programmed desorption (NOx-TPD) and NH3 temperature-programmed desorption (NH3-TPD) were performed for the catalysts (Figure 6).
The NOx-TPD profiles are presented in Figure 6A. The first NOx peak of CeZr0.5Ti2Ox was at ca. 110 °C, mainly due to the desorption of physisorbed NOx, while the other NOx peak was at ca. 300 °C and was associated with the decomposition of chemsorbed NOx species [42,43]. On the other hand, two weak peaks were observed for CeZr0.5Ox at ca. 270 °C and ca. 410 °C, respectively, which were due to the decomposition of different types of chemsorbed NOx species. With the addition of Ti, the adsorbed NOx on CeZr0.5Ti2Ox was obviously more than that of CeZr0.5Ox. Particularly, the desorbed NO2 of CeZr0.5Ti2Ox was much higher, owing to the enhanced low-temperature activity for NO oxidation (as shown by the separated NO oxidation results), which could facilitate the conversion of NOx in NH3-SCR.
Surface acidity plays a dominate role in the high-temperature SCR activity due to its effects on the adsorption and activation of NH3 [35,41]. Previous studies have revealed that Ti species of NH3-SCR catalysts mainly act as acid sites in the reaction for NH3 adsorption [4]. Therefore, the adsorbed NH3 of CeZr0.5Ti2Ox was much more than that of CeZr0.5Ox, which might be an important reason for the better NH3-SCR activity of CeZr0.5Ti2Ox at high temperatures.

2.6. XPS

The X-ray photoelectron spectroscopy (XPS) results for Ce 3d of the CeZr0.5Ox and CeZr0.5Ti2Ox are shown in Figure 7. The sub-bands labeled with u’/v’ and u0/v0 represent the 3d104f1 initial electronic state corresponding to Ce3+ and the 3d94f2 state of Ce3+, respectively [44]. The sub-bands labeled with u’’’ and v’’’ represent the 3d104f0 state of Ce4+, and the sub-bands labeled with u, u’’, v and v’’ represent the 3d94f1 state corresponding to Ce4+ [44]. The presence of Ce3+ would induce a charge imbalance, which could lead to unsaturated chemical bonds and oxygen vacancies. The calculated Ce3+ ratio of CeZr0.5Ti2Ox (36.0%) was higher than that of CeZr0.5Ox (33.8%), indicating that more surface oxygen vacancies presented in CeZr0.5Ti2Ox. In addition, the Ce3+ ratio of the catalyst could influence the redox ability and reactant adsorption and activation functions, and thereby contribute to NH3-SCR performance.
The surface oxygen vacancies of the catalysts might generate weakly-adsorbed oxygen species or additional chemisorbed oxygen on the surface of the catalyst [27,45]. The XPS results of O 1s of the CeZr0.5Ox and CeZr0.5Ti2Ox are shown in Figure 8. The O 1s peak was fit into two sub-bands. The sub-bands at 531.2–531.5 eV and 529.1–529.6 eV were assigned to the surface adsorbed oxygen (Oα), such as the O22− and O belonging to defect-oxide or a hydroxyl-like group, and the lattice oxygen O2− (Oβ), respectively [46]. The Oα ratios of the catalysts were calculated by Oα/(Oα + Oβ), and the CeZr0.5Ti2Ox showed higher Oα ratio than CeZr0.5Ox. The results confirmed that the addition of Ti indeed induced more surface-adsorbed oxygen, which would facilitate NO oxidation to NO2 (as shown by the separated NO oxidation and NOx-TPD results), and thus facilitates the conversion of NO by fast SCR effects.

2.7. Formation Process Analysis of the CeZr0.5Ti2Ox Catalyst

Figure 9 shows the pH variations of the mixed solutions for the preparation of the CeZr0.5Ox and CeZr0.5Ti2Ox catalysts. During the preparation of CeZr0.5Ox, the initial pH value of the solution was 1.6. With the hydrolysis of urea, the pH increased gradually to be 7.6 after heating for 12 h. Due to the increase in pH, suspended particles began to appear in the solution in the second hour. The particles with the precipitation time of 2 h, 4 h, 6 h, and 12 h were collected and then calcined to be catalyst samples. The activity tests of these samples showed similar NOx conversions with each other.
Due to the acidity induced by the added Ti(SO4)2, the initial pH value of the mixed solution during the preparation of CeZr0.5Ti2Ox dropped to be 1.1. With the hydrolysis of urea, the pH increased gradually after heating, and some white particles generated in the first hour and suspended in the solution. With the increase of time, the particles gradually turned yellow. The pH reached ca. 7.0 after 12 h of reaction. The particles with the precipitation times of 1 h, 4 h, 6 h, and 12 h were collected and then calcined to be catalyst samples. Interestingly, the activity test showed a remarkable enhancement of NOx conversions for the four samples with the increase in precipitation time.
The surface metal atomic concentrations of the CeZr0.5Ti2Ox samples with different precipitation times were analyzed using XPS, and the variations in Ce, Zr, and Ti concentrations with precipitation time are shown in Figure 10. For the 1-h precipitation sample, only Ti and Zr, without Ce, were detected. With the increase in precipitation time, surface Ce concentration increased gradually in the samples. At the same time, Ti and Zr concentrations gradually decreased with the increase in precipitation time. A TEM-EDS mapping image showed that Ce was highly dispersed in the CeZr0.5Ti2Ox catalyst (Figure 11).
Considering the variations in the solution pH value when preparing the CeZr0.5Ti2Ox, the formation process of the catalyst can be proposed as follows: The Ti and Zr species were first co-precipitated with the increase in solution pH. Then, the Ce species uniformly precipitated onto the precipitated Zr-Ti species with the further increase in pH. Finally, a CeZr0.5Ti2Ox catalyst with a higher surface Ce concentration than Ti and Zr was obtained. Through control of the hydrolysis of urea, the variations in the solution pH can be controlled, and then we can control the precipitation process of the catalyst, which is very important for the formation of highly-dispersed CeO2 on ZrO2-TiO2. Thus, the obtained catalyst can present excellent NH3-SCR performance.

3. Experimental Section

3.1. Catalyst Preparation and Activity Test

The CeZr0.5TiaOx (a = Ti/Ce molar ratio = 0, 1, 2, 5, 10), with a Zr/Ce molar ratio fixed to be 0.5, was prepared using a precipitation method. Desired precursors of Ce(NO3)3·6H2O (>99%, Sinopharm Chemical Reagent Co., Ltd., Shanghai, China), Zr(NO3)4·5H2O (>99%, Sinopharm Chemical Reagent Co., Ltd., Shanghai, China) and Ti(SO4)2 (>98%, Sinopharm Chemical Reagent Co., Ltd., Shanghai, China) were dissolved together in distilled water, and urea (>99%, Sinopharm Chemical Reagent Co., Ltd., Shanghai, China) was added to the mixed solution as a slowly-releasing precipitator. Then, the solution was heated to 90 °C to facilitate the release of NH3 and thereby raise the pH value gradually. The temperature of the mixed solution was held at 90 °C for 12 h under vigorous stirring (some samples with shorter precipitation times were also prepared). After that, the precipitated powders were collected via filtration, washed using distilled water, and dried for 12 h at 100 °C. Finally, the catalyst was obtained after calcination at 500 °C for 5 h.
The SCR activity of the catalysts (40–60 mesh) were tested in a fixed-bed quartz flow reactor. The reaction conditions were controlled as follows: 500 ppm NO, 500 ppm NH3, 5 vol.% O2, N2 balance, and 400 mL/min total flow rate. Different gas hourly space velocities (GHSVs) were obtained by changing the volume of catalysts, i.e., 0.24 mL catalyst for a GHSV = 100,000 h−1 and 0.12 mL catalyst for a GHSV = 200,000 h−1. The concentrations of effluent N-containing gases (NO, NH3, NO2 and N2O) were continuously measured by an online FTIR gas analyzer (Nicolet Antaris IGS analyzer, Thermo-Fisher Scientific, Waltham, MA, USA). NOx conversion and N2 selectivity were calculated using the following equations, respectively:
NO x   conversion = ( 1 [ NO ] o u t + [ NO 2 ] o u t [ NO ] i n + [ NO 2 ] i n ) × 100 %
N 2   selectivity = ( 1 2 [ N 2 O ] o u t [ NO x ] i n + [ NH 3 ] i n [ NO x ] o u t [ NH 3 ] o u t ) × 100 %

3.2. Characterizations

X-ray diffraction (XRD) measurements were carried out on a computerized AXS D8 diffractometer (Bruker, GER), with Cu Kα (λ = 0.15406 nm) radiation, from 20 to 80° at 8°/min.
Surface areas were tested using an ASAP 2020 (Micromeritics, Norcross, GA, USA) at −196 °C by N2 adsorption/desorption and calculated using a BET equation in the 0.05–0.35 partial pressure range.
The X-ray photoelectron spectroscopy (XPS) results of Ce 3d and O 1s were measured on an ESCALAB 250Xi Scanning X-ray Microprobe (Thermo-Fisher Scientific, Waltham, MA, USA) using Al Ka radiation (1486.7 eV) and a C 1 s peak, with BE = 284.8 eV as the calibration standard.
The transmission electron microscopy (TEM) image and energy-dispersive X-ray spectroscopy (EDS) mapping of Ce were obtained using a JEM-2100F equipment (JEOL, Tokyo, Japan), combined with a specimen tilting beryllium holder for energy dispersive spectroscopy. The accelerating voltage was 200 kV.
The H2 temperature-programmed reduction (H2-TPR) was tested using an AutoChem_II_2920 chemisorption analyzer (Micromeritics, Norcross, GA, USA), and the temperature-programmed desorption of NH3 and NOx (NOx-TPD and NH3-TPD) were tested using the same reaction system as the activity tests. Experiment details can be found in Reference [42].

4. Conclusions

A series of Ce-Zr-Ti oxide catalysts were prepared using a stepwise precipitation approach for NH3-SCR. CeZr0.5Ox without Ti just showed a relatively low NOx conversion. When Ti was introduced, Ce-Zr-Ti catalysts showed much better activities and N2 selectivity. A CeZr0.5Ti2Ox catalyst, which contains moderate Ti amounts, showed the best performance, which is associated with its optimal ratios for the redox (CeOx) and acidic (TiO2) components.
CeZr0.5Ox and CeZr0.5Ti2Ox catalysts were characterized using various methods and the formation process during preparation was investigated. CeZr0.5Ti2Ox catalyst showed superior redox properties (by H2-TPR), good adsorption and NOx/NH3 activation functions (by NOx-TPD and NH3-TPD, respectively), and enhanced charge imbalance (by XPS).
During preparation, the Ti and Zr species were first co-precipitated with an increase in solution pH. Then, the Ce species uniformly precipitated onto the precipitated Zr-Ti species with the further increase in pH. As a result, CeZr0.5Ti2Ox catalyst with a surface Ce concentration higher than those of Ti and Zr was obtained. This preparation process resulted in the formation of highly-dispersed CeO2 on ZrO2-TiO2, and thus the catalyst can present excellent NH3-SCR performance.

Author Contributions

W.S. and H.H. conceived the project; Y.G. and Y.Z. performed the experiments; W.S. and Z.L. carried out the data analysis; W.S. and Y.G. wrote the paper; H.H. supervised the study.

Funding

This work was supported by the National Key R&D Program of China (2017YFC0212502, 2017YFC0211101), the National Natural Science Foundation of China (201637005), and the Key Research Program of the Chinese Academy of Sciences (ZDRW-ZS-2017-6-2-3).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Parvulescu, V.I.; Granger, P.; Delmon, B. Catalytic removal of NO. Catal. Today 1998, 46, 233–316. [Google Scholar] [CrossRef]
  2. Granger, P.; Parvulescu, V.I. Catalytic NOx abatement systems for mobile sources: From three-way to lean burn after-treatment technologies. Chem. Rev. 2011, 111, 3155–3207. [Google Scholar] [CrossRef] [PubMed]
  3. Liu, Z.; Woo, S.I. Recent advances in catalytic DeNOx science and technology. Catal. Rev. 2006, 48, 43–89. [Google Scholar] [CrossRef]
  4. 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 1998, 18, 1–36. [Google Scholar] [CrossRef]
  5. Liu, F.; Yu, Y.; He, H. Environmentally-benign catalysts for the selective catalytic reduction of NOx from diesel engines: Structure-activity relationship and reaction mechanism aspects. Chem. Commun. 2014, 50, 8445–8463. [Google Scholar] [CrossRef] [PubMed]
  6. Shan, W.; Song, H. Catalysts for the selective catalytic reduction of NOx with NH3 at low temperature. Catal. Sci. Technol. 2015, 5, 4280–4288. [Google Scholar] [CrossRef]
  7. Guan, B.; Zhan, R.; Lin, H.; Huang, Z. Review of state of the art technologies of selective catalytic reduction of NOx from diesel engine exhaust. Appl. Therm. Eng. 2014, 66, 395–414. [Google Scholar] [CrossRef]
  8. Gao, F.; Kwak, J.; Szanyi, J.; Peden, C.F. Current understanding of Cu-exchanged chabazite molecular sieves for use as commercial diesel engine DeNOx catalysts. Top. Catal. 2013, 56, 1441–1459. [Google Scholar] [CrossRef]
  9. Brandenberger, S.; Kröcher, O.; Tissler, A.; Althoff, R. The state of the art in selective catalytic reduction of NOx by ammonia using metal-exchanged zeolite catalysts. Catal. Rev. 2008, 50, 492–531. [Google Scholar] [CrossRef]
  10. Li, J.; Chang, H.; Ma, L.; Hao, J.; Yang, R.T. Low-temperature selective catalytic reduction of NOx with NH3 over metal oxide and zeolite catalysts-a review. Catal. Today 2011, 175, 147–156. [Google Scholar] [CrossRef]
  11. Deka, U.; Lezcano-Gonzalez, I.; Weckhuysen, B.M.; Beale, A.M. Local environment and nature of Cu active sites in zeolite-based catalysts for the selective catalytic reduction of NOx. ACS Catal. 2013, 413–427. [Google Scholar] [CrossRef]
  12. Li, Y.; Cheng, H.; Li, D.; Qin, Y.; Xie, Y.; Wang, S. WO3/CeO2-ZrO2, a promising catalyst for selective catalytic reduction (SCR) of NOx with NH3 in diesel exhaust. Chem. Commun. 2008, 1470–1472. [Google Scholar] [CrossRef] [PubMed]
  13. Can, F.; Berland, S.; Royer, S.; Courtois, X.; Duprez, D. Composition-dependent performance of CexZr1–xO2 mixed-oxide-supported WO3 catalysts for the NOx storage reduction–selective catalytic reduction coupled process. ACS Catal. 2013, 3, 1120–1132. [Google Scholar] [CrossRef]
  14. Long, R.Q.; Yang, R.T. Superior Fe-ZSM-5 catalyst for selective catalytic reduction of nitric oxide by ammonia. J. Am. Chem. Soc. 1999, 121, 5595–5596. [Google Scholar] [CrossRef]
  15. Chen, L.; Li, J.; Ge, M. Promotional effect of Ce-doped V2O5-WO3/TiO2 with low vanadium loadings for selective catalytic reduction of NOx by NH3. J. Phys. Chem. C 2009, 113, 21177–21184. [Google Scholar] [CrossRef]
  16. Wu, Z.; Jin, R.; Liu, Y.; Wang, H. Ceria modified MnOx/TiO2 as a superior catalyst for NO reduction with NH3 at low-temperature. Catal. Commun. 2008, 9, 2217–2220. [Google Scholar] [CrossRef]
  17. Niu, C.; Shi, X.; Liu, K.; You, Y.; Wang, S.; He, H. A novel one-pot synthesized CuCe-SAPO-34 catalyst with high NH3-SCR activity and H2O resistance. Catal. Commun. 2016, 81, 20–23. [Google Scholar] [CrossRef]
  18. Carja, G.; Delahay, G.; Signorile, C.; Coq, B. Fe-Ce-ZSM-5 a new catalyst of outstanding properties in the selective catalytic reduction of NO with NH3. Chem. Commun. 2004, 1404–1405. [Google Scholar] [CrossRef] [PubMed]
  19. Gao, X.; Jiang, Y.; Fu, Y.; Zhong, Y.; Luo, Z.; Cen, K. Preparation and characterization of CeO2/TiO2 catalysts for selective catalytic reduction of NO with NH3. Catal. Commun. 2010, 11, 465–469. [Google Scholar] [CrossRef]
  20. Liu, Z.; Yi, Y.; Li, J.; Woo, S.I.; Wang, B.; Cao, X.; Li, Z. A superior catalyst with dual redox cycles for the selective reduction of NOx by ammonia. Chem. Commun. 2013, 49, 7726–7728. [Google Scholar] [CrossRef] [PubMed]
  21. 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 2014, 144, 90–95. [Google Scholar] [CrossRef]
  22. Peng, Y.; Qu, R.; Zhang, X.; Li, J. The relationship between structure and activity of MoO3-CeO2 catalysts for NO removal: Influences of acidity and reducibility. Chem. Commun. 2013, 49, 6215–6217. [Google Scholar] [CrossRef] [PubMed]
  23. Qu, R.; Gao, X.; Cen, K.; Li, J. Relationship between structure and performance of a novel cerium-niobium binary oxide catalyst for selective catalytic reduction of NO with NH3. Appl. Catal. B 2013, 142, 290–297. [Google Scholar] [CrossRef]
  24. Shan, W.; Liu, F.; He, H.; Shi, X.; Zhang, C. Novel cerium-tungsten mixed oxide catalyst for the selective catalytic reduction of NOx with NH3. Chem. Commun. 2011, 47, 8046–8048. [Google Scholar] [CrossRef] [PubMed]
  25. Krishna, K.; Seijger, G.B.F.; Bleek, C.M.; Calis, H.P.A. Very active CeO2-zeolite catalysts for NOx reduction with NH3. Chem. Commun. 2002, 18, 2030–2031. [Google Scholar] [CrossRef]
  26. Shan, W.; Liu, F.; Yu, Y.; He, H. The use of ceria for the selective catalytic reduction of NOx with NH3. Chin. J. Catal. 2014, 35, 1251–1259. [Google Scholar] [CrossRef]
  27. Gu, T.; Liu, Y.; Weng, X.; Wang, H.; Wu, Z. 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]
  28. Zhang, L.; Pierce, J.; Leung, V.L.; Wang, D.; Epling, W.S. Characterization of ceria’s interaction with NOx and NH3. J. Phys. Chem. C 2013, 117, 8282–8289. [Google Scholar] [CrossRef]
  29. Si, Z.; Weng, D.; Wu, X.; Yang, J.; Wang, B. Modifications of CeO2–ZrO2 solid solutions by nickel and sulfate as catalysts for NO reduction with ammonia in excess O2. Catal. Commun. 2010, 11, 1045–1048. [Google Scholar] [CrossRef]
  30. Si, Z.; Weng, D.; Wu, X.; Ran, R.; Ma, Z. NH3-SCR activity, hydrothermal stability, sulfur resistance and regeneration of Ce0.75Zr0.25O2–PO43− catalyst. Catal. Commun. 2012, 17, 146–149. [Google Scholar] [CrossRef]
  31. Shen, B.; Wang, Y.; Wang, F.; Liu, T. The effect of Ce-Zr on NH3-SCR activity over MnOx(0.6)/Ce0.5Zr0.5O2 at low temperature. Chem. Eng. J. 2014, 236, 171–180. [Google Scholar] [CrossRef]
  32. Ko, J.H.; Park, S.H.; Jeon, J.-K.; Kim, S.-S.; Kim, S.C.; Kim, J.M.; Chang, D.; Park, Y.-K. Low temperature selective catalytic reduction of NO with NH3 over Mn supported on Ce0.65Zr0.35O2 prepared by supercritical method: Effect of Mn precursors on NO reduction. Catal. Today 2012, 185, 290–295. [Google Scholar] [CrossRef]
  33. Liu, Z.; Su, H.; Li, J.; Li, Y. Novel MoO3/CeO2–ZrO2 catalyst for the selective catalytic reduction of NOx by NH3. Catal. Commun. 2015, 65, 51–54. [Google Scholar] [CrossRef]
  34. Ding, S.P.; Liu, F.D.; Shi, X.Y.; Liu, K.; Lian, Z.H.; Xie, L.J.; He, H. Significant promotion effect of Mo additive on a novel Ce-Zr mixed oxide catalyst for the selective catalytic reduction of NOx with NH3. ACS Appl. Mater. Interfaces 2015, 7, 9497–9506. [Google Scholar] [CrossRef] [PubMed]
  35. Shan, W.; Liu, F.; He, H.; Shi, X.; Zhang, C. A superior Ce-W-Ti mixed oxide catalyst for the selective catalytic reduction of NOx with NH3. Appl. Catal. B 2012, 115–116, 100–106. [Google Scholar] [CrossRef]
  36. Chen, A.; Zhou, Y.; Ta, N.; Li, Y.; Shen, W. Redox properties and catalytic performance of ceria-zirconia nanorods. Catal. Sci. Technol. 2015, 5, 4184–4192. [Google Scholar] [CrossRef]
  37. Liu, Z.; Zhang, S.; Li, J.; Zhu, J.; Ma, L. Novel V2O5-CeO2/TiO2 catalyst with low vanadium loading for the selective catalytic reduction of NOx by NH3. Appl. Catal. B 2014, 158–159, 11–19. [Google Scholar] [CrossRef]
  38. Vuong, T.H.; Radnik, J.; Rabeah, J.; Bentrup, U.; Schneider, M.; Atia, H.; Armbruster, U.; Grünert, W.; Brückner, A. Efficient VOx/Ce1–xTixO2 catalysts for low-temperature NH3-SCR: Reaction mechanism and active sites assessed by in situ/operando spectroscopy. ACS Catal. 2017, 7, 1693–1705. [Google Scholar] [CrossRef]
  39. Wang, Z.; Qu, Z.; Quan, X.; Wang, H. Selective catalytic oxidation of ammonia to nitrogen over ceria–zirconia mixed oxides. Appl. Catal. A 2012, 411–412, 131–138. [Google Scholar] [CrossRef]
  40. 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]
  41. Lietti, L.; Forzatti, P.; Berti, F. Role of the redox properties in the SCR of NO by NH3 over V2O5-WO3/TiO2 catalyst. Catal. Lett. 1996, 41, 35–39. [Google Scholar] [CrossRef]
  42. Geng, Y.; Huang, H.; Chen, X.; Ding, H.; Yang, S.; Liu, F.; Shan, W. The effect of Ce on a high-efficiency CeO2/WO3-TiO2 catalyst for the selective catalytic reduction of NOx with NH3. RSC Adv. 2016, 6, 64803–64810. [Google Scholar] [CrossRef]
  43. Liu, F.; He, H. Structure-activity relationship of iron titanate catalysts in the selective catalytic reduction of NOx with NH3. J. Phys. Chem. C 2010, 114, 16929–16936. [Google Scholar] [CrossRef]
  44. Bêche, E.; Charvin, P.; Perarnau, D.; Abanades, S.; Flamant, G. Ce 3d XPS investigation of cerium oxides and mixed cerium oxide (CexTiyOz). Surf. Interface Anal. 2008, 40, 264–267. [Google Scholar] [CrossRef]
  45. Liu, C.; Chen, L.; Chang, H.; Ma, L.; Peng, Y.; Arandiyan, H.; Li, J. Characterization of CeO2-WO3 catalysts prepared by different methods for selective catalytic reduction of NOx with NH3. Catal. Commun. 2013, 40, 145–148. [Google Scholar] [CrossRef]
  46. Shan, W.; Liu, F.; He, H.; Shi, X.; Zhang, C. An environmentally-benign CeO2-TiO2 catalyst for the selective catalytic reduction of NOx with NH3 in simulated diesel exhaust. Catal. Today 2012, 184, 160–165. [Google Scholar] [CrossRef]
Figure 1. (A) NOx conversions and (B) N2 selectivity over the CeZr0.5Ox and Ce-Zr-Ti oxide catalysts. Reaction conditions: [NO] = [NH3] = 500 ppm, [O2] = 5 vol.%, N2 balance, and GHSV = 200,000 h−1.
Figure 1. (A) NOx conversions and (B) N2 selectivity over the CeZr0.5Ox and Ce-Zr-Ti oxide catalysts. Reaction conditions: [NO] = [NH3] = 500 ppm, [O2] = 5 vol.%, N2 balance, and GHSV = 200,000 h−1.
Catalysts 08 00592 g001
Figure 2. NOx conversion over CeZr0.5Ti2Ox catalyst under different reaction conditions. Reaction conditions: [NO] = [NH3] = 500 ppm, [O2] = 5 vol.%, [H2O] = 5 vol.% (when used), N2 balance, and GHSV = 100,000 or 200,000 h−1.
Figure 2. NOx conversion over CeZr0.5Ti2Ox catalyst under different reaction conditions. Reaction conditions: [NO] = [NH3] = 500 ppm, [O2] = 5 vol.%, [H2O] = 5 vol.% (when used), N2 balance, and GHSV = 100,000 or 200,000 h−1.
Catalysts 08 00592 g002
Figure 3. (A) NO2 productions during separate NO oxidation reaction and (B) NH3 conversions during separate NH3 oxidation reaction over the CeZr0.5Ox and CeZr0.5Ti2Ox catalysts. Reaction conditions: (A) [NO] _ENREF_30= 500 ppm, (B) [NH3] = 500 ppm, [O2] = 5 vol.%, N2 balance and GHSV = 200,000 h−1.
Figure 3. (A) NO2 productions during separate NO oxidation reaction and (B) NH3 conversions during separate NH3 oxidation reaction over the CeZr0.5Ox and CeZr0.5Ti2Ox catalysts. Reaction conditions: (A) [NO] _ENREF_30= 500 ppm, (B) [NH3] = 500 ppm, [O2] = 5 vol.%, N2 balance and GHSV = 200,000 h−1.
Catalysts 08 00592 g003
Figure 4. XRD patterns of the CeZr0.5Ox and Ce-Z-Ti oxide catalysts.
Figure 4. XRD patterns of the CeZr0.5Ox and Ce-Z-Ti oxide catalysts.
Catalysts 08 00592 g004
Figure 5. H2-TPR profiles of the CeZr0.5Ox and CeZr0.5Ti2Ox catalysts.
Figure 5. H2-TPR profiles of the CeZr0.5Ox and CeZr0.5Ti2Ox catalysts.
Catalysts 08 00592 g005
Figure 6. (A) NOx-TPD and (B) NH3-TPD profiles of the CeZr0.5Ox and CeZr0.5Ti2Ox catalysts.
Figure 6. (A) NOx-TPD and (B) NH3-TPD profiles of the CeZr0.5Ox and CeZr0.5Ti2Ox catalysts.
Catalysts 08 00592 g006
Figure 7. XPS results of Ce 3d of the CeZr0.5Ox and CeZr0.5Ti2Ox catalysts.
Figure 7. XPS results of Ce 3d of the CeZr0.5Ox and CeZr0.5Ti2Ox catalysts.
Catalysts 08 00592 g007
Figure 8. XPS results of O 1s of the CeZr0.5Ox and CeZr0.5Ti2Ox catalysts.
Figure 8. XPS results of O 1s of the CeZr0.5Ox and CeZr0.5Ti2Ox catalysts.
Catalysts 08 00592 g008
Figure 9. The pH variation of the mixed solution during the preparation of the (A) CeZr0.5Ox and (B) CeZr0.5Ti2Ox catalysts, and the NOx conversions of the obtained samples at different precipitation time.
Figure 9. The pH variation of the mixed solution during the preparation of the (A) CeZr0.5Ox and (B) CeZr0.5Ti2Ox catalysts, and the NOx conversions of the obtained samples at different precipitation time.
Catalysts 08 00592 g009
Figure 10. Surface metal atomic concentrations of the CeZr0.5Ti2Ox samples with different precipitation times.
Figure 10. Surface metal atomic concentrations of the CeZr0.5Ti2Ox samples with different precipitation times.
Catalysts 08 00592 g010
Figure 11. TEM image (A) and the corresponding EDS mapping (B) for the Ce of the CeZr0.5Ti2Ox catalyst.
Figure 11. TEM image (A) and the corresponding EDS mapping (B) for the Ce of the CeZr0.5Ti2Ox catalyst.
Catalysts 08 00592 g011

Share and Cite

MDPI and ACS Style

Shan, W.; Geng, Y.; Zhang, Y.; Lian, Z.; He, H. A CeO2/ZrO2-TiO2 Catalyst for the Selective Catalytic Reduction of NOx with NH3. Catalysts 2018, 8, 592. https://doi.org/10.3390/catal8120592

AMA Style

Shan W, Geng Y, Zhang Y, Lian Z, He H. A CeO2/ZrO2-TiO2 Catalyst for the Selective Catalytic Reduction of NOx with NH3. Catalysts. 2018; 8(12):592. https://doi.org/10.3390/catal8120592

Chicago/Turabian Style

Shan, Wenpo, Yang Geng, Yan Zhang, Zhihua Lian, and Hong He. 2018. "A CeO2/ZrO2-TiO2 Catalyst for the Selective Catalytic Reduction of NOx with NH3" Catalysts 8, no. 12: 592. https://doi.org/10.3390/catal8120592

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