Next Article in Journal
Effect of CO Concentration on the α-Value of Plasma-Synthesized Co/C Catalyst in Fischer-Tropsch Synthesis
Next Article in Special Issue
Comparison of Efficiencies and Mechanisms of Catalytic Ozonation of Recalcitrant Petroleum Refinery Wastewater by Ce, Mg, and Ce-Mg Oxides Loaded Al2O3
Previous Article in Journal
Development of Ni-Based Catalysts Derived from Hydrotalcite-Like Compounds Precursors for Synthesis Gas Production via Methane or Ethanol Reforming
Previous Article in Special Issue
Synergically Improving Light Harvesting and Charge Transportation of TiO2 Nanobelts by Deposition of MoS2 for Enhanced Photocatalytic Removal of Cr(VI)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Novel Fe‐W‐Ce Mixed Oxide for the Selective Catalytic Reduction of NOx with NH3 at Low Temperatures

1
Key Laboratory of Biofuels, Qingdao Institute of Bioenergy and Bioprocess Technology, Chinese Academy of Sciences, Qingdao 266101, China
2
Institute of Combustion and Power Plant Technology, University of Stuttgart, Stuttgart 70569, Germany
3
School of Chemistry and Environmental Engineering, Wuhan Institute of Technology, Wuhan 430205, China
*
Authors to whom correspondence should be addressed.
Catalysts 2017, 7(2), 71; https://doi.org/10.3390/catal7020071
Submission received: 16 January 2017 / Accepted: 16 February 2017 / Published: 20 February 2017
(This article belongs to the Special Issue Heterogeneous Catalysis for Environmental Remediation)

Abstract

:
A set of novel iron doped cerium-tungsten catalysts were prepared by sol-gel method with a view to their application for low temperature selective catalytic reduction (SCR) of NOx with NH3 in power plants. With a molar ratio Fe/W/Ce of 0.5:1:1, a NOx reduction of >90% at 200 °C was achieved. In Fe-W-Ce catalysts with low iron oxide content, it was found that the iron compounds were highly dispersed and formed a solid solution within the cerium oxide lattice, which promoted the SCR activity. Large amounts of iron in the catalysts might form a layer of Fe2O3 on the catalyst surface, which induced the synergistic inhibition effect among Fe, Ce and W species. Moreover, the Fe-W-Ce catalysts possessed a high resistance to changed operation parameters as well as to deactivation by SO2 and/or H2O. The novel catalyst showed to be competitive among recently developed low-temperature SCR catalysts.

1. Introduction

During combustion processes, many different kinds of pollutants can be generated, such as NOx. NOx emissions can have a strong negative impact on the environment and on human health, which result in photochemical smog and acid rain [1]. To avoid these damages, several techniques for the abatement of NOx (DeNOx) have been developed in the past. A widely applied technique of NOx abatement is the selective catalytic reduction (SCR) with ammonia (NH3) as reduction agent. Mostly vanadium catalysts on titanium support doped with tungsten or molybdenum oxide are commercially used (denoted as V2O5-WO3/TiO2 and V2O5-MoO3/TiO2, respectively) in a relatively narrow temperature window of 300–400 °C [2,3]. However this system has some problems, such as the volatility and toxicity of vanadium species, poisoning by K2O and PbO, low N2 selectivity at high temperatures and high conversion of SO2 to SO3 [4]. Therefore, it is necessary to develop novel catalysts that operate at lower temperatures [5].
With the aim to find a replacement for conventional catalysts, diverse catalysts have been studied extensively over the past years, such as Mn-oxides [6] and other transition metals oxides such as Ce-Fe/TiO2 [7], CuO/Ti0.95Ce0.05O2 [8], Ce-Nb binary oxide [9], Ba/CeO2-MOx (M = Mn, Fe) mixed oxides [10], Ce-Sn-Ox [11] or Ce-Ta mixed oxides [12] have been used as additives. Iron oxide is reported to especially enhance the NOx conversion at low temperatures. Moreover, Fe-doped Mn-Ce/TiO2 catalysts can be used since iron is said to enhance the resistance to H2O and SO2 [13,14]. Additionally, the shift between Ce3+ and Ce4+ is also hold responsible for the enhancement of NO to NO2 oxidation by cerium oxide, which favors the fast SCR reaction [15,16]. In addition, tungsten oxide is reported to widen the operation temperature window. For example, Shan et al. describe a NOx conversion of almost 100% from 250 to 450 °C over Ce-W mixed oxides [17,18]. The WO3-ZrO2 catalyst exhibited >80% NOx conversion at temperatures as low as 150 °C [19]. A reason may be that WO3 is said to increase the strength and amount of Brönsted acid sites on catalyst surface [20]. Chen et al. assume that tungsten promotes the NO to NO2 oxidation, which facilitates the fast SCR reaction that can occur at low temperatures [21]. Furthermore, tungsten seems to have the ability to inhibit the oxidation of NH3 to N2O and therefore increases the N2 selectivity [19]. Besides, in combination with Ce, tungsten addition can lead to an improved dispersion of Ce and the amount of Ce3+ which leads to an increased chemisorption of oxygen [20].
Thus the combination of cerium and tungsten with iron might promise a high NOx reduction ability at lower temperatures with a high N2 selectivity and good resistance to deactivation by diverse compounds. This novel catalyst is developed and characterized, and the catalytic performance is also tested under diverse conditions, such as varying gas hourly space velocities (GHSV), SO2 and H2O flue gas content and NH3/NO ratio. The results of these tests shall facilitate an estimation of the catalyst’s behavior under “real” conditions in a power plant.

2. Results

2.1. Activity Measurement

Figure 1 shows the NH3-SCR performance for all tested catalysts over a temperature range of 100 to 500 °C. Without Fe doping, the W-Ce catalyst showed more than 80% NOx conversion at above 190 °C. The addition of Fe to W-Ce mixed oxide has obvious influence on the catalytic activity. When the Fe/W/Ce molar ratio was controlled at 0.5:1:1, the low temperature SCR activity was significantly enhanced with 80% NOx conversion at 160 °C. Further increasing the Fe/W/Ce molar ratio to 1:1:1 and 2:1:1 resulted in the lower SCR activity, and the 80% NOx conversion could be obtained at as high as 250 °C. Moreover, all catalysts also showed the higher N2 selectivity (>95%) in the whole temperature range (Figure 1B). Therefore, the optimal Fe addition could contribute to improving the NH3-SCR activity.

2.2. BET and XRD

All the samples show type IV isotherms according to the International Union of Pure and Applied Chemistry (IUPAC) classification (shown in Figure S1), which are typical for mesoporous materials (pore diameter 2–50 nm) [8]. The results of the BET analysis for this study are listed in Table 1. It could be found that compared to W-Ce catalyst, the specific surface area was almost tripled at the highest iron loading. This also indicated that there was no link between a large specific surface area and a good catalytic performance. In addition, the molar ratio of Fe/W/Ce in samples (determined by inductively coupled plasma and atomic emission spectrometry (ICP-AES), Table 1) is more or less the same to the theoretical value. The XRD patterns of the Fe containing catalysts are compared to the W-Ce catalyst in Figure 2. All peaks of W-Ce catalyst can be assigned to cerianite (CeO2, 2θ = 28.7°, 33.8°, 47.4°, 56.0°, 69.5° and 76.6°) and tungsten oxide (WO3, 2θ = 23.7°, 28.7°, 33.8° 41.6°, 47.4°, 49.9°, 53.8° and 60.2°). This accorded with other studies in which CeO2 and WO3 were also the dominant species [21]. Meanwhile, the peaks were not very sharp, which pointed at low crystallinity of the sample [16]. This phenomenon might result from the addition of WO3 to CeO2 as reported in literature, which was regarded to be an indication of strong interactions between the metal oxides [22]. For Fe0.5-W-Ce catalyst, no new peaks appear, but the CeO2 and WO3 peaks became less intense and broader. This was reflected by the small crystallite size of the sample, which was usually associated with a high catalytic performance [22]. In addition, two possible explanations were given for the absence of new peaks from Fe compounds in Fe0.5-W-Ce. First, Fe oxides existed in an amorphous state and were highly dispersed over the surface of the WO3 and CeO2 lattice [14]. Second, smaller Fe ions (atomic radius 140 pm) can substitute larger Ce ions (atomic radius 185 pm) in the CeO2 lattice [23], Fe/Ti oxides [24] as well as Fe/Mn oxides [25].
For Fe1-W-Ce and Fe2-W-Ce catalysts, the larger crystallite size was observed compared with W-Ce and Fe0.5-W-Ce samples (Table 1). Nevertheless the peak at 35° allowed a clear identification as γ-Fe2O3, which has previously been reported to form in Fe catalysts [13]. Probably, the appearance of the peak meant that the Fe ions can no longer be incorporated into the Ce lattice. Instead, a layer of Fe2O3 formed on the catalyst surface. This phenomenon already occurred for other catalysts with high iron loadings [26]. It might be concluded that this Fe2O3 layer covered the catalyst surface and hindered the interaction between the W-Ce, which induced the lower NH3-SCR activity.

2.3. Raman Characterization

Figure 3 depicts the Raman spectrum of the samples investigated in this work. The spectrum of W-Ce shows a sharp peak at 460 cm−1, which belongs to the F2g mode of the symmetric breathing mode of oxygen atoms surrounding cerium ions in the cubic fluorite phase CeO2 [27]. In addition, typical WO3 crystallization peaks (272, 711, and 810 cm−1) occur on the W-Ce catalysts, which agrees with the XRD results. However, it is worth mentioning that the peak that belonged to the CeO2 becomes weaker significantly for Fe0.5-W-Ce catalyst, which may be related with the decreased CeO2 particle size (Table 1) [28]. As for Fe1-W-Ce and Fe2-W-Ce catalyst, the spectrum of CeO2 becomes stronger than that of Fe0.5-W-Ce, suggesting that only proper adding of Fe inhibited the crystallization of CeO2. The conclusion is highly consistent with the XRD results.

2.4. UV-Vis Diffuse Reflectance Spectroscopy

Figure 4 shows the UV-vis absorption spectra of W-Ce, Fe0.5-W-Ce, Fe1-W-Ce and Fe2-W-Ce. All samples contain three absorption maxima centered at ~230 nm, 290 nm and 320 nm. The former maxima corresponded to Ce3+ → O2− charge transfer transitions [29], which implied the occurrence of oxygen vacancy defects [30]. The latter two adsorption maxima were ascribed to oxygen-to-Ce4+ charge transfer and interband transitions, respectively [31]. In addition, an additional peak shoulder can be identified around 380 nm, which is attributed to WO3 [32]. An obvious blue shift of the peaks corresponding to tungsten and ceria compounds can be seen compared with pure WO3. This may be a result of a strong interaction between tungsten and ceria species [33]. These findings are in accordance with the XRD results, which predicted also strong interaction between ceria and tungsten.
Meanwhile, Fe compounds show a broad peak centered at 530 nm. In literature the peak has been assigned to large Fe2O3 particles [34]. The intensity of the peaks can be correlated with the degree of dispersion of metal oxides. The Fe1-W-Ce catalyst shows a higher intensity peak of Fe2O3, suggesting that a relatively lower dispersion of Fe2O3 in Fe1-W-Ce catalyst. The peak intensity of Fe2-W-Ce is higher than Fe1-W-Ce, which approves that Fe2-W-Ce shows the lowest dispersion of Fe2O3. The results confirm that the Fe compounds in Fe0.5-W-Ce are highly dispersed [24]. Overall, the UV-vis results are in full agreement with the conclusions drawn from the XRD. This includes the strong interaction between CeO2 and WO3, the high degree of iron dispersion in Fe0.5-W-Ce and the occurrence of large amounts of Fe2O3 in the Fe1-W-Ce and Fe2-W-Ce catalyst.

2.5. Influence of GHSV and NH3/NO Ratio

The NOx reduction with different NH3/NO ratios is shown in Figure 5. Previous experiments have been carried out with a NH3/NO ratio of 1.2:1. The plot shows that there is a significant decrease of the overall NOx reduction with a NH3/NO ratio of 0.9:1. A maximal NOx conversion of 91% is reached at 350 °C. Furthermore, the NOx conversion decreases at temperatures >400 °C on NH3/NO ratio of 0.9:1. This is not observed with higher NH3/NO ratios. The NH3/NO ratio has a strong influence on the NOx conversion when it drops below the stoichiometric value of 1. In literature the SCR reaction is described to proceed with zero order concerning NH3 at NH3/NO > 1 and with first order for NH3/NO < 1 [35]. This fits with the results shown in Figure 5. For NH3/NO > 1 the adsorption process of NH3 to the catalyst surface is said to limit the reaction progress [36]. This is reflected by the minimal difference of the results for NH3/NO 1.1 and 1.2. The strong dependence is not seen as an obstacle for practical application, since the NH3/NO ratio can be adapted while the process is running.
The influence of GHSV on NOx reduction is tested. Figure 6 shows the slight decrease of NOx conversion from 99% to 96% at 500 °C, when the GHSV is doubled from 20,000 to 40,000 h−1. The GHSV has a less pronounced influence on the NOx conversion compared to the NH3/NO ratio. A variation of the GHSV of 50% (20,000 to 40,000 h−1) leads to a NOx conversion of 96%, while a change of the NH3/NO ratio by only 25% (1.2 to 0.9) decreases the NOx conversion to 91%. Hence, the SCR reaction is more sensitive to changes in the NH3/NO ratio.
From Figure 6, it can be seen that the impact of GHSV also depends on the reaction temperature. At 500 °C the difference between the NOx conversion with high and low GHSV is only 3.2%, while at lower temperatures, e.g., 170 °C, the difference is 13.1%. This connection was also found by other researchers [37]. The temperature dependence is problematic as far as the practical application is concerned, since it is intended to use the catalysts for low-temperature SCR.

2.6. Influence of SO2 and H2O

The influence of additional gas streams of 200 ppm SO2 and/or 5% H2O on NOx reduction is investigated. The temperature is kept constant at 350 °C. The NOx reduction is registered some minutes before and after the additional gas streams are turned on and off. The results are displayed in Figure 7. After SO2 addition, the NOx conversion decreases from 99% to 95%. When water is added, the NOx reduction decreases to 91% and when SO2 and water are added simultaneously, it becomes about 84%. Furthermore, it is observed that the fluctuations of the measured NOx concentrations increase in the following order: SO2 < H2O < SO2 and H2O. After the additional gases are turned off, the NOx reduction efficiency returns in all three cases to a value very similar to the initial value.
The SO2 and H2O resistance of the Fe0.5-W-Ce catalyst is cross compared to some experimental values found in literature. The applied operation conditions vary widely. Especially the reaction temperature is reported to have a strong influence. At higher temperatures the impact of SO2 and H2O is less severe than at lower temperatures [38]. As mentioned above the SCR performance of the Fe0.5-W-Ce catalyst is investigated only at 350 °C. Other temperatures should be tested in future experiments. It is often stated that the influence of H2O is reversible, while the impact of SO2 on the reaction is permanent and only partial recovery is possible [39]. Therefore the terms “H2O inhibition” and “SO2 deactivation” are used. This strong influence is not observed in the present study. The NOx conversion decrease is almost 100% reversible after H2O and/or SO2 treatment (Figure 7). Reasons for the good deactivation resistance can be found in the literature. Tungsten [19] and iron [14] are said to enhance the resistance of catalysts. For tungsten oxide, this fact is attributed to its ability to lower the thermal stability of cerium sulfates, which would otherwise block the active sites of the catalyst [21]. Since tungsten and iron are present in Fe0.5-W-Ce, it is assumed that in the current study these effects come into operation. Thus, the Fe0.5-W-Ce catalyst shows a very good resistance to SO2 and H2O poisoning, which supports its applicability in a power plant.
In the stability test, the SCR reaction showed a stable behavior over 80 h (Figure 8). During the whole period, the NOx reduction stays almost constant. It fluctuates between 99.0% and 98.4% with an average of 98.7%. Even though the Fe0.5-W-Ce catalyst shows a very good stability, the desired catalyst lifetime in practical application is thousands of hours-a timespan that can not be tested in laboratory scale experiments. Some researchers predict the deactivation of SCR catalysts by ammonium nitrate (NH4NO3) [40]. In the present experiment heating of the gas tubes is applied to exclude liquid water from the reaction and avoid NH4NO3 formation. A physicochemical characterization of the spend catalysts could reveal if ammonium nitrate is nevertheless formed in significant amounts.

3. Discussion

It can be concluded that this Fe2O3 layer covers the catalyst surface and hinders the interaction between the CeO2-WO3 and the reactants in the gas phase (NO, NH3). Hints to a highly dispersed form of iron oxide in Fe0.5-W-Ce are given by the XRD, Raman and UV-vis results. The formation of a solid solution of Fe and Ce as a new active phase can lead to a higher NOx reduction efficiency [25]. This can be an explanation why the Fe0.5-W-Ce performs better compared to the W-Ce catalyst. The idea that a large amount of Fe2O3 may cover the active sides of WOx and CeO2 in Fe1-W-Ce and Fe2-W-Ce is supported by the XRD and UV-vis results.
The XRD patterns of the samples after reaction are shown in Figure S2. The characteristic peaks of Fe-W-Ce catalysts are similar with that of the fresh samples (Figure 2), indicating the crystal structure of FeW-Ce are not destroyed after reaction. Furthermore, no significant changes for the characteristic peaks of Fe0.5-W-Ce after the experiments of SO2 and H2O resistance are observed (Figures S3 and S4), suggesting the Fe0.5-W-Ce could possess the high stability. Nevertheless, the catalyst after the experiments of SO2 and H2O resistance results in a decrease in surface area (Table S1). The decrease in BET surface area is primarily due to the surface sulfate species formation.
At the first glance, the weak SCR performance of Fe1-W-Ce and Fe2-W-Ce samples seem to be contradictory to its high specific surface area. However, as mentioned before, the direct neighbors of each atom in a crystal lattice are more important for the SCR than the overall crystallinity [21]. In the next step, the SCR performance of the novel catalysts shall be compared to previous studies. For this task, three criteria are chosen: the NOx conversion, the lowest operation temperature, and the size of the operation temperature window. It is important to define the term “operation temperature” as it is used in this study. Two definitions for low temperature SCR applications exist in literature: some researchers claim that the operation temperature window comprises the temperature range where the NOx reduction is >80% [41]. Other research groups set a threshold of >90% NOx reduction [25]. Because stricter legal regulations for NOx emissions can be expected in the future the value of 90% is used in this study. The lowest operation temperature should be in the ideal case <200 °C, since this is the temperature of the flue gases at the tail-end of a power plant.
Firstly, catalysts that contain only cerium and tungsten oxides shall be compared. Our W-Ce catalyst reaches the threshold of 90% NOx reduction at 230 °C and slightly misses the ideal operation temperature of <200 °C. The maximum NOx reduction for these catalysts is 99%. The catalysts of other research groups show a similarly high efficiency (Table 2). Peng et al. reached the 90% mark at 185 °C [42], Ma et al. at 190 °C [33], Shan et al. at 215 °C [17] and Liu et al. at 290 °C [43]. It has to be kept in mind that different preparation techniques and operation conditions have been used. Roughly speaking, the W-Ce catalyst achieves average performance results. Considering the iron oxide catalysts, Fe0.5-W-Ce achieves a NOx reduction >90% at 200 °C and a maximal NOx reduction of 99%. It therefore fulfills the criteria set for a low temperature SCR catalyst. In literature one finds iron oxide catalysts with >90% NOx reduction at 150 °C [13] and at 350 °C [35]. Another Fe catalysts prepared by the sol-gel method only reaches 55% maximal NOx reduction [14].

4. Materials and Methods

4.1. Catalyst Preparation

Four W-Ce catalysts doped with different amounts of Fe were prepared by the sol-gel method. First an appropriate amount of ammonium tungstate was added to deionized water under continuous stirring. To dissolve the ammonium tungstate, an equal amount of oxalic acid dihydrate was added and the solution was mixed until it became clear. The molar ratios were Fe/W/Ce = 0:1:1, 0.5:1:1, 1:1:1 and 2:1:1. The Fe(NO3)3·9H2O and Ce(NO3)3·6H2O was the source of iron and Cerium, respectively. As complexing agent citric acid was added in a molar ratio of 1.3:1 to all metal ions. Last polyethylene glycol (PEG) was added in an amount corresponding to 50% (w/w) of citric acid. Afterwards, the samples were heated to 80 °C under ongoing stirring until the gel formation was completed. The gel was then put to a furnace at 120 °C until the gel was dry. Subsequently, the samples were calcined at 500 °C for 3 h. After calcination, the catalysts were pressed, crushed and sieved to 20–40 mesh. The catalysts are denoted as Fex-W-Ce, and x stands for the molar ratio of Fe to W.

4.2. Catalyst Characterization

The specific surface areas were obtained from N2 adsorption/desorption isotherms measured on a Micromeritics ASAP 2020 M (Quantachrome, Boynton Beach, FL, USA) at −196 °C, then are calculated using the Brunauer–Emmett–Teller (BET) method. The ICP-AES was used to determine the molar ratio of Fe/W/Ce in the synthesized samples, which was performed on an OPTIMA 2000 (PerkinElmer, Waltham, MA, USA). The XRD patterns of catalysts were collected on a Bruker AXS-D8 Advance powder diffractometer (Bruker, Karlsruhe, Germany) with a Cu Kα radiation source of wavelength 1.5406 Å, and the crystallite size was calculated by the main peak at 28.7° according to Scherrer’s equation. UV Raman spectra were recorded on a DXR Microscope Raman spectrograph (ThermoFisher Scientific, Waltham, MA, USA) with He-Cd laser of 325 nm excitation wavelength. The UV-vis diffuse reflectance spectroscopy (DRS) was carried out on a Hitachi U-4100 UV/Vis/NIR spectrophotometer (Hitachi, Tokyo, Japan), and wavelengths from 200 to 800 nm were scanned for light absorption.

4.3. Activity Measurement

The NH3-SCR activity was performed in a fixed bed reactor using a MRU vario plus industrial flue gas measurement device. The Fe-W-Ce catalyst (0.3 g, 20–40 mesh) was mixed with silica sand (0.4 g, 20–40 mesh), and then filled into a quartz reactor with 8 mm inner diameter. The reaction conditions were controlled as follows: 450–500 ppm NOx, 450–500 ppm NH3, 2.5% O2, N2 as balance. Before the activity test, the measured values of the gas concentrations were allowed to stabilize at room temperature for around 1.5 h. Each temperature was kept until the measured values stabilize (40–60 min).

5. Conclusions

The investigation of the SCR performance reveals a good NOx conversion and high resistance to differences in operation parameters as well as to poisoning by flue gas compounds. In comparison to the literature, the very wide operation temperature window of the W-Ce and the Fe0.5-W-Ce catalyst is remarkable. A NOx reduction of >90% is reached over a span of 260 °C for W-Ce and over a span of 290 °C for Fe0.5-W-Ce. This indicates that the newly developed catalysts can be used in low temperature SCR as well as at higher temperatures. They are very flexible in application.

Supplementary Materials

The following are available online at www.mdpi.com/2073-4344/7/2/71/s1, Figure S1: The N2 adsorption-desorption isotherms of Fe-W-Ce catalysts, Figure S2: XRD patterns of Fe-W-Ce catalysts after reaction, Figure S3: XRD patterns of Fe0.5-W-Ce catalysts after reaction (a: SO2 + NOx + NH3 + O2, b: SO2 + NOx + NH3 + H2O + O2), Figure S4: Raman patterns of Fe0.5-W-Ce catalysts after reaction (a: SO2 + NOx + NH3 + O2, b: SO2 + NOx + NH3 + H2O + O2), Table S1: Specific surface area and crystallite size of Fe0.5-W-Ce catalysts after reaction.

Acknowledgments

The work described above was supported by the National Natural Science Foundation of China (No: 21607162, 21276263 and 21506239), the Doctoral Fund of Shandong Province (BS2015HZ003), Qingdao Indigenous Innovation Program (16-5-1-27-jch) and the Open Foundation of Key Laboratory of Industrial Ecology and Environmental Engineering, MOE (KLIEEE-15-05).

Author Contributions

Zhong Wang and Xuebing Li conceived and designed the experiments; Anna Stahl performed the experiments; Anna Stahl and Zhong Wang analyzed the data; Anna Stahl and Zhong Wang wrote the paper; and Jun Ke and Tobias Schwämmle revised the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Parvulescu, V.; Grange, P.; Delmon, B. Catalytic removal of NO. Catal. Today 1998, 46, 233–316. [Google Scholar] [CrossRef]
  2. Qi, G.; Yang, R. Low-temperature selective catalytic reduction of NO with NH3 over iron and manganese oxides supported on titania. Appl. Catal. B 2003, 44, 217–225. [Google Scholar] [CrossRef]
  3. Alemany, J.; Lietti, L.; Ferlazzo, N.; Forzatti, P.; Busca, G.; Ramis, G.; Giamello, E.; Bregani, F. Reactivity and physicochemical characterisation of V2O5-WO3/TiO2 De-NOx catalysts. J. Catal. 1995, 155, 117–130. [Google Scholar] [CrossRef]
  4. Kamata, H.; Takahashi, K.; Odenbrand, C. The role of K2O in the selective reduction of NO with NH3 over a V2O5(WO3)/TiO2 commercial selective catalytic reduction catalyst. J. Mol. Catal. A 1999, 139, 189–198. [Google Scholar] [CrossRef]
  5. Lee, S.; Park, K.; Kim, S.; Kwon, D.; Hong, S. Effect of the Mn oxidation state and lattice oxygen in Mn-based TiO2 catalysts on the low-temperature selective catalytic reduction of NO by NH3. J. Air Waste Manag. Assoc. 2012, 62, 1085–1092. [Google Scholar] [CrossRef] [PubMed]
  6. Eigenmann, F.; Maciejewski, M.; Baiker, A. Selective reduction of NO by NH3 over manganese-cerium mixed oxides: Relation between adsorption, redox and catalytic behavior. Appl. Catal. B 2006, 62, 311–318. [Google Scholar] [CrossRef]
  7. Shu, Y.; Tanana, A.; 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 2014, 150–151, 630–635. [Google Scholar] [CrossRef]
  8. Guo, R.T.; Zhen, W.L.; Pan, W.G.; Zhou, Y.; Hong, J.N.; Xu, H.J.; Jin, Q.; Ding, C.G.; Guo, S.Y. Effect of Cu doping on the SCR activity of CeO2 catalyst prepared by citric acid method. J. Ind. Eng. Chem. 2014, 20, 1577–1580. [Google Scholar] [CrossRef]
  9. 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, 142–143, 290–297. [Google Scholar] [CrossRef]
  10. Mousavi, S.; Niaei, A.; Gómez, M.; Salari, D.; Panahi, P.; Abaladejo-Fuentes, V. Characterization and activity of alkaline earth metals loaded CeO2-MOx (M = Mn, Fe) mixed oxides in catalytic reduction of NO. Mater. Chem. Phys. 2014, 3, 921–928. [Google Scholar] [CrossRef]
  11. Li, X.; Li, Y.; Deng, S.; Rong, T. A Ce-Sn-Ox catalyst for the selective catalytic reduction of NOx with NH3. Catal. Commun. 2013, 40, 47–50. [Google Scholar] [CrossRef]
  12. Zhang, T.; Qu, R.; Su, W.; Li, J. A novel Ce-Ta mixed oxide catalyst for the selective catalytic reduction of NOx with NH3. Appl. Catal. B 2015, 176–177, 338–346. [Google Scholar] [CrossRef]
  13. Wang, X.; Gui, K. Fe2O3 particles as superior catalysts for low temperature selective catalytic reduction of NO with NH3. J. Environ. Sci. 2013, 12, 2469–2475. [Google Scholar] [CrossRef]
  14. Shen, B.; Liu, T.; Zhao, N.; Yang, X.; Deng, L. Iron-doped Mn-Ce/TiO2 catalyst for low temperature selective catalytic reduction of NO with NH3. J. Environ. Sci. 2010, 9, 1447–1454. [Google Scholar] [CrossRef]
  15. Liu, C.; Chen, L.; Li, J.; Ma, L.; Arandiyan, H.; Du, Y.; Xu, J.; Hao, J. Enhancement of activity and sulfur resistance of CeO2 supported on TiO2-SiO2 for the selective catalytic reduction of NO by NH3. Environ. Sci. Technol. 2012, 11, 6182–6189. [Google Scholar] [CrossRef] [PubMed]
  16. Liu, J.; Li, Y.; Ke, J.; Wang, Z.; Xiao, H.N. Synergically improving light harvesting and charge transportation of TiO2 nanobelts by deposition of MoS2 for enhanced photocatalytic removal of Cr(VI). Catalysts 2017, 7, 30. [Google Scholar] [CrossRef]
  17. 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]
  18. Nie, J.; Wu, X.; Ma, Z.; Xu, T.; Si, Z.; Chen, L.; Weng, D. Tailored temperature window of MnOx-CeO2 SCR catalyst by addition of acidic metal oxides. Chin. J. Catal. 2014, 35, 1281–1288. [Google Scholar] [CrossRef]
  19. Kwon, D.; Nam, K.; Hong, S. Influence of tungsten on the activity of a Mn/Ce/W/Ti catalyst for the selective catalytic reduction of NO with NH3 at low temperatures. Appl. Catal. A 2015, 497, 160–166. [Google Scholar] [CrossRef]
  20. Jiang, Y.; Xing, Z.; Wang, X.; Huang, S.; Wang, X.; Liu, Q. Activity and characterization of a Ce-W-Ti oxide catalyst prepared by a single step sol-gel method for selective catalytic reduction of NO with NH3. Fuel 2015, 151, 124–129. [Google Scholar] [CrossRef]
  21. Li, P.; Xin, Y.; Li, Q.; Wang, Z.; Zhang, Z.; Zheng, L. Ce-Ti amorphous oxides for selective catalytic reduction of NO with NH3: Confirmation of Ce-O-Ti active sites. Environ. Sci. Technol. 2012, 17, 9600–9605. [Google Scholar] [CrossRef] [PubMed]
  22. Ke, J.; Liu, J.; Sun, H.Q.; Zhang, H.Y.; Duan, X.G.; Liang, P.; Li, X.Y.; Moses, O.T.; Liu, S.M.; Wang, S.B. Facile assembly of Bi2O3/Bi2S3/MoS2n-p heterojunction with layered n-Bi2O3 and p-MoS2 for enhanced photocatalytic water oxidation and pollutant degradation. Appl. Catal. B 2017, 200, 47–55. [Google Scholar] [CrossRef]
  23. Xing, W.; Qian, K.; Dao, L. Catalytic combustion of chlorobenzene over MnOx-CeO2 mixed oxide catalysts. Appl. Catal. B 2009, 86, 166–175. [Google Scholar] [CrossRef]
  24. Liu, F.; He, H.; Zhang, C.; Feng, Z.; Zheng, L.; Xie, Y.; Hu, T. Selective catalytic reduction of NO with NH3 over iron titanate catalyst: Catalytic performance and characterization. Appl. Catal. B 2010, 96, 408–420. [Google Scholar] [CrossRef]
  25. Liu, F.; He, H.; Ding, Y.; Zhang, C. 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 2009, 93, 194–204. [Google Scholar] [CrossRef]
  26. Peña, D.; Uphade, B.; Smirniotis, P. TiO2-supported metal oxide catalysts for low-temperature selective catalytic reduction of NO with NH3. I. Evaluation and characterization of first row transition metals. J. Catal. 2004, 221, 421–431. [Google Scholar] [CrossRef]
  27. Wang, Z.; Qu, Z.P.; 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]
  28. Xu, J.; Li, P.; Song, X.; He, C.; Yu, J.; Han, Y. Operando Raman spectroscopy for determining the active phase in one-dimensional Mn1−xCexOy nanorod catalysts during methane combustion. J. Phys. Chem. Lett. 2010, 1, 1648–1654. [Google Scholar] [CrossRef]
  29. Bensalem, A.; Verduraz, F.B.; Delamar, M.; Bugli, G. Preparation and characterization of highly dispersed silica-supported ceria. Appl. Catal. A 1995, 121, 81–93. [Google Scholar] [CrossRef]
  30. Reddy, B.M.; Rao, K.N.; Bharali, P. Copper promoted cobalt and nickel catalysts supported on ceria-alumina mixed oxide: Structural characterization and CO oxidation activity. Ind. Eng. Chem. Res. 2009, 48, 8478–8486. [Google Scholar] [CrossRef]
  31. Liu, L.; Yao, Z.; Deng, Y.; Gao, F.; Liu, B.; Dong, L. Morphology and crystal-plane effects of nanoscale ceria on the activity of CuO/CeO2 for NO reduction by CO. ChemCatChem 2011, 3, 978–989. [Google Scholar] [CrossRef]
  32. Si, Z.; Wu, X.; Weng, D.; Ma, Z.; Ma, J. Tailored temperature window of CuOx/WOx-ZrO2 for NOx reduction via adjusting the calcination temperature of WOx-ZrO2. Mater. Chem. Phys. 2013, 138, 399–404. [Google Scholar] [CrossRef]
  33. Ma, Z.; Weng, D.; Wu, X.; Si, Z. Effects of WOx modification on the activity, adsorption and redox properties of CeO2 catalyst for NOx reduction with ammonia. J. Environ. Sci. 2012, 7, 1305–1316. [Google Scholar] [CrossRef]
  34. Hong, W.J.; Ueda, M.; Iwamoto, S.; Hosokawa, S.; Wada, K.; Kanai, H.; Deguchi, H.; Inoue, M. Effect of Fe content on physical properties of BaO-CeOx-FeOy catalysts for direct NO decomposition. Appl. Catal. B 2011, 106, 142–148. [Google Scholar] [CrossRef] [Green Version]
  35. Wang, C.; Yang, S.; Chang, H.; Peng, Y.; Li, J. Structural effects of iron spinel oxides doped with Mn, Co, Ni and Zn on selective catalytic reduction of NO with NH3. J. Mol. Catal. A 2013, 376, 13–21. [Google Scholar] [CrossRef]
  36. Yang, J.; Ma, H.; Yamamoto, Y.; Yu, J.; Xu, G.; Zhang, Z.; Suzuki, Y. SCR catalyst coated on low-cost monolith support for flue gas denitration of industrial furnaces. Chem. Eng. J. 2013, 230, 513–521. [Google Scholar] [CrossRef]
  37. 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]
  38. Yu, M.; Li, C.; Zeng, G.; Zhou, Y.; Zhang, X.; Xie, Y. The selective catalytic reduction of NO with NH3 over a novel Ce-Sn-Ti mixed oxides catalyst: Promotional effect of SnO2. Appl. Surf. Sci. 2015, 342, 174–182. [Google Scholar] [CrossRef]
  39. Lee, S.; Park, K.; Hong, S. MnOx/CeO2-TiO2 mixed oxide catalysts for the selective catalytic reduction of NO with NH3 at low temperature. Chem. Eng. J. 2012, 195–196, 323–331. [Google Scholar] [CrossRef]
  40. Ma, Z.; Wu, X.; Feng, Y.; Si, Z.; Weng, D. Effects of WO3 doping on stability and N2O escape of MnOx-CeO2 mixed oxides as a low-temperature SCR catalyst. Catal. Commun. 2015, 69, 188–192. [Google Scholar] [CrossRef]
  41. Chen, L.; Si, Z.; Wu, X.; Weng, D.; Ran, R.; Yu, J. Rare earth containing catalysts for selective catalytic reduction of NOx with ammonia: A review. J. Rare Earths 2014, 10, 907–917. [Google Scholar] [CrossRef]
  42. Peng, Y.; Liu, Z.; Niu, X.; Zhou, L.; Fu, C.; Zhang, H.; Li, J.; Han, W. Manganese doped CeO2-WO3 catalysts for the selective catalytic reduction of NOx with NH3: An experimental and theoretical study. Catal. Commun. 2012, 19, 127–131. [Google Scholar] [CrossRef]
  43. 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]
Figure 1. (A) NOx conversion; and (B) N2 selectivity of Fe-W-Ce catalysts. Reaction conditions: [NOx] = 450 ppm, [NH3] = 450 ppm, 2.5% O2, N2 as balance, GHSV = 20,000 h−1.
Figure 1. (A) NOx conversion; and (B) N2 selectivity of Fe-W-Ce catalysts. Reaction conditions: [NOx] = 450 ppm, [NH3] = 450 ppm, 2.5% O2, N2 as balance, GHSV = 20,000 h−1.
Catalysts 07 00071 g001
Figure 2. XRD patterns of Fe-W-Ce catalysts.
Figure 2. XRD patterns of Fe-W-Ce catalysts.
Catalysts 07 00071 g002
Figure 3. Raman patterns of Fe-W-Ce catalysts.
Figure 3. Raman patterns of Fe-W-Ce catalysts.
Catalysts 07 00071 g003
Figure 4. UV-vis of different catalysts.
Figure 4. UV-vis of different catalysts.
Catalysts 07 00071 g004
Figure 5. Influence of different NH3/NO ratios on NOx conversion over Fe0.5-W-Ce catalyst. Reaction conditions: [NH3] = 450 ppm, 2.5% O2, N2 as balance, GHSV = 20,000 h−1.
Figure 5. Influence of different NH3/NO ratios on NOx conversion over Fe0.5-W-Ce catalyst. Reaction conditions: [NH3] = 450 ppm, 2.5% O2, N2 as balance, GHSV = 20,000 h−1.
Catalysts 07 00071 g005
Figure 6. Influence of GHSV variation on NOx conversion over Fe0.5-W-Ce catalyst. Reaction conditions: [NOx] = 450 ppm, [NH3] = 450 ppm, 2.5% O2, N2 as balance.
Figure 6. Influence of GHSV variation on NOx conversion over Fe0.5-W-Ce catalyst. Reaction conditions: [NOx] = 450 ppm, [NH3] = 450 ppm, 2.5% O2, N2 as balance.
Catalysts 07 00071 g006
Figure 7. Influence of H2O and/or SO2 on NOx conversion over Fe0.5-W-Ce catalyst. Reaction conditions: [NOx] = 450 ppm, [NH3] = 450 ppm, 2.5% O2, N2 as balance, GHSV = 20,000 h−1.
Figure 7. Influence of H2O and/or SO2 on NOx conversion over Fe0.5-W-Ce catalyst. Reaction conditions: [NOx] = 450 ppm, [NH3] = 450 ppm, 2.5% O2, N2 as balance, GHSV = 20,000 h−1.
Catalysts 07 00071 g007
Figure 8. The long-term stability test on Fe0.5-W-Ce catalyst. Reaction conditions: [NOx] = 450 ppm, [NH3] = 450 ppm, 2.5% O2, N2 as balance, GHSV = 20,000 h−1.
Figure 8. The long-term stability test on Fe0.5-W-Ce catalyst. Reaction conditions: [NOx] = 450 ppm, [NH3] = 450 ppm, 2.5% O2, N2 as balance, GHSV = 20,000 h−1.
Catalysts 07 00071 g008
Table 1. Specific surface area and crystallite of different catalysts.
Table 1. Specific surface area and crystallite of different catalysts.
SamplesBET Surface Area (m2/g)Crystallite Size (nm)Fe/W/Ce Molar Ratio a
W-Ce24.810.51.09:0.96
Fe0.5-W-Ce39.45.40.61:1.03:1.06
Fe1-W-Ce53.27.31.08:0.95:1.03
Fe2-W-Ce77.012.82.12:1.03:1.10
a Detected by ICP-AES.
Table 2. Comparison of the activity for NH3-SCR on different catalysts.
Table 2. Comparison of the activity for NH3-SCR on different catalysts.
SamplesReaction ConditionsTemperature (°C)NOx Conversion (%)References
Fe2O3500 ppm NO, 500 ppm NH3, 3% O2, and balance N215090[13]
Fe-Mn-Ce/TiO20.06 vol % NO, 0.06 vol % NH3, 3 vol % O2 and pure N2 in balance18096[14]
Ce-W500 ppm NO, 500 ppm NH3, 5 vol % O2, balance N221590[17]
WO3/CeO2500 ppm NO, 500 ppm NH3, 5% O2 and N2 in balance19090[33]
Co/Fe2O3[NO] = 500 ppm, [NH3] = 500 ppm, [O2] = 2 vol %, balance gas: N235090[35]
Mn-CeO2-WO30.05% NO, 0.05% NH3, 3% O2, and in a N2 stream.18590[42]
CeO2-WO3500 ppm NO, 500 ppm NH3, 3 vol % O2 and balance in N229090[43]

Share and Cite

MDPI and ACS Style

Stahl, A.; Wang, Z.; Schwämmle, T.; Ke, J.; Li, X. Novel Fe‐W‐Ce Mixed Oxide for the Selective Catalytic Reduction of NOx with NH3 at Low Temperatures. Catalysts 2017, 7, 71. https://doi.org/10.3390/catal7020071

AMA Style

Stahl A, Wang Z, Schwämmle T, Ke J, Li X. Novel Fe‐W‐Ce Mixed Oxide for the Selective Catalytic Reduction of NOx with NH3 at Low Temperatures. Catalysts. 2017; 7(2):71. https://doi.org/10.3390/catal7020071

Chicago/Turabian Style

Stahl, Anna, Zhong Wang, Tobias Schwämmle, Jun Ke, and Xuebing Li. 2017. "Novel Fe‐W‐Ce Mixed Oxide for the Selective Catalytic Reduction of NOx with NH3 at Low Temperatures" Catalysts 7, no. 2: 71. https://doi.org/10.3390/catal7020071

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