Next Article in Journal
Nanostructured Inorganic Materials at Work in Electrochemical Sensing and Biofuel Cells
Next Article in Special Issue
Aqueous and Surface Chemistries of Photocatalytic Fe-Doped CeO2 Nanoparticles
Previous Article in Journal
Visualization of Gas Distribution in a Model AP-XPS Reactor by PLIF: CO Oxidation over a Pd(100) Catalyst
Previous Article in Special Issue
N2O Direct Dissociation over MgxCeyCo1−xyCo2O4 Composite Spinel Metal Oxide
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Effect of Sr Addition in Cu- and Fe-Modified CeO2 and ZrO2 Soot Combustion Catalysts

by
Verónica Rico-Pérez
,
Eleonora Aneggi
* and
Alessandro Trovarelli
Dipartimento Politecnico di Ingegneria e Architettura, Università degli Studi di Udine, via Cotonificio 108, 33100 Udine, Italy
*
Author to whom correspondence should be addressed.
Catalysts 2017, 7(1), 28; https://doi.org/10.3390/catal7010028
Submission received: 27 December 2016 / Revised: 11 January 2017 / Accepted: 11 January 2017 / Published: 17 January 2017
(This article belongs to the Special Issue Ceria-based Catalysts)

Abstract

:
This study investigates the activity of transition and alkaline-earth metal-doped catalysts supported on ceria or zirconia for the NOx-assisted oxidation of diesel particulate. A series of Cu- and Fe-impregnated catalysts over CeO2 and ZrO2 supports were prepared by incipient wetness impregnation and characterized by BET, X-ray diffraction (XRD), and temperature-programmed reduction (TPR) experiments while their catalytic activity was investigated in NOx-assisted reaction by means of temperature programmed oxidation experiments (TPO). Higher activity was achieved by copper modified catalysts; the addition of Sr positively affected the performance of the materials, suggesting a synergic effect between transition metals and alkaline-earth metal. The role of copper is correlated to the oxidation of NO to NO2, while strontium seems to be mainly involved in the storage of NOx species.

Graphical Abstract

1. Introduction

Diesel engines have increased popularity in recent years due to their higher efficiency compared to gasoline engines. Particulate matter (PM), or soot particles, are formed as undesired by-products of the combustion process, being one of the main pollutants emitted by diesel engines, together with NOx, CO, and unburned hydrocarbons [1]. Particulate, which is composed of aggregated carbonaceous soot and soluble organic fraction (SOF) of condensed hydrocarbons on soot, is a potential carcinogen and contributes to respiratory issues [2,3]. Regulations on diesel exhaust emissions have become always more and more stringent and this has stimulated researchers to develop emission-reduction technologies. Among the various methodologies developed, the diesel particulate filter (DPF) loaded with a soot oxidation catalyst is one of the most efficient devices to reduce soot emissions [4,5,6,7].
Two different strategies could be adopted for catalytic regeneration of the filter: one approach is to increase the contact points between the soot particles and the catalysts (i.e., by using fuel-borne catalyst additives or molten salt catalysts that can “wet” the soot surface and, therefore, decrease the oxidation temperature [8,9,10]). Another strategy is to use an oxidation catalyst that can promote the formation of mobile compounds (i.e., active oxygen species, or NO2 from NO) which then act as the true oxidant for soot in a more active and efficient way compared to O2. An example is the oxidation of soot over Pt-based catalysts where Pt promotes NO oxidation to NO2, and the latter oxidizes carbon. Subsequently, the nitrogen dioxide will convert soot into CO and CO2 and back to NO, which can participate in the next soot oxidation event [11].
Several catalytic materials were developed in the last decade [12] and the most promising are based on transition metals (Ag, Co, Cu, Mn, Fe) [13,14,15,16,17,18,19,20,21,22,23,24,25,26,27], ceria-doped materials [28,29,30,31,32,33,34,35,36], alkaline or alkaline-earth metals [37,38,39,40,41], and perovskites [42,43]. The redox capacity of transition metals, cycling between more oxidation states, can efficiently enhanced soot oxidation activity. The mechanism of the reaction is based on the formation of “active oxygen” species that can easily react with soot, decreasing the temperature of combustion. Moreover, their redox cycle can also successfully oxidize NO to NO2, thus activating the NOx-assisted reaction as an alternative soot oxidation route, where nitrogen dioxide is directly involved in soot combustion due to its more powerful oxidation capacity. Among transition metals, Fe and Cu have often been employed as dopants/promoters of soot oxidation catalysts supported on metal oxides with different properties, such as alumina, titania, zirconia, and ceria [16,19,22,44,45,46,47]. The mechanism path commonly proposed for Cu- and Fe-based materials is correlated to the presence of metal oxide particles and their capability to be cyclically reduced and then reoxidized by O2 gas phase or lattice oxygen from the support [45,48,49].
Cerium and zirconium oxide are widely used as supports in several catalytic reactions. Zirconia is a “non-reducible” oxide (in the typical reaction conditions used for soot oxidation) that displays useful physical and chemical properties, i.e., chemical stability, low thermal conductivity, and high corrosion resistance. On the other hand, ceria-based materials are successfully used in several catalytic applications due to their high availability of surface oxygen and high surface reducibility [28,29]. It is believed that the activity of ceria on soot combustion is due to its redox activity and its ability to deliver oxygen from the lattice to the carbon particles [50].
The introduction of alkali or alkaline-earth metals, on catalytic formulations, results in an enhanced catalytic activity in soot oxidation [37,40,41]. Alkaline-earth metals can promote soot combustion activity, working either as oxidation catalysts or NOx storage traps [51,52,53,54]. Indeed, on alkaline-earth metal NOx storage occurs suggesting their use as a promising component in LNT (lean NOx trap) applications where NOx is stored during the lean phase and then it is reduced in the rich interval. Recently, a synergic effect by simultaneous doping of CeO2 by alkaline or alkaline-earth metals (K and Ba) and copper has been reported [54,55]. The lowering of the oxidation temperatures is obtained by the increased NOx storage capacity brought about by K (or Ba) and Cu.
In light of these considerations, in this study we proposed to combine the potential capacity of copper and iron supported catalysts with the promotional effect of an alkaline metal like strontium. A series of Cu- and Fe-impregnated catalysts over CeO2 and ZrO2 supports were prepared and the effect of the presence of Sr in the formulation was investigated. The catalysts have been characterized and their catalytic activity was investigated in NOx-assisted soot oxidation reaction by means of temperature-programmed oxidation experiments (TPO).

2. Results and Discussion

2.1. Textural and Structural Characterization

Composition and BET surface area of the catalysts are reported in Table 1. Similar BET surface areas have been found for all catalysts (in the range 17–35 m2/g). The addition of a single component to the support (CeO2 or ZrO2) does not affect the surface area that remains almost stable, while a small decrease was observed after addition of the two dopants in the Sr-doped materials. In accordance with the BET results, crystal size values are in the range 13–22 nm.
The structural features of all the materials were analysed by powder X-ray diffraction (XRD). Diffraction profiles of ceria-based catalysts (Figure 1A) exhibit reflections characteristic of a pure fluorite phase, while zirconia-based materials show a monoclinic structure (Figure 1B). In ceria-based catalysts, two very weak peaks, characteristic of a CuO phase, were also detected at 2θ = 35.5° and 38.8° for CeCu and CeSrCu samples, which is in agreement with other studies in the literature [20,54]. SrCO3 phase was also detected for Sr impregnated materials (CeSr, CeSrFe, and CeSrCu) and it may occur as a consequence of the exposure of the catalysts to ambient air conditions; indeed, the basic oxides such as the oxides of alkaline-earth elements are readily carbonated when exposed in air [56,57]. No evidence of the presence of Fe2O3 or other iron-containing phases was obtained, in accordance with the literature [22,44,58]. XRD features do not indicate formation of any ceria solid solution with copper or iron, suggesting that Fe or Cu are dispersed on the surface. It is known that lower valence ions, such as Fe3+ and Cu2+, do not easily dissolve into the ceria lattice using conventional impregnation methods; dissolution, if present, is limited to a small fraction of the overall loading [22,59]. To facilitate formation of solid solution alternative preparation methods are required [60,61]. In zirconia-based samples no evidence for any copper or iron phase was found, while formation of SrCO3 and SrZrO3 phases was observed.

2.2. Reduction Behaviour

In order to characterize the reduction behaviour of the materials, temperature-programmed reduction experiments with H2 (H2-TPR) have been carried out on all samples (Figure 2). Figure 2A shows the temperature programmed reduction profiles of ceria and ceria modified catalysts. The reduction feature of pure ceria is well known and it shows the characteristic bimodal profile with two peaks at low (ca. 525 °C) and high (ca. 840 °C) temperature attributable, respectively, to the reduction of small crystallites and/or surface ceria and to the reduction of bulk and large ceria crystallites [62]. For pure zirconia, a TPR feature of a typical ‘‘non-reducible’’ support was found (Figure 2B). The reduction profiles of the two supports are affected by the addition of metals.
ZrFe shows two reduction signals that could be ascribed to the existence of free Fe2O3 on the support surface. The hydrogen reduction profile for Fe2O3 occurs in two steps, with Fe2O3 first converting to Fe3O4 (with maximum at around 340 °C) and then to Fe (with maximum at ca. 550 °C) [63,64]. In CeFe, reduction of iron species overlap with the reduction features of ceria and four peaks are distinguished: the first one at around 350 °C was related to the reduction of hematite to magnetite state; the second and third peaks at around 450 °C and 600 °C can be attributed to the surface reduction of Ce4+ and the reduction of Fe2+ to Fe, while the last signal at 825 °C can be associated with the reduction of bulk ceria [65,66].
Materials loaded with Sr evidenced a main peak at around 700–750 °C that could be correlated to the desorption of superficial carbonate species, as confirmed by the analysis of outlet gas composition followed by an online quadrupole mass-spectrometer. For CeSr, the peak overlaps with the bulk reduction feature of ceria.
For the CeSrFe, the H2-TPR profile undergoes some modifications respect CeFe and only three reduction features are detected due to the merging of the first two signals of CeFe. The low temperature peak (375 °C) is correlated to the reduction of Fe3+ to Fe2+ and to the reduction of surface Ce4+. The second signal (600 °C), corresponding to the third peak in CeFe, is not affected by the presence of Sr in the formulation. Moreover, the high temperature peak band (790 °C) is due to the overlapping of bulk reduction of ceria and the desorption of carbonate species from the Sr phase.
ZrSrFe is less affected by Sr addition and the TPR profile is the result of the combination of ZrSr and ZrFe features. The first two peaks are due to the iron oxide reduction (as in the ZrFe sample) while the high temperature signal is due to the surface carbonate adsorbed on Sr (as in ZrSr).
The addition of Cu on ceria modifies the redox properties of both species, as a consequence of the CuO-CeO2 interaction at the oxide interface [59,67,68,69]. TPR profile of CeCu exhibits three main reduction peaks at 130, 160, and 840 °C. The two low-temperature signals can be assigned to the reduction of CuOx species and surface Ce4+. It is well known that bulk CuO reduction takes place at around 315–380 °C [20,67,68] while in our material a shift to lower temperature of the CuOx reduction features occurs due to metal support interaction. In addition, we can observe that the low reduction signal of ceria at around 525 °C is not present, indicating that the hydrogen spillover process promotes surface ceria reduction at a much lower temperature [67,70]. The quantitative analysis of the TPR profile reveals that a part of Ce4+ is reduced at 100–300 °C; the amount of hydrogen consumption in this temperature range (1.10 mmol/gcat) is larger than that required for the complete reduction of CuOx species (0.79 mmol/gcat), in agreement with other studies [20,70]. The presence of copper can, therefore, promote the reduction of surface ceria at much lower temperatures. The presence of Sr slightly modifies the shape and position of the low temperature signals, and only one peak is visible in CeSrCu, shifted at a slightly higher temperature. The modification of the reduction profile at low temperature can also be associated to a detrimental effect on the reduction properties of the catalyst, confirmed by the decrease of the H2 consumption at low temperature compared to CeCu (0.83 vs. 1.10 mmol/gcat).
Generally, CuO-CeO2 materials exhibit more than one reduction peaks correlated to Cu [20,68,70], suggesting the presence of more than one copper oxide species. A great amount of literature data is available for TPR studies of Cu-based catalysts; although shape and position highly depend on the experimental conditions and on the preparation of the catalyst, the shift to lower temperature of the reduction peaks suggests a synergic interaction between CuO and CeO2. The low temperature peak (typically denoted as peak α) is usually attributed to highly-dispersed CuO species closely interacting with CeO2 (more easily reduced species), while the higher temperature peak (typically denoted as peak β) is assigned to the overlapping of the reduction of larger CuO particles (still highly dispersed and strongly interacting with the support) and surface Ce4+. Usually a third reduction peak at higher temperature is found in TPR profile of CuO-CeO2 catalyst (not present in our materials) and is correlated with reduction of segregated crystalline CuO [67,68].
The low temperature region of ZrCu sample is comparable to CeCu, indeed, two peaks at 160 °C and 230 °C due to the reduction of copper species are displayed. The main difference between CeCu and ZrCu is the consumption of H2 in the low-temperature range, and that, for the latter, corresponds only to the reduction from Cu2+ to Cu°, as we can expect for a “non-reducible” support like zirconia (0.70 mmol/g). These peaks are found also in the ZrSrCu sample, where a peak at high temperature (710 °C) due to the contribution of adsorbed strontium carbonate is also detected.

2.3. Catalytic Tests

Figure 3 summarizes the results of soot combustion studies carried out under O2/N2 and NO/O2/N2 atmosphere in terms of peak-top temperature (Tp). A representative oxidation profile is shown in Figure 4 for the CeSrCu sample.
In O2 atmosphere (Figure 3A) all of the catalyst formulation are slightly active compared to the oxidation of soot without a catalyst, displaying Tp in the range 580–595 °C with the exception of CeSr, Zr, and ZrCu, which show temperatures of oxidation higher than 600 °C.
Several studies investigated the promotional effects of copper and iron in soot combustion catalysts under O2 atmosphere [20,22,45,48]. The mechanism of reaction commonly proposed is correlated to the presence of MxOy particles and the ability of the metal to be reduced to M(n−1)+ and then reoxidized to Mn+, producing active oxygen species that can easily react with soot [45,50].
The addition of NO in the reaction mixture (Figure 3B) causes a decrease in the temperature of combustion for all catalysts (Tp lower than 560 °C). The addition of Cu and Fe results in a beneficial effect on soot combustion compared to bare supports, while the introduction of Sr has a detrimental influence on the performances. The fact that Sr acts as an efficient NOx trap [52,71,72], with the formation of stable nitrates species, is likely to cause a loss of soot combustion activity, in agreement with previous literature results [15].
With transition metals and Sr containing formulations, the oxidation is shifted at temperatures lower than 500 °C, suggesting a synergic effect of the two components. The best performances are obtained with Cu-Sr combination, with a Tp of 468 and 482 °C for Zr and Ce sample, respectively, compared to 501 °C and 496 °C for SrFeZr and SrFeCe.
As shown in Figure 4 when soot oxidation is carried out with a mixture of NO/O2, an enhancement in catalytic combustion was found, compared to the reaction performed under oxygen environment and the reason could be ascribed to two different mechanisms that could take part in a NO/O2 atmosphere: (i) from one side soot can be oxidized by active oxygen species (O*) that are generated over Cu- or Fe-containing materials, i.e., transition metals exhibit the capability to cycle between two states of oxidation contributing to soot combustion [45,48,49]; (ii) on the other hand, the NOx-assisted mechanism can improve the catalytic activity forming NO2, a more oxidant and mobile species [71].
In the case of NOx-assisted reaction it is important that the catalyst can promote the oxidation of NO to NO2 at low temperature. If NO2 is formed at temperature lower than that of soot oxidation, it can participate to reaction and offer an alternative route in the combustion of particulate. CeCu is the most active catalyst in NO oxidation as shown in Figure 5, with a Tm lower than 400 °C. All of the other samples containing Cu and Fe exhibit a Tm in the range 425–470 °C, with the exception of pure zirconia, CeSr and ZrSr samples that are only moderately active in NO oxidation (Tm is higher than 500 °C). In the presence of NO, Cu- and Fe-loaded ceria and zirconia promote soot combustion at lower temperature, showing that catalysts that are more efficient in NO oxidation are more active in soot combustion. This can be clearly seen in Figure 6, where the Tp vs. Tm temperatures are reported and a correlation between soot combustion (Tp) and NO oxidation (Tm) temperatures can be found for both CeO2- and ZrO2-based materials.
With the only exception of CeCu catalyst, Figure 6A shows a tendency of soot oxidation temperature against NO oxidation activity (i.e., the highest is the NO oxidation activity, the highest is the soot combustion temperature). The best catalysts for soot oxidation are the formulations doped with both transition and alkaline-earth metals (in particular CeSrCu and ZrSrCu), which suggests a synergic effect between Cu and Sr, independently from the support. It is well known that, in the presence of NO, alkali-earth metals are mainly involved in the storage of NOx species [53,54]. Table 2 shows NOx adsorption/desorption properties obtained from NOx-TPD experiments.
During the NOx-TPD measurements all of the materials investigated release a significant amount of NOx species, where the amount released with Ce-based catalysts is higher if compared to Zr-based formulations. The incorporation of Sr further enhances the desorption of NOx indicating that when Sr is present in the formulation nitrite/nitrate species can be efficiently stored on the sample and then released when increasing the temperature.
To summarize, the transition metal is favourably involved in the oxidation of NO to NO2, while strontium is involved in the storage of NOx species. The nitrite/nitrate species stored on strontium, when the temperature increases, start to decompose, releasing NOx that can easily react with particulate-forming CO and CO2. At this point, the presence of a transition metal is crucial to recycling NO to NO2. When Sr is added to the catalyst, the NOx involved in soot oxidation can originate from the oxidation of NO present in the reaction mixture and from the nitrates stored on the system, positively influencing the particulate combustion activity.

3. Materials and Methods

3.1. Catalyst Preparation

A series of 5 wt % Cu- and Fe-loaded CeO2 and ZrO2 doped with 10 wt % of Sr has been prepared according to the following methodology: Supports were synthesized by calcination of cerium nitrates (Treibacher Industrie AG, Althofen, Austria) and zirconium hydroxides (Mel Chemicals, Manchester, UK) at 500 °C for 3 h. After that, aqueous solutions with appropriate amounts of Cu (Copper(II) nitrate hemi (pentahydrate), Sigma Aldrich, Saint Louis, MO, USA), Fe (Iron(III) nitrate nonahydrate, Sigma Aldrich), and/or Sr (strontium nitrate, Strem chemicals, Newburyport, MA, USA) were added by incipient wetness impregnation, and dried overnight at 100 °C. Cu- and Fe-loaded materials were prepared by a single impregnation step, while for catalysts containing Sr, two successive impregnation steps have been applied. After drying the samples were finally calcined at 700 °C for 3 h.

3.2. Catalyst Characterization

Textural characteristics of all fresh samples were measured according to the BET method by nitrogen adsorption at −196 °C, using a Tristar 3000 gas adsorption analyser (Micromeritics, Norcross, GA, USA).
Structural features of the catalysts were characterized by X-ray diffraction (XRD). XRD patterns were recorded on a Philips X’Pert diffractometer (PANalytical B.V., Almelo, The Netherlands) operated at 40 kV and 40 mA with nickel-filtered Cu-Kα radiation. Diffractograms were collected using a step size of 0.02° and a counting time of 40 s per angular abscissa in the range 20–80°. The Philips X’Pert HighScore software (PANalytical B.V., Almelo, The Netherlands, 2002) was used for phase identification. The mean crystalline size was estimated from the full width at the half maximum (FWHM) of the X-ray diffraction peak using the Scherrer equation [73] with a correction for instrument line broadening.
Redox activity was measured by temperature-programmed reduction (TPR) experiments; catalysts (50 mg) were heated at a constant rate (10 °C/min) in a U-shaped quartz reactor from room temperature to 1000 °C under a 4.5% H2/N2 flow (35 mL/min). Previously, a pre-treatment was carried out by heating the samples up to 500 °C during 1 h (10 °C/min) in air flow. The hydrogen consumption was monitored using a thermal conductivity detector (TCD) (Autochem II 2920, Micromeritics, Norcross, GA, USA).
The adsorption/desorption of NOx were investigated by temperature programmed desorption experiments under N2 after NOx adsorption at 250 °C. Fifty milligrams of catalyst were exposed for 30 min at 250 °C to a 500 mL/min gas flow with 500 ppm NOx/10%O2/N2. Then, the gas mixture was replaced by pure N2 and the temperature was cooled down for 30 min. Finally, the desorption experiments were performed in N2 (500 mL/min) raising the temperature from 100 to 550 °C at 10 °C/min and NOx species were monitored by a Fourier Transform Infrared Spectroscopy (FT-IR) gas analysers (MultiGas 2030, MKS Instruments, Inc., Andover, MA USA).

3.3. Soot Oxidation Tests

Samples for catalytic measurements were prepared by mixing, in a loose contact condition, known amounts of soot (Printex-U by Degussa) and catalysts, in order to adopt a catalyst/soot weight ratio of 20:1.
Soot oxidation activity was determined from peak-top temperature (Tp) during temperature programmed oxidation (TPO) of catalyst-soot mixtures. During the TPO measurements 20 mg of mixture were heated at a constant rate (10 °C/min) in a quartz reactor, while the gas flow (10% of O2/N2 or 500 ppm NO/10% O2/N2) was kept fixed at 500 mL/min. The catalyst temperature was checked by a chromel-alumel thermocouple, located on the catalyst bed.
Outlet composition was monitored by a FT-IR gas analysers (MultiGas 2030, MKS) by recording the percentages of NO, NO2, CO, and CO2. Reproducibility of results was verified by running several TPO experiments on similar samples and the results in terms of Tp were always within 5 °C.
In order to evaluate the NO oxidation capacity of the materials, TPO experiments in a NO/O2/N2 atmosphere on the as-prepared catalyst were also performed and as a measure of activity the temperature of the maximum production of NO2 (Tm) was used.

4. Conclusions

This study focuses on the catalytic behaviour of strontium modified Cu- and Fe-based catalysts in soot oxidation. The addition of Sr positively affected the catalytic activity of the materials, suggesting a synergic effect between transition metals (Cu or Fe) and alkaline-earth metal (Sr). Among all the formulations the most active is CeSrCu; it is proposed that the role of copper is to convert NO into NO2 following a redox cycle where the copper active sites are alternately oxidized and reduced. The role of strontium seems to be correlated to its ability to store NOx species and then release NOx enhancing the NOx-assisted mechanism. Therefore, this class of materials could be considered as an interesting alternative to Pt-based catalysts for the simultaneous removal of soot and NOx in LNT applications. Further studies should be performed on their reactivity in the reduction of the stored NOx.

Acknowledgments

The authors thank financial support from MIUR (Futuro in ricerca, FIRB 2012, project SOLYST) and Regione FVG, through project LR 14/2010.

Author Contributions

E.A. and A.T. conceived and designed the experiments; V.R.-P. and E.A. performed the experiments; V.R.-P. and E.A. analyzed the data; A.T. and E.A. contributed reagents/materials/analysis tools; E.A. wrote the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Neeft, J.P.A.; Makkee, M.; Moulijn, J.A. Diesel particulate emission control. Fuel Process. Technol. 1996, 47, 1–69. [Google Scholar] [CrossRef]
  2. Summers, J.C.; VanHoutte, S.; Psaras, D. Simultaneous control of particulate and NOx emissions from diesel engines. Appl. Catal. B-Environ. 1996, 10, 139–156. [Google Scholar] [CrossRef]
  3. Qian, Z.Q.; Siegmann, K.; Keller, A.; Matter, U.; Scherrer, L.; Siegmann, H.C. Nanoparticle air pollution in major cities and its origin. Atmos. Environ. 2000, 34, 443–451. [Google Scholar]
  4. Fino, D. Diesel emission control: Catalytic filters for particulate removal. Sci. Technol. Adv. Mater. 2007, 8, 93–100. [Google Scholar] [CrossRef]
  5. Majewski, W.A.; Khair, M.K. Diesel Emissions and Their Control; SAE International: Warrendale, PA, USA, 2006. [Google Scholar]
  6. Twigg, M.V. Controlling automotive exhaust emissions: Successes and underlying science. Philos. Trans. A Math. Phys. Eng. Sci. 2005, 363, 1013–1033. [Google Scholar] [CrossRef] [PubMed]
  7. Van Setten, B.A.A.L.; Makkee, M.; Moulijn, J.A. Science and technology of catalytic diesel particulate filters. Catal. Rev. 2001, 43, 489–564. [Google Scholar] [CrossRef]
  8. Van Setten, B.A.A.L.; Schouten, J.M.; Makkee, M.; Moulijn, J.A. Realistic contact for soot with an oxidation catalyst for laboratory studies. Appl. Catal. B-Environ. 2000, 28, 253–257. [Google Scholar] [CrossRef]
  9. Saracco, G.; Badini, C.; Russo, N.; Specchia, V. Development of catalysts based on pyrovanadates for diesel soot combustion. Appl. Catal. B-Environ. 1999, 21, 233–242. [Google Scholar] [CrossRef]
  10. Ciambelli, P.; Palma, V.; Russo, P.; Vaccaro, S. Redox properties of a TiO2 supported Cu-V-K-Cl catalyst in low temperature soot oxidation. J. Mol. Catal. A-Chem. 2003, 204, 673–681. [Google Scholar] [CrossRef]
  11. Oi-Uchisawa, J.; Obuchi, A.; Enomoto, R.; Xu, J.Y.; Nanba, T.; Liu, S.T.; Kushiyama, S. Oxidation of carbon black over various Pt/MOx/SiC catalysts. Appl. Catal. B-Environ. 2001, 32, 257–268. [Google Scholar] [CrossRef]
  12. Hernandez-Gimenez, A.M.; Castello, D.L.; Bueno-Lopez, A. Diesel soot combustion catalysts: Review of active phases. Chem. Pap. 2014, 68, 1154–1168. [Google Scholar] [CrossRef]
  13. Shimizu, K.; Katagiri, M.; Satokawa, S.; Satsuma, A. Sintering-resistant and self-regenerative properties of Ag/SnO2 catalyst for soot oxidation. Appl. Catal. B-Environ. 2011, 108, 39–46. [Google Scholar] [CrossRef]
  14. Aneggi, E.; Llorca, J.; de Leitenburg, C.; Dolcetti, G.; Trovarelli, A. Soot combustion over silver-supported catalysts. Appl. Catal. B-Environ. 2009, 91, 489–498. [Google Scholar] [CrossRef]
  15. Peralta, M.A.; Zanuttini, M.S.; Querini, C.A. Activity and stability of BaKCo/CeO2 catalysts for diesel soot oxidation. Appl. Catal. B-Environ. 2011, 110, 90–98. [Google Scholar] [CrossRef]
  16. Fu, M.L.; Yue, X.H.; Ye, D.Q.; Ouyang, J.H.; Huang, B.C.; Wu, J.H.; Liang, H. Soot oxidation via cuo doped CeO2 catalysts prepared using coprecipitation and citrate acid complex-combustion synthesis. Catal. Today 2010, 153, 125–132. [Google Scholar] [CrossRef]
  17. Tikhomirov, K.; Krocher, O.; Elsener, M.; Wokaun, A. MnOx-CeO2 mixed oxides for the low-temperature oxidation of diesel soot. Appl. Catal. B-Environ. 2006, 64, 72–78. [Google Scholar] [CrossRef]
  18. Wagloehner, S.; Kureti, S. Modelling of the kinetics of the catalytic soot oxidation on Fe2O3. Appl. Catal. B-Environ. 2013, 129, 501–508. [Google Scholar] [CrossRef]
  19. Reddy, B.M.; Rao, K.N. Copper promoted ceria-zirconia based bimetallic catalysts for low temperature soot oxidation. Catal. Commun. 2009, 10, 1350–1353. [Google Scholar] [CrossRef]
  20. Gimenez-Manogil, J.; Bueno-Lopez, A.; Garcia-Garcia, A. Preparation, characterisation and testing of CuO/Ce0.8Zr0.2O2 catalysts for NO oxidation to NO2 and mild temperature diesel soot combustion. Appl. Catal. B-Environ. 2014, 152, 99–107. [Google Scholar] [CrossRef]
  21. Reichert, D.; Bockhorn, H.; Kureti, S. Study of the reaction of NOx and soot, on Fe2O3 catalyst in excess of O2. Appl. Catal. B-Environ. 2008, 80, 248–259. [Google Scholar] [CrossRef]
  22. Aneggi, E.; de Leitenburg, C.; Dolcetti, G.; Trovarelli, A. Promotional effect of rare earths and transition metals in the combustion of diesel soot over CeO2 and CeO2-ZrO2. Catal. Today 2006, 114, 40–47. [Google Scholar] [CrossRef]
  23. Mei, D.Q.; Yue, S.; Wu, H.; Zhao, X.; Yuan, Y.N. Effect of Co3O4 on the kinetics of thermal decomposition of diesel particulate matter. Emerg. Mater. Res. 2016, 5, 100–109. [Google Scholar] [CrossRef]
  24. Li, H.C.; Li, K.Z.; Wang, H.; Zhu, X.; Wei, Y.G.; Yan, D.X.; Cheng, X.M.; Zhai, K. Soot combustion over Ce1−xFexO2-δ and CeO2/Fe2O3 catalysts: Roles of solid solution and interfacial interactions in the mixed oxides. Appl. Surf. Sci. 2016, 390, 513–525. [Google Scholar] [CrossRef]
  25. Liu, S.; Wu, X.D.; Liu, W.; Chen, W.M.; Ran, R.; Li, M.; Weng, D. Soot oxidation over CeO2 and Ag/CeO2: Factors determining the catalyst activity and stability during reaction. J. Catal. 2016, 337, 188–198. [Google Scholar] [CrossRef]
  26. Rico-Perez, V.; Aneggi, E.; Bueno-Lopez, A.; Trovarelli, A. Synergic effect of Cu/Ce0.5Pr0.5O2−δ and Ce0.5Pr0.5O2−δ in soot combustion. Appl. Catal. B-Environ. 2016, 197, 95–104. [Google Scholar] [CrossRef]
  27. Wang, J.G.; Cheng, L.; An, W.; Xu, J.L.; Men, Y. Boosting soot combustion efficiencies over CuO-CeO2 catalysts with a 3DOM structure. Catal. Sci. Technol. 2016, 6, 7342–7350. [Google Scholar] [CrossRef]
  28. Aneggi, E.; Leitenburg, C.D.; Trovarelli, A. Ceria-Based Formulations for Catalysts for Diesel Soot Combustion. In Catalysis by Ceria and Related Materials, 2nd ed.; Trovarelli, A., Fornasiero, P., Eds.; Imperial College Press: London, UK, 2013; pp. 565–621. [Google Scholar]
  29. Bueno-Lopez, A. Diesel soot combustion ceria catalysts. Appl. Catal. B-Environ. 2014, 146, 1–11. [Google Scholar] [CrossRef]
  30. Oliveira, C.F.; Garcia, F.A.C.; Araujo, D.R.; Macedo, J.L.; Dias, S.C.L.; Dias, J.A. Effects of preparation and structure of cerium-zirconium mixed oxides on diesel soot catalytic combustion. Appl. Catal. A-Gen. 2012, 413, 292–300. [Google Scholar] [CrossRef]
  31. Kumar, P.A.; Tanwar, M.D.; Russo, N.; Pirone, R.; Fino, D. Synthesis and catalytic properties of CeO2 and Co/CeO2 nanofibres for diesel soot combustion. Catal. Today 2012, 184, 279–287. [Google Scholar] [CrossRef]
  32. Aneggi, E.; de Leitenburg, C.; Trovarelli, A. On the role of lattice/surface oxygen in ceria-zirconia catalysts for diesel soot combustion. Catal. Today 2012, 181, 108–115. [Google Scholar] [CrossRef]
  33. Aneggi, E.; Wiater, D.; de Leitenburg, C.; Llorca, J.; Trovarelli, A. Shape-dependent activity of ceria in soot combustion. ACS Catal. 2014, 4, 172–181. [Google Scholar] [CrossRef]
  34. Aneggi, E.; Divins, N.J.; de Leitenburg, C.; Llorca, J.; Trovarelli, A. The formation of nanodomains of Ce6O11 in ceria catalyzed soot combustion. J. Catal. 2014, 312, 191–194. [Google Scholar] [CrossRef]
  35. Di Sarli, V.; Landi, G.; Lisi, L.; Saliva, A.; Di Benedetto, A. Catalytic diesel particulate filters with highly dispersed ceria: Effect of the soot-catalyst contact on the regeneration performance. Appl. Catal. B-Environ. 2016, 197, 116–124. [Google Scholar] [CrossRef]
  36. Andana, T.; Piumetti, M.; Bensaid, S.; Russo, N.; Fino, D.; Pirone, R. Nanostructured ceria-praseodymia catalysts for diesel soot combustion. Appl. Catal. B-Environ. 2016, 197, 125–137. [Google Scholar] [CrossRef]
  37. Aneggi, E.; de Leitenburg, C.; Dolcetti, G.; Trovarelli, A. Diesel soot combustion activity of ceria promoted with alkali metals. Catal. Today 2008, 136, 3–10. [Google Scholar] [CrossRef]
  38. Shan, W.J.; Yang, L.H.; Ma, N.; Yang, J.L. Catalytic activity and stability of K/CeO2 catalysts for diesel soot oxidation. Chin. J. Catal. 2012, 33, 970–976. [Google Scholar] [CrossRef]
  39. Peralta, M.A.; Zanuttini, M.S.; Ulla, M.A.; Querini, C.A. Diesel soot and NOx abatement on K/La2O3 catalyst: Influence of k precursor on soot combustion. Appl. Catal. A-Gen. 2011, 399, 161–171. [Google Scholar] [CrossRef]
  40. Galvez, M.; Ascaso, S.; Moliner, R.; Lazaro, M. Influence of the alkali promoter on the activity and stability of transition metal (Cu, Co, Fe) based structured catalysts for the simultaneous removal of soot and NOx. Top. Catal. 2013, 56, 493–498. [Google Scholar] [CrossRef]
  41. Matarrese, R.; Castoldi, L.; Lietti, L.; Forzatti, P. Soot combustion: Reactivity of alkaline and alkaline earth metal oxides in full contact with soot. Catal. Today 2008, 136, 11–17. [Google Scholar] [CrossRef]
  42. Russo, N.; Furfori, S.; Fino, D.; Saracco, G.; Specchia, V. Lanthanum cobaltite catalysts for diesel soot combustion. Appl. Catal. B-Environ. 2008, 83, 85–95. [Google Scholar] [CrossRef]
  43. Fan, G.L.; Zhao, L.; Gong, C.R.; Ma, J.; Xue, G. Effect of supports on soot oxidation of copper catalysts: BaTiO3 versus Fe2O3@BaTiO3 core/shell microsphere. Nano 2016, 11. [Google Scholar] [CrossRef]
  44. Galvez, M.E.; Ascaso, S.; Stelmachowski, P.; Legutko, P.; Kotarba, A.; Moliner, R.; Lazaro, M.J. Influence of the surface potassium species in Fe-K/Al2O3 catalysts on the soot oxidation activity in the presence of nox. Appl. Catal. B-Environ. 2014, 152, 88–98. [Google Scholar] [CrossRef] [Green Version]
  45. Laversin, H.; Courcot, D.; Zhilinskaya, E.A.; Cousin, R.; Aboukais, A. Study of active species of Cu-K/ZrO2 catalysts involved in the oxidation of soot. J. Catal. 2006, 241, 456–464. [Google Scholar] [CrossRef]
  46. Lick, I.D.; Carrascull, A.L.; Ponzi, M.I.; Ponzi, E.N. Zirconia-supported Cu-KNO3 catalyst: Characterization and catalytic behavior in the catalytic combustion of soot with a NO/O2 mixture. Ind. Eng. Chem. Res. 2008, 47, 3834–3839. [Google Scholar] [CrossRef]
  47. Wu, X.D.; Liang, Q.; Weng, D.; Lu, Z.X. The catalytic activity of CuO-CeO2 mixed oxides for diesel soot oxidation with a NO/O2 mixture. Catal. Commun. 2007, 8, 2110–2114. [Google Scholar] [CrossRef]
  48. Wagloehner, S.; Kureti, S. Study on the mechanism of the oxidation of soot on Fe2O3 catalyst. Appl. Catal. B-Environ. 2012, 125, 158–165. [Google Scholar] [CrossRef]
  49. Lopez-Suarez, F.E.; Bueno-Lopez, A.; Illan-Gomez, M.J.; Adamski, A.; Ura, B.; Trawczynski, J. Copper catalysts for soot oxidation: Alumina versus perovskite supports. Environ. Sci. Technol. 2008, 42, 7670–7675. [Google Scholar] [CrossRef] [PubMed]
  50. Soler, L.; Casanovas, A.; Escudero, C.; Perez-Dieste, V.; Aneggi, E.; Trovarelli, A.; Llorca, J. Ambient pressure photoemission spectroscopy reveals the mechanism of carbon soot oxidation in ceria-based catalysts. Chemcatchem 2016, 8, 2748–2751. [Google Scholar] [CrossRef]
  51. Milt, V.G.; Querini, C.A.; Miro, E.E.; Ulla, M.A. Abatement of diesel exhaust pollutants: NOx adsorption on Co,Ba,K/ CeO2 catalysts. J. Catal. 2003, 220, 424–432. [Google Scholar] [CrossRef]
  52. Pisarello, M.L.; Milt, V.; Peralta, M.A.; Querini, C.A.; Miro, E.E. Simultaneous removal of soot and nitrogen oxides from diesel engine exhausts. Catal. Today 2002, 75, 465–470. [Google Scholar] [CrossRef]
  53. Matarrese, R.; Castoldi, L.; Lietti, L.; Forzatti, P. Simultaneous removal of NOx and soot over Pt-Ba/Al2O3 and Pt-K/Al2O3 DPRN catalysts. Top. Catal. 2009, 52, 2041–2046. [Google Scholar] [CrossRef]
  54. Lin, F.; Wu, X.D.; Weng, D. Effect of barium loading on CuOx-CeO2 catalysts: NOx storage capacity, NO oxidation ability and soot oxidation activity. Catal. Today 2011, 175, 124–132. [Google Scholar] [CrossRef]
  55. Weng, D.A.; Li, J.; Wu, X.D.; Si, Z.C. NOx-assisted soot oxidation over K/CuCe catalyst. J. Rare Earth 2010, 28, 542–546. [Google Scholar] [CrossRef]
  56. Fierro, J.L.G. Metal Oxides: Chemistry and Applications; CRC Press: Boca Raton, FL, USA, 2005. [Google Scholar]
  57. Castoldi, L.; Aneggi, E.; Matarrese, R.; Bonzi, R.; Llorca, J.; Trovarelli, A.; Lietti, L. Silver-based catalytic materials for the simultaneous removal of soot and NOx. Catal. Today 2015, 258, 11. [Google Scholar] [CrossRef] [Green Version]
  58. Shen, Q.; Lu, G.Z.; Du, C.H.; Guo, Y.; Wang, Y.Q.; Guo, Y.L.; Gong, X.Q. Role and reduction of NOx in the catalytic combustion of soot over iron-ceria mixed oxide catalyst. Chem. Eng. J. 2013, 218, 164–172. [Google Scholar] [CrossRef]
  59. Martinez-Arias, A.; Gamarra, D.; Fernandez-Garcia, M.; Hornes, A.; Bera, P.; Koppany, Z.; Schay, Z. Redox-catalytic correlations in oxidised copper-ceria CO-PROX catalysts. Catal. Today 2009, 143, 211–217. [Google Scholar] [CrossRef]
  60. Li, G.S.; Smith, R.L.; Inomata, H. Synthesis of nanoscale Ce1−xFexO2 solid solutions via a low-temperature approach. J. Am. Chem. Soc. 2001, 123, 11091–11092. [Google Scholar] [CrossRef] [PubMed]
  61. Wang, H.; Qu, Z.P.; Xie, H.B.; Maeda, N.; Miao, L.; Wang, Z. Insight into the mesoporous FexCe1−xO2−δ catalysts for selective catalytic reduction of NO with NH3: Regulable structure and activity. J. Catal. 2016, 338, 56–67. [Google Scholar] [CrossRef]
  62. Giordano, F.; Trovarelli, A.; de Leitenburg, C.; Giona, M. A model for the temperature-programmed reduction of low and high surface area ceria. J. Catal. 2000, 193, 273–282. [Google Scholar] [CrossRef]
  63. Magnacca, G.; Cerrato, G.; Morterra, C.; Signoretto, M.; Somma, F.; Pinna, F. Structural and surface characterization of pure and sulfated iron oxides. Chem. Mater. 2003, 15, 675–687. [Google Scholar] [CrossRef]
  64. Soni, K.C.; Shekar, S.C.; Singh, B.; Gopi, T. Catalytic activity of Fe/ZrO2 nanoparticles for dimethyl sulfide oxidation. J. Colloid Interface Sci. 2015, 446, 226–236. [Google Scholar] [CrossRef] [PubMed]
  65. Pojanavaraphan, C.; Luengnaruemitchai, A.; Gulari, E. Effect of support composition and metal loading on Au catalyst activity in steam reforming of methanol. Int. J. Hydrogen Energ. 2012, 37, 14072–14084. [Google Scholar] [CrossRef]
  66. Sang, X.L.; Li, K.Z.; Wang, H.; Wei, Y.G. Selective oxidation of methane and carbon deposition over Fe2O3/Ce1-xZrxO2 oxides. Rare Met. 2014, 33, 230–238. [Google Scholar] [CrossRef]
  67. Tang, X.L.; Zhang, B.C.; Li, Y.; Xu, Y.D.; Xin, Q.; Shen, W.J. CuO/CeO2 catalysts: Redox features and catalytic behaviors. Appl. Catal. A-Gen. 2005, 288, 116–125. [Google Scholar] [CrossRef]
  68. Arango-Diaz, A.; Cecilia, J.A.; Moretti, E.; Talon, A.; Nunez, P.; Marrero-Jerez, J.; Jimenez-Jimenez, J.; Jimenez-Lopez, A.; Rodriguez-Castellon, E. Comparative study of CuO supported on CeO2, Ce0.8Zr0.2O2 and Ce0.8Al0.2O2 based catalysts in the CO-PROX reaction. Int. J. Hydrogen Energ. 2014, 39, 4102–4108. [Google Scholar] [CrossRef] [Green Version]
  69. Zabilskiy, M.; Djinović, P.; Tchernychova, E.; Tkachenko, O.P.; Kustov, L.M.; Pintar, A. Nanoshaped Cuo/CeO2 materials: Effect of the exposed ceria surfaces on catalytic activity in N2O decomposition reaction. ACS Catal. 2015, 5, 5357–5365. [Google Scholar] [CrossRef]
  70. Muroyama, H.; Hano, S.; Matsui, T.; Eguchi, K. Catalytic soot combustion over CeO2-based oxides. Catal. Today 2010, 153, 133–135. [Google Scholar] [CrossRef]
  71. Kustov, A.L.; Makkee, M. Application of NOx storage/release materials based on alkali-earth oxides supported on Al2O3 for high-temperature diesel soot oxidation. Appl. Catal. B-Environ. 2009, 88, 263–271. [Google Scholar] [CrossRef]
  72. Kustov, A.L.; Ricciardi, F.; Makkee, M. NOx storage and high temperature soot oxidation on Pt-Sr/ZrO2 catalyst. Top. Catal. 2009, 52, 2058–2062. [Google Scholar] [CrossRef]
  73. Jenkins, R.; Snyder, R. Introduction to X-ray Powder Diffractometry; Wiley: New York, NY, USA, 1996. [Google Scholar]
Figure 1. XRD diffraction profiles for (A) Ce- and (B) Zr-based catalysts.
Figure 1. XRD diffraction profiles for (A) Ce- and (B) Zr-based catalysts.
Catalysts 07 00028 g001
Figure 2. H2-TPR profiles for (A) Ce- and (B) Zr-based catalysts.
Figure 2. H2-TPR profiles for (A) Ce- and (B) Zr-based catalysts.
Catalysts 07 00028 g002
Figure 3. Catalytic activity as Tp (°C) for soot combustion under loose contact conditions in (A) O2/N2 and (B) NO/O2/N2 atmosphere for CeO2 (blue)- and ZrO2 (red)-based catalysts. The broken line indicates the soot oxidation temperature for uncatalyzed reactions.
Figure 3. Catalytic activity as Tp (°C) for soot combustion under loose contact conditions in (A) O2/N2 and (B) NO/O2/N2 atmosphere for CeO2 (blue)- and ZrO2 (red)-based catalysts. The broken line indicates the soot oxidation temperature for uncatalyzed reactions.
Catalysts 07 00028 g003
Figure 4. Soot oxidation carried out in a conventional flow reactor by monitoring evolution of CO2 in O2/N2 and NO/O2/N2 atmosphere for a CeSrCu sample.
Figure 4. Soot oxidation carried out in a conventional flow reactor by monitoring evolution of CO2 in O2/N2 and NO/O2/N2 atmosphere for a CeSrCu sample.
Catalysts 07 00028 g004
Figure 5. NO2 production profile for (A) Ce- and (B) Zr-based catalysts; (C) NO oxidation activity: for CeO2 support (blue), ZrO2 support (red).
Figure 5. NO2 production profile for (A) Ce- and (B) Zr-based catalysts; (C) NO oxidation activity: for CeO2 support (blue), ZrO2 support (red).
Catalysts 07 00028 g005
Figure 6. Correlation between Tp and Tm for (A) Ce- and (B) Zr-based catalysts.
Figure 6. Correlation between Tp and Tm for (A) Ce- and (B) Zr-based catalysts.
Catalysts 07 00028 g006
Table 1. Composition and textural characterization of investigated samples.
Table 1. Composition and textural characterization of investigated samples.
SampleCompositionSurface Area (m2/g)Particle Size (nm)
CeCeO23519
CeCuCu(5%)/CeO23513
CeFeFe(5%)/CeO23015
CeSrSr(10%)/CeO23214
CeSrCuCu(5%)-Sr(10%)/CeO22419
CeSrFeFe(5%)-Sr(10%)/CeO21713
ZrZrO22722
ZrCuCu(5%)/ZrO23218
ZrFeFe(5%)/ZrO22618
ZrSrSr(10%)/ZrO22917
ZrSrCuCu(5%)-Sr(10%)/ZrO22118
ZrSrFeFe(5%)-Sr(10%)/ZrO22217
Table 2. Desorption of NOx based on NOx-TPD experiments.
Table 2. Desorption of NOx based on NOx-TPD experiments.
Samplea Desorbed NOx (μmol/gcat)SampleDesorbed NOx (μmol/gcat)
Ce30Zr12
CeCu55ZrCu41
CeFe46ZrFe46
CeSr64ZrSr35
CeSrCu65ZrSrCu54
CeSrFe65ZrSrFe63
a Amount of NO and NO2 desorbed in the range of temperature 100–550 °C.

Share and Cite

MDPI and ACS Style

Rico-Pérez, V.; Aneggi, E.; Trovarelli, A. The Effect of Sr Addition in Cu- and Fe-Modified CeO2 and ZrO2 Soot Combustion Catalysts. Catalysts 2017, 7, 28. https://doi.org/10.3390/catal7010028

AMA Style

Rico-Pérez V, Aneggi E, Trovarelli A. The Effect of Sr Addition in Cu- and Fe-Modified CeO2 and ZrO2 Soot Combustion Catalysts. Catalysts. 2017; 7(1):28. https://doi.org/10.3390/catal7010028

Chicago/Turabian Style

Rico-Pérez, Verónica, Eleonora Aneggi, and Alessandro Trovarelli. 2017. "The Effect of Sr Addition in Cu- and Fe-Modified CeO2 and ZrO2 Soot Combustion Catalysts" Catalysts 7, no. 1: 28. https://doi.org/10.3390/catal7010028

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