Next Article in Journal
Oxygen Adsorption and Activation on Cobalt Center in Modified Keggin Anion-DFT Calculations
Next Article in Special Issue
Evoked Methane Photocatalytic Conversion to C2 Oxygenates over Ceria with Oxygen Vacancy
Previous Article in Journal
A Review on Quantum Dots Modified g-C3N4-Based Photocatalysts with Improved Photocatalytic Activity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

In Situ Investigations on the Facile Synthesis and Catalytic Performance of CeO2-Pt/Al2O3 Catalyst

1
Sino-Precious Metals Holding Co., Ltd., 988 Keji Road, Hi-tech Development Zone, Kunming 650106, China
2
State Key Laboratory of Advanced Technologies for Comprehensive Utilization of Platinum Metals, Kunming Institute of Precious Metals, Kunming 650106, China
3
State-Local Joint Engineering Laboratory of Precious Metal Catalytic Technology and Application, Kunming Sino-Platinum Metals Catalysts Co. Ltd., Kunming 650106, China
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(2), 143; https://doi.org/10.3390/catal10020143
Submission received: 24 December 2019 / Revised: 17 January 2020 / Accepted: 18 January 2020 / Published: 21 January 2020

Abstract

:
Ceria-modified Pt/Al2O3 catalyst has been commonly prepared by the impregnation of platinum on ceria-modified alumina and widely applied in the chemical industry and automotive industry. The in situ diffuse reflectance infrared Fourier transformed spectroscopy (DRIFTS), and thermogravimetric (TG) analysis techniques were employed to investigate the typical mechanisms of the bis(ethanolammonium)hexahydroxyplatinate(IV) and cerium nitrate decomposition catalyzed by Ptδ+ species for the facile synthesis of CeO2-Pt/Al2O3 catalyst. It was found that Pt4+-catalyzed decomposition of cerium nitrate leads to the higher dispersity of ceria and forming more active oxygen species, on the basis of X-ray diffraction (XRD) and H2 temperature-programmed reduction (H2-TPR) results. The in situ activity measurements were also performed to investigate the reaction mechanisms and the specific activities for the catalytic CO, NO, C3H6 and C3H8 co-oxidation. The results indicate that undesirable N2O by-product is formed by the selective catalytic reduction (SCR) of NO by C3H6 below 350 °C. The cerium addition effectively improves the activity of catalytic oxidation, but exhibits an increased N2O yield, due to the increased reducibility.

Graphical Abstract

1. Introduction

Platinum-based catalysts have been widely used in the chemical industry for the catalytic dehydrogenation of light alkanes, such as ethane and propane [1,2], as well as the automotive industry to catalytically oxidize harmful pollutants, including CO, NO, and hydrocarbons [3,4,5,6]. Ceria has been used in automotive catalysts, due to its oxygen storage capacity and the ability to stabilize active metals [7,8,9]. The recognized ability of ceria to undergo rapid reduction/oxidation cycles (4CeO2 ⇌ 2Ce2O3 + O2) plays a significant role in providing oxygen mobility and oxygen storage/release [8]. In our previous investigations, it was reported that the synergy between ceria additive and active Pt species has a positive impact on the stability of coordinately unsaturated Pt2+ and Pt4+ species, thereby drives C–C and C–H bonds activation and promotes low-temperature oxidation of propane [10]. Ceria can also stabilize atomically dispersed ionic platinum (Pt2+ species) to accelerate CO oxidation at low temperature [11]. The shape-dependent ability to anchor single Pt2+ species increases in the order: Ceria cubes < ceria nanorods < polyhedral ceria [9]. Bruix [12] et al. suggested that the unsaturated Pt2+ species is effectively stabilized by bonding to O2– sites in square pockets of CeO2 surfaces.
The ceria addition into Pt-based diesel oxidation catalyst (DOC) showed the positive impact on the improvement of platinum dispersion, and generation of more active oxygen species and oxygen vacancies over the catalyst, which strengthens the interactions between NO molecules and the catalyst surfaces to promote NO oxidation into NO2 and form nitrate species [13]. The introduction of cerium into Pt/Al2O3 catalyst leads to enhancing the dispersity of active Pt nanoparticles and reducibility of catalyst, which was reported by Ma et al. [14]. They also found that the Ce-containing Pt/Al2O3 catalyst exhibits more active for propane dehydrogenation and higher coking-resistant ability than Ce-free Pt/Al2O3 catalyst. As a doping element, cerium addition to alumina increases the thermal stability of alumina and improves the ability to remove carbon species during partial oxidation of methane [15]. Improved effectiveness of the carbon cleaning promotes the reactivity of partial oxidation of methane. Promotion of Pt/Al2O3 by cerium for CO oxidation was also found by Son and his co-workers [16,17], due to the enhanced oxygen supply to the Pt sites.
Some methods for cerium addition into Al2O3 supported catalyst have been reported, for instance, incipient wetness impregnation [10], controlled surface reaction (CSR) technique [18], and atomic layer deposition (ALD) technique [19]. The latter two methods can effectively improve the dispersity of ceria, but the use of expensive organometallics, such as Ce(acac)3 (cerium acetylacetonate) and Ce(TMHD)4 (tetrakis(2,2,6,6-tetramethyl-3,5-heptanedionato)cerium) is required to prepare the CeO2-x/Pt/Al2O3 catalyst. By contrast, the incipient wetness impregnation method was considered as the most practical method, resulting from requiring the use of low-cost inorganic precursors, such as Ce(NO3)3 (cerium nitrate) and (NH4)2[Ce(NO3)6] (diammonium hexanitrate cerate) [10,20]. The cerium addition by incipient wetness impregnation using Ce(NO3)3 has been successfully employed to prepare the highly-efficient automotive catalysts [13]. However, the structure evolution of precursors is still unknown during the process for preparing the cerium-modified Pt/Al2O3 catalyst.
It is worth mentioning that the choice of platinum precursors is also very important for the preparation of Pt catalysts by the incipient wetness impregnation method [21]. The typical Pt-precursors, such as H2PtCl6 (hexachloroplatinic(IV) acid), Pt(NO3)2 (platinum(II) nitrate) and Pt(NH3)4(NO3)2 (tetraammineplatinum(II) nitrate), were used to prepare those Pt-supported catalysts [22,23]. However, the use of H2PtCl6 as a precursor will lead to the introduction of chlorine, although the tetravalent state of platinum readily interacts with non-stoichiometric oxides. By contrast, using a new [HOCH2CH2NH3]2Pt(OH)6 (bis(ethanolammonium)hexahydroxyplatinate(IV)) as the Pt-precursor can refrain from chlorine incorporation and its hydroxyls can bond to the surface of oxide support with hydrogen bonds to strengthen the interactions between platinum precursors and support. In the work, therefore, some in situ characterization techniques were used to investigate the facile synthesis of CeO2-Pt/Al2O3 catalyst by impregnation method using Ce(NO3)3 and [HOCH2CH2NH3]2Pt(OH)6 as the precursors and test the catalytic performance for CO, NO, C3H6 and C3H8 co-oxidation.

2. Results and Discussion

2.1. Thermogravimetric Results

The thermogravimetric (TG) analysis of the different samples was carried out, and the results were displayed in Figure 1. The pure Al2O3 sample shows a slight, but consecutive weight loss up to 550 °C, which can be ascribed to the dehydration of sample and the decomposition of chemically adsorbed hydroxyl species [24]. By contrast, the introduction of Pt-precursor induces the additional weight loss that is reasonably related to the decomposition of chemical fragments from [HOCH2CH2NH3]2Pt(OH)6. As for the Ce-containing samples, the weight loss is largely higher compared to those of Ce-free samples, which can be assigned to the transformation and thermal decomposition of nitrate precursors.

2.2. Decomposition of Pt-Precursor

The TG curve for the decomposition of Pt-precursor on Al2O3 powder was achieved through subtracting TG data of pure Al2O3 (Figure 1) from the data of the pristine Pt/alumina powder without calcination. The obtained TG curve and the corresponding derivative thermogravimetric (DTG) curve were displayed in Figure 2a. The TG curve can be divided into three weight loss steps according to the DTG curve. Step 1 with weight loss of 0.41 mg per 100 mg sample, where a temperature ranges from 50 °C to 135 °C, represents the weight loss by the removal of physically adsorbed water [25]. Step 2 removes H2O molecules from Pt(OH)62– species [26]. In the TG experiment, the theoretical weight loss for the complete decomposition of Pt(OH)62– species is 0.27 mg. However, the actual weight loss in this step reaches to 0.51 mg, indicating that partial HOCH2CH2NH3+ species starts to be transformed in the step. Increasing the temperature to 482 °C, the total weight loss reaches 1.16 mg. However, no significant change in weight loss was observed with the further temperature increments. It is assumed that Pt-precursor was completely decomposed into PtO2 and gas; the total weight loss would be 1.38 mg. The gapped results between actual and theoretical weight loss can be contributed to the formation of coke during decomposition of Pt-precursor, as well as some errors during the data processing.
Further evidence for the transformation of Pt complexes during the programmed temperature procedure in N2 steam was provided by diffuse reflectance infrared Fourier transformed spectroscopy (DRIFTS), as displayed in Figure 2b. The band centered at 1454 cm−1 can be assigned to deformation vibration of the methylene group δ(CH2) from HOCH2CH2NH3+ species, and the corresponding symmetric and asymmetric vibrations were respectively observed at 2888 cm−1, vs(CH2), and 2950 cm−1, vas(CH2) [27]. The shoulder at 1533 cm−1 should be related with O–H bending mode in water [28]. The local peak centered at 1587 cm−1 corresponds to the bending mode of the Pt–NH3 species [29]. Increasing temperature from 100 °C to 200 °C only led to a significant decrease in the intensity of the bending vibration hydroxyl group δ(OH) in water, which illustrates water loss in the catalyst sample. Carbonyl (bands at 2070 ~ 2040 cm−1, linear bonded Pt–CO species) and isocyanate surface complexes (bands at 2150 ~ 2100 cm−1, Pt–NCO species) [30,31] first clearly appeared at 250 °C. With further increments in temperature, the Pt–NCO vibrational bands showed an increase in intensity and a shift towards higher wavenumbers, reaching 2150 cm−1 at 500 °C, while the band at 2048 cm−1 assigned to Pt–CO species shifted toward 2040 cm−1 at 450 °C and disappeared at 500 °C. The blue-shift of vs(NCO) reflects increasing NCO coverage, due to the further transformation of HOCH2CH2NH3+ species. In addition, three were also the formation of some surface oxy-carbon species, including enolates (1392 cm−1, vs(CH2=CH–O)) [32], and formates (1377 cm−1, vs(COO)) [33] in the temperature regions of 250 ~ 400 °C. According to TG and in situ DRIFTS results, the thermal decomposition process of Pt-precursor can be summarized as the following reaction (I):
[HOCH2CH2NH3]2Pt(OH)6 → Pt–CO + Pt–NCO + oxy-carbon → Pt–NCO + Pt sites.

2.3. Decomposition of Ce-Precursor

In Figure 3a,c, the weight loss at the temperature range of 50 ~ 138 °C is assigned to the removal of physically adsorbed water, while the other weight loss at higher temperature corresponds to removing chemically bound water and NO3 species [24]. As showed in Figure 3a, a significant weight loss of 10.54 mg was observed in the temperature regions of 138 ~ 296 °C, primarily, due to the transformation and decomposition of NO3 species. The TG curve showed in Figure 3c exhibited a total weight loss of 11.06 mg up to 248 °C and a maximum weight loss rate of 0.243 mg/°C at 195 °C. The decomposition temperature of Ce-precursor on Pt/alumina was lower, and the maximum weight loss rate was higher compared to that of Ce-precursor on Al2O3, which indicates the occurrence of Pt-catalyzed decomposition of cerium nitrate. The total weight losses in Figure 3a,c were 14.03 mg and 15.11 mg per 100 mg sample, respectively, showing the weight gap of 1.08 mg that is closely related to the decomposition of Pt-precursor.
From Figure 3b, it can be seen that Ce-precursor on Al2O3 started to transform bridging nitrates (1650 ~ 1600 cm−1, (Ce–O)2=NO) and monodentate nitrates (1530 ~ 1480 cm−1, Ce–ONO2) into chelating bidentate nitrates (bands at 1585 ~ 1530 cm−1, Ce–O2NO) [34,35,36] above 150 °C, and the vibrational bands of nitrates disappeared with the temperature increment up to 450 °C. This indicates that the following reactions for the thermal decomposition of cerium(III) nitrate are accomplished, as showed in the reaction (II) reported by Wendlandt et al. [37]. Compared with the DRIFTS spectra for the thermal decomposition process of Ce-precursor on Al2O3, the ones of Ce-precursor on Pt/alumina (Figure 3d) exhibited a weaker intensity in stretching vibrations of bridging nitrates at lower wavenumber (centered at 1616 cm−1) and bands in the 2200 ~ 2100 cm−1 region assigned to N2O, N2, NO2+, and NO+ [36]. In addition, the DRIFTS spectrum at 400 °C in Figure 3d exhibits a significantly weaker intensity in Ce–O2NO mode at 1552 cm−1 than the decomposition of Ce-precursor on Al2O3. The results can be ascribed to Pt-catalyzed decomposition of nitrates that promotes the removal of NO3 species and the formation of CeO2–x. These promotional effects would be conducive to the enhancement of Pt–CeO2–x interactions, due to the in situ reduction of Ptδ+ species by Ce3+ ions in an inert atmosphere,
Ce(NO3)3·xH2O(s) → Ce(NO3)3(s) → CeONO3(s) → CeO2–x(s).

2.4. Catalytic Performance

The activities of catalytic CO, NO, C3H6 and C3H8 co-oxidation over the Pt-supported catalysts were performed on a continuous flow fixed-bed micro-reactor. The results of activity measurements were showed in Figure 4, and the corresponding characteristic temperatures were summarized in Table 1. The activity of different reactions over the Pt/Al2O3 catalyst was easily distinguished, decreasing in the order: CO oxidation > C3H6 oxidation > NO oxidation > C3H8 oxidation according to the light-off curves (showed in Figure 4a). The catalytic oxidation of hydrocarbons required the higher reaction temperature than CO oxidation, in particular for C3H8 oxidation, owing to the higher reactivity of CO oxidation on Pt active sites compared to that of hydrocarbons [3]. The activity of C3H6 oxidation is higher than NO oxidation, due to the NO oxidation inhibition by C3H6 over the Pt/Al2O3 catalyst [38].
In the case of NO conversion, no NH3 product was detected, but significant N2O (below 350 °C) and NO2 (above 250 °C) products were formed. NO molecule started to be consumed when CO oxidation reached near to complete conversion, showing the maximum NO conversion rate of 83% at 289 °C, maximum N2O yield of 32% at 260 °C and maximum NO2 yield of 38% at 340 °C. The results indicate that NO was transferred into undesirable N2O via the selective catalytic reduction (SCR) of NO by C3H6 at low temperature and oxidized into desirable NO2 in the higher temperature regions. The related reactions include:
2NO + C3H6 + 4O2 → N2O + 3CO2 + 3H2O,
2NO + O2 → 2NO2.
By comparison with Pt/Al2O3, the NO conversion curve almost entirely overlapped with the one of C3H6 with a higher N2O yield over the CeO2-Pt/Al2O3 catalyst up to 210 °C, illustrating that the low-temperature C3H6-SCR activity was improved, due to the increased reducibility by the addition of ceria [39]. In addition, the CeO2-Pt/Al2O3 catalyst exhibited a lower temperature to achieve the maximum NO conversion rate, N2O and NO2 yield. From Table 1, it can also be observed that the ceria addition led to a significant T50 decrease of 32.9 °C, 52.7 °C, 34.5 °C, 48.5 °C, respectively, for the CO, NO, C3H6 and C3H8 conversion. The T90 gaps between Pt/Al2O3 and CeO2-Pt/Al2O3 catalyst for CO and C3H8 conversions are higher than the corresponding T50 gaps, reaching 41.2 °C and 92.0 °C, respectively. These results indicate that the ceria addition significantly promotes low-temperature catalytic oxidation of CO, NO, C3H6 and C3H8, which agree well with the literature [10,13,16,20]. In addition, the CeO2-Pt/Al2O3 catalyst showed the lower characteristic temperatures for CO, NO and C3H6 conversion than that of the reference catalyst, although there was no obvious superiority for catalytic oxidation of C3H8. Particularly, the activity for catalytic NO oxidation into NO2 was significantly improved by using [HOCH2CH2NH3]2Pt(OH)6. This indicates that [HOCH2CH2NH3]2Pt(OH)6 is a better Pt-precursor compared to Pt(NH3)4(NO3)2.

2.5. XRD and H2 Temperature-Programmed Reduction (H2-TPR) Analysis

Figure 5a showed the XRD patterns of the different catalyst samples. The diffraction peaks associated with ceria in the CeO2/Al2O3 and CeO2-Pt/Al2O3 catalysts can be easily observed and exhibited well-crystallized fluorite-type phase [40]. The characteristic peaks appeared at 2θ = 28.54°, 33.16°, 47.58° and 56.52° are respectively assigned to the (111), (200), (220) and (311) surfaces of ceria nanocrystals. The domain size was calculated using the Scherrer Equation based on XRD line-width broadening data of the (111) surface. The average sizes of CeO2 nanoparticles in the CeO2/Al2O3 and CeO2-Pt/Al2O3 catalysts were 7.0 nm and 4.1 nm, respectively, indicating that the catalytic decomposition of cerium nitrate by Ptδ+ species show a positive impact on promoting the dispersity of ceria. Co-existence of Pt4+ and Pt2+ species in Pt/alumina powder using [HOCH2CH2NH3]2Pt(OH)6 as the Pt-precursor has been confirmed in our previous investigations [10]. The average size of CeO2 nanoparticles in the reference catalyst was bigger by comparison with the CeO2-Pt/Al2O3 catalyst, reaching 5.4 nm. The results illustrate that the presence of Pt4+ species has a better promotional effect on improving the dispersity of ceria than Pt2+ species. Besides, the ceria addition resulted in a significant decrease in the intensity of the peaks assigned to alumina supports, which can be ascribed to the Ce-promoted stability of alumina structures [16,19].
The reducibility of the catalysts was also investigated by the H2-TPR measurement, and the results were shown in Figure 5b. In the case of the Pt/Al2O3 catalyst, three H2 consumption peaks appeared at 111 °C, 154 °C and 214 °C, respectively. The slight peaks centered at 111 °C and 154 °C are closely related to the reduction of adsorbed oxygen on Pt nanocrystals and O2– species bonded to surface platinum atoms, whereas, the peak centered at 214 °C should be assigned to the reduction of oxygen species that are adjacent to active Pt sites. As for the CeO2/Al2O3 catalyst, the reduction peaks at 100 °C and 139 °C correspond to adsorbed oxygen species on ceria nanocrystals. The peaks at 252 °C and higher temperatures are respectively associated with the reduction of activated surface oxygen species and shallow lattice oxygen in ceria by the following reaction (V) [10,40]. By contrast, the intensity of H2 consumption peaks below 250 °C strongly increased, and the peaks showed an obvious shift toward lower temperature, which indicates the significant increments of activated surface oxygen species and the increased reducibility of the CeO2-Pt/Al2O3 catalyst. The H2-TPR profile of the reference catalyst exhibits the lower intensity of H2 consumption peaks below 200 °C, and these peaks appeared at higher temperature compared with the CeO2-Pt/Al2O3 catalyst, suggesting that the choice of Pt-precursors plays an important role in the reducibility of Ce-containing catalysts and [HOCH2CH2NH3]2Pt(OH)6 is an optimal precursor,
2CeO2 + H2 → Ce2O3 + H2O.

3. Materials and Methods

3.1. Materials and Catalyst Preparation

Bis(ethanolammonium)hexahydroxyplatinate(IV), [HOCH2CH2NH3]2Pt(OH)6, was supplied by Sino-platinum metals Co., Ltd., Kunming, China. The nano-alumina was purchased from Aladdin, Shanghai, China. Cerium nitrate and aluminium nitrate were purchased from Xilong Chemical Co., Ltd., Guangzhou, China.
The Pt/alumina powder with the target Pt loading of 1.176% was prepared by incipient wetness impregnation of nano-alumina with an aqueous solution of [HOCH2CH2NH3]2Pt(OH)6, followed by dry at 50 °C under vacuum condition overnight and calcination at 200 °C in 10%O2/N2 stream for 30 min. The Pt/Al2O3 catalyst was prepared by incipient wetness impregnation of the collected Pt/alumina powder. 17 g of the as-prepared Pt/alumina powder was added into Al(NO3)3 aqueous solution containing 29.42 mmol Al under magnetic stirring at room temperature for 2 h. The obtained slurries were dried at 50 °C under vacuum condition overnight and calcined at 550 °C in 10%O2/N2 stream for 2 h with a heating rate of 10 °C/min in N2 stream to obtain the Pt/Al2O3 sample. To synthesize the CeO2-Pt/Al2O3 catalyst, 17 g of the prepared Pt/alumina powder was dispersed in an aqueous solution of Ce(NO3)3 containing 17.43 mmol Ce, and then the obtained slurries were stirred, dried and calcined by the same procedures with the Pt/Al2O3 sample. A reference catalyst was also prepared using tetraammineplatinum (II) nitrate as Pt-precursor by the same procedures with the CeO2-Pt/Al2O3 catalyst.

3.2. Characterization

The in situ diffuse reflectance infrared Fourier transformed spectroscopy (DRIFTS) spectra were collected on a Nicolet 6700 FT-IR spectrometer equipped with an MCT detector. As for in situ detection of the thermal decomposition process of Pt-precursor and Ce-precursor, 2 g of the pristine sample without calcination was firstly pretreated at 50 °C under vacuum condition for 10 h. Then, background spectra were collected at every 50 °C from 100 °C to 500 °C in a flow of nitrogen (50 mL/min) with a heating rate of 10 °C/min. After cooling to room temperature, 100 mg of as-pretreated catalyst sample was placed in the in situ cell, and each transient experiment was measured in N2 stream with a total flow rate of 50 mL/min. The same methods were taken with the background spectrum collection procedure to acquire DRIFTS spectra.
The thermal decomposition characteristics of Pt-precursor and Ce-precursor in N2 stream with a heating rate of 10 °C/min were examined by thermogravimetric (TG) analysis, using an STA 449 F3 thermoanalyzer. The pristine samples without calcination were firstly pretreated at 50 °C under vacuum condition for 10 h. 100 mg of as-pretreated sample was immediately placed in the pot of thermoanalyzer and heated at a rate of 10 °C/min from 50 °C to 550 °C in a flow of nitrogen (50 mL/min).
X-ray diffraction (XRD) patterns were recorded on a Philips X’pert Pro diffractometer in the 2θ range of 5–90° with an increment step of 0.02°, using a Ni filtered Cu Kα radiation (λ = 0.15418 nm) source. The X-ray tube was operated at 36 kV and 30 mA.
H2 temperature-programmed reduction (H2-TPR) was carried out on CHEMBET 3000 automated chemisorption instrument equipped with a thermal conductivity detector (TCD). 100 mg of sample was placed in a quartz U-tube reactor and pretreated in an airstream with a total flow rate of 50 mL/min at 450 °C for 30 min. After cooled to room temperature, H2-TPR experiments were performed in an H2-Ar (9.91% of H2 by volume) stream with a total flow rate of 50 mL/min from 50 °C to 620 °C at a heating rate of 10 °C/min.

3.3. Catalytic Performance Tests

The catalytic oxidation activity of the catalysts was performed under atmospheric pressure in a continuous flow fixed-bed micro-reactor system. 500 mg of the catalyst sample with an average diameter of 40–60 mesh was placed in a quartz tube with an inner diameter of 8 mm for activity measurements. The total flow rate was 500 mL/min to make the gas hourly space velocity of (GHSV) value of 60,000 mL/(g·h). The outlet concentrations of reaction gases and gaseous products were analyzed online using a 2030DBG2EZKS13T MultiGas FT—IR Analyzer purchased from MKS Instruments, Inc. Prior to the activity measurement, each catalyst sample was pretreated at 500 °C in a flow of O2-N2 (10% of O2 by volume) mixture for 30 min. The catalytic performance was tested in a continuous flow of the reaction mixture gas. The inlet composition of the reaction mixture gas was 500 ppm CO, 200 ppm NO, 660 ppm C3H6, 290 ppm C3H8, 10 vol.% O2, 8 vol.% CO2, 5 vol.% H2O and ~77 vol.% N2. NO conversion, α(NO, %), was calculated by the following Equation (1):
α ( NO , % ) = C NO , i n C NO , o u t C NO , i n × 100 % ,
Y ( NH 3 , % ) = C NH 3 , o u t C NO , i n × 100 % ,
Y ( N 2 O , % ) = 2 × C N 2 O , o u t C NO , i n × 100 % ,
Y ( NO 2 , % ) = C NO 2 , o u t C NO , i n × 100 % ,
where, CNO,in and CNO,out represent the concentrations of the inlet and outlet NO gas, respectively. The NH3, N2O and NO2 yields were calculated according to the Equations (2–4) on the basis of the outlet concentrations of NH3, N2O and NO2 products, respectively. The values of CO, C3H6 and C3H8 conversion rate were calculated by the same method associated with the calculation for NO conversion.

4. Conclusions

It was observed that Pt-catalyzed decomposition of cerium nitrate effectively increases the dispersity of ceria and significantly promotes the formation of activated surface oxygen species. The presence of Pt4+ species effectively improve the dispersity of ceria and readily increases the reducibility of Ce-containing catalysts. The undesirable N2O by-product was formed, due to the selective catalytic reduction of NO by C3H6 below 350 °C, according to the results of in situ activity measurements. The N2O yield was increased by the addition of ceria, resulting from the increased reducibility of Ce-containing catalysts. The improved activity of CeO2-Pt/Al2O3 catalyst for the catalytic CO, NO, C3H6 and C3H8 co-oxidation were attributed to its high reducibility and the high surface concentrations of activated oxygen species from the interactions between active Pt sites and ceria.

Author Contributions

Conceptualization and methodology, C.W. and Y.Z.; investigation and formal analysis, C.W. and J.D.; experiments and data curation, C.W., D.R. and Q.Q. and A.Z.; writing—original draft preparation and editing, C.W. and A.Z.; writing—review and project administration, L.C., H.C. and Y.Z.; funding acquisition, C.W., J.C. and Y.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Natural Science Foundation of China (21862010, 51676017), the Provincial Applied Fundamental Research Program of Yunnan (2018FD140), the Major Science and Technology Programs of Yunnan (2019ZE001), and the Technology Talents and Platform Project of Yunnan (2018IC091).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Marcinkowski, M.D.; Darby, M.T.; Liu, J.; Wimble, J.M.; Lucci, F.R.; Lee, S.; Michaelides, A.; Flytzani-Stephanopoulos, M.; Stamatakis, M.; Sykes, E.C.H. Pt/Cu single-atom alloys as coke-resistant catalysts for efficient C–H activation. Nat. Chem. 2018, 10, 325–332. [Google Scholar] [CrossRef] [PubMed]
  2. Sattler, J.J.H.B.; Ruiz-Martinez, J.; Santillan-Jimenez, E.; Weckhuysen, B.M. Catalytic dehydrogenation of light alkanes on metals and metal oxides. Chem. Rev. 2014, 114, 10613–10653. [Google Scholar] [CrossRef] [PubMed]
  3. Herreros, J.M.; Gill, S.S.; Lefort, I.; Tsolakis, A.; Millington, P.; Mossb, E. Enhancing the low temperature oxidation performance over a Pt and a Pt–Pd diesel oxidation catalyst. Appl. Catal. B Environ. 2014, 147, 835–841. [Google Scholar] [CrossRef] [Green Version]
  4. Gu, L.; Chen, X.; Zhou, Y.; Zhu, Q.; Huang, H.; Lu, H. Propene and CO oxidation on Pt/Ce-Zr-SO42− diesel oxidation catalysts: Effect of sulfate on activity and stability. Chin. J. Catal. 2017, 38, 607–616. [Google Scholar] [CrossRef]
  5. Mathieu, O.; Lavy, J.; Jeudy, E. Investigation of hydrocarbons conversion vver a Pt-based automotive diesel oxidation catalyst: Application to exhaust port fuel injection. Top. Catal. 2009, 52, 1893–1897. [Google Scholar] [CrossRef]
  6. AL-Harbi, M.; Hayes, R.; Votsmeier, M.; Epling, W.S. Competitive no, co and hydrocarbon oxidation reactions over a diesel oxidation catalyst. Can. J. Chem. Eng. 2012, 90, 1527–1538. [Google Scholar] [CrossRef]
  7. Wang, C.; Tan, J.; Harle, G.; Gong, H.; Xia, W.; Zheng, T.; Yang, D.; Ge, Y.; Zhao, Y. Ammonia formation over Pd/Rh three-way catalysts during lean-to-rich fluctuations: The effect of the catalyst aging, exhaust temperature, Lambda, and duration in rich conditions. Environ. Sci. Technol. 2019, 53, 12621–12628. [Google Scholar] [CrossRef]
  8. He, J.; Wang, C.; Zheng, T.; Zhao, Y. Thermally induced deactivation and the corresponding strategies for improving durability in automotive three-way catalysts. Johns. Matthey Technol. Rev. 2016, 60, 196–203. [Google Scholar] [CrossRef]
  9. Jones, J.; Xiong, H.; DeLaRiva, A.T.; Peterson, E.J.; Pham, H.; Challa, S.R.; Qi, G.; Oh, S.; Wiebenga, M.H.; Hernández, X.I.P.; et al. Thermally stable single-atom platinum-on-ceria catalysts via atom trapping. Science 2016, 353, 150–154. [Google Scholar] [CrossRef] [Green Version]
  10. Wang, C.; Feng, F.; Du, J.; Zheng, T.; Pan, Z.; Zhao, Y. Activation of surface lattice oxygen in ceria supported Pt/Al2O3 catalyst for low-temperature propane oxidation. ChemCatChem 2019, 11, 2054–2057. [Google Scholar] [CrossRef]
  11. Nie, L.; Mei, D.; Xiong, H.; Peng, B.; Ren, Z.; Hernandez, X.I.P.; DeLaRiva, A.; Wang, M.; Engelhard, M.H.; Kovarik, L.; et al. Activation of surface lattice oxygen in single-atom Pt/CeO2 for low-temperature CO oxidation. Science 2017, 358, 1419–1423. [Google Scholar] [CrossRef] [Green Version]
  12. Bruix, A.; Lykhach, Y.; Matolínová, I.; Neitzel, A.; Skála, T.; Tsud, N.; Vorokhta, M.; Stetsovych, V.; Ševčíková, K.; Mysliveček, J.; et al. Maximum noble-metal efficiency in catalytic materials: Atomically dispersed surface platinum. Angew. Chem. Int. Ed. 2014, 53, 10525–10530. [Google Scholar] [CrossRef] [PubMed]
  13. Liang, Y.; Ding, X.; Dai, J.; Zhao, M.; Zhong, L.; Wang, J.; Chen, Y. Active oxygen-promoted NO catalytic on monolithic Pt-based diesel oxidation catalyst modified with Ce. Catal. Today 2019, 327, 64–72. [Google Scholar] [CrossRef]
  14. Ma, Z.; Wang, J.; Li, J.; Wang, N.; An, C.; Sun, L. Propane dehydrogenation over Al2O3 supported Pt nanoparticles: Effect of cerium addition. Fuel Process. Technol. 2014, 128, 283–288. [Google Scholar] [CrossRef]
  15. Silva, F.A.; Martinez, D.S.; Ruiz, J.A.C.; Mattos, L.V.; Hori, C.E.; Noronha, F.B. The effect of the use of cerium-doped alumina on the performance of Pt/CeO2/Al2O3 and Pt/CeZrO2/Al2O3 catalysts on the partial oxidation of methane. Appl. Catal. A Gen. 2008, 335, 145–152. [Google Scholar] [CrossRef]
  16. Son, I.H.; Lane, A.M. Promotion of Pt/γ-Al2O3 by Ce for preferential oxidation of CO in H2. Catal. Lett. 2001, 76, 151–154. [Google Scholar] [CrossRef]
  17. Son, I.H. Study of Ce-Pt/γ-Al2O3 for the selective oxidation of CO in H2 for application to PEFCs: Effect of gases. J. Power Source 2006, 159, 1266–1273. [Google Scholar] [CrossRef]
  18. Wells, P.P.; Crabb, E.M.; King, C.R.; Fiddy, S.; Amieiro-Fonseca, A.; Thompsett, D.; Russell, A.E. Reduction properties of Ce in CeOx/Pt/Al2O3 catalysts. Catal. Struct. React. 2015, 1, 88–94. [Google Scholar] [CrossRef]
  19. Onn, T.M.; Zhang, S.; Arroyo-Ramirez, L.; Xia, Y.; Wang, C.; Pan, X.; Graham, G.W.; Gorte, R.J. High-surface-area ceria prepared by ALD on Al2O3 support. Appl. Catal. B Environ. 2017, 201, 430–437. [Google Scholar] [CrossRef] [Green Version]
  20. Damyanova, S.; Bueno, J.M.C. Effect of CeO2 loading on the surface and catalytic behaviors of CeO2-Al2O3-supported Pt catalysts. Appl. Catal. A Gen. 2003, 253, 135–150. [Google Scholar] [CrossRef]
  21. Dawody, J.; Skoglundh, M.; Wall, S.; Fridell, E. Role of Pt-precursor on the performance of Pt/BaCO3/Al2O3 NOx storage catalysts. Appl. J. Mol. Catal. A Chem. 2005, 225, 259–269. [Google Scholar] [CrossRef]
  22. van den Tillaart, J.A.A.; Leyrer, J.; Eckhoff, S.; Lox, E.S. Effect of support oxide and noble metal precursor on the activity of automotive diesel catalysts. Appl. Catal. B Environ. 1996, 10, 53–68. [Google Scholar] [CrossRef]
  23. Falco, M.G.; Canavese, S.A.; Fígoli, N.S. The calcination temperature after platinum addition and its effect on Pt/WOx–ZrO2 properties. Catal. Commun. 2001, 2, 207–211. [Google Scholar] [CrossRef]
  24. Gao, Y.; Wu, X.; Nord, R.; Härelind, H.; Weng, D. Sulphation and ammonia regeneration of Pt/MnOx-CeO2/Al2O3 catalyst for NOx-assisted soot oxidation. Catal. Sci. Technol. 2018, 8, 1621–1631. [Google Scholar] [CrossRef]
  25. Mueller, R.; Kammler, H.K.; Wegner, K.; Pratsinis, S.E. OH surface density of SiO2 and TiO2 by thermogravimetric analysis. Langmuir 2003, 19, 160–165. [Google Scholar] [CrossRef]
  26. Dou, D.; Liu, D.J.; Williamson, W.B.; Kharas, K.C.; Robota, H.J. Structure and chemical properties of Pt nitrate and application in three-way automotive emission catalysts. Appl. Catal. B Environ. 2001, 30, 11–24. [Google Scholar] [CrossRef]
  27. Aldiyarov, A.; Aryutkina, M.; Drobyshev, A.; Kurnosov, V. IR spectroscopy of ethanol in nitrogen cryomatrices with different concenteration ratios. Low Temp. Phys. 2011, 37, 524–531. [Google Scholar] [CrossRef]
  28. Cabritaa, J.F.; Viana, A.S.; Mourato, A.; Montforts, F.P.; Abrantes, L.M. Self-assembled monolayers of metalloporphyrin phosphonates on electrochemically modified HOPG. Catal. Today 2012, 187, 70–76. [Google Scholar] [CrossRef]
  29. Wallin, M.; Byberg, M.; Grönbeck, H.; Skoglundh, M.; Eriksson, M.; Spetz, A.L. Vibrational analysis of H2 and NH3 adsorption on Pt/SiO2 and Ir/SiO2 model sensors. In Proceedings of the IEEE Sensors 2007, Atlanta, GA, USA, 28–31 October 2007; pp. 1315–1317. [Google Scholar]
  30. Carlsson, P.A.; Österlund, L.; Thormählen, P.; Palmqvist, A.; Fridell, E.; Jansson, J.; Skoglundh, M. A transient in situ FTIR and XANES study of CO oxidation over Pt/Al2O3 catalysts. J. Catal. 2004, 226, 422–434. [Google Scholar] [CrossRef]
  31. Solymosi, F.; Sárkány, J.; Schauer, A. Study of the formation of isocyanate surface complexes on Pt/AI2O3 catalysts. J. Catal. 1977, 46, 297–307. [Google Scholar] [CrossRef]
  32. O’Brien, C.P.; Lee, I.C. A detailed spectroscopic analysis of the growth of oxy-carbon species on the surface of Pt/Al2O3 during propane oxidation. J. Catal. 2017, 347, 1–8. [Google Scholar] [CrossRef]
  33. Ermini, V.; Finocchio, E.; Sechi, S.; Busca, G. An FT-IR and flow reactor study of the conversion of propane on γ-Al2O3 in oxygen-containing atmosphere. Appl. Catal. A Gen. 2000, 190, 157–167. [Google Scholar] [CrossRef]
  34. Azambre, B.; Zenboury, L.; Koch, A.; Weber, J.V. Adsorption and desorption of NOx on commercial ceria-zirconia (CexZr1-xO2) mixed oxides: A combined TGA, TPD-MS, and DRIFTS study. J. Phys. Chem. C 2009, 113, 13287–13299. [Google Scholar] [CrossRef]
  35. Azambre, B.; Zenboury, L.; Delacroix, F.; Weber, J.V. Adsorption of NO and NO2 on ceria–zirconia of composition Ce0.69Zr0.31O2: A DRIFTS study. Catal. Today 2008, 137, 278–282. [Google Scholar] [CrossRef]
  36. Hadjiivanov, K.I. Identification of neutral and charged NxOy surface species by IR spectroscopy. Catal. Rev. 2007, 42, 71–144. [Google Scholar] [CrossRef]
  37. Wendlandt, W.W.; Bear, J.L. Thermal decomposition of the heavier rare-earth metal nitrate hydrates: Thermobalance and differential thermal analysis studies. J. Inorg. Nucl. Chem. 1960, 12, 276–280. [Google Scholar] [CrossRef]
  38. Oh, H.; Luo, J.; Epling, W.S. NO oxidation inhibition by hydrocarbons over a diesel oxidation catalyst: Reaction between surface nitrates and hydrocarbons. Catal. Lett. 2011, 141, 1746–1751. [Google Scholar] [CrossRef]
  39. Han, L.; Cai, S.; Gao, M.; Hasegawa, J.; Wang, P.; Zhang, J.; Shi, L.; Zhang, D. Selective catalytic reduction of NOx with NH3 by using novel catalysts: State of the art and future prospects. Chem. Rev. 2019, 119, 10916–10976. [Google Scholar] [CrossRef]
  40. Wang, C.; Zheng, T.; Lu, J.; Wu, X.; Hochstadt, H.; Zhao, Y. Three-way catalytic reactions on Rh-based catalyst: Effect of Rh/ceria interfaces. Appl. Catal. A Gen. 2017, 544, 30–39. [Google Scholar] [CrossRef]
Figure 1. Thermogravimetric (TG) curves of the different samples at a heating rate of 10 ºC/min in N2 steam.
Figure 1. Thermogravimetric (TG) curves of the different samples at a heating rate of 10 ºC/min in N2 steam.
Catalysts 10 00143 g001
Figure 2. TG and derivative thermogravimetric (DTG) curves (a) and diffuse reflectance infrared Fourier transformed spectroscopy (DRIFTS) spectra (b) collected during thermal decomposition of Pt-precursor on Al2O3 powder at a heating rate of 10 ºC/min in N2 steam.
Figure 2. TG and derivative thermogravimetric (DTG) curves (a) and diffuse reflectance infrared Fourier transformed spectroscopy (DRIFTS) spectra (b) collected during thermal decomposition of Pt-precursor on Al2O3 powder at a heating rate of 10 ºC/min in N2 steam.
Catalysts 10 00143 g002
Figure 3. TG and DTG curves (a,c) and DRIFTS spectra (b,d) collected during thermal decomposition of Ce-precursor on Al2O3 (ab) and Pt/Al2O3 (cd) powders at a heating rate of 10 °C/min in N2 steam.
Figure 3. TG and DTG curves (a,c) and DRIFTS spectra (b,d) collected during thermal decomposition of Ce-precursor on Al2O3 (ab) and Pt/Al2O3 (cd) powders at a heating rate of 10 °C/min in N2 steam.
Catalysts 10 00143 g003aCatalysts 10 00143 g003b
Figure 4. Conversion (a,c) and yield (b,d) curves during catalytic oxidation reactions over the Pt/Al2O3 (a,b), CeO2-Pt/Al2O3 (c,d) and reference (e,f) catalysts.
Figure 4. Conversion (a,c) and yield (b,d) curves during catalytic oxidation reactions over the Pt/Al2O3 (a,b), CeO2-Pt/Al2O3 (c,d) and reference (e,f) catalysts.
Catalysts 10 00143 g004
Figure 5. XRD patterns (a) and H2 temperature-programmed reduction (H2-TPR) profiles (b) of the different catalyst samples.
Figure 5. XRD patterns (a) and H2 temperature-programmed reduction (H2-TPR) profiles (b) of the different catalyst samples.
Catalysts 10 00143 g005
Table 1. The characteristic temperatures for catalytic oxidation of CO, NO, C3H6 and C3H8 over the Pt/Al2O3 and CeO2-Pt/Al2O3 catalysts.
Table 1. The characteristic temperatures for catalytic oxidation of CO, NO, C3H6 and C3H8 over the Pt/Al2O3 and CeO2-Pt/Al2O3 catalysts.
SamplesCONOC3H6C3H8
T50aT90bT50T90T50T90T50T90
Pt/Al2O3210.6241.3262.2nd c243.3252.0384.0456.8
CeO2-Pt/Al2O3177.7200.1209.5nd208.8217.6335.5364.8
Reference catalyst201.2244.7ndnd245.2264.0329.1390.4
1 d32.941.252.7nd34.534.448.592.0
2 e23.544.6ndnd36.446.4-6.425.6
aT50 is the specific temperature reached the conversion of 50%, so-called “light-off temperature”. b T90 is the specific temperature reached the conversion of 90%. c nd = no data. d1 represents the characteristic temperature gap between Pt/Al2O3 and CeO2-Pt/Al2O3 catalyst. e2 represents the characteristic temperature gap between the reference catalyst and CeO2-Pt/Al2O3 catalyst.

Share and Cite

MDPI and ACS Style

Wang, C.; Ren, D.; Du, J.; Qin, Q.; Zhang, A.; Chen, L.; Cui, H.; Chen, J.; Zhao, Y. In Situ Investigations on the Facile Synthesis and Catalytic Performance of CeO2-Pt/Al2O3 Catalyst. Catalysts 2020, 10, 143. https://doi.org/10.3390/catal10020143

AMA Style

Wang C, Ren D, Du J, Qin Q, Zhang A, Chen L, Cui H, Chen J, Zhao Y. In Situ Investigations on the Facile Synthesis and Catalytic Performance of CeO2-Pt/Al2O3 Catalyst. Catalysts. 2020; 10(2):143. https://doi.org/10.3390/catal10020143

Chicago/Turabian Style

Wang, Chengxiong, Dezhi Ren, Junchen Du, Qinggao Qin, Aimin Zhang, Li Chen, Hao Cui, Jialin Chen, and Yunkun Zhao. 2020. "In Situ Investigations on the Facile Synthesis and Catalytic Performance of CeO2-Pt/Al2O3 Catalyst" Catalysts 10, no. 2: 143. https://doi.org/10.3390/catal10020143

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