Next Article in Journal
Accelerating Kinetics with Time-Reversal Path Sampling
Next Article in Special Issue
PVP Passivated δ-CsPbI3: Vacancy Induced Visible-Light Absorption and Efficient Photocatalysis
Previous Article in Journal
Study on Physicochemical Properties of Food Protein
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Monometallic and Bimetallic Catalysts Supported on Praseodymium-Doped Ceria for the Water–Gas Shift Reaction

by
Weerayut Srichaisiriwech
and
Pannipa Tepamatr
*
Department of Chemistry, Faculty of Science and Technology, Thammasat University, Pathumthani 12120, Thailand
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(24), 8146; https://doi.org/10.3390/molecules28248146
Submission received: 13 October 2023 / Revised: 8 December 2023 / Accepted: 14 December 2023 / Published: 18 December 2023
(This article belongs to the Special Issue New Materials and Catalysis in Environmental Protection)

Abstract

:
The water–gas shift (WGS) performance was investigated over 5%Ni/CeO2, 5%Ni/Ce0.95Pr0.05O1.975, and 1%Re4%Ni/Ce0.95Pr0.05O1.975 catalysts to decrease the CO amount and generate extra H2. CeO2 and Pr-doped CeO2 mixed oxides were synthesized using a combustion method. After that, Ni and Re were loaded onto the ceria support via an impregnation method. The structural and redox characteristics of monometallic Ni and bimetallic NiRe materials, which affect their water–gas shift performance, were investigated. The results show that the Pr addition into Ni/ceria increases the specific surface area, decreases the ceria crystallite size, and improves the dispersion of Ni on the CeO2 surface. Furthermore, Re addition results in the enhancement of the WGS performance of the Ni/Ce0.95Pr0.05O1.975 catalyst. Among the studied catalysts, the ReNi/Ce0.95Pr0.05O1.975 catalyst showed the highest catalytic activity, reaching 96% of CO conversion at 330°. It was established that the occurrence of more oxygen vacancies accelerates the redox process at the ceria surface. In addition, an increase in the Ni dispersion, Ni surface area, and surface acidity has a positive effect on hydrogen generation during the water–gas shift reaction due to favored CO adsorption.

1. Introduction

Water–gas shift (WGS) reaction is an industrial technology that involves the reaction of water vapor and carbon monoxide to generate H2 and CO2. The water–gas shift reaction takes place according to the following equation:
CO + H2O ⇆ H2 + CO2         ΔH298 = −41.2 kJ/mol
The development of H2 and fuel cell technologies provides numerous advantages such as high-efficiency power, environmentally friendliness, and sustainability. WGS reactions have received widespread attention to increase the H2 concentration in the syngas. Hydrocarbon reforming can produce syngas that consists of CO, H2, H2O, and CO2. However, a trace concentration of CO poisons the catalysts utilized in fuel cells [1]. Precious metal-based proton exchange membrane fuel cell anodes require a carbon monoxide amount in the inlet gas below 10–20 ppm; otherwise, the anode is poisoned [2]. A purification process is required to decrease the CO concentration lower than the cell tolerance level after hydrogen is produced via the reforming process of carbon-containing molecules (such as hydrocarbons or alcohols). The advantage of using the WGS reaction is to reduce CO content while generating more hydrogen as fuel for the H2 fuel cell. The appearance of a suitable catalyst in the WGS reaction can reduce the CO content to 10 ppm.
The support plays a critical role in oxidation reactions such as water–gas shift reactions or CO oxidation. The utilization of redox-active oxides such as CeO2 as a support material leads to superior catalytic efficiency compared to the use of other oxide support such as alumina or silica [3,4,5,6]. Due to its redox characteristics, CeO2 can promote vacancy generation and water dissociation, which plays an important role in the catalytic performance of water–gas shift reactions. However, pure CeO2 has some disadvantages, like deactivation of the Ce4+/Ce3+ redox couple and thermal sintering, resulting in a reduction in its oxygen storage capacity (OSC) and catalytic performance [7]. Thus, many efforts have been dedicated to improving the reducibility and catalytic performance of CeO2 by doping it with other cations to generate more oxygen vacancies and develop its resistance to thermal sintering [8,9]. In this part, doping CeO2 with aliovalent (such as Eu3+, La3+, Gd3+, and Sm3+) cations [10,11,12] or isovalent (such as Zr4+, Hf4+, and Ti4+) cations [13,14] has been well described in the literature. The incorporation of an aliovalent cation into the CeO2 lattice produces oxygen vacancies by charge compensation on the final oxide materials [15]. Conversely, the CeO2 doping with isovalent cations also results in the enhancement of the redox properties of CeO2. Therefore, the partial substitution of Ce4+ by Zr4+ causes a deformation in the CeO2 lattice because of the lower ionic radius of isovalent cations such as Zr4+, deriving an enhancement in the OSC of CeO2. Moreover, these two effects can also be combined in a single study when doping CeO2 with variable oxidation states of lanthanide elements. In such cases, Tb [16] and Pr [17,18] are the most studied ones. The Pr addition into the CeO2 lattice enhances both oxygen desorption and oxygen vacancy generation compared to pure CeO2. This result is due to the lower binding energy of oxygen anions in Ce-Pr mixed oxides and the higher reduction potential of Pr4+/Pr3+ compared to Ce4+/Ce3+. Therefore, the Zr dopant induces a limited amount of oxygen vacancies into the CeO2 lattice [19,20], whereas the Pr dopant leads to a greater concentration of oxygen vacancies [21]. These behaviors result in redox properties in the Ce-Pr mixed oxides being superior to those of other mixed oxides. In previous works, Ce-Pr-O mixed oxide supports were studied for WGS reaction. It was found that the optimum promoting effect of Pr appears at a low loading of 5 %wt. [22,23]. In addition, the catalytic activity of such oxides can be obviously improved by adding a small amount of transition metals (cobalt, chromium, copper, and nickel) [24,25,26]. Nickel combined with CeO2-based oxides are cost-effective alternatives to expensive noble metal catalysts and are often more reactive than noble metals [27]. However, CO and CO2 methanation are very common side reactions for these catalysts [28,29]. This behavior is usually tempered by the incorporation of a second active metal like platinum [30,31]. Rhenium might be a good choice for replacing Pt catalyst because of its good electrochemical properties, its cheaper price compared to platinum, and its sustainable source. Re is widely used as a second metal to form bimetallic catalysts. In recent years, the use of bimetallic catalysts has attracted much attention due to their excellent efficiency and capability [32,33,34]. The effect of rhenium on the WGS performance of Pt/TiO2 and Pt/ZrO2 catalysts has been investigated, and the results exhibited that rhenium addition induces an increase in the WGS activity of Pt catalysts. Rhenium acts as an anchor for the platinum particles to enhance the Pt dispersion. In addition, the redox process between Re4+ and Re7+ over the WGS reaction would promote CO oxidation on the Pt catalyst [35,36]. Our initial studying found that rhenium also enhanced the WGS activity when it was doped on Ni/CeO2–based oxides [11,37,38]. Apart from enhancing the performance of the WGS reaction, Pt–Re/carbon has also been reported to be active in several reactions, such as the reforming process [23] and glycerol to syngas conversion [39]. Furthermore, in studying the catalytic activity of ReCo/Al2O3 for the Fischer–Tropsch reaction found that Re addition increased the reaction rate of the Co catalyst. Re has been shown to be a good promoter by facilitating the reduction rate of cobalt species and producing more available active Co metal sites to participate in the reaction [40].
In this work, the performance of Ni/CeO2 and Ni/Ce0.95Pr0.05O1.975 catalysts for the water–gas shift reaction was studied. Additionally, the role of rhenium addition on the water–gas shift performance of Ni/Ce0.95Pr0.05O1.975 catalyst was also observed. Therefore, the utilization of Pr as a dopant and Re as a metal additive in this work to maximize the WGS performance. The physicochemical properties of monometallic and bimetallic catalysts were examined to clarify the key factors in increasing the catalytic activity using the following techniques: X-ray diffraction, BET surface area, NH3 temperature programmed desorption, H2 temperature programmed reduction, Raman spectroscopy, and chemisorption techniques.

2. Results and Discussion

2.1. Catalysts Characterization

X-ray diffraction patterns of CeO2, 5%Ni/CeO2, 5%Ni/Ce0.95Pr0.05O1.975, and 1%Re4%Ni/Ce0.95Pr0.05O1.975 were illustrated in Figure 1. It suggests that the diffraction peaks of all catalysts correspond well to CeO2 phases with a cubic structure (Joint Committee on Powder Diffraction File No. 43-1002). In addition, the weak peaks at 2θ about 37.1°, 43.2° and 63.2° attributed to the NiO phases of Ni-based catalysts, suggesting that there was a small proportion of nickel oxide. The CeO2 crystallite size of supported Ni catalysts was determined using the Debye–Scherrer equation (Table 1). The calcination of Ni/CeO2 catalyst at a high temperature (650 °C) after impregnation of Ni onto ceria support leads to the aggregation of ceria crystallites; thereby, the surface area decreases with a growth in CeO2 crystallite size. However, Pr addition into 5%Ni/CeO2 results in a decrease in the ceria crystallite size together with an increase in a specific surface area. The diffraction peaks of Ni/Ce0.95Pr0.05O1.975 and ReNi/Ce0.95Pr0.05O1.975 appeared at lower diffraction angles compared with the diffraction peaks of Ni/CeO2, indicating that Pr incorporation in the CeO2 lattice enlarged unit cell. The enhancement of the unit cell for Ni/Ce0.95Pr0.05O1.975 when compared with Ni/CeO2 is due to Ce4+ ions (0.097 nm) being replaced by larger Pr3+ ions (0.112 nm). Therefore, an oxygen vacancy is expected to be formed because Pr3+ incorporation in the CeO2 lattice produces unbalanced charges and strain. On the other hand, the diffraction peaks of Ni/CeO2 appeared at higher diffraction angles compared with the diffraction peaks of pure CeO2 because of the lattice contraction after calcination at high temperatures. Nickel could not be incorporated into CeO2 lattice due to the nature of impregnation synthesis but the reduction in the cell dimension because of the decomposition of surface hydroxyls during calcination at 650 °C.
H2 chemisorption analysis was used to determine the Ni dispersion of supported Ni catalysts. It was found that an addition of Pr to Ni/CeO2 increases Ni dispersion on the catalyst surface. Moreover, rhenium impregnation onto Ni/Ce0.95Pr0.05O1.975 tremendously enhanced the dispersion and surface area of metallic nickel. This result may be due to the movement of electrons between Re, Ni, and CeO2, which results in the formation of strong interaction between Ni metal and support; thereby, the metal dispersion and metal surface coverage enhances, whereas particle size reduces. 1%Re4%Ni/Ce0.95Pr0.05O1.975 exhibited the highest Ni surface area and dispersion among all the catalysts. Usually, a greater metal surface area provides more surface active sites exposed to reactants [36,37].
The chemical analysis by SEM micrographs with the corresponding elemental mapping was conducted to investigate the elemental distribution and the homogeneity of the supported Ni catalyst. As presented in Figure 2, Ce, Pr, and Ni elements were uniformly distributed on 5%Ni/Ce0.95Pr0.05O1.975 catalysts. The highly dispersed Ni suggests a strong metal–support interaction, and the enhancement in Ni dispersion would provide more active sites that are exposed to reactants, which is beneficial to the increase in WGS performance.
Raman spectroscopy was performed to quantify oxygen vacancies in the catalyst. CeO2 catalyst initiates the water–gas shift process via a redox mechanism at high temperatures. CO adsorbs on the catalyst surface and subsequently oxidizes it with CeO2 lattice oxygen to generate carbon dioxide and oxygen vacancy. H2O oxidizes reduced CeO2 again to produce H2. A mechanism for increasing the catalytic performance of CeO2 is the incorporation of dopant ions, with Pr3+ as a promising candidate dopant [41]. It is widely regarded that when doping CeO2 with a trivalent cation, two Ce4+ ions in the CeO2 lattice are substituted by the dopants, and then an O ion is eliminated to conserve the charge [15]. Therefore, oxygen vacancy is directly related to catalytic activity in water–gas shift reaction.
As shown in Figure 3, a Raman peak near 460 cm−1 was attributed to a triple degeneracy active mode (F2g peak), which represents the symmetrical stretching vibration generated by eight O atoms bound to one Ce atom. Secondary peaks at around 240 and 320 cm−1 characteristics of CeO2 nanostructures are also found in 1%Re4%Ni/Ce0.95Pr0.05O1.975. In addition, another broad peak near 570 cm−1 (denoted by D peak) was associated with oxygen vacancies in CeO2 [42,43]. The oxygen vacancies concentration can be represented by the ratio of ID/IF2g [44]. The intensity of the D peak in 1%Re4%Ni/Ce0.95Pr0.05O1.975 catalyst is stronger than that of other catalysts, indicating that higher oxygen vacancy concentration can be obtained by the addition of Re onto Ni/Ce0.95Pr0.05O1.975 catalyst. Moreover, the presence of Ce3+ in the CeO2-based catalyst can be demonstrated by a red shift of the F2g peak, which is due to the lattice expansion when Ce4+ ions (ionic radius 0.097 nm) are replaced by Ce3+ ions (ionic radius 0.114 nm) for oxygen vacancy formation [43,45]. The enhancement of oxygen vacancy concentration in the 1%Re4%Ni/Ce0.95Pr0.05O1.975 catalyst enables the interaction between rhenium, nickel, and CeO2 to drive the metal dispersion and prevent the sintering of metal particles. This result indicates that the addition of Re to the Ni/Ce0.95Pr0.05O1.975 catalyst improves the reducibility and stability.
The H2-TPR profiles of CeO2, 5%Ni/CeO2, 5%Ni/Ce0.95Pr0.05O1.975, and 1%Re4%Ni/Ce0.95Pr0.05O1.975 catalysts are shown in Figure 4. The H2-TPR of CeO2 exhibits two broad peaks at 500 °C and 750 °C. The peak at 500 °C is assigned to the reduction of surface-capping oxygen of CeO2, and the peak at 750 °C is assigned to the bulk CeO2 reduction. The TPR profile of the Ni/CeO2 catalyst is characterized by a low-temperature peak at 272 °C, medium temperature at 345 °C, and bulk reduction at 830 °C. The reduction peak at 272 °C is assigned to the reduction of nickel oxide species. The consumption peak at 345 °C is assigned to the Ni-catalyzed reduction of the CeO2 surface [46,47]. It is interesting to note that the incorporation of Ni to CeO2 support significantly shifts the reduction peak of surface CeO2 from 500 °C to 345 °C. The H2-TPR of Ni/Ce0.95Pr0.05O2−δ presented two nickel oxide reduction peaks at 220 °C and 278 °C, which was due to the different environments of Ni. The peak at 220 °C is probably due to the reduction of Ni in the vicinity of CeO2, whereas the consumption peak at 278 °C is due to the presence of Pr. The reduction peak of surface and bulk species of Ni/Ce0.95Pr0.05O1.975 appeared at the same position as the reduction of Ni/CeO2. This indicated that the addition of Pr to Ni/CeO2 alters the NiO reduction behavior. The H2-TPR profile of the bimetallic ReNi is different from those of monometallic Ni supported on Pr-doped CeO2. In this case, electron density transfers between Re, Ni, Ce, and Pr may occur. As the result of electron density transfer, a concurrent reduction of metal oxide species was found, and reduction of metal oxide is easier. A stronger interaction between nickel and CeO2 is expected to tune the nickel dispersion. The reduction peak appearing at a higher temperature normally means that it is more difficult to reduce with stronger metal–support interactions [48]. For the 1%Re4%Ni/Ce0.95Pr0.05O1.975 catalyst, stronger metal-support interactions are presented, which proved that the reduction of the surface shell of CeO2 occurred at a higher temperature. The stronger interaction between metal and support is beneficial to maintain the metal dispersion and hinder its aggregation.
The surface acidity of the prepared catalysts was studied using temperature-programmed desorption of ammonia (Figure 5), and the total acidity was estimated from the area under the NH3 desorption peak. NH3-TPD analysis was carried out in order to clarify the effect of the acidity of ReNi/Ce0.95Pr0.05O1.975 and Ni/Ce0.95Pr0.05O1.975 on the catalytic performance in the water–gas shift reaction. The peaks were assigned to weak, medium, or strong acid sites when falling in the 100–200 °C, 200–450 °C, or 450–700 °C temperature range, respectively. Increased surface acidity enabled a higher content of CO adsorption on the catalyst surface since a CO reactant in the WGS process is a weak base, explaining the observed increase in catalyst activity. Furthermore, the acidic character of the Ni catalyst surface proved to be beneficial for CO2 desorption, leaving behind free active sites for carbon monoxide and H2O adsorption in subsequent reaction cycles [49]. The result from NH3-TPD analysis indicates that the addition of Re onto Ni/Ce0.95Pr0.05O1.975 increases the concentration of weak-strength acid sites (peak area increases at <200 °C). In addition, the total concentration of surface acid sites can be estimated by integrating the NH3-TPD curves, and it was found to be 41 and 28 mols/g for 1%Re4%Ni/Ce0.95Pr0.05O1.975 and 5%Ni/Ce0.95Pr0.05O1.975, respectively. The obtained results could imply a higher tendency for carbon monoxide adsorption and subsequently easier CO2 desorption on the bimetallic ReNi supported by Pr-doped CeO2 surface, thereby the overall water–gas shift reaction rate over ReNi/Ce0.95Pr0.05O1.975 may be enhanced.
XPS characterization was used to investigate the surface chemical states of the catalysts. Figure 6a shows the O 1s XPS spectra of 5%Ni/Ce0.95Pr0.05O1.975 and 1%Re4%Ni/Ce0.95Pr0.05O1.975 catalysts. Three different types of oxygen species were detected in all samples. The detected peaks near 529 eV (OL), 532 eV (OA), and 533 eV (OH) are attributed to lattice oxygen in metal oxide, chemically adsorbed oxygen on the surface, and a surface hydroxyl oxygen species, respectively. The ratios of OA/OL, which are calculated from the area of each peak, are an indicator of active oxygen vacancies on the surface [50,51]. XPS results indicated that active oxygen vacancies were higher for the 1%Re4%Ni/Ce0.95Pr0.05O1.975 (OA/OL = 0.28) compared to 5%Ni/Ce0.95Pr0.05O1.975 catalysts (OA/OL = 0.24). Therefore, the bimetallic ReNi catalyst tended to display greater activity due to it producing more vacancies or defects.
5%Ni/Ce0.95Pr0.05O1.975 and 1%Re4%Ni/Ce0.95Pr0.05O1.975 illustrated Ni 2p spectra mainly contributed by Ni2+ species at around 855 and 856 eV with a minor content of Ni0 species at around 853 eV (Figure 6b). All samples were reduced with 5%H2/N2 at 400 °C for 1 hour before XPS measurement. Ni species in the reduced catalysts were in the form of metallic Ni. Furthermore, the different Ni species co-existed due to the interaction with CeO2-based materials. Metallic Ni0 was indicated to be the dominant active species in accelerating the reactants with content of 36.6% and 23.3% in 1%Re4%Ni/Ce0.95Pr0.05O1.975 and 5%Ni/Ce0.95Pr0.05O1.975 catalysts, respectively. Therefore, the increase in metallic Ni amount in the bimetallic ReNi catalyst implies a superior catalytic performance of Ni catalyst by the addition of Re.
SEM analysis (shown in Figure 7) confirms there was almost no carbon deposition on the surface for the used 1%Re4%Ni/Ce0.95Pr0.05O1.975 catalyst. Wang et al. [52] reported that more carbon was deposited and accumulated on the surface of monometallic Ni catalysts during steam reforming of biomass tar, whereas bimetallic NiFe catalysts suppressed the carbon deposition on the surface of the reacted catalyst. Therefore, using bimetallic catalysts could prevent coke formation on the catalysts by providing oxidation of the accumulated carbon.
The carbon deposition of the spent catalysts was evaluated by TG analysis (Figure 8). The oxidation of the carbon deposition in the air leads to weight loss. Small weight loss at low temperatures (below 200 °C) was ascribed to the elimination of moisture and volatile species [53]. The mass loss in the range of 200–400 °C was ascribed to the thermal decomposition of physisorbed carbonaceous species or soft-coke. A major weight reduction between 400 and 600 °C was due to the bulky carbonaceous products or hard coke on the used catalysts. The weight loss percentages of the bulky carbonaceous species on monometallic Ni and bimetallic NiRe catalysts were 11.6% and 6.2% for 5%Ni/CeO2 and 1%Re4%Ni/Ce0.95Pr0.05O1.975, respectively, indicating that carbon decomposition decreases when use Pr as dopant and Re as metal additives.

2.2. Water–Gas Shift Activity and Stability

Figure 9 exhibits the %CO conversion of Ni/CeO2, Ni/Ce0.95Pr0.05O1.975, and ReNi/Ce0.95Pr0.05O1.975. From previous studies, it was found that further addition of Re does not further raise the rate of water–gas shift reaction, and the optimal content of Re that is enough to maximize the water–gas shift rate is 1 %wt. [54]. When attention is drawn to the variation in Sm amount [12], it appears that the Ni catalyst with 5%Sm-doped CeO2 gives the highest water–gas shift activity. The enhancement of Sm content to 15 wt.% leads to a lowering of nickel dispersion. This result is due to the agglomeration of samarium at a high amount. Therefore, 5% doping amount and 1% of Re metal additives were used in this work to maximize the WGS performance.
For Ni/CeO2, the CO conversion started above 150 °C and ascended slowly to reach the maximum conversion of 84% at 350 °C. As observed, the highest CO conversion was achieved over a bimetallic NiRe/CePrO catalyst, reaching 96% CO conversion at 330 °C with a WGS activity higher than the activities of the monometallic catalysts. NiRe/Ce0.95Pr0.05O1.975 has been determined as an excellent catalyst due to its high surface acidity, nickel metal dispersion, and nickel surface area which can enhance the concentration of CO adsorption on the catalyst surface. Furthermore, the NiRe/Ce0.95Pr0.05O1.975 catalyst produced more oxygen vacancies, which could increase the redox ability, causing higher WGS activity.
Figure 10 presents the CO2 and CH4 selectivity of Ni/Ce0.95Pr0.05O1.975 and ReNi/Ce0.95Pr0.05O1.975 in the temperature range of 300–500 °C. 1%Re4%Ni/Ce0.95Pr0.05O1.975 was exhibited to be an excellent catalyst in terms of WGS activity and selectivity of CO2 and CH4. Methane is an unwanted product because it is a precursor for coke formation and competes against H2 generation. As observed, 5%Ni/Ce0.95Pr0.05O1.975 generates CH4 at low temperatures, whereas 1%Re4%Ni/Ce0.95Pr0.05O1.975 is highly selective toward the WGS reaction throughout the investigated temperature. Thus, the incorporation of Re onto Ni/Ce0.95Pr0.05O1.975 increased CO conversion at the same time that it suppressed CH4 formation.
The water–gas shift stability has been performed on the most active catalyst, ReNi/Ce0.95Pr0.05O1.975, under the feed mixture containing 5% CO, 10% H2O, and 85% N2 at 300 °C. As shown in Figure 11, the ReNi/Ce0.95Pr0.05O1.975 catalyst retained a high CO conversion of about 89% during the first 20 h on stream. Then, the CO conversion slightly decreases to 82% after 60 h of reaction. Hence, the bimetallic ReNi supported by Pr-doped CeO2 is resistant toward deactivation during a water–gas shift reaction.

3. Experimental Procedure

3.1. Catalysts Preparation

A combustion technique was used to synthesize pure CeO2, and ceria–praseodymia mixed oxides supports with Ce(NO3)3.6H2O (Sigma-Aldrich, Pte. Ltd., Singapore) and Pr(NO3)3.6H2O (Sigma-Aldrich, Pte. Ltd., Singapore) as starting materials. Urea was utilized as a fuel to ignite the reaction. The redox reactions between NH2CONH2 and metal nitrates provide the exothermicity essential for the nucleation and growth of the metal oxide powders [55]. Metal nitrate was mixed with urea using a stoichiometry between urea and metal nitrates as 1:2.5. Stirring a mixture obtained a homogeneous solution and then heating with a Bunsen burner until autoignition occurred. CeO2 and ceria–praseodymia mixed oxides powders were obtained by the thermal decomposition of nitrate and other organic compounds [56].
Ni(NO₃)₂.6H₂O (Alfa Aesar, Thermo Fisher Scientific Inc, Seoul, Republic of Korea) and NH4ReO4 (Sigma-Aldrich, Pte. Ltd., Singapore) were used as the metal precursors for the preparation of Ni/CeO2, Ni/Ce0.95Pr0.05O1.975, and ReNi/Ce0.95Pr0.05O1.975 via impregnation method. A solution of nickel and rhenium was added to ceria and ceria–praseodymia mixed oxides powders. All catalysts were dried at 100 °C for 12 h and calcined at 650 °C for 8 h.

3.2. Catalyst Characterization

The specific surface areas of all catalysts were measured by N2 adsorption–desorption isotherms at 77 K using the BELSORP-MAX instrument (ITS Co. Ltd., Bangkok, Thailand). The samples were outgassed at 300 °C for 3 hours before the analysis. The Brunauer–Emmett –Teller method was utilized to estimate the specific surface areas of the catalysts.
X-ray powder diffraction (XRD) was performed using a PANalytical X’Pert Pro diffractometer (Malvern Panalytical Ltd., Malvern, UK) with the filtered radiation of a copper anode in the range temperature of 20–80°. The X-ray diffractograms were collected using the current of 40 mA and 40 kV with 0.02° per step and 0.5 s per step. The crystallite sizes of CeO2 were estimated from the full width at half maximum of the strongest (111) reflection using the Debye–Sherrer equation.
Raman spectra were collected on Perkin-Elmer System 2000 FTIR/FT-Raman (Perkin Elmer, Rodgau, Germany) with argon ion laser irradiation at 532 nm wavelength and 10 mW maximum power. The spectra were recorded over the range of 100–1000 cm−1 using an operating spectra resolution of 1.0 cm−1 of Raman shift.
The H2 chemisorption, Temperature Programmed Desorption of Ammonia (NH3-TPD), and H2-Temperature Programmed Reduction (H2-TPR) were performed using a catalyst analyzer BELCAT-B instrument (ITS Co. Ltd., Bangkok, Thailand) equipped with a thermal conductivity detector. The reduction behavior of the samples was studied by H2-TPR. The catalyst was first heated from room temperature to 120 °C in the He flow, maintained at 120 °C for 30 min, and cooled down to 50 °C under the He flow. The TPR measurement was performed from 50 °C up to 1000 °C with the rate of 10 °C/min under 5%H2 in argon flow. NH3-TPD analysis was performed to investigate catalyst acidity. The catalyst was first heated from room temperature to 500 °C in argon flow and cooled down to 50 °C under argon. The catalyst was then exposed to pulse titration by using a loop of a known volume of NH3 in Ar flow until saturation. NH3-TPD was finally carried out from 50 to 800 °C with a heating rate of 10 °C/min under argon flow. The H2 chemisorption was performed to determine the surface area, particle size, and dispersion of Ni metal. The sample was evacuated in the He flow at 40 °C and then reduced at 400 °C for 1 h under H2 flow (30 mL/min). The reduced catalyst was cooled down to 40 °C under helium flow and followed by volumetric H2 chemisorption with pure hydrogen. The Ni surface area (SNi), Ni dispersion (D), and Ni particle size (dNi) were obtained from the instrument software based on the calculation by the following equation [57,58].
D (%) = VH × MNi × F/Vm × W × 100
SNi (m2/g) = Vm × NA × F × ANi/Vmolar
dNi = 60 × W/ρ × SNi
where VH is the chemisorbed H2 volume (mL/g), Vm is the molar volume of H2 (mL/mol), W is % wt. of nickel, MNi is the atomic weight of nickel (g/mol), ANi is the cross-sectional area of nickel atom (m2/atom), ρ is the density of Ni (g/mL), NA is Avogadro’s number, and F is the stoichiometry factor (the number of active metal atoms to which one adsorbate gas molecule can attach).
Scanning electron microscopy (SEM) was performed on an FE-SEM (HITACHI SU-8030, Hitachi High-Technologies Corporation, Tokyo, Japan) with high vacuum mode using secondary electrons and an acceleration of 30 kV. Energy dispersive X-ray spectroscopy (EDX) was used in conjunction with scanning electron microscopy for the elemental analysis.
Thermo-gravimetric analysis was performed using a Perkin-Elmer TGA/DTA 6300 instrument (Perkin Elmer, Rodgau, Germany) under an airflow rate of 100 mL/min. The content of carbon deposition on the used catalysts was investigated. The mass change in Ni-based catalysts was measured as a function of temperatures up to 800 °C with a heating rate of 20 °C/min.

3.3. Water–Gas Shift Activity

The water–gas shift activity was measured at the temperature from 100 to 500 °C. The catalyst (150 mg) was placed inside a fixed bed flow reactor (310 stainless steel, 0.6 cm outside diameter) between two layers of quartz wool. The catalyst was reduced under 5% H2 in N2 flow at 300 °C for an hour before the WGS activity testing. H2O was fed through a pre-heater using a syringe pump, whereas the flow rates of CO and N2 were controlled by a mass flow controller. A mixed gas containing 5% CO, 10% H2O, and 85% N2 was fed into the reactor. The total flow rate was maintained at 100 mL/min in all testing conditions. The composition of the gas mixture leaving the reactor was determined using an online Shimadzu GC-14B gas chromatography equipped with a thermal conductivity detector (TCD) and a ShinCarbon ST column. Argon is used as the eluent for a ShinCarbon ST column to detect the H2, CO, and CH4 at the rate of 50 mL/min. The concentration of CO, CO2, and CH4 at the outlet was repeated at least four times for each analysis. The water–gas shift activities can be calculated according to the following equation:
% CO   c o n v e r s i o n = C O i n C O o u t C O i n × 100
where COin is the inlet molar flow rate of CO (mol s−1) and COout is the outlet molar flow rate of CO (mol s−1).

4. Conclusions

The influence of Re and Pr on the catalytic activity of Ni/CeO2 was studied. The incorporation of Re and Pr into Ni/CeO2 increased the WGS efficiency when compared with Ni/CeO2. An addition of Pr to Ni/CeO2 reduced the crystallite size of CeO2, increased the BET surface area, and promoted higher dispersion of nickel on the CeO2 surface. Furthermore, the role of rhenium on the water–gas shift performance of supported Ni catalyst was also considered. The results revealed that the addition of rhenium onto Ni/Ce0.95Pr0.05O1.975 increased the catalytic performance toward the water–gas shift reaction and suppressed CH4 formation. The role of rhenium in improving the catalytic activity was due to an increase in surface acidity, Ni surface area, and Ni dispersion, which facilitate CO adsorption on the catalyst surface. Additionally, the acidic character of the catalyst can accelerate CO2 desorption, leaving behind free active sites for the adsorption of CO and H2O reactants in subsequent reaction cycles. Moreover, the enhancement of oxygen vacancy concentrations alerts the redox processes at the catalyst surface, which contributes to improving the WGS rate.

Author Contributions

Conceptualization, W.S. and P.T.; methodology, W.S. and P.T.; validation, P.T.; formal analysis, W.S. and P.T.; investigation, P.T.; resources, W.S. and P.T.; data curation, P.T.; writing—original draft preparation, W.S. and P.T.; writing—review and editing, P.T.; supervision, P.T.; funding acquisition, W.S. and P.T. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by Thammasat University Research Fund, Contract No. TUFT29/2566 and Thammasat University Research Unit in smart materials from biomass.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, H.; Lei, Q.; Li, P.; Liu, C.; Xue, Y.; Zhang, X.; Li, C.; Yang, Z. Key CO2 capture technology of pure oxygen exhaust gas combustion for syngas-fueled high-temperature fuel cells. Int. J. Coal Sci. Technol. 2021, 8, 383–393. [Google Scholar] [CrossRef]
  2. Molochas, C.; Tsiakaras, P. Carbon Monoxide Tolerant Pt-Based Electrocatalysts for H2-PEMFC Applications: Current Progress and Challenges. Catalysts 2021, 11, 1127. [Google Scholar] [CrossRef]
  3. Kim, K.; Yoo, J.D.; Lee, S.; Bae, M.; Bae, J.; Jung, W.; Han, J.W. A Simple Descriptor to Rapidly Screen CO Oxidation Activity on Rare-Earth Metal-Doped CeO2: From Experiment to First-Principles. ACS Appl. Mater. Interfaces 2017, 9, 15449–15458. [Google Scholar] [CrossRef] [PubMed]
  4. Kayaalp, B.; Lee, S.; Klauke, K.; Seo, J.; Nodari, L.; Kornowski, A.; Jung, W.; Mascotto, S. Template-free mesoporous La0.3Sr0.7Ti1−xFexO3±δ for CH4 and CO oxidation catalysis. Appl. Catal. B Environ. 2019, 245, 536–545. [Google Scholar] [CrossRef]
  5. Montini, T.; Melchionna, M.; Monai, M.; Fornasiero, P. Fundamentals and Catalytic Applications of CeO2-Based Materials. Chem. Rev. 2016, 116, 5987–6041. [Google Scholar] [CrossRef] [PubMed]
  6. Rodriguez, J.A.; Grinter, D.C.; Liu, Z.; Palomino, R.M.; Senanayake, S.D. Ceria-based model catalysts: Fundamental studies on the importance of the metal–ceria interface in CO oxidation, the water–gas shift, CO2 hydrogenation, and methane and alcohol reforming. Chem. Soc. Rev. 2017, 46, 1824–1841. [Google Scholar] [CrossRef] [PubMed]
  7. Atribak, I.; Such-Basáñez, I.; Bueno-López, A.; García, A. Comparison of the catalytic activity of MO2 (M = Ti, Zr, Ce) for soot oxidation under NOx/O2. J. Catal. 2007, 250, 75–84. [Google Scholar] [CrossRef]
  8. 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]
  9. Damyanova, S.; Pawelec, B.; Arishtirova, K.; Huerta, M.M.; Fierro, J. Study of the surface and redox properties of ceria–zirconia oxides. Appl. Catal. A Gen. 2008, 337, 86–96. [Google Scholar] [CrossRef]
  10. Hernández, W.Y.; Centeno, M.A.; Romero-Sarria, F.; Odriozola, J.A. Synthesis and Characterization of Ce1−xEuxO2−x/2 Mixed Oxides and Their Catalytic Activities for CO Oxidation. J. Phys. Chem. C 2009, 113, 5629–5635. [Google Scholar] [CrossRef]
  11. Lomonaco, J.G.; Sesuk, T.; Charojrochkul, S.; Tepamatr, P. Effect of Re Addition on the Water–Gas Shift Activity of Ni Catalyst Supported by Mixed Oxide Materials for H2 Production. Catalysts 2023, 13, 959. [Google Scholar] [CrossRef]
  12. Tojira, O.; Tepamatr, P. Catalytic Activity of Ni Based Materials Prepared by Different Methods for Hydrogen Production via the Water Gas Shift Reaction. Catalysts 2023, 13, 176. [Google Scholar] [CrossRef]
  13. Atribak, I.; Bueno-López, A.; García-García, A. Thermally stable ceria–zirconia catalysts for soot oxidation by O2. Catal. Commun. 2008, 9, 250–255. [Google Scholar] [CrossRef]
  14. Atribak, I.; Bueno-López, A.; García-García, A. Role of yttrium loading in the physico-chemical properties and soot combustion activity of ceria and ceria–zirconia catalysts. J. Mol. Catal. A Chem. 2009, 300, 103–110. [Google Scholar] [CrossRef]
  15. Kırkgeçit, R.; Torun, H.; Dokan, F.K.; Öztürk, E. Optical and electrical conductivity properties of rare earth elements (Sm, Y, La, Er) co-doped CeO2. J. Rare Earths 2021, 40, 1619–1627. [Google Scholar] [CrossRef]
  16. Reddy, B.M.; Saikia, P.; Bharali, P.; Yamada, Y.; Kobayashi, T.; Muhler, M.; Grünert, W. Structural Characterization and Catalytic Activity of Nanosized Ceria-Terbia Solid Solutions. J. Phys. Chem. C 2008, 112, 16393–16399. [Google Scholar] [CrossRef]
  17. Somacescu, S.; Parvulescu, V.; Calderon-Moreno, J.M.; Suh, S.-H.; Osiceanu, P.; Su, B.-L. Uniform nanoparticles building Ce1−xPrxO2−δ mesoarchitectures: Structure, morphology, surface chemistry, and catalytic performance. J. Nanoparticle Res. 2012, 14, 885. [Google Scholar] [CrossRef]
  18. Zhang, G.; Zhao, Z.; Xu, J.; Zheng, J.; Liu, J.; Jiang, G.; Duan, A.; He, H. Comparative study on the preparation, characterization and catalytic performances of 3DOM Ce-based materials for the combustion of diesel soot. Appl. Catal. B Environ. 2011, 107, 302–315. [Google Scholar] [CrossRef]
  19. Rodriguez, J.A.; Hanson, J.C.; Kim, J.-Y.; Liu, G.; Iglesias-Juez, A.; Fernández-García, M. Properties of CeO2 and Ce1−xZrxO2 Nanoparticles: X-ray Absorption Near-Edge Spectroscopy, Density Functional, and Time-Resolved X-ray Diffraction Studies. J. Phys. Chem. B 2003, 107, 3535–3543. [Google Scholar] [CrossRef]
  20. Wang, X.; Hanson, J.C.; Liu, G.; Rodriguez, J.A.; Iglesias-Juez, A.; Fernández-García, M. The behavior of mixed-metal oxides: Physical and chemical properties of bulk Ce1−xTbxO2 and nanoparticles of Ce1−xTbxOy. J. Chem. Phys. 2004, 121, 5434–5444. [Google Scholar] [CrossRef]
  21. Ryan, K.M.; McGrath, J.P.; Farrell, R.A.; O’Neill, W.M.; Barnes, C.J.; Morris, M.A. Morris, Measurements of the lattice constant of ceria when doped with lanthana and praseodymia-the possibility of local defect ordering and the observation of extensive phase separation. J. Phys. Condens. Matter 2003, 15, L49. [Google Scholar] [CrossRef]
  22. Poggio-Fraccari, E.; Irigoyen, B.; Baronetti, G.; Mariño, F. Ce-Pr mixed oxides as active supports for Water-gas Shift reaction: Experimental and density functional theory characterization. Appl. Catal. A Gen. 2014, 485, 123–132. [Google Scholar] [CrossRef]
  23. Poggio-Fraccari, E.; Baronetti, G.; Mariño, F. Pr3+ surface fraction in CePr mixed oxides determined by XPS analysis. J. Electron Spectrosc. Relat. Phenom. 2018, 222, 1–4. [Google Scholar] [CrossRef]
  24. Mariño, F.; Descorme, C.; Duprez, D. Supported base metal catalysts for the preferential oxidation of carbon monoxide in the presence of excess hydrogen (PROX). Appl. Catal. B Environ. 2005, 58, 175–183. [Google Scholar] [CrossRef]
  25. Hung, C.-M. The effect of the calcination temperature on the activity of Cu–La–Ce composite metal catalysts for the catalytic wet oxidation of ammonia solution. Powder Technol. 2009, 191, 21–26. [Google Scholar] [CrossRef]
  26. Chen, Y.-Z.; Liaw, B.-J.; Chen, H.-C. Selective oxidation of CO in excess hydrogen over CuO/CexZr1−xO2 catalysts. Int. J. Hydrogen Energy 2006, 31, 427–435. [Google Scholar] [CrossRef]
  27. Ratnasamy, C.; Wagner, J.P. Water Gas Shift Catalysis. Catal. Rev. 2009, 51, 325–440. [Google Scholar] [CrossRef]
  28. Kim, S.H.; Chung, J.H.; Kim, Y.T.; Han, J.; Yoon, S.P.; Nam, S.W.; Lim, T.-H.; Lee, H.-I. SiO2/Ni and CeO2/Ni catalysts for single-stage water gas shift reaction. Int. J. Hydrogen Energy 2010, 35, 3136–3140. [Google Scholar] [CrossRef]
  29. Hilaire, S.; Wang, X.; Luo, T.; Gorte, R.; Wagner, J. A comparative study of water-gas-shift reaction over ceria supported metallic catalysts. Appl. Catal. A Gen. 2001, 215, 271–278. [Google Scholar] [CrossRef]
  30. Poggio-Fraccari, E.; Mariño, F.; Laborde, M.; Baronetti, G. Copper and nickel catalysts supported on praseodymium-doped ceria (PDC) for the water-gas shift reaction. Appl. Catal. A Gen. 2013, 460–461, 15–20. [Google Scholar] [CrossRef]
  31. Wang, T.; Porosoff, M.D.; Chen, J.G. Effects of oxide supports on the water-gas shift reaction over PtNi bimetallic catalysts: Activity and methanation inhibition. Catal. Today 2014, 233, 61–69. [Google Scholar] [CrossRef]
  32. Wang, D.; Li, Y.; Wen, L.; Xi, J.; Liu, P.; Hansen, T.W.; Li, P. Ni-Pd-incorporated Fe3O4 yolk-shelled nanospheres as efficient magnetically recyclable catalysts for reduction of n-containing unsaturated compounds. Catalysts 2023, 13, 190. [Google Scholar] [CrossRef]
  33. Wen, L.; Wang, D.; Xi, J.; Tian, F.; Liu, P.; Bai, Z.-W. Heterometal modified Fe3O4 hollow nanospheres as efficient catalysts for organic transformations. J. Catal. 2022, 413, 779–785. [Google Scholar] [CrossRef]
  34. Guo, X.; Li, M.; Qiu, L.; Tian, F.; He, L.; Geng, S.; Liu, Y.; Song, Y.; Yang, W.; Yu, Y. Engineering electron redistribution of bimetallic phosphates with CeO2 enables high-performance overall water splitting. Chem. Eng. J. 2023, 453, 139796. [Google Scholar] [CrossRef]
  35. Iida, H.; Igarashi, A. Difference in the reaction behavior between Pt–Re/TiO2 (Rutile) and Pt–Re/ZrO2 catalysts for low-temperature water gas shift reactions. Appl. Catal. A Gen. 2006, 303, 48–55. [Google Scholar] [CrossRef]
  36. Lomonaco, J.G.; Tojira, O.; Charojrochkul, S.; Tepamatr, P. Structure-Activity Relationship of Ceria Based Catalyst for Hydrogen Production. Chiang Mai J. Sci. 2022, 49, 1129–1134. [Google Scholar] [CrossRef]
  37. Tepamatr, P.; Laosiripojana, N.; Charojrochkul, S. Water gas shift reaction over monometallic and bimetallic catalysts supported by mixed oxide materials. Appl. Catal. A Gen. 2016, 523, 255–262. [Google Scholar] [CrossRef]
  38. Tojira, O.; Lomonaco, J.G.; Sesuk, T.; Charojrochkul, S.; Tepamatr, P. Enhancement of hydrogen production using Ni catalysts supported by Gd-doped ceria. Heliyon 2021, 7, e08202. [Google Scholar] [CrossRef]
  39. Saw, E.T.; Oemar, U.; Ang, M.L.; Kus, H.; Kawi, S. High-temperature water gas shift reaction on Ni–Cu/CeO2 catalysts: Effect of ceria nanocrystal size on carboxylate formation. Catal. Sci. Technol. 2016, 6, 5336–5349. [Google Scholar] [CrossRef]
  40. Tsakoumis, N.E.; Walmsley, J.C.; Rønning, M.; van Beek, W.; Rytter, E.; Holmen, A. Evaluation of Reoxidation Thresholds for γ-Al2O3-Supported Cobalt Catalysts under Fischer–Tropsch Synthesis Conditions. J. Am. Chem. Soc. 2017, 139, 3706–3715. [Google Scholar] [CrossRef]
  41. D’angelo, A.M.; Chaffee, A.L. Correlations between Oxygen Uptake and Vacancy Concentration in Pr-Doped CeO2. ACS Omega 2017, 2, 2544–2551. [Google Scholar] [CrossRef] [PubMed]
  42. Taniguchi, T.; Watanabe, T.; Sugiyama, N.; Subramani, A.K.; Wagata, H.; Matsushita, N.; Yoshimura, M. Identifying Defects in Ceria-Based Nanocrystals by UV Resonance Raman Spectroscopy. J. Phys. Chem. C 2009, 113, 19789–19793. [Google Scholar] [CrossRef]
  43. Lee, Y.; He, G.; Akey, A.J.; Si, R.; Flytzani-Stephanopoulos, M.; Herman, I.P. Raman Analysis of Mode Softening in Nanoparticle CeO2−δ and Au-CeO2−δ during CO Oxidation. J. Am. Chem. Soc. 2011, 133, 12952–12955. [Google Scholar] [CrossRef] [PubMed]
  44. Guo, M.; Lu, J.; Wu, Y.; Wang, Y.; Luo, M. UV and Visible Raman Studies of Oxygen Vacancies in Rare-Earth-Doped Ceria. Langmuir 2011, 27, 3872–3877. [Google Scholar] [CrossRef] [PubMed]
  45. Hong, S.J.; Virkar, A.V. Lattice Parameters and Densities of Rare-Earth Oxide Doped Ceria Electrolytes. J. Am. Ceram. Soc. 1995, 78, 433–439. [Google Scholar] [CrossRef]
  46. Guo, J.; Lou, H.; Zhao, H.; Chai, D.; Zheng, X. Dry reforming of methane over nickel catalysts supported on magnesium aluminate spinels. Appl. Catal. A Gen. 2004, 273, 75–82. [Google Scholar] [CrossRef]
  47. Akiki, E.; Akiki, D.; Italiano, C.; Vita, A.; Abbas-Ghaleb, R.; Chlala, D.; Ferrante, G.D.; Laganà, M.; Pino, L.; Specchia, S. Production of hydrogen by methane dry reforming: A study on the effect of cerium and lanthanum on Ni/MgAl2O4 catalyst performance. Int. J. Hydrogen Energy 2020, 45, 21392–21408. [Google Scholar] [CrossRef]
  48. Parastaev, A.; Muravev, V.; Osta, E.H.; van Hoof, A.J.F.; Kimpel, T.F.; Kosinov, N.; Hensen, E.J.M. Boosting CO2 hydrogenation via size-dependent metal–support interactions in cobalt/ceria-based catalysts. Nat. Catal. 2020, 3, 526–533. [Google Scholar] [CrossRef]
  49. Sartoretti, E.; Novara, C.; Chiodoni, A.; Giorgis, F.; Piumetti, M.; Bensaid, S.; Russo, N.; Fino, D. Nanostructured ceria-based catalysts doped with La and Nd: How acid-base sites and redox properties determine the oxidation mechanisms. Catal. Today 2021, 390–391, 117–134. [Google Scholar] [CrossRef]
  50. Sultana, S.; Mansingh, S.; Scurrell, M.; Parida, K.M. Controlled Synthesis of CeO2NS-Au-CdSQDs Ternary Nanoheterostructure: A Promising Visible Light Responsive Photocatalyst for H2 Evolution. Inorg. Chem. 2017, 56, 12297–12307. [Google Scholar] [CrossRef]
  51. Mansingh, S.; Padhi, D.K.; Parida, K.M. Enhanced visible light harnessing and oxygen vacancy promoted N, S co-doped CeO2 nanoparticle: A challenging photocatalyst for Cr(vi) reduction. Catal. Sci. Technol. 2017, 7, 2772–2781. [Google Scholar] [CrossRef]
  52. Wang, L.; Li, D.; Koike, M.; Koso, S.; Nakagawa, Y.; Xu, Y.; Tomishige, K. Catalytic performance and characterization of Ni-Fe catalysts for the steam reforming of tar from biomass pyrolysis to synthesis gas. Appl. Catal. A Gen. 2011, 392, 248–255. [Google Scholar] [CrossRef]
  53. Sahoo, S.K.; Ray, S.S.; Singh, I. Structural characterization of coke on spent hydroprocessing catalysts used for processing of vacuum gas oils. Appl. Catal. A Gen. 2004, 278, 83–91. [Google Scholar] [CrossRef]
  54. Chayakul, K.; Srithanratana, T.; Hengrasmee, S. Effect of Re addition on the activities of Co/CeO2 catalysts for water gas shift reaction. J. Mol. Catal. A Chem. 2011, 340, 39–47. [Google Scholar] [CrossRef]
  55. Takeda, T.; Kato, K.; Kikkawa, S. Gel Combustion Synthesis of Rare Earth Aluminate Using Glycine or Urea. J. Ceram. Soc. Jpn. 2007, 115, 588–591. [Google Scholar] [CrossRef]
  56. Tepamatr, P.; Laosiripojana, N.; Sesuk, T.; Charojrochkul, S. Effect of samarium and praseodymium addition on water gas shift performance of Co/CeO2 catalysts. J. Rare Earths 2020, 38, 1201–1206. [Google Scholar] [CrossRef]
  57. Tuti, S.; Luisetto, I.; Laverdura, U.P.; Marconi, E. Dry Reforming of Methane on Ni/Nanorod-CeO2 Catalysts Prepared by One-Pot Hydrothermal Synthesis: The Effect of Ni Content on Structure, Activity, and Stability. Reactions 2022, 3, 333–351. [Google Scholar] [CrossRef]
  58. Di Michele, A.; Dell’Angelo, A.; Tripodi, A.; Bahadori, E.; Sànchez, F.; Motta, D.; Dimitratos, N.; Rossetti, I.; Ramis, G. Steam reforming of ethanol over Ni/MgAl2O4 catalysts. Int. J. Hydrogen Energy 2018, 44, 952–964. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of Ni catalysts and CeO2 support.
Figure 1. XRD patterns of Ni catalysts and CeO2 support.
Molecules 28 08146 g001
Figure 2. SEM image and elemental mapping of 5%Ni/Ce0.95Pr0.05O1.975.
Figure 2. SEM image and elemental mapping of 5%Ni/Ce0.95Pr0.05O1.975.
Molecules 28 08146 g002
Figure 3. Raman spectra of CeO2 support and Ni catalysts in the wide range of 100–1000 cm−1 (a) and the narrow range of 400–800 cm−1 (b).
Figure 3. Raman spectra of CeO2 support and Ni catalysts in the wide range of 100–1000 cm−1 (a) and the narrow range of 400–800 cm−1 (b).
Molecules 28 08146 g003
Figure 4. H2-TPR profiles of monometallic Ni and bimetallic NiRe supported by pure CeO2 and Pr-doped CeO2.
Figure 4. H2-TPR profiles of monometallic Ni and bimetallic NiRe supported by pure CeO2 and Pr-doped CeO2.
Molecules 28 08146 g004
Figure 5. NH3-TPD of monometallic Ni and bimetallic NiRe supported by Pr-doped ceria.
Figure 5. NH3-TPD of monometallic Ni and bimetallic NiRe supported by Pr-doped ceria.
Molecules 28 08146 g005
Figure 6. XPS spectra of reduced catalysts before the reaction in 5%Ni/Ce0.95Pr0.05O1.975 and 1%Re4%Ni/Ce0.95Pr0.05O1.975 for O 1s (a) and Ni 2p (b).
Figure 6. XPS spectra of reduced catalysts before the reaction in 5%Ni/Ce0.95Pr0.05O1.975 and 1%Re4%Ni/Ce0.95Pr0.05O1.975 for O 1s (a) and Ni 2p (b).
Molecules 28 08146 g006
Figure 7. SEM results of (a) fresh and (b) used 1%Re4%Ni/Ce0.95Pr0.05O1.975 catalysts.
Figure 7. SEM results of (a) fresh and (b) used 1%Re4%Ni/Ce0.95Pr0.05O1.975 catalysts.
Molecules 28 08146 g007
Figure 8. Thermo-gravimetric analysis of the used catalysts.
Figure 8. Thermo-gravimetric analysis of the used catalysts.
Molecules 28 08146 g008
Figure 9. %CO conversion of supported Ni catalysts.
Figure 9. %CO conversion of supported Ni catalysts.
Molecules 28 08146 g009
Figure 10. Selectivity to CO2 and CH4 as a function of temperature over supported Ni catalysts.
Figure 10. Selectivity to CO2 and CH4 as a function of temperature over supported Ni catalysts.
Molecules 28 08146 g010
Figure 11. Long-term stability test at 300 °C of 1%Re4%Ni/Ce0.95Pr0.05O1.975.
Figure 11. Long-term stability test at 300 °C of 1%Re4%Ni/Ce0.95Pr0.05O1.975.
Molecules 28 08146 g011
Table 1. BET surface area, crystallite size of CeO2, Ni dispersion, and Ni surface area of Ni-based catalysts.
Table 1. BET surface area, crystallite size of CeO2, Ni dispersion, and Ni surface area of Ni-based catalysts.
CatalystsCrystallite Size a (nm)BET Surface Area b (m2/g)Ni Dispersion c (%)Ni Particle Size c (nm)Ni Surface Area c (m2/g)
CeO29.868---
5%Ni/CeO213.35450.1735.10.95
5%Ni/CePrO8.55640.3019.61.70
1%Re4%Ni/CePrO8.01601.155.506.08
a Calculated from the 111 diffraction peak broadening. b Estimated from N2 adsorption at −196 °C. c Estimated from H2-chemisorption.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Srichaisiriwech, W.; Tepamatr, P. Monometallic and Bimetallic Catalysts Supported on Praseodymium-Doped Ceria for the Water–Gas Shift Reaction. Molecules 2023, 28, 8146. https://doi.org/10.3390/molecules28248146

AMA Style

Srichaisiriwech W, Tepamatr P. Monometallic and Bimetallic Catalysts Supported on Praseodymium-Doped Ceria for the Water–Gas Shift Reaction. Molecules. 2023; 28(24):8146. https://doi.org/10.3390/molecules28248146

Chicago/Turabian Style

Srichaisiriwech, Weerayut, and Pannipa Tepamatr. 2023. "Monometallic and Bimetallic Catalysts Supported on Praseodymium-Doped Ceria for the Water–Gas Shift Reaction" Molecules 28, no. 24: 8146. https://doi.org/10.3390/molecules28248146

Article Metrics

Back to TopTop