1. Introduction
Fuel cell technology is rapidly developing for both power plant and transportation applications because of its high energy efficiency and ability in converting the chemical energy into electricity through an electrochemical reaction. O2 from air and H2 fuel, which can be produced from gasoline, CH4, or CH3OH, combine electrochemically in a fuel cell to generate electricity without combustion. Steam reforming of CH4 is a common method for hydrogen production. Normally, the reformate gas includes CO, CO2, H2, and H2O. However, the presence of CO in the fuel stream poisons the Pt electrodes, thereby hindering the performance of PEM fuel cell. There has been a renewed interest in the water–gas shift (WGS) reaction in recent years, primarily because of the increasing demand for H2 and the requirement for more efficient and sustainable methods for its generation. The WGS reaction, which converts CO and steam into CO2 and H2, is a crucial step in many industrial processes, including H2 production for fuel cells and other applications.
Ceria acts as a stabilizer for both noble and non-noble metals in various catalytic applications. It achieves this through several mechanisms, including metal–support interactions, oxygen storage capacity, and the generation of defects like oxygen vacancies. CeO
2 loaded with Pt metal is a very good catalyst for WGS reaction. CeO
2-based WGS catalysts are very attractive because of their high oxygen storage capacity of ceria and modified CeO
2 [
1,
2]. Furthermore, CeO
2 material serves as a stabilizer of both noble and non-noble metals and as a promoter of many reactions, including the WGS reaction and CO oxidation. In previous studies, the WGS kinetics over Rh/CeO
2-Al
2O
3 were studied [
3]. The reaction mechanism of this catalyst is a regenerative mechanism in which CO is inserted into the reactive oxygen species of Rh/CeO
2-based materials to give CO
2 and oxygen vacancies, which are then immediately filled by H
2O to give H
2. Similar results over CeO
2-supported Pt, Rh, and Pd [
4] demonstrated the importance of OSC and oxygen mobility in these catalysts. The authors noted that a regenerative mechanism of WGS was controlled by two steps; first, oxygen from CeO
2 is transferred to a metal catalyst at the interface and then the CeO
2 is re-oxidized [
4]. Ample evidence supports a regenerative mechanism in oxidation reactions over ceria-based materials. This mechanism involves the exchange of electrons between the ceria support and the supported metal oxides, resulting in the oxidation of a reactant and then the reduction in the CeO
2.
The introduction of a dopant with a lower oxidation state, such as M
3+, into the Ce
4+ lattice of CeO
2 increases the oxygen vacancies concentration in the CeO
2 structure and improves the thermal stability of the catalysts by hindering the agglomeration of CeO
2-based materials [
5,
6,
7,
8]. Increasing oxygen vacancies in a ceria-based catalyst is a rational strategy to improve the efficiency of the oxidation reaction (such as the water–gas shift reaction). This can be achieved through methods like aliovalent doping with lower-valence cations (e.g., La
3+, Pr
3+, or Sm
3+), which creates a charge imbalance and defects [
5,
6,
7]. The superior redox behavior of Pr- or Sm-doped CeO
2 means that it is reduced at a lower temperature compared to pure CeO
2 and typically reduces more easily than La-doped CeO
2. Since La is fixed in the La
3+ state, the generation of oxygen vacancies relies solely on the reduction in nearby Ce
4+ sites. Whereas a Pr or Sm can cycle between oxidation states, this process is dynamic and creates highly mobile vacancies. Therefore, the presence of Pr or Sm leads to a higher concentration of oxygen vacancies compared to La-doped CeO
2. Furthermore, some studies indicate that while La can improve catalyst activity and stability, Sm and Pr may show even better performance for certain reactions. For example, one study found that Sm doping improved the performance of a CuO-CeO
2 catalyst, while La doping either had no effect or lowered the performance [
9].
In this work, we are interested in examining non-noble metal support on CeO
2-based material as catalysts for the WGS reaction. Recent findings found that ceria-supported Ni catalysts are very active for CO oxidation [
10,
11,
12]. Furthermore, Cu- and Ni-loaded CeO
2 catalysts were studied for the water–gas shift reaction and compared to CuO–ZnO commercial catalysts. It was found that CeO
2-based catalysts present better thermal stability and extend the operating temperature range. Therefore, we have examined the WGS reaction over non-noble metal (Co and Ni) supported on CeO
2 and X-doped CeO
2 (X = Sm and Pr) catalysts to clarify the characteristics of these catalysts. The influence of each metal on the redox behavior and reactivity of CeO
2-based material was compared.
3. Results and Discussion
Figure 1 presents CO conversions over Co- or Ni-containing CeO
2. Pure CeO
2 is not an active catalyst for WGS reaction. Its activity is significantly enhanced when modified with Ni and Co, or when doped with Sm. The CO conversion of all catalysts increases with increasing reaction temperature. 1%Ni/Ce5%SmO is a superior WGS catalyst compared to other catalysts. Notably, almost 90% CO conversion can be achieved at 450 °C over 1%Ni/Ce5%SmO catalyst, and the conversion slightly decreases at temperatures higher than 450 °C. In the case of Co-containing ceria catalysts, the CO conversion of Co on Ce5%SmO support is much higher than that of Co on CeO
2 and Ce5%PrO supports. The 1%Co/Ce5%SmO catalyst shows a maximum CO conversion of 80% at 450 °C, with a subsequent decrease in conversion at higher temperatures.
BET surface areas, average pore size, CeO
2 crystallite sizes, and metal loading of M/CeO
2 and M/doped CeO
2 (M = Ni, Co) are shown in
Table 1. Doping of CeO
2 with Sm or Pr increases the specific surface area, decreases average pore size and reduces the CeO
2 crystallite size of the catalyst. On the other hand, an addition of M (M = Co, Ni) on CeO
2 and doped CeO
2 supports leads to a decrease in BET surface area and an increase in CeO
2 crystallite size. This result is indicated to be due to the growth and agglomeration of CeO
2 particles after calcination at high temperature (650 °C). Sm and Pr ions in M/Sm-doped CeO
2 and M/Pr-doped CeO
2 stabilize the catalyst and hamper the catalyst from sintering by maintaining small particles and high surface area. Furthermore, on BET surface areas of the catalysts after water–gas shift reaction testing up to 600 °C for 10 h, very little loss of surface area took place. The actual metal loadings (Ni, Co, Sm, and Pr) in different ceria-based catalysts were determined using X-ray fluorescence (XRF), and which were all measured to be close to the initial target value used in the preparation.
The N
2 adsorption–desorption isotherms and pore size distribution curves for CeO
2-based catalysts are exhibited in
Figure 2. According to the IUPAC classification, Ni and Co catalysts show type IV isotherms with H
3 type hysteresis loop, featured by a sharp uptake at low pressure which indicates the presence of small pores (micropore), while a large hysteresis loop at higher pressure points to larger pores (mesopore). The rising (adsorption) and falling (desorption) portions of the loop do not coincide, suggesting significant hysteresis and the presence of mesoporous structures [
16]. The pore size distribution curves of Ni catalysts showed a narrow peak centered at 0.6 nm (
Figure 2c), suggesting a high degree of uniformity in the pore diameters. This uniformity and small pore size can enhance catalytic activity by providing a high density of active sites and creating channels that are optimal for the rapid diffusion of reactants and products [
17]. Therefore, the smallest average pore size of the Ni/Ce5%SmO catalyst contributes to increased catalyst stability and decreased sintering.
The X-ray diffraction patterns of calcined catalysts are shown in
Figure 3. The XRD patterns of all catalysts were found to be fluorite-type cubic crystal structures of ceria (JCPDS No. 43-1002). The presence of a very small content of 1% Ni can be difficult to detect. In addition, the absence of Sm
2O
3 or Pr
2O
3 peaks strongly suggests that Sm and Pr have successfully replaced the Ce ions in the CeO
2 crystal structure to form a solid solution [
18,
19]. When doping a CeO
2 material with larger ions (Pr
3+ and Sm
3+ ions), the crystal structure expands to accommodate the larger dopant, and thereby oxygen vacancies are expected to generate in the CeO
2 lattice to keep the charge equilibrium [
20]. This leads to a shift in the XRD peaks towards lower 2θ values.
The crystallite sizes of CeO
2 on the calcined catalysts in
Table 1 were estimated from the (111) plane using Scherrer’s equation. The addition of Sm and Pr into Ni/CeO
2 and Co/CeO
2 causes a decrease in CeO
2 crystallite size together with high surface area, indicating a greater distribution of the catalysts. The absence of Sm and Pr in the preparation of CeO
2 leads to the preferential growth of CeO
2. This is because Sm and Pr influence the crystal structure and growth behavior of CeO
2, potentially leading to different morphologies or the formation of solid solutions with CeO
2.
In essence, maximizing Ni dispersion is essential for optimizing Ni-based catalysts for efficient and selective WGS reactions. The choice of support material influences Ni dispersion and stability. To quantify the Ni dispersion, CO chemisorption pulse was operated on supported Ni and Co catalysts (
Table 2). As expected, the content of CO chemisorbed to the surface enhanced when doping CeO
2 with Sm and Pr. The result presented that a 1%Ni/Ce5%SmO catalyst exhibits the highest Ni dispersion compared to other catalysts. This enhanced dispersion promotes CO adsorption, which is crucial for improving the WGS activity. Furthermore, higher dispersion means smaller metal particles, leading to a larger surface area and more active sites.
H
2-TPR measurements were performed to study the reducibility of the catalysts (
Figure 4). The intensity in arbitrary units indicates the relative amount of consumed hydrogen during the reduction in a catalyst. A higher intensity corresponds to a greater rate of hydrogen consumption. The increased reducibility in doped ceria materials is evidenced by the H
2 consumption, which is summarized in
Table 2. The content of consumed H
2 during the reduction in metal oxides is evaluated from the areas under the TPR peaks corresponding to each metal oxide reduction. The H
2 consumption of Co and Ni catalysts increased in the order of 1%Co/CeO
2 (1.17 mmol/g) < 1%Co/Ce5%PrO (1.35 mmol/g) < 1%Co/Ce5%SmO (1.43 mmol/g) < 1%Ni/CeO
2 (1.93 mmol/g) < 1%Ni/Ce5%SmO (2.15 mmol/g) < 1%Ni/Ce5%PrO (2.19 mmol/g). Accordingly, the reduction temperature and H
2 consumption data demonstrate that Sm addition significantly enhances the reduction in Ni/CeO
2 and Co/CeO
2 by lower reduction temperatures and higher H
2 consumption. This process results in the creation of oxygen vacancies on the catalyst surface which serve as active sites for the dissociation of water molecules, thereby increasing H
2 production [
21]. From H
2-TPR profiles, CeO
2 and doped CeO2 supports show two main characteristics at around 450 °C and 850 °C, ascribed to the reduction in surface CeO
2 and bulk CeO
2, respectively [
22]. The addition of Co and Ni improves the reduction of CeO
2 supports because hydrogen atoms spill over from the metal particles onto the CeO
2 support, facilitating its reduction at lower temperatures. Furthermore, Sm and Pr doping into Ni/CeO
2 and Co/CeO
2 can lower the reduction temperature and enhance the overall redox activity. The Ni/CeO
2 profile exhibits the reduction peak in the range of 300–400 °C, assigned to the reduction in NiO species with varying degrees of interaction with the support [
23]. Similarly to the reduction profile of the bare support, the reduction peak at around 850 °C related to the bulk reduction of CeO
2. Upon the addition of Sm or Pr to Ni/CeO
2 catalysts, the reduction peaks feature of NiO shifts to a lower temperature. This can be attributed to the synergistic interaction between nickel and CeO
2-based supports. This synergistic effect leads to smaller and more evenly dispersed nickel particles [
24]. The fine dispersion inhibits Ni sintering, a process where particles clump together at high temperatures, and enhances the catalyst’s resistance to oxidation during reactions.
The TPR patterns for Co/CeO
2 catalysts exhibited three TPR peaks. The first peak at 350 °C was attributed to the reduction of Co
3+ to Co
2+. The second peak at 400 °C was assigned to the reduction of Co
2+ to metallic Co. The third peak at around 800 °C corresponds to the reduction in bulk CeO
2 [
25]. In contrast, four TPR peaks were observed for the Co/Ce5%SmO and Co/Ce5%PrO catalysts. A peak detected at about 320 °C resulted from the reduction of Co
3+ to Co
2+. In addition, the peaks were observed in the range of 350–500 °C, corresponding to the reduction of Co
2+ interacting with CeO
2 to metallic Co, whereas the peak at about 800 °C corresponds to the reduction in bulk CeO
2. Therefore, the addition of Sm or Pr into Co/CeO
2 shifts the reduction peak to a lower temperature and increases the H
2 consumption of Co catalysts when compared to Co supported by pure CeO
2.
Raman spectroscopy for Ni catalysts was conducted to study oxygen vacancies in the catalyst (
Figure 5). The Raman peak of CeO
2 support at about 460 cm
−1 was attributed to the F
2g vibration for the Ce–O bond in cubic fluorite ceria [
26]. Compared to pure ceria, doping ceria with Sm drastically decreased the F
2g peak intensity and the peak position shifted to lower wavelengths. This result was due to the lattice distortion of ceria which led to the reduction in the Ce-O bond symmetry [
27]. Besides the F
2g peak, the Raman peak was observed at 580 cm
−1 (D peak) corresponding to the vibration induced by defect sites on the ceria structure [
27]. In comparison to Ni/CeO
2 and Ni/Ce5%SmO, the D peak intensity of the 1%Ni/Ce5%SmO catalyst was higher than that of 1%Ni/CeO
2, suggesting that the presence of Sm promoted the oxygen vacancies production on the ceria structure. The ratio of the peak intensities for the D and F
2g peaks serves as an indicator of surface oxygen vacancy concentration. Notably, the estimated I
D/I
F2g values follow a trend of Ni/Ce5%SmO (0.50)∼Ce5%SmO (0.52) > Ni/CeO
2 (0.09) > CeO
2 (0.06). This implies that the addition of Sm into Ni/CeO
2 significantly enhances the number of surface oxygen vacancies.
The WGS reaction converts CO and water to CO
2 and H
2. However, nickel can also catalyze the methanation reaction (3H
2 + CO → CH
4 + H
2O), leading to the generation of CH
4 as a byproduct, which reduces the efficiency of hydrogen production.
Figure 6 shows the selectivity to CO
2 and CH
4 on the Ni/CeO
2 and Ni/Ce5%SmO catalysts. On the Ni/CeO
2 catalyst, CH
4 was the only byproduct. The Ni/Ce5%SmO catalyst presented 100% CO
2 selectivity and 0% CH
4 selectivity at all temperatures studied. This is due to the high oxygen vacancy amount of the Ni/Ce5%SmO catalyst, which disrupts the methanation reaction by increasing the supply of oxygen atoms. The 1%Ni/Ce5%SmO catalyst clearly exhibited higher activity than the Ni/CeO
2 catalyst with no methanation activity, making it a promising catalyst for the WGS reaction. Therefore, introducing Sm to the Ni-based catalyst can enhance its activity for the WGS reaction, potentially suppressing methane formation.
The carbon deposition on the surface of used catalysts after the WGS test was analyzed by temperature programmed oxidation (
Figure 7). The thermal conductivity detector (TCD) responds to changes in the thermal conductivity of the gas exiting the reactor as a function of temperature. The TCD measures the concentration of reaction products by comparing the thermal conductivity of the carrier gas with the gas mixture exiting the heated sample. The oxygen consumption curves represent the carbon oxidation of spent catalysts. Carbon deposition on the used catalysts surface could be amorphous carbon, graphitic carbon, or atomic in nature depending on the oxidation temperature. It has been reported that atomic carbon oxidizes at a temperature below 250 °C, while the oxidation peaks in the temperature ranges of 250–600 °C can be ascribed to the combustion of amorphous carbon. Lastly, the oxidation peak appearing at a temperature above 600 °C is graphitic carbon [
28,
29]. The TPO curves for the spent Ni/CeO
2 and Ni/Ce5%SmO catalysts had higher intensity peaks appearing between 250 and 600 °C, corresponding to the presence of amorphous carbon on these catalysts. In addition, a small peak can be seen at above 600 °C for the spent Ni/CeO
2 catalyst. This indicated that there appeared to be very small deposits of graphitic carbon on the used Ni/CeO
2. The highest amount of oxygen consumed during the oxidation found on the used Ni/CeO
2 catalyst indicates the largest content of carbon deposited on it. More carbon deposition on catalyst surface means more carbon to oxidize, leading to a higher oxygen consumption. Therefore, carbon deposition on catalyst surface which happened during the WGS reaction causes carbon balance to be less than 100%. As can be seen from
Figure 8, the carbon balance did not reach 100% toward the Ni/CeO
2 catalyst for most of the temperatures and decreased gradually with increasing temperature. For the Ni/Ce5%SmO catalyst, the carbon balance is ~100% for all testing in the temperature range of 350–600 °C. The carbon balance in the water–gas shift reaction maintains the principle of mass conservation: carbon is neither created nor destroyed, so the total amount of carbon in the reactants (CO and water) equals the total amount of carbon in the products (CO
2 and any potential byproducts, like methane). The overall reaction is CO + H
2O ⇌ CO
2 + H
2, where all carbon atoms are conserved and converted from CO to CO
2. However, undesirable side reactions can occur, such as the methanation reaction (CO + 3H
2 → CH
4 + H
2O). Therefore, while the ideal carbon balance is achieved by converting CO to CO
2, the actual balance must account for the loss of carbon in methane if it forms.
The CO conversion of the Ni/CeO
2 and Ni/Ce5%SmO catalysts measured with respect to time at a reaction temperature of 400 °C is shown in
Figure 9. The Ni/Ce5%SmO catalyst maintained consistent CO conversion in the whole period of 40 h, demonstrating its good stability. Notably, the Ni/CeO
2 catalyst exhibited signs of deactivation by a decrease in CO conversion around 10% after WGS reaction for 40 h. Generally, carbon deposition is a significant factor in the deactivation of catalysts used in the water–gas shift reaction. The deposited carbon blocks active sites of the catalyst, hindering reactant access and ultimately reducing the catalyst’s activity. From the TPO result of the used catalysts, a large amount of the carbon deposition on the surface was found in the used Ni/CeO
2 catalyst, whereas the excellent resistance to coke formation in the Ni/Ce5%SmO catalyst, which was influenced by higher oxygen vacancy, resulted in stable catalyst performance.
The WGS rates were performed using 0.02 g of catalyst and a feed gas mixture of 10%H
2O, 5%CO, and 85%N
2. The activation energy (Ea) of Ni catalysts was calculated from the slope of the Arrhenius plot in the temperature range of 200–400 °C (
Figure 10) and the estimated parameters are presented in
Table 3. Generally, the apparent Ea values reported from the literature are in the range of 60–80 kJ/mol for transition metals supported on CeO
2 [
30], while the activation energies obtained from this work are in the range of 57–66 kJ/mol. Among the studied catalysts, 1%Ni/Ce5%SmO catalyst exhibited the highest WGS rate at 400 °C and the lowest of the Ea values (57 kJ/mol).
In addition, the calculated TOF values are summarized in
Table 3. The 1%Ni/Ce5%SmO catalyst exhibited a superior performance (in terms of TOF and WGS rate) as compared to other Ni-based catalysts at 400 °C. A TOF value of 0.40 s
−1 and a WGS rate of 253 mmol/kg·s were achieved with the 1%Ni/Ce5%SmO catalyst. Therefore, the 1%Ni/Ce5%SmO catalyst is a promising and highly active catalyst which can be used for the water–gas shift reaction conducted at high temperatures of approximately 400 °C.