Next Article in Journal
Color-Coordinated Photocatalysis of the One-Pot Synthesis of Schiff Bases from Benzyl Alcohol and Nitro Compounds Using a Hybrid Magnetic Catalyst
Next Article in Special Issue
Transition Metal-Promoted LDH-Derived CoCeMgAlO Mixed Oxides as Active Catalysts for Methane Total Oxidation
Previous Article in Journal
Highly Efficient 1-Iodination of Terminal Alkynes Catalyzed by Inorganic or Organic Bases
Previous Article in Special Issue
Facile Synthesis of Sodium Alginate (SA)-Based Quaternary Bio-Nanocomposite (SA@Co-Zn-Ce) for Antioxidant Activity and Photocatalytic Degradation of Reactive Red 24
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Comparative Assessment of First-Row 3d Transition Metals (Ti-Zn) Supported on CeO2 Nanorods for CO2 Hydrogenation

by
Maria Lykaki
1,*,
Sofia Stefa
1,2,
Georgios Varvoutis
3,
Vassilios D. Binas
2,4,
George E. Marnellos
5,6 and
Michalis Konsolakis
1,*
1
School of Production Engineering and Management, Technical University of Crete, 73100 Chania, Greece
2
Institute of Electronic Structure and Laser, Foundation for Research and Technology-Hellas (FORTH–IESL), Vasilika Vouton, 71110 Heraklion, Greece
3
Department of Mechanical Engineering, University of Western Macedonia, 50100 Kozani, Greece
4
School of Chemistry, Faculty of Sciences, Aristotle University of Thessaloniki, 54124 Thessaloniki, Greece
5
Chemical Process & Energy Resources Institute, Centre for Research & Technology Hellas, 57001 Thessaloniki, Greece
6
School of Chemical Engineering, Faculty of Engineering, Aristotle University of Thessaloniki, 54124 Thessaloniki, Greece
*
Authors to whom correspondence should be addressed.
Catalysts 2024, 14(9), 611; https://doi.org/10.3390/catal14090611
Submission received: 7 August 2024 / Revised: 10 September 2024 / Accepted: 10 September 2024 / Published: 11 September 2024

Abstract

:
Herein, motivated by the excellent redox properties of rod-shaped ceria (CeO2-NR), a series of TM/CeO2 catalysts, employing the first-row 3d transition metals (Ti, V, Cr, Mn, Fe, Co, Ni, Cu, and Zn) as active metal phases, were comparatively assessed under identical synthesis and reaction conditions to decipher the role of active metal in the CO2 hydrogenation process. Notably, a volcano-type dependence of CO2 hydrogenation activity/selectivity was disclosed as a function of metal entity revealing a maximum for the Ni-based sample. Ni/CeO2 is extremely active and fully selective to methane (YCH4 = 90.8% at 350 °C), followed by Co/CeO2 (YCH4 = 45.2%), whereas the rest of the metals present an inferior performance. No straightforward relationship was disclosed between the CO2 hydrogenation performance and the textural, structural, and redox properties, whereas, on the other hand, a volcano-shaped trend was established with the relative concentration of oxygen vacancies and partially reduced Ce3+ species. The observed trend is also perfectly aligned with the previously reported volcano-type dependence of atomic hydrogen adsorption energy and CO2 activation as a function of 3d-orbital electron number, revealing the key role of intrinsic electronic features of each metal in conjunction to metal–support interactions.

Graphical Abstract

1. Introduction

Global warming has emerged as a pressing environmental issue due to the observed continuous increase in global temperature of 0.2 °C per decade, which is directly linked to the escalation of greenhouse gas emissions [1,2]. As the concentration of greenhouse gasses in the atmosphere continues to rise, the detrimental impact on the environment is greatly enhanced. Specifically, carbon dioxide (CO2), which originates mainly from fossil fuel energy sources, is a major contributor to the accumulation of greenhouse gasses. To mitigate greenhouse gas emissions, it is imperative to optimize the utilization of fossil fuels and implement effective Carbon Capture and Utilization (CCU) technologies [1,3,4].
A sustainable CCU strategy involves the reduction of captured CO2 via its reaction with hydrogen, a process known as CO2 hydrogenation. This method presents an effective means of valorizing CO2 emissions and efficiently storing surplus power from non-intermittent renewable energy sources (such as solar and wind) in the form of “green” hydrogen [5,6,7]. Coupling a CCU process with a water electrolysis unit producing “green” hydrogen inhibits the release of CO2 emissions, while at the same time converting CO2 into an energy carrier, such as synthetic natural gas (SNG), which consists mainly of methane and can be integrated in the present natural gas grid network, thus closing the carbon cycle [8,9]. Under atmospheric pressure conditions, it can be achieved by either the mildly endothermic reverse water–gas shift (rWGS) reaction (Equation (1)), which yields CO, or the highly exothermic methanation reaction, commonly known as the “Sabatier reaction” (Equation (2)), which produces CH4 [10,11,12].
C O 2 + H 2 C O + H 2 O , Δ H = + 41.2   k J / m o l
C O 2 + 4 H 2 C H 4 + 2 H 2 O , Δ H = 164.7   k J / m o l
Extensive research has explored various catalytic systems, often summarized in the literature covering catalysts for reactions like rWGS [13,14,15,16,17,18] and CO2 methanation [17,19,20,21,22,23]. Notably, composite catalytic systems involving metals supported on reducible metal oxides (e.g., CeO2 and ZrO2) or combinations of them have been extensively studied. Among the investigated oxide materials, ceria (CeO2) has gained significant attention as a supporting carrier due to its exceptional properties such as oxygen storage capacity, oxygen mobility, strong metal–support interactions, and rapid change between its two oxidation states (Ce3+ and Ce4+) [15,24,25,26,27,28]. Although ceria-based noble metals (Ru, Rh, and Pd) have demonstrated excellent catalytic activity, their high cost and limited availability make them less preferable from a techno-economic point of view [29,30,31,32,33,34]. Therefore, recent efforts have focused on the rational design of cost-efficient and highly active non-noble metal catalysts, with particular emphasis on the 3d transition metals, which can adequately adsorb and activate CO2 molecules [18,25,35,36,37,38,39,40,41]. In this direction, we recently showed that by appropriately adjusting the size and shape of ceria carrier via the use of appropriate synthetic routes, extremely active TMs/CeO2 catalysts can be obtained for the CO2 hydrogenation reaction with similar reactivity to that of noble metals [42]. For instance, it was shown that Ni catalysts supported on ceria nanoparticles of rod-like morphology exhibit excellent hydrogenation performance, ascribed mainly to the enhanced redox properties of ceria nanorods in conjunction with the synergistic interactions between nickel and ceria nanoparticles [25,43,44]. On the other hand, Cu/CeO2 samples demonstrate excellent rWGS performance, revealing the key role of the metal entity in CO2 adsorption/activation and in turn in activity and selectivity towards CO or CH4 [12,42].
In the case of the rWGS reaction, two main reaction pathways have been considered, namely the redox and the associative mechanism. In the first case, CO2 is initially activated via its adsorption on metal sites or oxygen vacancies, followed by the formation of carbonyl species that are finally desorbed as CO(g). On the other hand, during the associative mechanism, the adsorbed CO2 reacts with H adatoms towards the formation of active intermediate species (such as formates) that finally decompose to CO and H2O [6,15,16,45,46,47,48]. The main difference between the two rWGS reaction mechanisms is whether the dissociated hydrogen atoms engage in the reaction and generate carbon-containing intermediates (e.g., carboxyls and formates) with the adsorbed CO2 on the catalyst surface [15,49,50]. Similarly, the CO and formate-mediated routes have been proposed for CH4 production. In the first route, the chemisorbed CO derived by CO2 dissociation is further hydrogenated to CH4, whereas, in the second one, formate species (HCOO*) react with hydrogen adatoms [51,52,53]. In view of these aspects, CO2 dissociation is considered a key step in the overall reaction process, affected to a great extent by the electronic nature of active sites and the charge transfer to CO2 molecule [47,51,52].
In view of the above discussion, various transition metals have been studied in the context of CO2 hydrogenation reactions. Remarkably, the first-row transition metals have shown a capacity to adsorb and activate CO2 by facilitating charge transfer from the metal phase to the CO2 molecule [40,54]. Furthermore, computational studies have played a crucial role in enhancing our comprehension of the structural aspects and energetic characteristics of the interplay between metal surfaces and CO2 chemistry. To this end, Ray and Deo [55] have proposed the local d-density of states (d-DOS) at the Fermi level of the pristine surface as a potential descriptor for CO2 methanation over alumina-supported Ni and Ni-based alloy catalysts. In particular, the catalytic activity was linearly correlated with the number of electronic excitations from occupied to unoccupied states at the lowest energy cost [55], with this electronic property demonstrating the surface’s ability to react to an outside perturbation (e.g., a reactant present above the surface) by displaying a population of unoccupied states [56,57]. Density functional theory (DFT) studies reveal that metals with less than 2 empty 3d orbitals exhibit a unique CO2 hydrogenation selectivity, favoring formate formation via a specific co-adsorption configuration of a hydrogen atom and CO2 [58]. In particular, a “volcano” curve on barrier energy in the formate pathway was established with Ni being located at the peak with the lowest barrier energy [58].
Despite the intense interest in the field, there is no—to the best of our knowledge—systematic experimental work on the impact of 3d transition metals on the key activity descriptors and in turn on the CO2 hydrogenation performance under identical synthesis and catalytic evaluation conditions. The latter could bypass the variability of the synthesis procedure and reaction conditions followed in the relevant literature, revealing solely the role of the metal entity. In this regard, a series of TM/CeO2 catalysts with a constant atomic ratio of M/Ce = 0.25 were prepared, using the first-row 3d transition metals (Ti, V, Cr, Mn, Fe, Co, Ni, Cu, and Zn) as active phases for ceria nanorods (CeO2-NR). The catalytic materials were thoroughly characterized in terms of their textural (N2 adsorption at –196 °C), structural (X-ray diffraction, XRD), morphological (Scanning electron microscopy/Energy dispersive spectroscopy, SEM/EDS, Transmission electron microscopy, TEM), surface (X-ray photoelectron spectroscopy, XPS) and redox (Temperature-programmed reduction, H2-TPR) properties in an attempt to gain insight into the role of metal entity in CO2 hydrogenation. Notably, no particular correlation was disclosed between the aforementioned textural, structural, and redox properties and the catalytic performance. However, based on the XPS analysis, a volcano-shaped trend among the relative concentration of oxygen vacancies, the amount of Ce3+ species, and the CO2 conversion rate and CH4 selectivity were established. Moreover, a volcano-type dependence of both activity and selectivity as a function of 3d-orbital electron number was disclosed, with Ni located at the peak of the volcano curve. These findings can be interpreted on the basis of the intrinsic electronic features of each metal in relation to their ability to activate CO2 and dissociate H2, in compliance with relevant theoretical studies.

2. Results

2.1. Textural and Structural Characterization (BET (Brunauer–Emmett–Teller), XRD)

The main textural and structural characteristics of bare CeO2-NR and TM/CeO2 samples (atomic ratio M/Ce = 0.25) are summarized in Table 1. Bare CeO2 demonstrates a BET surface area of 79 m2/g while a slight decrease in the BET surface area is observed upon the addition of the transition metals into ceria support.
XRD patterns of bare CeO2-NR and TM/CeO2 samples are presented in Figure 1. The main diffraction peaks of bare CeO2 correspond to the following planes: (111), (200), (220), (311), (222), (400), (331), and (420) and are attributed to the face-centered cubic (fcc) fluorite structure of ceria (Fm3m symmetry, no. 225) [59,60]. Furthermore, in the case of TM/CeO2 samples (M: Ti, Fe, Co, Ni, Cu, and Zn), smaller peaks of their corresponding oxides are detected, demonstrating the existence of TiO2, Fe2O3, Co3O4, NiO, CuO, and ZnO, respectively, with no other crystal phases existing, apart from ceria. In contrast, Mn/CeO2 and Cr/CeO2 samples do not present any diffraction peaks, except from the ceria crystal phase, a fact that can be attributed to the low metal loadings in conjunction with the high metal dispersion, rendering the calculation of their crystallite sizes impossible. In the XRD pattern of V/CeO2, apart from ceria characteristic peaks, a series of XRD peaks located at 2θ values around 18.2°, 24.0°, 32.4°, 34.2°, 39.0°, 43.5°, 46.4°, 47.9°, 49.2°, 55.5°, and 60.2° were observed, which are related to the (101), (200), (112), (220), (301), (103), (321), (312), (400), (420) and (332) planes of CeVO4 [61,62], respectively. Therefore, it is evident that in the case of V/CeO2, a mixed crystal phase has been formed via the impregnation method, including CeO2 and CeVO4.
The average crystallite sizes of CeO2 and MxOy phases for both bare ceria and TM/Ceria samples are calculated by Scherrer’s equation (Equation (4)) and are presented in Table 1. The crystallite size of ceria does not significantly change with the impregnation of the metal entity (11–15 nm), indicating that ceria structural properties remain unaltered upon metal addition. The crystallite sizes of the active metal phases are following the order: CeVO4 (45 nm) > ZnO (44 nm) > CuO (43 nm) > NiO (23 nm) > TiO2 (20 nm) > Co3O4 (16 nm) > Fe2O3 (7 nm).

2.2. Morphological Characterization (SEM/TEM)-Elemental Analysis (SEM/EDS)

Elemental analysis was performed by SEM along with energy dispersive X-ray spectrometry (SEM/EDS) on the TM/CeO2 samples and the results (atomic ratio of M/Ce, metal content in weight percentage) are presented in Table 1. The atomic ratio of M/Ce is in good agreement with the nominal one. Additionally, the SEM images of the TM/CeO2 samples are depicted in Figure 2. From SEM/EDS analysis, it is revealed that there is a uniform distribution of ceria and metal phases on the entire surface of the mixed oxides.
The morphological characteristics of all samples were investigated in more detail by TEM analysis, as shown in Figure 3. It is apparent that CeO2 support retains its nanorod morphology upon the addition of the metal entity, with isolated MxOy particles of irregular shape being detected in all TM/CeO2 samples. Moreover, the MxOy particle sizes (Table 1), along with the particle size distribution (Figure 3) were determined by TEM analysis. MxOy particle size varied widely, ranging between 10 and 41 nm (Table 1).

2.3. Reducibility Studies (H2-TPR)

In order to gain insight into the impact of the metal entity on the reducibility, TPR measurements were performed in the temperature range of 100–1000 °C, using H2 as a reducing agent. More particularly, indicative catalytic systems (vide infra), in terms of conversion and selectivity, were chosen for the reducibility studies, namely, the Ni/CeO2 and Co/CeO2 catalysts, as highly active for the Sabatier reaction and the Fe/CeO2 and Cu/CeO2 ones, as representatives for the rWGS reaction. As depicted in Figure 4, bare CeO2-NR exhibits two broad reduction peaks centered at 545 °C and 788 °C which are attributed to the surface (Os) and bulk (Ob) oxygen reduction in ceria, respectively [63,64]. The Fe/CeO2 sample exhibits four reduction peaks at 390 °C, 465 °C, 588 °C, and 759 °C with the peaks at 465 °C and 759 °C being ascribed to ceria surface oxygen and bulk oxygen reduction, respectively, while the peaks at 390 °C and 588 °C corresponding to the reduction in the iron species, namely Fe2O3 to Fe3O4 and Fe3O4 to Fe0 [65]. As for the Co/CeO2 sample, it exhibits two main reduction peaks at 318 °C and 388 °C, ascribed to the stepwise reduction of Co3O4 to CoO and CoO to metallic Co, respectively [66]. Concerning the Cu/CeO2 sample, the low-temperature peak at 181 °C is attributed to the reduction of finely dispersed CuOx species strongly interacting with the ceria surface, while the peak at 217 °C can be ascribed to the reduction of Cu2O to Cu0 due to the interaction of the Cu–[Ox]–Ce structure [67,68,69]. Nickel–ceria nanorods exhibit three peaks; the peak at 220 °C can be attributed to the reduction of surface oxygen located at the defects of the ceria surface, the peak at 288 °C corresponds to the reduction of adsorbed oxygen species associated with the formation of Ni–O–Ce structure, while the peak at 353 °C is ascribed to the reduction of the well-dispersed NiO phase interacting strongly with the ceria support [70,71]. All TM/CeO2 samples exhibit a high-temperature peak in the range of 747–793 °C which is attributed to the ceria sub-surface oxygen reduction. As illustrated in Figure 4, the addition of the metal phase facilitates ceria surface oxygen reduction.
Furthermore, as depicted in Table 2, hydrogen consumption surpasses the theoretical amount for all TM/CeO2 samples, apart from the Fe/CeO2 sample, revealing that the addition of the metallic phase enhances the reducibility of ceria nanorods, due to synergistic metal–support interactions.

2.4. Surface Analysis (XPS)

In order to assess the impact of surface chemistry on the intrinsic properties and catalytic performance, XPS analysis was conducted over the optimum catalytic materials (TM/CeO2, TM: Fe, Cu, Co, Ni) in their reduced states. In Figure 5a, the Ce 3d XPS spectra are depicted with the Ce 3d3/2 and Ce 3d5/2 states corresponding to the u and v lines, respectively. The three pairs of peaks u (902.1 eV), u″ (908.5 eV), u‴ (917.3 eV) and v (883.2 eV), v″ (889.8 eV), v‴ (899.5 eV) characterize the Ce4+ species, while the u′ (904.7 eV) and v′ (886.2 eV) peaks are attributed to the Ce3+ species [72,73,74].
The concentration of Ce3+ species can be calculated from the following equation (Equation (3)) [72]:
C e 3 + = C e I I I C e I I I + C e I V
where Ce(III) and Ce(IV) denote the corresponding sums of the integrated peak areas associated with the Ce3+ and Ce4+ XPS signals, respectively (Table 3). Ni catalyst exhibits the highest amount of Ce3+ species (30.4%), followed by Co (29.9%), Fe (26.9%) and Cu (25.5%). The significant role of Ce3+ species in the CO2 hydrogenation process (activation of CO2 and formation of active intermediates) has been revealed in the literature [75,76,77], with these sites being associated with the abundance of oxygen vacancies and the extent of metal–support interactions [78]. Cao et al. [79] demonstrated a volcano-type curve between activity and surface Ce3+ fractions for CO2 hydrogenation to CO over ceria catalysts. Moreover, under CO2 hydrogenation reaction conditions, a Ni catalyst supported on mesoporous ceria nanofiber assembly was found to lead to the formation of active Ce3+ sites located at the interface, thus boosting the catalytic performance and promoting the formate pathways [80].
Figure 5b illustrates the O 1s XPS spectra with two peaks being observed; the peak at 529 eV represents the lattice oxygen (Olat), while the peak at 531 eV is ascribed to surface adsorbed oxygen species (Oads) [81,82]. Analysis of the O 1s XPS spectra of the reduced samples led to the determination of the Oads/Olat ratio which constitutes a useful tool for the estimation of the relative oxygen vacancy concentration [54,81]. As shown in Table 3, the Ni catalyst presents the highest Oads/Olat ratio, with the following order being observed: Ni/CeO2 > Co/CeO2 > Cu/CeO2 > Fe/CeO2, in good agreement with the order of Ce3+ species (Table 3).
Figure 6a presents the Cu 2p XPS spectrum of the reduced Cu/CeO2 catalyst which includes two main peaks at 953.7 eV and 933.8 eV, ascribed to Cu 2p1/2 and Cu 2p3/2 states, respectively. The existence of satellite peaks at 943 and 962 eV confirms the presence of Cu2+ species [69,83]. The Co 2p XPS spectrum for the reduced Co/CeO2 sample (Figure 6b) exhibits two main peaks of Co 2p1/2 (796.6 eV) and Co 2p3/2 (780.6 eV), as well as shake-up satellites at 786 eV which are characteristics of Co2+ [84,85]. Moreover, Figure 6c depicts the Fe 2p XPS spectrum of the reduced Fe/CeO2 catalyst. There are two peaks at 711 and 724 eV, assigned to the Fe 2p3/2 and Fe 2p1/2, respectively, and two satellite peaks at 719 and 734 eV, designating the presence of Fe3+ [86,87]. The satellite peaks overlap with the Auger Ce MNN peaks. The formation of Fe2+ species results from the iron oxide–ceria interaction via an interfacial redox process: xFe2O3 + (2 − y)CeO2−x → xFe2O3 − y + (2 − y)CeO2 [86,88,89]. In addition, Figure 6d illustrates the Ni 2p XPS spectrum of the reduced Ni/CeO2 sample, with two peaks at 855 eV (Ni 2p3/2) and at 873 eV (Ni 2p1/2), along with a broad satellite peak at 861 eV, associated with the presence of Ni3+ species (Ni2O3) [90,91]. According to the XPS results of the reduced catalysts, metal species exist in different oxidation states which can be related to the strength of the metal–support interactions, as indicated in the literature [85,90]. Nevertheless, the difficulty in establishing reliable relationships between metal oxidation states and catalytic performance should be underscored, on the ground of ex situ nature of XPS in conjunction with different metal entities employed as active phases in the present work.

2.5. Catalytic Performance

Figure 7 depicts the CO2 conversion and CO selectivity performance in the temperature range of 200–500 °C of all TM/CeO2 samples. For comparison purposes, the corresponding performance of bare ceria nanorods is also included. The main conclusions that can be drawn by the comparative assessment of the first-row 3d transition metals supported on CeO2 nanorods are the following:
(i)
The CO2 conversion of TM/CeO2 is strongly dependent on the metal entity, following the general trend: Ni > Co > Cu > Fe > Zn > Cr ≈ Ti ≈ V ≈ Mn.
(ii)
The Ni/CeO2 catalyst exhibits by far the best performance overall, in terms of maximum conversion and light-off temperature, offering ~90% conversion at 300 °C, which is close to equilibrium conversion (94%) under the employed experimental conditions. Equally importantly, Cu/CeO2 was the most active rWGS material, attaining a CO2-to-CO conversion of 47% at 450 °C, approaching equilibrium conversion (48.5%).
(iii)
The early 3d transition metals (Ti, V, Cr, Mn) are almost inactive (i.e., XCO2 < 10% at T < 400 °C), exhibiting an even lower conversion performance as compared to bare ceria nanorods.
(iv)
All TM/CeO2 systems, except for Ni and Co, are mainly selective to CO. In particular, the Ni-based catalyst is completely selective to CH4 in the whole temperature range investigated, whereas the Co-based sample exhibits intermediate selectivity values depending on temperature. All other TM/CeO2 catalysts are completely selective to CO.
(v)
In terms of CO2 conversion and CO/CH4 selectivity, no catalyst deactivation occurred during short-term (24 h) stability experiments over the optimum catalysts (Fe, Co, Ni, Cu) supported on ceria nanorods (not shown for brevity).
In order to gain a more thorough insight into the particular role of the metal entity in CO2 hydrogenation performance, the normalized reaction rate of CO2 consumption (rCO2) and the corresponding selectivity to methane (SCH4) are plotted at a specific reaction temperature (350 °C), as at this temperature noticeable differences can be unveiled among the samples (Figure 8). Markedly, a volcano-type dependence of CO2 conversion and CH4 selectivity was disclosed for first-row 3d transition metals, i.e., Ti, V, Cr, Mn, Fe, Co, Ni, Cu, Zn with a maximum obtained for Ni. In particular, the Ni/CeO2 catalyst exhibits by far the optimum hydrogenation performance in terms of CO2 consumption rate (~800 μmol gNi−1s−1) and CH4 selectivity (100%), followed by Co (rCO2 = 445 μmol gCo−1s−1, SCH4 = 90%). The 3d early transition metals (Ti–Fe), as well as the 3d late metals (Cu and Zn), exhibit much lower reactivity (rCO2 < 200 μmol gM−1s−1), while being selective to CO (Figure 8).

3. Discussion

In the present study, the CO2 hydrogenation performance of first-row 3d transition metals (Ti–Zn) supported on CeO2 nanorods (atomic ratio M/Ce = 0.25) were comparatively assessed employing identical synthesis and catalytic evaluation procedures, in an attempt to decipher the role of metal entity and metal–support interactions in the catalytic activity and selectivity. The as-prepared samples were characterized by various physicochemical methods to gain insight into the impact of metal nature on the structural, textural, morphological, and redox properties. On the basis of the obtained results, it could be argued that Ni/CeO2—and to a lower extent Co/CeO2—is active for the Sabatier reaction, whereas the rest of the transition metals exhibit a much inferior reactivity while being selective to CO (rWGS). Remarkably, a volcano-type dependence of CO2 hydrogenation activity and selectivity was disclosed as a function of metal entity (Ti–Zn) revealing a maximum for the Ni-based sample. In particular, Ni/CeO2 is highly active and fully selective to CH4 (YCH4 = 90.8% at 350 °C), followed by Co (YCH4 = 45.2% at 350 °C). The rest of the transition metals exhibit a much lower CO2 conversion, being also selective to CO instead of CH4. In an attempt to disclose possible structure-performance relationships and to reveal key activity descriptors, the obtained catalytic findings are discussed below in relation to the present characterization results and the intrinsic characteristics of each metal, taking into account the main reaction pathways suggested for Sabatier and rWGS reactions.
Firstly, the rWGS and Sabatier reaction mechanisms are considered, both constituting a topic of intensive debate. For the rWGS reaction, two main pathways have been in general considered, i.e., the redox and the associative mechanism, depending on the direct or not participation of H2 in CO2 activation [47,48]. In the redox mechanism, the gaseous CO2 is initially activated via its adsorption on an electron donor site, such as a metal or oxygen vacancy, followed by the formation of carbonyl species that are finally desorbed as CO(g). In this specific route, the role of H2 is limited to reducing the oxidized sites and sustaining the reduction/oxidation cycle on the catalyst surface [6,49]. Conversely, within the associative mechanism, CO2 chemisorbed species react with H adatoms derived from H2 dissociation for the formation of intermediate species (such as formates) that eventually decompose to CO and H2O [6,13,41].
Similarly, two main paths have been proposed for CO2 methanation, i.e., the CO and formate route. The former is related to the dissociation of CO2(g) to CO(ad) which is fully hydrogenated to CH4(ad) [51,52]. In the second route, formate species (HCOO*) instead of CO is the main intermediate, followed by a reaction with chemisorbed hydrogen adatoms [51,52,92]. With regard to the role of metal–ceria catalysts in the mechanism of hydrogen activation, it is considered that gas-phase H2 is activated via dissociation on metal particles that are increasingly active towards hydrogen activation in the vicinity of ceria surfaces. The adsorbed hydrogen adatoms rapidly migrate to the ceria support via spillover phenomena and can react with low-energy and abundant C-containing intermediates onto CeO2–x and result in gaseous methane molecules [93]. In this regard, in situ DRIFT experiments have revealed the involvement of the direct formate hydrogenation pathway in CO2 methanation over a nickel–ceria catalyst prepared via sol–gel, with this synthetic procedure promoting oxygen vacancy formation and exposure of the Ni (111) crystal plane [94].
In principle, the CO2 activation pathway is strongly affected by the metal entity and metal–oxygen interaction [95]. In this regard, it has also been found that the adsorption strength of CO2 on metal surfaces is affected by the electronic nature of active sites, i.e., the d-band center of the metal as well as the charge transfer to a CO2 molecule [96]. According to the d-band center model, the adsorption energy of a given molecule can uniformly decrease (increase) from one transition metal surface to another in which the d-electron number increases (decreases). Nevertheless, exceptions from the model can occur possibly due to the adsorbate’s large electronegativity and/or the substrate’s nearly full d-band [97].
Moreover, it is considered that M–CO bond strength determines the selectivity of CO2 hydrogenation to either CO or CH4; weak interaction results in the desorption of CO as a final product, whereas strong interaction facilitates its further hydrogenation to CH4. In this regard, it has been found via computational studies that the CO adsorption energy can be considered as a selectivity descriptor [6]. In specific, metal atoms with high CO desorption energy (such as Pd, Pt, Ni, and Rh) tend to mainly catalyze the methanation reaction, whereas group 11 metals (Cu, Ag, Au) with lower CO desorption energy produce CO as the main product. These findings are in complete agreement with the present ones (Figure 8), revealing the key role of metal nature in CO2 activation.
In light of the above discussion, CO2 activation in conjunction with H2 dissociation on metal sites could be considered as the main factors determining activity and selectivity. In view of this fact, in a comprehensive work, Sun et al. [58] theoretically explored 3d transition metals in relation to their activity and selectivity in CO2 hydrogenation by taking into account the adsorption configurations of H2, CO2, and the co-adsorption of H atom and CO2 on metal sites. In Figure 9, adopted from the aforementioned article, the atomic hydrogen adsorption energy (Figure 9a), the barrier energy (Figure 9b), and the CO2 charge change (Figure 9c) are depicted as functions of 3d-orbital electron transfer during the CO2 hydrogenation.
It is evident that there is a volcano-type dependence of barrier energy in relation to the number of 3d-orbital electrons, with Ni located at the peak with the lowest energy. This implies the superiority of Ni for CO2 hydrogenation among the other 3d metals. Moreover, the change in charge from IS (initial state) to TS (transition states) on adsorbed CO2 also exhibits a volcano-shaped relationship in agreement with the barrier energy, confirming that the barrier energy is closely correlated to electron transfer. Besides, the adsorption energy of H atoms is perfectly correlated with the barrier energy of CO2 hydrogenation, being again the lowest for Ni [58].
Notably, a perfect agreement between the theoretically predicted trend for CO2 hydrogenation via the formate pathway (Figure 9) and the present experimental findings over supported catalysts (Figure 8) was revealed for the first time, implying the key role of the energy barrier for CO2 activation and H2 dissociation. In other words, the intrinsic characteristics of each metal related to the d-band center and the charge transfer to CO2 or H2 molecules can be considered as the main factors, affecting the activation of CO2 and its consequent hydrogenation to methane or other products.
The above is further supported by the lack of straightforward relationships between the structural and redox properties of TM/CeO2 samples explored in the present work. In specific, the majority of TM/CeO2 samples exhibit similar textural and structural properties (Table 1), despite their widely separated catalytic performance. Moreover, both the metal and ceria crystallite sizes are not affected, to a great extent, by metal entities, suggesting similar interfacial interactions. For instance, although Ni, Co, and Cu possess similar surface areas (72–75 m2/g) and crystallite sizes for ceria support (12–14 nm) and metal active phase (10–16 nm), they exhibit completely different performance, in terms of activity and selectivity. Similarly, taking into account the present H2-TPR results, it can be inferred that the reducibility cannot be considered by itself as the key activity/selectivity descriptor during the CO2 hydrogenation process. In other words, the differences in the redox behavior, in terms of H2 uptake and TPR peak temperatures (Table 2), cannot be solely accounted for the dissimilar hydrogenation performance of explored samples.
However, taking into account the XPS results, it is noteworthy that a volcano-shaped trend was revealed among the oxygen vacancy and Ce3+ species concentrations and the CO2 conversion rate and CH4 selectivity at 350 °C for the optimum rod-shaped reduced catalysts (Fe, Cu, Co, and Ni), following the order: Fe/CeO2 < Cu/CeO2 < Co/CeO2 < Ni/CeO2 (Figure 10).
These results clearly demonstrate the fundamental role of the metal entity and in turn of metal–support interactions towards determining the relative abundance of oxygen vacancies and partially reduced cerium species, both essential for CO2 activation. In particular, Lee et al. [98] proposed that Ce3+ sites with oxygen vacancies over nickel–ceria catalyst led to the activation of CO2 towards the formation of carbonate species that are hydrogenated to formate and finally to methane. In a similar manner, nickel–ceria nanorods exhibited higher methanation activity than their nanocubes counterparts in the low-temperature range (200–250 °C) due to their abundance in oxygen vacancies which have a direct effect on the formation of active intermediates [36]. The {110} facets mainly exposed on ceria nanorods have the ability to generate and stabilize the oxygen vacancy sites which are essential in CO2 to CH4 conversion [99]. Nevertheless, the pivotal role of metal–support interactions in catalytic activity should be underscored, as the experimental results of the present study clearly reveal the significance of the metal entity and its interaction with the rod-shaped ceria support in generating oxygen vacancies and Ce3+ species. In a similar manner, it was demonstrated that oxygen vacancies and moderate basic sites function as contributing factors in CO2 activation, with the Ni metal having a pivotal role in H2 activation and dissociation [100].
Collectively, on the basis of the present results, it can be inferred that Ni is the best metal for methanation reaction, due to its specific d-band center and its ability to activate the CO2 and H2 molecules. In this regard, any attempt to optimize the CO2 hydrogenation performance of Ni-based samples should be focused on the selection of appropriate supporting carriers, and promoters, and fine-tuning engineering approaches towards optimizing the ability of Ni sites (via metal–support interactions) to promote the aforementioned process. In this regard, we recently showed that a compromise between the extent of Ni–ceria perimeter and the competitive presence of larger Ni particles is a prerequisite for maximizing the CO2 hydrogenation to methane [43]. The latter was ascribed, on the ground of a structure sensitivity analysis, to the presence of highly active edge and kink sites instead of inactive terrace sites. Hence, on the basis of the aforementioned discussion, it could be argued that CO2 activation could be more favorably promoted on under-coordination sites with an “optimum” d-band center and charge transfer ability.

4. Materials and Methods

4.1. Materials Synthesis

The chemicals used in the present work were Ce(NO3)3·6H2O (Acros Organics, Geel, Belgium, purity 99.5%), tetrabutyl titanate (TBOT, Sigma-Aldrich, St. Louis, MO, USA, purity ≥ 97%), NH4VO3 (Supelco, Merck, Darmstadt, Germany), Cr(NO3)3·9H2O (Sigma-Aldrich, St. Louis, MO, USA, purity 99%), Mn(NO3)2·4H2O (Sigma-Aldrich, St. Louis, MO, USA, purity ≥ 97.0%), Fe(NO3)3·9H2O (Sigma-Aldrich, St. Louis, MO, USA, purity ≥ 98%), Co(NO3)2·6H2O (Sigma-Aldrich, St. Louis, MO, USA, purity ≥ 98%), Ni(NO3)2·6H2O (Alfa Aesar, ThermoFisher, Kandel, Germany, purity 98%), Cu(NO3)2·3H2O (Sigma-Aldrich, St. Louis, MO, USA, purity 99-104%), Zn(CH3COO)2·2H2O (Sigma-Aldrich, St. Louis, MO, USA, purity ≥ 99%) NaOH (Honeywell Fluka, Seelze, Germany, purity ≥ 98%) and ethanol (Supelco, Merck, Darmstadt, Germany).
Bare ceria nanorods (CeO2-NR) were hydrothermally synthesized, as thoroughly described in our previous work [101]. Ceria-based transition metal samples (TM/CeO2, TM: Ti, V, Cr, Mn, Fe, Co, Ni, Cu, and Zn) were synthesized by the wet impregnation method, using aqueous solutions of metal nitrate precursors to obtain a nominal metal/cerium atomic ratio of 0.25. The obtained suspensions were heated under stirring until water evaporation, dried at 90 ºC for 12 h, and finally calcined at 500 °C for 2 h (heating ramp 5 °C/min) [44].

4.2. Materials Characterization

The textural properties of samples were assessed by N2 adsorption–desorption isotherms at −196 °C (Nova 2200e Quantachrome flow apparatus, Boynton Beach, FL, USA). X-ray diffraction patterns of as-prepared samples were obtained by a Rigaku diffractometer (model RINT 2000, Tokyo, Japan). The crystallite size of the samples was calculated by applying Scherrer’s equation (Equation (4)):
D X R D n m = Κ λ c o s θ
where λ is the X-ray wavelength in nm, K is Scherrer’s constant, B is the line broadening and θ is the Bragg angle. By using Scherrer’s equation (Equation (4)), on the most intense diffraction peak, i.e., for CeO2 (2θ: 28.5°), the primary particle size of this crystal phase was determined. Morphological/elemental analysis was carried out by scanning electron microscopy (SEM, JEOL JSM-6390LV, JEOL Ltd., Akishima, Tokyo, Japan) operating at 20 keV, equipped with an energy dispersive X-ray spectrometry (EDS) system and transmission electron microscopy (TEM) on a JEM-2100 instrument (JEOL, Tokyo, Japan). Temperature Programmed Reduction (H2-TPR) experiments were carried out in a fully automated AMI-200 Catalyst Characterization Instrument (Altamira Instruments, Pittsburgh, PA, USA) under an H2 atmosphere, in order to assess the reducibility of the samples. XPS analysis was conducted on Kratos AXIS Ultra HSA, Manchester, United Kingdom with VISION software (http://www.casaxps.com/kratos/, accessed on 10 September 2024) for data acquisition and CASAXPS software (http://www.casaxps.com/, accessed on 10 September 2024) for data analysis.

4.3. Catalytic Evaluation Studies

Catalytic studies were carried out in a quartz fixed-bed U-shaped reactor (i.d. = 1 cm), loaded with 200 mg of catalyst diluted with 200 mg of inert SiO2. The sample’s pre-treatment included in situ reduction at 400 °C for 1 h under pure H2 flow (40 cm3 min−1), followed by flushing with He (10 cm3 min−1) until room temperature. The experiments were conducted at 1 bar and in the temperature range of 200–500 °C at intervals of 15–20 °C and a heating rate of 1 °C/min. The total volumetric feed flow was 100 cm3 min−1, corresponding to a weight hourly space velocity (WHSV) of 30 L·g−1·h−1. The gas feed constituted a pure H2/CO2 mixture at a molar ratio of 80/20. Calculations for the thermodynamic equilibrium were derived using the mathematical model RGibbs in Aspen Plus software (https://www.aspentech.com/en/products/engineering/aspen-plus, accessed on 10 September 2024. Aspen Technology, Inc., Bedford, MA, USA). The analysis of the reactor outlet mixture was performed in a gas chromatograph (Shimadzu GC-14B, Shimadzu, Kyoto, Japan) with He as the carrier gas, equipped with a thermal conductivity detector (TCD) for the detection of CO and CO2, a flame ionization detector (FID) for monitoring CH4 and two separation columns (Molecular Sieve 13X, Restek GmbH, Hesse, Germany and Porapak QS, Restek GmbH, Hesse, Germany).
The reactor effluent was passed through a cold trap submerged in an ice bath in order to condensate the H2O produced by the reactions. The only carbonaceous products detected in the reactor outlet stream were CH4 and CO. The conversion of carbon dioxide (ΧCO2), product selectivities (SCO and SCH4), and yields (YCO and YCH4) were calculated as follows (Equations (5)–(9)), where [i] stands for the concentration of the respective gas at the outlet of the reactor:
X C O 2 = 100 × C O + [ C H 4 ] [ C O 2 ] + C O + [ C H 4 ]
S C O = 100 × C O C O + [ C H 4 ]
S C H 4 = 100 × [ C H 4 ] C O + [ C H 4 ]
Y C O = X C O 2 × S C O
Y C H 4 = X C H 4 × S C H 4
The macroscopic reaction rates were defined in terms of the rate of moles of CO2 consumed per mass of the catalyst, rm;
r m m o l   C O 2 · g 1 · s 1 = [ C O 2 ] i n × F i n × X C O 2 100 × 60 × m c a t × V m
where mcat is the mass of the catalyst in grams, and Vm is the gas molar volume at 25 °C and 1 bar (24,436 cm3/mol).

5. Conclusions

In the present study, a series of TM/CeO2 catalysts (atomic ratio M/Ce = 0.25) employing the first-row 3d transition metals (Ti, V, Cr, Mn, Fe, Co, Ni, Cu, and Zn) as active metals were catalytically evaluated in CO2 hydrogenation. The superiority of the Ni/CeO2 sample was clearly revealed in terms of CO2 conversion rate and CH4 selectivity, followed by Co/CeO2. The 3d early transition metals (Ti–Fe) and the 3d late metals (Cu and Zn) exhibit much lower activity while being selective to CO. The significance of the metal entity in the CO2 hydrogenation performance was clearly established via a volcano-type dependence of both activity and selectivity to methane as a function of 3d-orbital electron number, with Ni located at the top of this volcano curve. Moreover, a distinct volcano-shaped trend among the relative abundance of oxygen vacancies and Ce3+ species with the CO2 conversion and CH4 selectivity was disclosed. The present experimental findings agree perfectly with the theoretically predicted trend for CO2 hydrogenation over 3d transition metals as a function of the energy barrier for CO2 activation and H2 dissociation, implying the key role of intrinsic electronic properties of metal entities in conjunction with metal–support interactions rather than of textural/structural characteristics.

Author Contributions

Data curation, formal analysis, investigation, validation, writing—original draft preparation, M.L.; investigation, validation, writing—original draft preparation, S.S.; data curation, investigation, G.V., V.D.B. and G.E.M.; conceptualization, data curation, formal analysis, funding acquisition, project administration, supervision, writing—review and editing, M.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research has been co-financed by the European Union NextGenerationEU under the call RESEARCH–CREATE–INNOVATE 16971 Recovery and Resilience Facility (project code: TAEDK-06169).

Data Availability Statement

The authors confirm that the data supporting the findings of this study are available within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Atsbha, T.A.; Yoon, T.; Seongho, P.; Lee, C.-J. A Review on the Catalytic Conversion of CO2 using H2 for Synthesis of CO, Methanol, and Hydrocarbons. J. CO2 Util. 2021, 44, 101413. [Google Scholar] [CrossRef]
  2. Hansen, J.; Sato, M.; Ruedy, R.; Lo, K.; Lea, D.W.; Medina-Elizade, M. Global temperature change. Proc. Natl. Acad. Sci. USA 2006, 103, 14288–14293. [Google Scholar] [CrossRef] [PubMed]
  3. Meylan, F.D.; Moreau, V.; Erkman, S. CO2 Utilization in the Perspective of Industrial Ecology, an Overview. J. CO2 Util. 2015, 12, 101–108. [Google Scholar] [CrossRef]
  4. Wang, H.; Liu, Y.; Laaksonen, A.; Krook-Riekkola, A.; Yang, Z.; Lu, X.; Ji, X. Carbon Recycling—An Immense Resource and Key to a Smart Climate Engineering: A Survey of Technologies, Cost and Impurity Impact. Renew. Sustain. Energy Rev. 2020, 131, 110010. [Google Scholar] [CrossRef]
  5. Saeidi, S.; Amin, N.A.S.; Rahimpour, M.R. Hydrogenation of CO2 to Value-Added Products—A Review and Potential Future Developments. J. CO2 Util. 2014, 5, 66–81. [Google Scholar] [CrossRef]
  6. Bahmanpour, A.M.; Signorile, M.; Kröcher, O. Recent Progress in Syngas Production via Catalytic CO2 Hydrogenation Reaction. Appl. Catal. B Environ. 2021, 295, 120319. [Google Scholar] [CrossRef]
  7. Garba, M.D.; Usman, M.; Khan, S.; Shehzad, F.; Galadima, A.; Ehsan, M.F.; Ghanem, A.S.; Humayun, M. CO2 towards Fuels: A Review of Catalytic Conversion of Carbon Dioxide to Hydrocarbons. J. Environ. Chem. Eng. 2021, 9, 104756. [Google Scholar] [CrossRef]
  8. Torres-Sempere, G.; Pastor-Perez, L.; Odriozola, J.A.; Yu, J.; Duran-Olivencia, F.J.; Bobadilla, L.F.; Reina, T.R. Recent Advances on Gas-Phase CO2 Conversion: Catalysis Design and Chemical Processes to Close the Carbon Cycle. Curr. Opin. Green Sustain. Chem. 2022, 36, 100647. [Google Scholar] [CrossRef]
  9. Rafiee, A.; Rajab Khalilpour, K.; Milani, D.; Panahi, M. Trends in CO2 Conversion and Utilization: A Review from Process Systems Perspective. J. Environ. Chem. Eng. 2018, 6, 5771–5794. [Google Scholar] [CrossRef]
  10. Romeo, L.M.; Bailera, M. Design Configurations to Achieve an Effective CO2 use and Mitigation through Power to Gas. J. CO2 Util. 2020, 39, 101174. [Google Scholar] [CrossRef]
  11. Quintino, F.M.; Nascimento, N.; Fernandes, E.C. Aspects of Hydrogen and Biomethane Introduction in Natural Gas Infrastructure and Equipment. Hydrogen 2021, 2, 301–318. [Google Scholar] [CrossRef]
  12. Konsolakis, M.; Lykaki, M.; Stefa, S.; Carabineiro, S.A.C.; Varvoutis, G.; Papista, E.; Marnellos, G.E. CO2 Hydrogenation over Nanoceria-Supported Transition Metal Catalysts: Role of Ceria Morphology (Nanorods versus Nanocubes) and Active Phase Nature (Co versus Cu). Nanomaterials 2019, 9, 1739. [Google Scholar] [CrossRef] [PubMed]
  13. Tawalbeh, M.; Javed, R.M.N.; Al-Othman, A.; Almomani, F. The Novel Contribution of Non-Noble Metal Catalysts for Intensified Carbon Dioxide Hydrogenation: Recent Challenges and Opportunities. Energy Convers. Manag. 2023, 279, 116755. [Google Scholar] [CrossRef]
  14. Tawalbeh, M.; Javed, R.M.N.; Al-Othman, A.; Almomani, F.; Ajith, S. Unlocking the Potential of CO2 Hydrogenation into Valuable Products Using Noble Metal Catalysts: A Comprehensive Review. Environ. Technol. Innov. 2023, 31, 103217. [Google Scholar] [CrossRef]
  15. Ebrahimi, P.; Kumar, A.; Khraisheh, M. A Review of CeO2 Supported Catalysts for CO2 Reduction to CO through the Reverse Water Gas Shift Reaction. Catalysts 2022, 12, 1101. [Google Scholar] [CrossRef]
  16. Chen, X.; Chen, Y.; Song, C.; Ji, P.; Wang, N.; Wang, W.; Cui, L. Recent Advances in Supported Metal Catalysts and Oxide Catalysts for the Reverse Water-Gas Shift Reaction. Front. Chem. 2020, 8, 709. [Google Scholar] [CrossRef]
  17. Lykaki, M.; Mandela, E.; Varvoutis, G.; Lampropoulos, A.; Marnellos, G.E.; Konsolakis, M. State-of-the-Art Thermocatalytic Systems for CH4 and CO Production via CO2 Hydrogenation: Critical Comparison, Mechanistic Considerations and Structure-Performance Insights. Discov. Chem. Eng. 2024, 4, 11. [Google Scholar] [CrossRef]
  18. Choi, Y.; Sim, G.D.; Jung, U.; Park, Y.; Youn, M.H.; Chun, D.H.; Rhim, G.B.; Kim, K.Y.; Koo, K.Y. Copper Catalysts for CO2 Hydrogenation to CO through Reverse Water–Gas Shift Reaction for e-Fuel Production: Fundamentals, Recent Advances, and Prospects. Chem. Eng. J. 2024, 492, 152283. [Google Scholar] [CrossRef]
  19. Gao, X.; Wang, Z.; Huang, Q.; Jiang, M.; Askari, S.; Dewangan, N.; Kawi, S. State-of-Art Modifications of Heterogeneous Catalysts for CO2 Methanation—Active Sites, Surface Basicity and Oxygen Defects. Catal. Today 2022, 402, 88–103. [Google Scholar] [CrossRef]
  20. Lee, W.J.; Li, C.; Prajitno, H.; Yoo, J.; Patel, J.; Yang, Y.; Lim, S. Recent Trend in Thermal Catalytic Low Temperature CO2 Methanation: A Critical Review. Catal. Today 2021, 368, 2–19. [Google Scholar] [CrossRef]
  21. Ashok, J.; Pati, S.; Hongmanorom, P.; Tianxi, Z.; Junmei, C.; Kawi, S. A Review of Recent Catalyst Advances in CO2 Methanation Processes. Catal. Today 2020, 356, 471–489. [Google Scholar] [CrossRef]
  22. Memon, M.A.; Jiang, Y.; Hassan, M.A.; Ajmal, M.; Wang, H.; Liu, Y. Heterogeneous Catalysts for Carbon Dioxide Methanation: A View on Catalytic Performance. Catalysts 2023, 13, 1514. [Google Scholar] [CrossRef]
  23. Bacariza, M.C.; Spataru, D.; Karam, L.; Lopes, J.M.; Henriques, C. Promising Catalytic Systems for CO2 Hydrogenation into CH4: A Review of Recent Studies. Processes 2020, 8, 1646. [Google Scholar] [CrossRef]
  24. Hussain, I.; Tanimu, G.; Ahmed, S.; Aniz, C.U.; Alasiri, H.; Alhooshani, K. A Review of the Indispensable Role of Oxygen Vacancies for Enhanced CO2 Methanation Activity over CeO2-Based Catalysts: Uncovering, Influencing, and Tuning Strategies. Int. J. Hydrog. Energy 2023, 48, 24663–24696. [Google Scholar] [CrossRef]
  25. Konsolakis, M.; Lykaki, M. Facet-Dependent Reactivity of Ceria Nanoparticles Exemplified by CeO2-Based Transition Metal Catalysts: A Critical Review. Catalysts 2021, 11, 452. [Google Scholar] [CrossRef]
  26. Yang, C.; Lu, Y.; Zhang, L.; Kong, Z.; Yang, T.; Tao, L.; Zou, Y.; Wang, S. Defect Engineering on CeO2-Based Catalysts for Heterogeneous Catalytic Applications. Small Struct. 2021, 2, 2100058. [Google Scholar] [CrossRef]
  27. Manan, W.N.; Wan Isahak, W.N.R.; Yaakob, Z. CeO2-Based Heterogeneous Catalysts in Dry Reforming Methane and Steam Reforming Methane: A Short Review. Catalysts 2022, 12, 452. [Google Scholar] [CrossRef]
  28. Dong, C.; Zong, X.; Jiang, W.; Niu, L.; Liu, Z.; Qu, D.; Wang, X.; Sun, Z. Recent Advances of Ceria-Based Materials in the Oxidation of Carbon Monoxide. Small Struct. 2021, 2, 2000081. [Google Scholar] [CrossRef]
  29. Fan, L.; Zhang, J.; Ma, K.; Zhang, Y.; Hu, Y.-M.; Kong, L.; Jia, A.; Zhang, Z.; Huang, W.; Lu, J.-Q. Ceria Morphology-Dependent Pd-CeO2 Interaction and Catalysis in CO2 Hydrogenation into Formate. J. Catal. 2021, 397, 116–127. [Google Scholar] [CrossRef]
  30. Jiang, F.; Wang, S.; Liu, B.; Liu, J.; Wang, L.; Xiao, Y.; Xu, Y.; Liu, X. Insights into the Influence of CeO2 Crystal Facet on CO2 Hydrogenation to Methanol over Pd/CeO2 Catalysts. ACS Catal. 2020, 10, 11493–11509. [Google Scholar] [CrossRef]
  31. Martin, N.M.; Velin, P.; Skoglundh, M.; Bauer, M.; Carlsson, P.-A. Catalytic Hydrogenation of CO2 to Methane over Supported Pd, Rh and Ni Catalysts. Catal. Sci. Technol. 2017, 7, 1086–1094. [Google Scholar] [CrossRef]
  32. Khobragade, R.; Roškarič, M.; Žerjav, G.; Košiček, M.; Zavašnik, J.; Van de Velde, N.; Jerman, I.; Tušar, N.N.; Pintar, A. Exploring the Effect of Morphology and Surface Properties of Nanoshaped Pd/CeO2 Catalysts on CO2 Hydrogenation to Methanol. Appl. Catal. A Gen. 2021, 627, 118394. [Google Scholar] [CrossRef]
  33. Wang, C.; Lu, Y.; Zhang, Y.; Fu, H.; Sun, S.; Li, F.; Duan, Z.; Liu, Z.; Wu, C.; Wang, Y.; et al. Ru-Based Catalysts for Efficient CO2 Methanation: Synergistic Catalysis between Oxygen Vacancies and Basic Sites. Nano Res. 2023, 16, 12153–12164. [Google Scholar] [CrossRef]
  34. Chen, M.; Liu, L.; Chen, X.; Qin, X.; Li, K.; Zhang, J.; Bao, X.; Ma, L.; Zhang, C. Effects of Ru Particle Size over TiO2 on the Catalytic Performance of CO2 Hydrogenation. Appl. Surf. Sci. 2024, 654, 159640. [Google Scholar] [CrossRef]
  35. Akter, N.; Zhang, S.; Lee, J.; Kim, D.H.; Boscoboinik, J.A.; Kim, T. Selective Catalytic Reduction of NO by Ammonia and NO Oxidation Over CoOx/CeO2 Catalysts. Mol. Catal. 2020, 482, 110664. [Google Scholar] [CrossRef]
  36. Bian, Z.; Chan, Y.M.; Yu, Y.; Kawi, S. Morphology Dependence of Catalytic Properties of Ni/CeO2 for CO2 Methanation: A Kinetic and Mechanism Study. Catal. Today 2020, 347, 31–38. [Google Scholar] [CrossRef]
  37. Konsolakis, M.; Lykaki, M. Recent Advances on the Rational Design of Non-Precious Metal Oxide Catalysts Exemplified by CuOx/CeO2 Binary System: Implications of Size, Shape and Electronic Effects on Intrinsic Reactivity and Metal-Support Interactions. Catalysts 2020, 10, 160. [Google Scholar] [CrossRef]
  38. Xiao, M.; Zhang, X.; Yang, Y.; Cui, X.; Chen, T.; Wang, Y. M (M = Mn, Co, Cu)-CeO2 Catalysts to Enhance Their CO Catalytic Oxidation at a Low Temperature: Synergistic Effects of the Interaction between Ce3+-Mx+-Ce4+ and the Oxygen Vacancy Defects. Fuel 2022, 323, 124379. [Google Scholar] [CrossRef]
  39. Cargnello, M.; Fornasiero, P.; Gorte, R.J. Opportunities for Tailoring Catalytic Properties through Metal-Support Interactions. Catal. Lett. 2012, 142, 1043–1048. [Google Scholar] [CrossRef]
  40. Liu, C.; Cundari, T.R.; Wilson, A.K. CO2 Reduction on Transition Metal (Fe, Co, Ni, and Cu) Surfaces: In Comparison with Homogeneous Catalysis. J. Phys. Chem. C 2012, 116, 5681–5688. [Google Scholar] [CrossRef]
  41. Li, Y.; Zhang, X.; Zheng, Z. A Review of Transition Metal Oxygen-Evolving Catalysts Decorated by Cerium-Based Materials: Current Status and Future Prospects. CCS Chem. 2022, 4, 31–53. [Google Scholar] [CrossRef]
  42. Varvoutis, G.; Lykaki, M.; Marnellos, G.E.; Konsolakis, M. Recent Advances on Fine-Tuning Engineering Strategies of CeO2-Based Nanostructured Catalysts Exemplified by CO2 Hydrogenation Processes. Catalysts 2023, 13, 275. [Google Scholar] [CrossRef]
  43. Varvoutis, G.; Lykaki, M.; Stefa, S.; Binas, V.; Marnellos, G.E.; Konsolakis, M. Deciphering the Role of Ni Particle Size and Nickel-Ceria Interfacial Perimeter in the Low-Temperature CO2 Methanation Reaction over Remarkably Active Ni/CeO2 Nanorods. Appl. Catal. B Environ. 2021, 297, 120401. [Google Scholar] [CrossRef]
  44. Varvoutis, G.; Lykaki, M.; Stefa, S.; Papista, E.; Carabineiro, S.A.C.; Marnellos, G.E.; Konsolakis, M. Remarkable Efficiency of Ni Supported on Hydrothermally Synthesized CeO2 Nanorods for Low-Temperature CO2 Hydrogenation to Methane. Catal. Commun. 2020, 142, 106036. [Google Scholar] [CrossRef]
  45. Podrojková, N.; Sans, V.; Oriňak, A.; Oriňaková, R. Recent Developments in the Modelling of Heterogeneous Catalysts for CO2 Conversion to Chemicals. ChemCatChem 2020, 12, 1802–1825. [Google Scholar] [CrossRef]
  46. Jangam, A.; Das, S.; Dewangan, N.; Hongmanorom, P.; Hui, W.M.; Kawi, S. Conversion of CO2 to C1 Chemicals: Catalyst Design, Kinetics and Mechanism Aspects of the Reactions. Catal. Today 2020, 358, 3–29. [Google Scholar] [CrossRef]
  47. Saeidi, S.; Najari, S.; Fazlollahi, F.; Nikoo, M.K.; Sefidkon, F.; Klemeš, J.J.; Baxter, L.L. Mechanisms and Kinetics of CO2 Hydrogenation to Value-Added Products: A Detailed Review on Current Status and Future Trends. Renew. Sustain. Energy Rev. 2017, 80, 1292–1311. [Google Scholar] [CrossRef]
  48. Su, X.; Yang, X.; Zhao, B.; Huang, Y. Designing of Highly Selective and High-Temperature Endurable RWGS Heterogeneous Catalysts: Recent Advances and the Future Directions. J. Energy Chem. 2017, 26, 854–867. [Google Scholar] [CrossRef]
  49. Zhang, W.; Sun, J.; Wang, H.; Cui, X. Recent Advances in Hydrogenation of CO2 to CO with Heterogeneous Catalysts Through the RWGS Reaction. Chem. Asian J. 2024, 19, e202300971. [Google Scholar] [CrossRef]
  50. Triviño, M.L.T.; Arriola, N.C., Jr.; Seok Kang, Y.; Gil Seo, J. Transforming CO2 to Valuable Feedstocks: Emerging Catalytic and Technological Advances for the Reverse Water Gas Shift Reaction. Chem. Eng. J. 2024, 487, 150369. [Google Scholar] [CrossRef]
  51. Xie, Y.; Wen, J.; Li, Z.; Chen, J.; Zhang, Q.; Ning, P.; Chen, Y.; Hao, J. Progress in Reaction Mechanisms and Catalyst Development of Ceria-Based Catalysts for Low-Temperature CO2 Methanation. Green Chem. 2023, 25, 130–152. [Google Scholar] [CrossRef]
  52. Alam, M.I.; Cheula, R.; Moroni, G.; Nardi, L.; Maestri, M. Mechanistic and Multiscale Aspects of Thermo-Catalytic CO2 conversion to C1 products. Catal. Sci. Technol. 2021, 11, 6601–6629. [Google Scholar] [CrossRef] [PubMed]
  53. Gao, X.; Cai, P.; Wang, Z.; Lv, X.; Kawi, S. Surface Acidity/Basicity and Oxygen Defects of Metal Oxide: Impacts on Catalytic Performances of CO2 Reforming and Hydrogenation Reactions. Top. Catal. 2023, 66, 299–325. [Google Scholar] [CrossRef]
  54. Etim, U.J.; Zhang, C.; Zhong, Z. Impacts of the Catalyst Structures on CO2 Activation on Catalyst Surfaces. Nanomaterials 2021, 11, 3265. [Google Scholar] [CrossRef] [PubMed]
  55. Ray, K.; Deo, G. A Potential Descriptor for the CO2 Hydrogenation to CH4 over Al2O3 Supported Ni and Ni-Based Alloy Catalysts. Appl. Catal. B Environ. 2017, 218, 525–537. [Google Scholar] [CrossRef]
  56. Feibelman, P.J.; Hamann, D.R. Electronic Structure of a “Poisoned” Transition-Metal Surface. Phys. Rev. Lett. 1984, 52, 61–64. [Google Scholar] [CrossRef]
  57. Escaño, M.C.; Nguyen, T.Q.; Nakanishi, H.; Kasai, H. Another way of looking at bonding on bimetallic surfaces: The role of spin polarization of surface metal d states. J. Phys. Condens. Matter 2009, 21, 492201. [Google Scholar] [CrossRef]
  58. Sun, J.; Zhao, H.; Fang, X.; Zhai, S.; Zhai, D.; Sun, L.; Deng, W. Theoretical Studies on the Catalytic Hydrogenation of Carbon Dioxide by 3d Transition Metals Single-Atom Catalyst Supported on Covalent Triazine Frameworks. Mol. Catal. 2021, 508, 111581. [Google Scholar] [CrossRef]
  59. Sebastian, R.; Swapna, M.S.; Sankararaman, S. Thermal Lens Study of Absolute Porosity in Ceria: A Sankar–Loeb Model Approach. SN Appl. Sci. 2020, 2, 1145. [Google Scholar] [CrossRef]
  60. Deori, K.; Gupta, D.; Saha, B.; Awasthi, S.K.; Deka, S. Introducing Nanocrystalline CeO2 as Heterogeneous Environmental Friendly Catalyst for the Aerobic Oxidation of Para-Xylene to Terephthalic Acid in Water. J. Mater. Chem. A 2013, 1, 7091–7099. [Google Scholar] [CrossRef]
  61. Peyrovi, P.; Gillot, S.; Dacquin, J.-P.; Granger, P.; Dujardin, C. The Activity of CeVO4-Based Catalysts for Ammonia-SCR: Impact of Surface Cerium Enrichment. Catal. Lett. 2021, 151, 1003–1012. [Google Scholar] [CrossRef]
  62. Mosleh, M.; Mahinpour, A. Sonochemical Synthesis and Characterization of Cerium Vanadate Nanoparticles and Investigation of Its Photocatalyst Application. J. Mater. Sci. Mater. Electron. 2016, 27, 8930–8934. [Google Scholar] [CrossRef]
  63. Rao, G.R. Influence of Metal Particles on the Reduction Properties of Ceria-Based Materials Studied by TPR. Bull. Mater. Sci. 1999, 22, 89–94. [Google Scholar] [CrossRef]
  64. Kammert, J.; Moon, J.; Wu, Z. A Review of the Interactions between Ceria and H2 and the Applications to Selective Hydrogenation of Alkynes. Chin. J. Catal. 2020, 41, 901–914. [Google Scholar] [CrossRef]
  65. Sudarsanam, P.; Hillary, B.; Amin, M.H.; Rockstroh, N.; Bentrup, U.; Brückner, A.; Bhargava, S.K. Heterostructured Copper-Ceria and Iron-Ceria Nanorods: Role of Morphology, Redox, and Acid Properties in Catalytic Diesel Soot Combustion. Langmuir 2018, 34, 2663–2673. [Google Scholar] [CrossRef] [PubMed]
  66. Bayram, B.; Soykal, I.I.; Von Deak, D.; Miller, J.T.; Ozkan, U.S. Ethanol Steam Reforming over Co-Based Catalysts: Investigation of Cobalt Coordination Environment under Reaction Conditions. J. Catal. 2011, 284, 77–89. [Google Scholar] [CrossRef]
  67. Guo, X.; Zhou, R. A New Insight into the Morphology Effect of Ceria on CuO/CeO2 Catalysts for CO Selective Oxidation in Hydrogen-Rich Gas. Catal. Sci. Technol. 2016, 6, 3862–3871. [Google Scholar] [CrossRef]
  68. Liu, H.-X.; Li, S.-Q.; Wang, W.-W.; Yu, W.-Z.; Zhang, W.-J.; Ma, C.; Jia, C.-J. Partially Sintered Copper–ceria as Excellent Catalyst for the High-Temperature Reverse Water Gas Shift Reaction. Nat. Commun. 2022, 13, 867. [Google Scholar] [CrossRef]
  69. Oo, W.; Park, J.H.; Sonia, Z.A.; Win, M.Z.; Cho, D.; Yi, K.B. Modification of Copper-Ceria Catalyst via Reverse Microemulsion Method and Study of the Effects of Surfactant on WGS Catalyst Activity. Catalysts 2023, 13, 951. [Google Scholar] [CrossRef]
  70. Zhao, P.; Qin, F.; Huang, Z.; Sun, C.; Shen, W.; Xu, H. Morphology-Dependent Oxygen Vacancies and Synergistic Effects of Ni/CeO2 Catalysts for N2O Decomposition. Catal. Sci. Technol. 2018, 8, 276–288. [Google Scholar] [CrossRef]
  71. Serafin, J.; Llorca, J. Nanoshaped Cerium Oxide with Nickel as a Non-Noble Metal Catalyst for CO2 Thermochemical Reactions. Molecules 2023, 28, 2926. [Google Scholar] [CrossRef] [PubMed]
  72. Zhang, F.; Wang, P.; Koberstein, J.; Khalid, S.; Chan, S.-W. Cerium Oxidation State in Ceria Nanoparticles Studied with X-Ray Photoelectron Spectroscopy and Absorption near Edge Spectroscopy. Surf. Sci. 2004, 563, 74–82. [Google Scholar] [CrossRef]
  73. Sohn, H.; Celik, G.; Gunduz, S.; Dogu, D.; Zhang, S.; Shan, J.; Tao, F.F.; Ozkan, U.S. Oxygen Mobility in Pre-Reduced Nano- and Macro-Ceria with Co Loading: An AP-XPS, In-Situ DRIFTS and TPR Study. Catal. Lett. 2017, 147, 2863–2876. [Google Scholar] [CrossRef]
  74. Xu, J.; Harmer, J.; Li, G.; Chapman, T.; Collier, P.; Longworth, S.; Tsang, S.C. Size Dependent Oxygen Buffering Capacity of Ceria Nanocrystals. Chem. Commun. 2010, 46, 1887–1889. [Google Scholar] [CrossRef]
  75. Alkhoori, A.A.; Elmutasim, O.; Dabbawala, A.A.; Vasiliades, M.A.; Petallidou, K.C.; Emwas, A.-H.; Anjum, D.H.; Singh, N.; Baker, M.A.; Charisiou, N.D.; et al. Mechanistic Features of the CeO2-Modified Ni/Al2O3 Catalysts for the CO2 Methanation Reaction: Experimental and Ab Initio Studies. ACS Appl. Energy Mater. 2023, 6, 8550–8571. [Google Scholar] [CrossRef]
  76. Szamosvölgyi, Á.; Rajkumar, T.; Sápi, A.; Szenti, I.; Ábel, M.; Gómez-Pérez, J.F.; Baán, K.; Fogarassy, Z.; Dodony, E.; Pécz, B.; et al. Interfacial Ni Active Sites Strike Solid Solutional Counterpart in CO2 Hydrogenation. Environ. Technol. Innov. 2022, 27, 102747. [Google Scholar] [CrossRef]
  77. Kuan, W.-F.; Chung, C.-H.; Lin, M.M.; Tu, F.-Y.; Chen, Y.-H.; Yu, W.-Y. Activation of Carbon Dioxide with Surface Oxygen Vacancy of Ceria Catalyst: An Insight from in-Situ X-Ray Absorption near Edge Structure Analysis. Mater. Today Sustain. 2023, 23, 100425. [Google Scholar] [CrossRef]
  78. Martin, N.M.; Hemmingsson, F.; Schaefer, A.; Ek, M.; Merte, L.R.; Hejral, U.; Gustafson, J.; Skoglundh, M.; Dippel, A.-C.; Gutowski, O.; et al. Structure-Function Relationship for CO2 Methanation over Ceria Supported Rh and Ni Catalysts under Atmospheric Pressure Conditions. Catal. Sci. Technol. 2019, 9, 1644–1653. [Google Scholar] [CrossRef]
  79. Cao, F.; Xiao, Y.; Zhang, Z.; Li, J.; Xia, Z.; Hu, X.; Ma, Y.; Qu, Y. Influence of Oxygen Vacancies of CeO2 on Reverse Water Gas Shift Reaction. J. Catal. 2022, 414, 25–32. [Google Scholar] [CrossRef]
  80. Jenkinson, K.; Spadaro, M.C.; Golovanova, V.; Andreu, T.; Morante, J.R.; Arbiol, J.; Bals, S. Direct Operando Visualization of Metal Support Interactions Induced by Hydrogen Spillover During CO2 Hydrogenation. Adv. Mater. 2023, 35, 2306447. [Google Scholar] [CrossRef]
  81. Ye, H.; Na, W.; Gao, W.; Wang, H. Carbon-Modified CuO/ZnO Catalyst with High Oxygen Vacancy for CO2 Hydrogenation to Methanol. Energy Technol. 2020, 8, 2000194. [Google Scholar] [CrossRef]
  82. Jiang, D.; Wang, W.; Zhang, L.; Zheng, Y.; Wang, Z. Insights into the Surface-Defect Dependence of Photoreactivity over CeO2 Nanocrystals with Well-Defined Crystal Facets. ACS Catal. 2015, 5, 4851–4858. [Google Scholar] [CrossRef]
  83. Hu, J.; Wei, F.; Hu, X.; Xu, J.; Deng, W. Synthesis of CuO-Loaded Ceria Hollow Spheres for Catalytic CO Oxidation. ChemistrySelect 2022, 7, e202103476. [Google Scholar] [CrossRef]
  84. Konsolakis, M.; Sgourakis, M.; Carabineiro, S.A.C. Surface and Redox Properties of Cobalt-Ceria Binary Oxides: On the Effect of Co Content and Pretreatment Conditions. Appl. Surf. Sci. 2015, 341, 48–54. [Google Scholar] [CrossRef]
  85. Zhan, Y.; Liu, Y.; Peng, X.; Zhao, W.; Zhang, Y.; Wang, X.; Au, C.; Jiang, L. Molecular-Level Understanding of Reaction Path Optimization as a Function of Shape Concerning the Metal-Support Interaction Effect of Co/CeO2 on Water-Gas Shift Catalysis. Catal. Sci. Technol. 2019, 9, 4928–4937. [Google Scholar] [CrossRef]
  86. Lykaki, M.; Stefa, S.; Carabineiro, S.A.C.; Pandis, P.K.; Stathopoulos, V.N.; Konsolakis, M. Facet-Dependent Reactivity of Fe2O3/CeO2 Nanocomposites: Effect of Ceria Morphology on CO Oxidation. Catalysts 2019, 9, 371. [Google Scholar] [CrossRef]
  87. Yogendra, K.; Sitaramulu, P.; Nazeer, S.; Kumar, P.M.; Reddy, B.M.; Rao, T.V. Influence of Iron Doping in Mesoporous Ceria on the Physicochemical Properties and Catalytic Activity in Styrene Oxidation. Appl. Surf. Sci. 2024, 676, 160971. [Google Scholar] [CrossRef]
  88. Wang, H.; Jin, B.; Wang, H.; Ma, N.; Liu, W.; Weng, D.; Wu, X.; Liu, S. Study of Ag Promoted Fe2O3@CeO2 as Superior Soot Oxidation Catalysts: The Role of Fe2O3 Crystal Plane and Tandem Oxygen Delivery. Appl. Catal. B Environ. 2018, 237, 251–262. [Google Scholar] [CrossRef]
  89. Liu, F.; Wang, Z.; Wang, D.; Chen, D.; Chen, F.; Li, X. Morphology and Crystal-Plane Effects of Fe/W-CeO2 for Selective Catalytic Reduction of NO with NH3. Catalysts 2019, 9, 288. [Google Scholar] [CrossRef]
  90. Liu, W.; Wang, W.; Tang, K.; Guo, J.; Ren, Y.; Wang, S.; Feng, L.; Yang, Y. The Promoting Influence of Nickel Species in the Controllable Synthesis and Catalytic Properties of Nickel-Ceria Catalysts. Catal. Sci. Technol. 2016, 6, 2427–2434. [Google Scholar] [CrossRef]
  91. Tang, K.; Liu, W.; Li, J.; Guo, J.; Zhang, J.; Wang, S.; Niu, S.; Yang, Y. The Effect of Exposed Facets of Ceria to the Nickel Species in Nickel-Ceria Catalysts and Their Performance in a NO + CO Reaction. ACS Appl. Mater. Interfaces 2015, 7, 26839–26849. [Google Scholar] [CrossRef] [PubMed]
  92. Musab Ahmed, S.; Ren, J.; Ullah, I.; Lou, H.; Xu, N.; Abbasi, Z.; Wang, Z. Ni-Based Catalysts for CO2 Methanation: Exploring the Support Role in Structure-Activity Relationships. ChemSusChem 2024, 17, e202400310. [Google Scholar] [CrossRef] [PubMed]
  93. Boaro, M.; Colussi, S.; Trovarelli, A. Ceria-Based Materials in Hydrogenation and Reforming Reactions for CO2 Valorization. Front. Chem. 2019, 7, 28. [Google Scholar] [CrossRef] [PubMed]
  94. Liu, Y.; Li, D.; Zhao, H.; Wang, C.; Xu, Y.; Li, L.; Li, Z.; Wang, H.; Li, K. Boosting CO2 Methanation Activity by Tuning Ni Crystal Plane and Oxygen Vacancy in Ni/CeO2 Catalyst. Chem. Eng. J. 2024, 494, 153004. [Google Scholar] [CrossRef]
  95. Dietz, L.; Piccinin, S.; Maestri, M. Mechanistic Insights into CO2 Activation via Reverse Water—Gas Shift on Metal Surfaces. J. Phys. Chem. C 2015, 119, 4959–4966. [Google Scholar] [CrossRef]
  96. Wang, S.-G.; Liao, X.-Y.; Cao, D.-B.; Huo, C.-F.; Li, Y.-W.; Wang, J.; Jiao, H. Factors Controlling the Interaction of CO2 with Transition Metal Surfaces. J. Phys. Chem. C 2007, 111, 16934–16940. [Google Scholar] [CrossRef]
  97. Bhattacharjee, S.; Waghmare, U.V.; Lee, S.-C. An improved d-band model of the catalytic activity of magnetic transition metal surfaces. Sci. Rep. 2016, 6, 35916. [Google Scholar] [CrossRef]
  98. Lee, S.M.; Lee, Y.H.; Moon, D.H.; Ahn, J.Y.; Nguyen, D.D.; Chang, S.W.; Kim, S.S. Reaction Mechanism and Catalytic Impact of Ni/CeO2-X Catalyst for Low-Temperature CO2 Methanation. Ind. Eng. Chem. Res. 2019, 58, 8656–8662. [Google Scholar] [CrossRef]
  99. Huang, W.; Gao, Y. Morphology-Dependent Surface Chemistry and Catalysis of CeO2 Nanocrystals. Catal. Sci. Technol. 2014, 4, 3772–3784. [Google Scholar] [CrossRef]
  100. Li, L.; Jiang, L.; Li, D.; Yuan, J.; Bao, G.; Li, K. Enhanced Low-Temperature Activity of CO2 Methanation over Ni/CeO2 Catalyst: Influence of Preparation Methods. Appl. Catal. O Open 2024, 192, 206956. [Google Scholar] [CrossRef]
  101. Lykaki, M.; Pachatouridou, E.; Carabineiro, S.A.C.; Iliopoulou, E.; Andriopoulou, C.; Kallithrakas-Kontos, N.; Boghosian, S.; Konsolakis, M. Ceria nanoparticles shape effects on the structural defects and surface chemistry: Implications in CO oxidation by Cu/CeO2 catalysts. Appl. Catal. B Environ. 2018, 230, 18–28. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of CeO2-NR and TM/CeO2 samples.
Figure 1. XRD patterns of CeO2-NR and TM/CeO2 samples.
Catalysts 14 00611 g001
Figure 2. SEM images of the samples: (a) Ti/CeO2, (b) V/CeO2, (c) Cr/CeO2, (d) Mn/CeO2, (e) Fe/CeO2, (f) Co/CeO2, (g) Ni/CeO2, (h) Cu/CeO2 and (i) Zn/CeO2.
Figure 2. SEM images of the samples: (a) Ti/CeO2, (b) V/CeO2, (c) Cr/CeO2, (d) Mn/CeO2, (e) Fe/CeO2, (f) Co/CeO2, (g) Ni/CeO2, (h) Cu/CeO2 and (i) Zn/CeO2.
Catalysts 14 00611 g002
Figure 3. TEM images of the samples: (a) CeO2-NR, (b) Ti/CeO2, (c) V/CeO2, (d) Cr/CeO2, (e) Mn/CeO2, (f) Fe/CeO2, (g) Co/CeO2, (h) Ni/CeO2, (i) Cu/CeO2 and (j) Zn/CeO2.
Figure 3. TEM images of the samples: (a) CeO2-NR, (b) Ti/CeO2, (c) V/CeO2, (d) Cr/CeO2, (e) Mn/CeO2, (f) Fe/CeO2, (g) Co/CeO2, (h) Ni/CeO2, (i) Cu/CeO2 and (j) Zn/CeO2.
Catalysts 14 00611 g003
Figure 4. H2-TPR profiles of CeO2-NR and representative TM/CeO2 samples.
Figure 4. H2-TPR profiles of CeO2-NR and representative TM/CeO2 samples.
Catalysts 14 00611 g004
Figure 5. XPS spectra of (a) Ce 3d and (b) O 1s of the reduced CeO2-NR and optimum catalysts (Fe, Co, Ni, Cu).
Figure 5. XPS spectra of (a) Ce 3d and (b) O 1s of the reduced CeO2-NR and optimum catalysts (Fe, Co, Ni, Cu).
Catalysts 14 00611 g005
Figure 6. XPS spectra of (a) Cu 2p, (b) Co 2p, (c) Fe 2p, and (d) Ni 2p of the reduced optimum catalysts (Fe, Co, Ni, Cu).
Figure 6. XPS spectra of (a) Cu 2p, (b) Co 2p, (c) Fe 2p, and (d) Ni 2p of the reduced optimum catalysts (Fe, Co, Ni, Cu).
Catalysts 14 00611 g006aCatalysts 14 00611 g006b
Figure 7. Comparative assessment of (a) CO2 conversion and (b) selectivity to CO for the first-row 3d transition metals supported on CeO2 nanorods. Reaction conditions: WHSV = 30 L·g−1·h−1, H2:CO2 = 4, P = 1 bar.
Figure 7. Comparative assessment of (a) CO2 conversion and (b) selectivity to CO for the first-row 3d transition metals supported on CeO2 nanorods. Reaction conditions: WHSV = 30 L·g−1·h−1, H2:CO2 = 4, P = 1 bar.
Catalysts 14 00611 g007
Figure 8. Metal entity-dependence of CO2 conversion (solid line) and CH4 selectivity (dashed line) at 350 °C for the first-row 3d transition metals supported on CeO2 nanorods.
Figure 8. Metal entity-dependence of CO2 conversion (solid line) and CH4 selectivity (dashed line) at 350 °C for the first-row 3d transition metals supported on CeO2 nanorods.
Catalysts 14 00611 g008
Figure 9. (a) Atomic H adsorption energy on P-2,5-DCP-CTF-M-CO2 ((2-pyridyl)-s-triazine is the simplified unit model of the covalent triazine frameworks (CTF), the unit model and its metal complex is P-2,5-DCP-CTF-M (M = Sc, Ti, V, Cr, Mn, Fe, Co, Ni, Cu, Zn)), (b) barrier energy, and (c) change in CO2 charge from IS to TS (C) in the formate pathway as a function of 3d-orbital electron number. Initial state (IS) of CO2 hydrogenation corresponding to the co-adsorption configuration of CO2 and H atom on the active metal site; transition states (TS); Sc, Ti, and V atoms with less than four 3d-orbital electrons were divided in the left region (orange). The metal atoms, including Mn, Fe, Co, Ni, and Cu, having more than four 3d-orbital electrons were divided in the right region (blue). Reproduced with permission from [58].
Figure 9. (a) Atomic H adsorption energy on P-2,5-DCP-CTF-M-CO2 ((2-pyridyl)-s-triazine is the simplified unit model of the covalent triazine frameworks (CTF), the unit model and its metal complex is P-2,5-DCP-CTF-M (M = Sc, Ti, V, Cr, Mn, Fe, Co, Ni, Cu, Zn)), (b) barrier energy, and (c) change in CO2 charge from IS to TS (C) in the formate pathway as a function of 3d-orbital electron number. Initial state (IS) of CO2 hydrogenation corresponding to the co-adsorption configuration of CO2 and H atom on the active metal site; transition states (TS); Sc, Ti, and V atoms with less than four 3d-orbital electrons were divided in the left region (orange). The metal atoms, including Mn, Fe, Co, Ni, and Cu, having more than four 3d-orbital electrons were divided in the right region (blue). Reproduced with permission from [58].
Catalysts 14 00611 g009
Figure 10. Relationship among the CO2 conversion rate and CH4 selectivity at 350 °C with the relative oxygen vacancy and Ce3+ concentrations for the optimum reduced catalysts (Fe, Co, Ni, Cu).
Figure 10. Relationship among the CO2 conversion rate and CH4 selectivity at 350 °C with the relative oxygen vacancy and Ce3+ concentrations for the optimum reduced catalysts (Fe, Co, Ni, Cu).
Catalysts 14 00611 g010
Table 1. Textural and structural characteristics of CeO2 and TM/CeO2 samples.
Table 1. Textural and structural characteristics of CeO2 and TM/CeO2 samples.
SampleNominal Metal LoadingEDS AnalysisBET AnalysisXRD AnalysisTEM Analysis
Atomic Ratio M/CeMetal Content (wt%)SBET (m2/g)Average Crystallite Size (nm)MxOy Particle Size (nm)
CeO2MxOy
CeO2-NR---7915--
Ti/CeO26.50.246.3-112016
V/CeO26.90.287.6-144528
Cr/CeO27.00.267.2-11- 110
Mn/CeO27.40.226.5-11- 115
Fe/CeO27.50.216.36910711
Co/CeO27.90.268.172141615
Ni/CeO27.90.257.872142310
Cu/CeO28.50.258.675124316
Zn/CeO28.70.248.376124441
1 Not calculated due to the absence of the corresponding XRD peaks.
Table 2. Redox properties of CeO2-NR and representative TM/CeO2 samples.
Table 2. Redox properties of CeO2-NR and representative TM/CeO2 samples.
SampleH2 Consumption (mmol H2 g−1) 1Theoretical H2 Consumption (mmol H2 g−1) 2Peak Temperature (°C)
CeO2-NR0.6--545-788
Fe/CeO21.61.9390465588759
Cu/CeO21.81.3181-217793
Ni/CeO21.81.3220288353747
Co/CeO22.41.7318-388789
1 Estimated by the area of the corresponding TPR peaks, which is calibrated against a known amount of CuO standard sample. 2 Calculated as the amount of H2 required for the complete reduction in fully oxidized MxOy to M0 on the basis of the metal nominal loading.
Table 3. XPS results of representative reduced samples.
Table 3. XPS results of representative reduced samples.
SampleOads/OlatCe3+ (%)
CeO2-NR0.5927.2
Fe/CeO20.4326.9
Cu/CeO20.5425.5
Co/CeO20.6329.9
Ni/CeO20.6530.4
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

Lykaki, M.; Stefa, S.; Varvoutis, G.; Binas, V.D.; Marnellos, G.E.; Konsolakis, M. Comparative Assessment of First-Row 3d Transition Metals (Ti-Zn) Supported on CeO2 Nanorods for CO2 Hydrogenation. Catalysts 2024, 14, 611. https://doi.org/10.3390/catal14090611

AMA Style

Lykaki M, Stefa S, Varvoutis G, Binas VD, Marnellos GE, Konsolakis M. Comparative Assessment of First-Row 3d Transition Metals (Ti-Zn) Supported on CeO2 Nanorods for CO2 Hydrogenation. Catalysts. 2024; 14(9):611. https://doi.org/10.3390/catal14090611

Chicago/Turabian Style

Lykaki, Maria, Sofia Stefa, Georgios Varvoutis, Vassilios D. Binas, George E. Marnellos, and Michalis Konsolakis. 2024. "Comparative Assessment of First-Row 3d Transition Metals (Ti-Zn) Supported on CeO2 Nanorods for CO2 Hydrogenation" Catalysts 14, no. 9: 611. https://doi.org/10.3390/catal14090611

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