Next Article in Journal
Spray Coated Colloidal Quantum Dot Films for Broadband Photodetectors
Next Article in Special Issue
Supported Gold Nanoparticles as Catalysts in Peroxidative and Aerobic Oxidation of 1-Phenylethanol under Mild Conditions
Previous Article in Journal
A Skin-Inspired Stretchable, Self-Healing and Electro-Conductive Hydrogel with a Synergistic Triple Network for Wearable Strain Sensors Applied in Human-Motion Detection
Previous Article in Special Issue
Facile Preparation of Pd/UiO-66-v for the Conversion of Furfuryl Alcohol to Tetrahydrofurfuryl Alcohol under Mild Conditions in Water
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

CO2 Hydrogenation over Nanoceria-Supported Transition Metal Catalysts: Role of Ceria Morphology (Nanorods versus Nanocubes) and Active Phase Nature (Co versus Cu)

by
Michalis Konsolakis
1,*,
Maria Lykaki
1,
Sofia Stefa
1,
Sόnia A. C. Carabineiro
2,
Georgios Varvoutis
3,4,
Eleni Papista
3 and
Georgios E. Marnellos
3,4
1
School of Production Engineering and Management, Technical University of Crete, GR-73100 Chania, Greece
2
Laboratório de Catálise e Materiais (LCM), Laboratório Associado LSRE-LCM, Faculdade de Engenharia, Universidade do Porto, 4200-465 Porto, Portugal
3
Department of Mechanical Engineering, University of Western Macedonia, GR-50100 Kozani, Greece
4
Chemical Process & Energy Resources Institute, Centre for Research & Technology Hellas, GR-57001 Thermi, Thessaloniki, Greece
*
Author to whom correspondence should be addressed.
Nanomaterials 2019, 9(12), 1739; https://doi.org/10.3390/nano9121739
Submission received: 4 November 2019 / Revised: 28 November 2019 / Accepted: 2 December 2019 / Published: 6 December 2019
(This article belongs to the Special Issue Catalysis by Metal-Oxide Nanostructures)

Abstract

:
In this work we report on the combined impact of active phase nature (M: Co or Cu) and ceria nanoparticles support morphology (nanorods (NR) or nanocubes (NC)) on the physicochemical characteristics and CO2 hydrogenation performance of M/CeO2 composites at atmospheric pressure. It was found that CO2 conversion followed the order: Co/CeO2 > Cu/CeO2 > CeO2, independently of the support morphology. Co/CeO2 catalysts demonstrated the highest CO2 conversion (92% at 450 °C), accompanied by 93% CH4 selectivity. On the other hand, Cu/CeO2 samples were very selective for CO production, exhibiting 52% CO2 conversion and 95% CO selectivity at 380 °C. The results obtained in a wide range of H2:CO2 ratios (1–9) and temperatures (200–500 °C) are reaching in both cases the corresponding thermodynamic equilibrium conversions, revealing the superiority of Co- and Cu-based samples in methanation and reverse water-gas shift (rWGS) reactions, respectively. Moreover, samples supported on ceria nanocubes exhibited higher specific activity (µmol CO2·m−2·s−1) compared to samples of rod-like shape, disclosing the significant role of support morphology, besides that of metal nature (Co or Cu). Results are interpreted on the basis of different textural and redox properties of as-prepared samples in conjunction to the different impact of metal entity (Co or Cu) on CO2 hydrogenation process.

Graphical Abstract

1. Introduction

It is widely accepted amongst the scientific community that the increasing trend of CO2 emissions in the Earth’s atmosphere since the onset of industrialization is the key attributor for the planet temperature rise over the last two centuries [1]. Global temperature is projected to rise by the year 2040 by 1.5 °C in comparison with the pre-industrial levels, according to the latest Intergovernmental Panel on Climate Change (IPCC) report on the impacts on global warming [2,3]. Efforts of mitigation of the aforementioned environmental issue can be simplified into three general approaches: (i) complete and/or partial replacement of carbon-based fuels with renewable energy sources (RESs), (ii) carbon dioxide capture and storage (CCS) technology and (iii) chemical conversion/utilization of CO2 toward value-added chemicals and fuels [4].
The latter approach has attracted intense interest over the past decades, with hydrogenation of CO2 being one of the most thoroughly investigated methods, owing to the wide variety of possible products [5]. This route can also provide an effective way to valorize CO2 emissions and efficiently store the surplus power from non-intermittent RESs (e.g., solar, wind) in the form of “green” hydrogen, providing either CO via the mildly endothermic reverse water-gas shift (rWGS) reaction (Equation (1)) or CH4 via the highly exothermic methanation reaction, often referred to as the “Sabatier reaction” (Equation (2)), discovered in 1902 by the French scientist Paul Sabatier [6].
CO2 + H2 ↔ CO + H2O, ΔH298K = +41.3 kJ/mol
CO2 + 4H2 ↔ CH4 + 2H2O, ΔH298K = −164.7 kJ/mol
Other possible products of potential value from CO2 hydrogenation include methanol [7,8,9,10], dimethyl ether [11], formic acid [12] and hydrocarbons [13].
Among the different CO2 hydrogenation products, carbon monoxide is a valuable feedstock for the C1 chemical industry, since various liquid synthetic hydrocarbons and chemicals can be produced by its subsequent upgrading, through the well-established Fischer–Tropsch synthesis [14,15]. Moreover, CO can be used in nickel purification [16]. Methane, on the other hand, as a major constituent of natural gas, is an energy carrier whose significance is applicable globally from household use to the industrial, energy, and transportation sectors. Given that the hydrogen needed for carbon dioxide methanation is provided via a carbon-neutral energy source (e.g., solar-powered water electrolysis), the overall scheme can be labeled as a power-to-gas (PtG) process [17,18]. Carbon dioxide methanation can effectively convert a less manageable energy vector, such as gaseous hydrogen into a high-energy source such as methane, since CH4 possesses three times the volumetric energy density of hydrogen [19]. Also, CH4 can be easily integrated into the existing natural gas storage and distribution network, especially throughout Europe [20,21].
Regarding the overall sustainability of the proposed process, it should be pointed out that the general scheme of the CO2 hydrogenation concept is associated with the use of “green” H2, originated by the excess energy provided by RESs and the concentrated amounts of CO2 emissions as feedstock. In this regard, the CO2 hydrogenation process could be implemented near a source of highly concentrated CO2 emissions, such as effluent streams of the steel industry or a CO2 capture plant. By employing highly active and inexpensive catalysts, large amounts of CO2 can be potentially mitigated with the concurrent production of value-added products, such as CH4 or CO, which can be used as fuels or feedstock in the chemical industry. Regarding the use of renewable hydrogen, it can be exploited entirely in the hydrogenation process or it can be partially converted to the electricity required for the reaction process. Also, excess hydrogen can be directly injected into the gas grid or used in fuel cell-powered vehicles. Various comprehensive studies have been devoted to the sustainability of the CO2 hydrogenation process, to which the reader can refer [22,23,24,25,26].
Whereas hydrogenation of CO2 can be a promising way to reduce the environmental carbon print, several limitations arise for the implementation of a technology based on either Equation (1) or Equation (2). The first is associated with the activation of carbon dioxide itself, a fully oxidized and thermodynamically stable compound whose reduction is not energetically favorable [27], and thus requires strong reductants (i.e., H2) or electrochemical-assisted reduction processes [28,29]. Secondly, CO2 hydrogenation is subjected to kinetic and equilibrium limitations, thus reaction rates need to be promoted [30]. Numerous catalytic systems have been employed in order to overcome these limitations and several reviews in the literature summarize the catalysts explored for either rWGS [2,31,32] or CO2 methanation [33,34,35,36] reactions. The most studied catalytic systems are composites with metals supported on reducible metal oxides (e.g., CeO2, ZrO2) or a combination of them. These systems have been employed as bi-functional catalysts, with oxide supports mainly providing oxygen vacancies to activate CO2, and metal active sites dissociating molecular hydrogen, the so-called hydrogen spillover process [31,37].
Among the oxides investigated, CeO2 has attracted much attention as a supporting carrier, due to its high oxygen mobility and unique redox properties, as cerium can rapidly change between its two oxidation states (Ce3+ and Ce4+) [38,39]. Moreover, ceria is a basic oxide promoting a strong interaction with CO2, facilitating its adsorption [40]. Besides bare ceria’s excellent redox properties, many studies have focused on the development of highly efficient and low-cost ceria-based catalytic composites, since the combination of various non-noble transition metals (TMs) (e.g., Cu, Co, Ni, Fe) with ceria, can further enhance the catalytic activity and/or selectivity due to the peculiar metal-support synergistic interactions [41,42,43]. Despite their adequate catalytic activity, the use of precious metals like Ru [44], Rh [45] and Pd [46] as active phases is generally not preferable, since their high cost and scarcity might render the process financially non-viable. Thus, from a techno-economical point of view, the use of TMs-based catalysts may be favorable, since these metals can achieve comparable activity to that of the most active noble metal catalysts albeit at a substantially lower cost [20,47]. In this regard, efforts from our group have recently focused on developing inexpensive TMs-based catalytic composites, with particular emphasis on middle-late 3d metals, i.e., Cu, Co, Ni, Fe, which were found to adsorb and consequently activate CO2 through a charge transfer from metal phase to CO2 moiety [48]. In particular, Density Functional Theory (DFT) calculations have revealed spontaneous chemisorption of CO2 and favorable thermodynamic properties for the aforementioned metals, with Cu, however, exhibiting a weaker interaction with CO2 [48].
In order to develop highly efficient ceria-based catalysts, much research has been devoted on the rational design of catalytic materials by means of advanced synthetic and/or promotional routes [49,50,51,52]. The ultimate goal would be to obtain catalytic systems with adequate stability, CO2 conversion activity and high product selectivity, in order to exclusively generate CO or CH4 in a real large-scale process. Of major importance towards fine-tuning of CeO2-based materials is the decrease of particles size in the nanoscale. Nano-materials exhibit abundance in surface atoms and defect sites such as oxygen vacancies, whereas the electronic perturbations between the metal and support nanoparticles greatly affect the catalytic performance [41,53,54]. Moreover, by tailoring the shape of nanoparticles by means of advanced synthetic routes (e.g., hydrothermal method), different crystal facets can be exposed leading to different oxygen storage capacity (OSC) and oxygen mobility [55,56]. For example, Ouyang et al. [57] have investigated ceria morphological effects during methanol synthesis from CO2 hydrogenation over CuO/CeO2 mixed oxides; copper-ceria nanorods exhibited the highest CO2 conversion and methanol yield due to the strongest metal–support interaction, as compared to nanocubes and nanospheres. Similarly, Au/CeO2 nanorods were found to exhibit a stronger gold-ceria interaction and higher activity in the forward WGS reaction than cubic and polyhedral CeO2 [58]. Also, Liu et al. [59] reported a better activity for ceria nanocubes in the rWGS reaction than nanorods and nanopolyhedra, with the nanocubic samples prepared, preferentially exposing (100) planes, a potentially more active surface than (110) and (111) planes.
Although there are several studies regarding the hydrogenation of CO2 over ceria-based composites, it should be noted that CO2 hydrogenation proceeds through a complex reaction pathway, being affected to a different extent by various factors, such as the metal–oxide interactions, the formation of oxygen vacancies, the reducibility, etc. [51,60,61,62,63,64]. Furthermore, selectivity towards CO, CH4 or other possible compounds can vary remarkably when using catalytic composites with various active metal phases supported onto ceria, depending thus on the metal entity employed [65,66,67].
In light of the above, the aim of the present study is to investigate the effect of non-noble metal phase nature (Co, Cu) and support morphology (nanorods, nanocubes) on the textural, structural, redox properties and, consequently, on the CO2 hydrogenation performance of mesoporous ceria-based nanocatalysts. The originality of the present work relies on the combined impact of transition metal nature (Cu, Co) and support morphology (ceria nanorods or nanocubes) on the CO2 hydrogenation performance over a wide range of H2:CO2 ratios (1–9) and temperatures (200–500 °C), in conjunction with the thermodynamic analysis performed under different reaction conditions. The as-prepared samples were synthesized hydrothermally and characterized by N2 adsorption-desorption, X-ray diffraction (XRD), transmission electron microscopy (TEM) and temperature programmed reduction (TPR) methods and they were catalytically evaluated in the CO2 hydrogenation reaction at atmospheric pressure. The obtained results are interpreted on the basis of a thermodynamic analysis at different reaction conditions, in conjunction to structural and surface characterization results that reveal the key role of metal entity and support morphology both on CO2 conversion and products selectivity.

2. Materials and Methods

2.1. Reagents

All chemicals used were of analytical reagent grade. Ce(NO3)3·6H2O (purity ≥ 99.0%, Fluka, Bucharest, Romania), Cu(NO3)2·2.5H2O (Fluka, Bucharest, Romania) and Co(NO3)2·6H2O (≥98%, Sigma-Aldrich, St. Louis, MO, USA) were used as precursors for the preparation of bare ceria, Cu/CeO2 and Co/CeO2 catalysts, respectively. Sodium hydroxide (≥98%, Honeywell Fluka, Bucharest, Romania), ethanol (purity 99.8%, ACROS Organics, Waltham, MA, USA) and double deionized water (DI) were also employed during catalysts preparation procedure.

2.2. Materials Synthesis

Initially, bare ceria nanoparticles were synthesized by the hydrothermal method, as described in our previous work [68]. Briefly, for the synthesis of bare ceria nanostructures, appropriate amounts of Ce(NO3)3·6H2O and NaOH were initially dissolved in double deionized water, then mixed under vigorous stirring for 1 h and aged for 24 h at 90 °C for ceria nanorods and at 180 °C for ceria nanocubes. Eventually, the resulting solids were recovered by centrifugation, washed thoroughly with double deionized water until pH reached a value of 7, in order to remove any co-precipitated salts, and finally washed with ethanol to avoid agglomeration of the nanoparticles. Then, the precipitate was dried at 90 °C for 12 h, followed by calcination at 500 °C for 2 h, under air flow (heating ramp 5 °C/min). The bare ceria samples are designated as CeO2-NX (NX: NR (nanorods), NC (nanocubes)).
Cu/CeO2-NX and Co/CeO2-NX catalysts were synthesized by the wet impregnation method, using aqueous solutions of Cu(NO3)2·2.5H2O and Co(NO3)2·6H2O, respectively, in order to obtain a metal/cerium atomic ratio of 0.25, corresponding to a Cu loading of ca. 8.5 wt.% and a Co loading of ca. 7.9 wt.%. Then, the obtained suspensions were heated under stirring until water evaporation, dried at 90 °C for 12 h and finally calcined at 500 °C for 2 h under air flow (heating ramp 5 °C/min).

2.3. Materials Characterization

The textural properties of the materials were evaluated by the N2-adsorption isotherms at −196 °C, using an ASAP 2010 (Micromeritics, Norcross, GA, USA) apparatus (ReQuimTe Analyses Laboratory, Universidade Nova de Lisboa, Portugal). Samples were previously degassed at 300 °C for 6 h. The specific surface area (SBET) was calculated by the Brunauer-Emmett-Teller (BET) equation. Structural characterization was carried out by means of X-ray diffraction (XRD) in a PAN’alytical X’Pert MPD (PANanalytical, Almelo, Netherlands) equipped with a X’Celerator detector and secondary monochromator (Cu Kα, λ = 0.154 nm, 50 kV, 40 mA; data recorded at a 0.017° step size, 100 s/step), located at Universidade de Trás-os-Montes e Alto Douro, Vila Real, Portugal. The collected spectra were analyzed by Rietveld refinement using PowderCell software (by Werner Kraus and Gert Nolze, http://www.ccp14.ac.uk), allowing the determination of crystallite sizes by means of the Williamson-Hall plot. The samples were imaged by transmission electron microscopy (TEM). The analyses were performed on a Leo 906E apparatus (Austin, TX, USA), at 120 kV. Samples were prepared by ultrasonic dispersion and a 400 mesh formvar/carbon copper grid (Agar Scientific, Essex, UK) was dipped into the solution for the TEM analysis. The redox properties were assessed by temperature programmed reduction (TPR) experiments in an AMI−200 Catalyst Characterization Instrument (Altamira Instruments, Pittsburgh, PA, USA), employing H2 as a reducing agent. In a typical H2-TPR experiment, 50 mg of the sample were heated up to 1100 °C (10 °C/min), under H2 flow (1.5 cm3) balanced with He (29 cm3).

2.4. Catalytic Evaluation Studies

Catalytic tests for CO2 hydrogenation were carried out in a quartz fixed-bed U-shaped reactor (internal diameter, ID = 1 cm). The reactor was placed inside a high temperature furnace, equipped with a thermocouple and a Proportional Integral Differential (PID) temperature controller. In each experiment, the catalyst bed consisted of a mixture of 200 mg catalyst diluted with 200 mg of inert SiO2. Prior to experiments, catalysts were reduced in situ at 400 °C for 1 h under pure H2 flow (50 cm3/min), followed by flushing with He (10 cm3/min). The experiments were conducted at atmospheric pressure and in the temperature range of 200–500 °C at intervals of 20–25 °C. The heating rate was 1 °C/min, adequate for the establishment of steady-state conditions before each measurement. To ensure the reproducibility of the obtained conversion/selectivity values, all tests were conducted twice, without noticeable differences (<5% difference between measurements at the same temperature).
All gases used were of 99.999% purity, provided and certified by Air Liquide Hellas S.A. The total flow rate of the feed gas was 100 cm3/min, corresponding to a gas hourly space velocity (GHSV) of 20,000 h−1. Gas feed comprised of H2/CO2 mixtures at different molar ratios (1–9). The analysis of gases was performed by a gas chromatograph with He as the carrier gas, equipped with a thermal conductivity detector (TCD) for detection of CO and CO2, a flame ionization detector (FID) for monitoring CH4 and two separation columns (Molecular Sieve 13X and Porapack QS). A cold trap submerged in a water bath was connected to the reactor effluent in order to condensate H2O produced by the reactions.
Carbon dioxide conversion, ΧCO2, and product selectivities, SCO and SCH4, were calculated as follows (Equations (3)–(5)):
X CO 2 = ( [ CO 2 ] in · F in )     ( [ CO 2 ] out · F out ) [ CO 2 ] in · F in · 100
S CO = [ CO ] out [ CO ] out + [ CH 4 ] out · 100
S CH 4 = [ CH 4 ] out [ CO ] out + [ CH 4 ] out · 100
where [i]in and [i]out represent the concentrations of reactants (i = CO2) or products (i = CO or CH4) at the inlet and outlet of the reactor, respectively. Fin and Fout are the total flow rates (cm3/min) at the inlet and outlet of the reactor, respectively. Detection of other carbonaceous products as well as elemental carbon was either negligible or nonexistent, thus only CO2, CH4 and CO were included in the calculations of carbon balance, which closes within ±6%.
Reaction rates were defined in terms of the rate of moles of CO2 consumed per both mass (rm) and surface area (rs) of the catalyst:
r m ( mol   CO 2 · g 1 · s 1 ) = [ CO 2 ] in · F in · X CO 2 100 · 60 · m cat · V m
r s ( mol   CO 2 · m 2 · s 1 ) = r m / S BET
where mcat (g) is the mass of the catalyst, SBET (m2/g) is the surface area of the catalyst and Vm is the gas molar volume at 25 °C and 1 bar (24,436 cm3/mol).
The thermodynamic equilibrium calculations were elaborated using the minimization of Gibbs free energy mathematical model (RGibbs) in Aspen Plus® software (Aspen Technology, Inc., Bedford, MA, USA) and the simulation results are included in the catalytic performance plots. The following components were included in the model: CO2, H2, CO, H2O, CH4, CH3OH, CH3OCH3, HCOOH, C2H4 and C2H6. Only the first five substances appeared to be formed in a significant amount in the equilibrium mixture, when all possible reactions between CO2 and H2 were considered. Properties of the substances were determined using the Soave–Redlich–Kwong (SRK) equation of state, a widely used method in gas-processing simulation.

3. Results and Discussion

3.1. Morphological Characterization (TEM)

In order to gain insight into the morphology of bare CeO2 and M/CeO2 samples, TEM analysis was performed. The NR samples (Figure 1a–c) display ceria in a uniform rod-like morphology, while the existence of nanocubic morphology is evident in the samples denoted as NC (Figure 1d–f). Apparently, the incorporation of the active metal phase into the CeO2 lattice has no effect on the support morphology, since distinctive rod and cubic particles are still clearly observed after the incorporation of the metal phase on the well-defined support structure. This also confirms the findings of XRD studies, where the structural features of catalysts remained essentially unaffected after the incorporation of cobalt and copper into the ceria support.

3.2. Textural and Structural Characterization (Brunauer–Emmett–Teller (BET), X-ray Diffraction (XRD))

The textural, structural and redox properties of bare ceria as well as of M/CeO2 catalysts are presented in Table 1. Bare ceria supports, i.e., CeO2-NR and CeO2-NC, exhibit a BET surface area of 79 and 37 m2/g, respectively. The incorporation of transition metals into ceria carriers slightly decreases the BET area. However, the order obtained for bare supports remained unaffected, that is, nanorod samples exhibit higher surface area than the corresponding nanocubic ones, regardless of the nature of the metal phase incorporated into the ceria carrier.
According to XRD results (Figure 2), the main peaks for all samples at 28.5°, 33.1°, 47.5°, 56.3°, 59.1°, 69.4°, 76.7°, and 79.1° correspond to the (111), (200), (220), (311), (222), (400), (331) and (420) planes of a face-centered cubic (FCC) fluorite structure of ceria (Fm3m symmetry, no. 225) [69] with (111), (220) and (311) planes mostly prevailing in the samples structure. Moreover, all M/CeO2 samples exhibit smaller peaks indexed to the corresponding oxide due to the incorporation of metal phase (~8 wt.% Co or Cu) into ceria support. Peaks corresponding to CuO crystal phases at 2θ = 35.3°, 38.2° and a less distinguishable peak at 62° are observed for Cu/CeO2 samples, indicating heterodispersion or aggregation of copper species on the surface of ceria [70]. For Cu/CeO2-NC, the double peak at 43–44° as well as the peak at 50° are present due to the sample holder. The Co/CeO2 samples show small peaks characteristic of Co3O4 at 2θ ~ 36°, 44° and 66° [71].
The average crystallite size of CeO2, CuO and Co3O4 phases for both bare ceria and M/Ceria samples of different morphology are summarized in Table 1. Ceria particles are following the order: CeO2-NC (27 nm) > Co/CeO2-NC (24 nm) > Cu/CeO2-NC (19 nm) > CeO2-NR (15 nm) > Co/CeO2-NR (14 nm) > Cu/CeO2-NR (12 nm). These values are close to those reported in literature for ceria-supported materials [52,72]. It is worth mentioning that ceria crystallite size of all samples with nanocubic morphology is higher than that of the corresponding samples with rod-like morphology. Moreover, no significant changes were observed on ceria particle sizes upon the incorporation of Cu or Co, implying that the structural features of ceria remain unaffected by metal addition, as further verified by TEM analysis (see above). Regarding the crystallite size of active metal phases, the CuO size is 52 and 43 nm for Cu/CeO2-NC and Cu/CeO2-NR, respectively. The corresponding values for Co3O4 phase are much lower, i.e., 19 and 16 nm, denoting a better dispersion of cobalt compared to copper on ceria support. Similar conclusions have been previously obtained in a series of CeO2 supported transition metal catalysts (Fe, Co, Ni, Cu) synthesized by the incipient wet impregnation method [73]. It is also worth mentioning that the crystallite size of both active phases (Co3O4 or CuO) and ceria nanoparticles follow the same trend, implying that the structural characteristics of CeO2 can determine the particle size of metal oxides (Co3O4 or CuO), in agreement with relevant literature studies [74,75]. Moreover, it should be noted that the crystallite size of both ceria and Co3O4 or CuO phases follows the reverse order of BET area, indicating an agglomeration upon the decrease of surface area.

3.3. Redox Properties (Hydrogen Temperature-Programmed Reduction (H2-TPR))

The redox properties of as-prepared catalysts were evaluated by means of hydrogen temperature-programmed reduction (H2-TPR) studies. Figure 3 presents the reduction profiles of all samples. In Table 1 the redox properties, in terms of H2 consumption (mmol/g) and maximum temperature of main TPR peaks, are summarized. The reduction profiles of bare CeO2 samples (Figure 3a) consist of two broad peaks centered at 545–590 °C and 790–810 °C, which are attributed to ceria surface (Os) and bulk oxygen (Ob) reduction, respectively [76,77]. Both peaks are exhibited in lower temperatures for the nanorod CeO2 with the effect being more intense for the ceria surface oxygen. Similar behavior is displayed also for the impregnated Co- and Cu-based samples, where both peaks are shifted to substantially lower temperatures, indicating the beneficial role of active phases in the overall redox properties of the as-prepared samples.
The corresponding TPR profiles of Co/CeO2 and Cu/CeO2 samples along with those of bare Co3O4 and CuO, are shown in Figure 3b,c, respectively. Interestingly, all M/CeO2 samples exhibit reduction peaks at significantly lower temperatures than bare oxides, which implies the synergy between the metal phase and CeO2 towards an improved reducibility. In particular, the Co/CeO2 samples exhibit two main reduction peaks at 318–335 °C (peak a) and 388–405 °C (peak b), attributed to the subsequent reduction of Co3O4 to CoO and CoO to metallic Co, respectively [78]. As for the Cu/CeO2 samples, the low-temperature peak (peak a) in the range of 181–194 °C is attributed to the reduction of finely dispersed CuOx species strongly interacting with the ceria surface [68]. The peak at higher temperature (peak b) can be attributed to the formation of larger CuO clusters on the ceria surface [79].
It is obvious that the reduction of M/CeO2 samples occurs at a significantly lower temperature than those of bare ceria samples, due to the synergy between the two oxide phases that weakens the metal-oxygen bonds [80]. It is also of worth pointing out the facilitation of the ceria surface oxygen reduction in the presence of metal phase (Co or Cu), which results in overlapping bands in the low-temperature range of the TPR profiles of M/CeO2 samples (Figure 3b,c).
In order to gain further insight into the impact of support morphology, as well as of the nature of the active phase on the redox properties of the as-prepared samples, the H2 consumption in the low temperature range (100–600 °C), corresponding to the surface oxygen reduction of both active phase and ceria, has been estimated (Table 1). CeO2-NR exhibits higher values (0.59 mmol H2·g−1) than CeO2-NC (0.41 mmol H2·g−1), implying the enhanced reducibility and oxygen mobility of ceria-nanorods. These findings are in complete agreement with the in situ Raman analysis performed in our previous work, in which nanorods exhibited the highest amount in defects and oxygen vacancies [68]. In a similar manner, the M/CeO2 samples of nanorod morphology exhibit higher H2 consumption than nanocubes in all cases, indicating their abundance in loosely-bound oxygen species. Regarding the impact of metal entity, the total hydrogen consumption follows, independently on support morphology, the order: Co/CeO2 > Cu/CeO2 > CeO2, matching the catalytic performance in terms of CO2 conversion (see below). It should also be noted that hydrogen consumption exceeds the theoretical amount required for the complete reduction of the oxidized transition metal phases for all M/CeO2 samples (Table 1). The latter implies the facilitation of ceria capping oxygen reduction in the presence of a metal entity, further corroborating the synergistic effect between the metal and ceria carrier.

3.4. Catalytic Evaluation Studies

3.4.1. CO2 Hydrogenation Activity

The catalytic performance of the as-prepared samples in the CO2 hydrogenation reaction was investigated in the temperature range of 200–500 °C. A commercial CeO2 sample (Fluka, SBET = 15 m2/g), denoted as CeO2-comm, was also tested as a reference sample. Figure 4 depicts CO2 conversion of all samples in the temperature range investigated, as compared to the thermodynamic equilibrium CO2 conversion profiles for methanation and rWGS reactions. In Table 2 the results from catalytic evaluation studies, in terms of both CO2 conversion and normalized rates at 400 °C, are presented comparatively for all samples.
The beneficial effect of the synthesis method on the catalytic activity is obvious, as all nano-ceria samples exhibit far better CO2 conversion values than CeO2-comm. Moreover, the incorporation of a metal phase into ceria leads to a dramatic improvement of the catalytic activity. Specifically, Co/CeO2-NX, Cu/CeO2-NX, CeO2-NX and CeO2-comm catalysts exhibited CO2 conversion values at 400 °C of approximately 86%, 52%, 20% and 8%, respectively. The observed order of CO2 conversion generally correlates with the amount of consumed hydrogen calculated by H2-TPR in Table 1.
The increase in temperature obviously increases the CO2 conversion, but to a different extent for each sample. As expected, CO2 conversion initially increases and reaches a plateau for the highly selective towards methane Co/CeO2-NX catalysts, favoring the exothermic CO2 methanation (Equation (2)) below 500 °C, in agreement with the thermodynamic calculations. A different trend was shown for bare ceria and Cu/CeO2-NX samples, as CO2 conversion increases steadily, but to a lesser extent, as compared to Co/CeO2-NX samples, approaching in the case of Cu-based catalysts the equilibrium for the rWGS reaction. Bare ceria carriers, although clearly favoring the rWGS reaction, demonstrated CO2 conversion values that are well below the corresponding values predicted by thermodynamics.
The impact of the nature of the metal phase (Cu, Co) on the CO2 hydrogenation performance is further evaluated on the basis of the selectivity towards CH4 and CO, depicted in Figure 5. Apparently, commercial ceria, nano-ceria and Cu/CeO2-NX samples are all highly selective to CO (>90%). Thus, maximum CO2 conversion values for these samples are expected to be closer to those for the rWGS reaction equilibrium, a much less favorable reaction than CO2 methanation, as indeed shown by the corresponding equilibrium curve (dotted line) in Figure 4. Similar results for CO production over CeO2 [59,81] and Cu-based catalysts are reported in literature (see Table 3). Interestingly, Cu/CeO2-NR reaches equilibrium conversion values at ca. 380 °C, well below that reported for many rWGS catalysts. On the contrary, the addition of cobalt into CeO2 leads to a completely different trend, since the selectivity towards CH4 for all Co/CeO2 samples is approximately 95% at temperatures higher than ca. 400 °C. However, at temperatures below 400 °C, selectivity towards CO is significant, decreasing rapidly for higher temperatures.
In light of the above results, it should be highlighted that the as-prepared cobalt catalysts are highly efficient in methane production at temperatures as low as 400 °C, under the reaction conditions employed, being superior, or at least comparable, to the most active Co-based catalysts reported in literature (see Table 3). On the other hand, the as-prepared copper catalysts are highly selective to CO, reaching rWGS equilibrium values at ca. 380°C. In a similar manner, Dai et al. [82] reported 100% selectivity to CO for Cu/CeO2, attributed to the presence of Cu0 active sites, whereas selectivity to CH4 for Co/CeO2 increased markedly between 350°C and 400 °C.
The obtained differences in activity/selectivity between Cu- and Co-based catalysts can be corroborated by taking into account the underlying mechanism of Cu- and Co-catalyzed CO2 hydrogenation reaction, in conjunction with the present characterization results. In particular, the dissociation of adsorbed CO, considered as the rate-determining step of the CO2 methanation process [92,93], has been proposed to proceed by two main pathways: direct COads dissociation and H-assisted COads dissociation. The first mechanism is proposed to occur over group VIII metal-based catalysts, such as Co [94,95]. In view of this fact, Liu et al. [48] demonstrated that Co showed more favorable thermodynamics and lower CO2 decomposition barriers for CO2 reduction compared to Cu. In a similar manner, a close relationship between the CO2 and H2 adsorption capacity of Co/KIT-6 catalysts and their CO2 conversion/selectivity performance has been revealed [87]. In particular, high H2 adsorption capacity can provide a large number of active H species for the further hydrogenation of intermediate species (such as HCOO) to methane, whereas the low H2 adsorption and activation capacity is favorable to CO formation. Considering the fact that as-prepared Co/CeO2-NX possess a higher reducibility than Cu/CeO2-NX in terms of H2 uptake (Table 1), this could be an explanation for the high methane selectivity exhibited for the Co-based samples.
By contrast, a redox mechanism has been proposed in the literature for rWGS reaction over Cu-based catalysts [29,96]. Specifically, Cu0 atoms can act as active sites for the dissociation of CO2 and the Cu2O formed is subsequently reduced by hydrogen to regenerate metallic Cu species. Hydrogen was proposed to be only a reducing agent without direct participation in the formation of intermediate species in the rWGS reaction [97]. The scheme can be simplified to the following reactions (Equations (8) and (9)):
CO2 + 2Cu → Cu2O + CO
H2 + Cu2O → 2Cu + H2O
The facile reduction of Cu/CeO2-NX catalysts towards reduced copper species at temperatures lower than ca. 300 °C, corroborated by the H2-TPR results (Figure 3), can possibly favor this redox mechanism, thus being the reason for the high selectivity towards CO even at low temperatures. However, hydrogenation of carbon dioxide proceeds through a rather complex pathway, thus more detailed work has to be undertaken in order to elucidate the differences in the mechanisms of both CO2 methanation and rWGS and in the product distribution depending on the metal active phase of the catalyst. To this end, mechanistic studies are currently under development in our laboratory.
Regarding the impact of support morphology, it can be concluded that the CO2 conversion and selectivity of the samples are not significantly affected by the support morphology (NR vs. NC), as shown in Figure 4 and Figure 5. In particular, a slightly better conversion performance is obtained for Cu/CeO2-NR samples as compared to Cu/CeO2-NC, whereas for bare CeO2 and Co/CeO2 minor differences in conversion performance between NC and NR were observed. This can also be demonstrated by the mass-normalized reactions rates, as values for NRs are very close to those for NCs (Table 2).
It should be mentioned, however, that NC samples exhibit two to three times lower BET surface area as compared to NR samples (Table 1), which should be further accounted for the different catalytic activity of NC vs. NR samples. In order to more precisely gain insight into the impact of support morphology on the intrinsic reactivity, the specific activity normalized per unit of surface area (µmol CO2·m−2·s−1) was thus calculated (Equation (7)) and is summarized in Table 2. It is evident that the area-normalized reaction rates of nanocubic samples are considerably higher compared to their nanorod counterparts. Specifically, Co/CeO2-NC exhibits the highest rate (1.07 µmol CO2·m−2·s−1), followed by Cu/CeO2-NC (0.51 µmol CO2·m−2·s−1) and CeO2-NC (0.18 µmol CO2·m−2·s−1). Samples with nanorod morphology exhibit almost half of the reaction rates shown by the nanocubic samples, with values of 0.40, 0.25 and 0.09 µmol CO2·m−2·s−1 for Co/CeO2-NR, Cu/CeO2-NR and CeO2-NR, respectively.
In a similar manner, it was found that Ru catalysts supported on ceria nanocubes showed higher CO2 methanation activity, in comparison with ceria nanorods and nanopolyhedra, on the basis of specific rate [98]. Kovacevic et al. [81] synthesized CeO2 nanoparticles with distinct morphology for the rWGS reaction and reported that nanocubes were more active than nanorods in terms of CO produced per surface area, owing to the greater inherent reactivity of (100) crystal planes enclosing cubes, in contrast with the inherently less reactive (111) facets exposed by rods and particles.

3.4.2. Effect of H2:CO2 Ratio

The effect of the H2:CO2 feed ratio (9:1, 6:1, 4:1 and 1:1) on the conversion and selectivity performance was next explored over Co/CeO2-NR and Cu/CeO2-NR samples. Reactant feed ratio is an important factor in the CO2 hydrogenation process, as it modifies the thermodynamic equilibrium of the system [99] and could increase selectivity of a specific product [10,100]. Moreover, knowledge on a catalyst performance under different reaction mixtures is important, especially with the scope of employing the catalytic system in an industrial-scale process, where hydrogen flow may vary, due to the inherent fluctuations associated with the available excess RES power [101].
The CO2 conversion along with the selectivities to CH4 and CO, as a function of temperature, for the different H2/CO2 feed molar ratios investigated, are depicted in Figure 6. The corresponding equilibrium curves for each reaction system are also shown for comparison. It is evident that the increase of H2:CO2 ratio exerted a beneficial effect on the CO2 conversion values, according to thermodynamics.
In particular, for Co/CeO2-NR catalysts, CO2 conversion values at 450 °C of 23, 61, 81, and 90% at a H2:CO2 ratio of 1, 4, 6 and 9, were obtained, respectively. When equal amounts of CO2 and H2 were fed into the reactor, CO2 conversion first increased up to 14% at 360 °C, decreased until 400 °C and then re-increased to a value of 26% at 500 °C. This behavior is expected thermodynamically, as sub-stoichiometric conditions for CO2 methanation were employed and rWGS was thus favored, allowing the conversion to exceed the theoretical quasi-limit calculated by including only CH4 as a product (green dotted line in Figure 6a). Similar results were found on thermodynamic analyses for CO2 methanation in literature [36,100,102]. It is worth mentioning that Co/CeO2-NR almost reaches maximum CO2 conversion values at ca. 400 °C for over-stoichiometric conditions and ca. 420 °C in the case of stoichiometric ratio. In addition, CO2 conversion approached values as high as 90% of the thermodynamic equilibrium ones, implying the superiority of Co/CeO2 catalysts for the Sabatier reaction.
Higher H2:CO2 feed ratios are likely to promote methane formation instead of the reverse WGS reaction, since Equation (2) is more dependent on hydrogen. Indeed, at 450 °C, selectivity values towards CH4 equal to 97.5%, 93.6%, 88.2%, and 38.1% were observed for a feed ratio of 9:1, 6:1, 4:1 and 1:1, respectively (Figure 6b). CH4 selectivity decreased significantly at temperatures higher than 380 °C when the H2:CO2 ratio was unity, reaching a value of 26% at 500 °C. The opposite trend was observed for CO selectivity. Traces of C2 hydrocarbons were also observed for Co/CeO2-NR, especially under H2-limiting conditions. In general, significant CO2 conversion to CH4 can be achieved on Co/CeO2-NR, even when hydrogen availability decreases, as long as the system is operating far from sub-stoichiometry.
For Cu/CeO2-NR, CO2 conversion at 450 °C increased from 24.2% to 58.1% as the feed ratio increased from 1:1 to 9:1, respectively, in agreement with the equilibrium CO2 conversion to CO. Considering that equimolar amounts of CO2 and H2 are required for the stoichiometric rWGS reaction to take place, any excess amount of hydrogen in the feed is expected to facilitate CO production. Regardless of the hydrogen inlet concentration, CO2 conversion approached the respective equilibrium values within ~5% at around 380 °C. Several catalytic systems employed for rWGS have been reported to function under thermodynamic limitation at the medium-high or high temperature regime [103,104,105]. As for the selectivity to CH4, it increased slightly, from about 1.0% for a H2:CO2 ratio of 1:1 to 5.1% for a ratio of 9:1. Correspondingly, CO selectivity decreased slightly with an increase in H2:CO2 ratio, with selectivity values at 500 °C equal to 95.3%, 97.1%, 98.2% and 99.5%, at H2:CO2 ratios of 9, 6, 4 and 1, respectively. Contrary to Co/CeO2-NR, no other carbonaceous products were identified in the case of Cu/CeO2, at the reaction conditions examined. Low methane selectivity values exhibited by Cu/CeO2-NR, even under excess hydrogen in the reactant stream, further substantiate the possibility of a redox mechanism over this catalyst. Excess hydrogen may directly re-reduce Cu2O species formed by CO2 dissociation over metallic Cu, rather than fully hydrogenating intermediate species, thus leading to high CO production, as discussed above.

3.4.3. Stability Tests

Further examination of the catalytic performance of Co/CeO2-NR and Cu/CeO2-NR was conducted in short-term stability experiments under isothermal conditions, in order to assess the catalysts lifetime characteristics and to detect possible deactivation phenomena. Herein, the aforementioned samples were further tested for 12 h at a constant temperature of 450 °C, with a feed ratio of H2:CO2 equal to 9:1 and a GHSV of 20,000 h−1. The results of CO2 conversion and selectivity values as a function of time are presented in Figure 7.
Copper-based catalysts are generally known for their inferior catalytic performance under high-temperature operation, due to the sintering of copper particles resulting in a decrease in active sites [106,107]. However, CO2 conversion on Cu/CeO2-NR remained relatively stable for the duration of the experiment, fluctuating between an average value of 52%, decreasing very slightly from the value of 56% found for the fresh catalyst during the corresponding light-off tests. It is also noted that the thermodynamic value for the rWGS reaction, under the employed conditions, is 63%. Moreover, the selectivity towards CO stabilized to ca. 96% after approximately 1 h. Thus, the Cu/CeO2-NR remained highly active towards CO production and exhibited very low methane yield even after 12 h at 450 °C, under hydrogen excess conditions, further corroborating the fact that CO2 methanation is largely unfavorable over this catalyst.
In the course of the methanation reaction, the stability of a catalyst is closely linked to coking and metal sintering [108,109]. For Co/CeO2-NR, it was demonstrated that CO2 conversion remained very stable at ~90%, whereas the thermodynamic value is calculated at 99.5% (Figure 7). At the same time, the selectivity towards CH4 remained stable at 96% for the whole duration of experiment, revealing the superiority of Co/CeO2-NR sample in CO2 methanation process, in terms of conversion, selectivity and stability.

4. Conclusions

The present results revealed the strong effect of metal phase (Cu or Cu) and reaction conditions on the CO2 hydrogenation performance of nanoceria-based M/CeO2 catalysts of different morphology (nanorods or nanocubes). It was shown that hydrothermally synthesized nano-ceria carriers exhibited better catalytic activity in the hydrogenation of CO2 than commercial ceria. More importantly, incorporating Co and Cu into the nanostructured ceria support led to a significant increase in the catalytic activity, with CO2 conversion following the order Co/CeO2 > Cu/CeO2 > CeO2. This order correlates with the hydrogen consumption estimated by H2-TPR, implying that the enhanced reducibility—linked to the different metal phase—probably favors the CO2 hydrogenation reactivity. The main product of the reaction was determined by the nature of the metal entity incorporated into cerium oxide, with Co/CeO2 exhibiting high selectivity towards CH4 and CO2 conversion values close to equilibrium. Cu/CeO2 exhibited remarkable CO selectivity even under hydrogen excess conditions and very close to equilibrium CO2 conversion values for the rWGS reaction. From a practical point of view, Co/CeO2 catalysts demonstrated an excellent conversion and selectivity performance, offering ~85% yield to methane at 450 °C. On the other hand, Cu/CeO2 samples were very selective for CO production, exhibiting 52% CO2 conversion and 95% CO selectivity at ca. 400 °C.
In both cases a stable conversion and selectivity performance (either for CO2 methanation or rWGS reaction) was attained in short-term (12 h) stability tests. Regarding the impact of ceria morphology, the samples supported on ceria nanocubes exhibited higher specific activity (µmol CO2·m−2·s−1), as compared to samples of rod-like shape, revealing the significant role of support morphology, besides that of metal nature (Co or Cu). These results are considered to be very promising in the sense of employing these catalysts in real-scale processes with variable H2:CO2 ratios, where renewable hydrogen from excess RES power can be used to efficiently and selectively convert CO2 to CO or CH4.

Author Contributions

M.L. contributed to materials synthesis and results interpretation; S.S. and S.A.C.C. contributed to materials characterizations; G.V. and E.P. contributed to catalytic evaluation studies; M.K. and G.E.M. conceptualized the overall idea of this research, validated the results and administered the project. G.V. wrote the original draft. Finally, M.K. reviewed, edited and submitted the manuscript in the final form. All authors read and approved the final version of the present manuscript.

Funding

This research has been co-financed by the European Union and Greek national funds through the Operational Program Competitiveness, Entrepreneurship and Innovation, under the call RESEARCH—CREATE—INNOVATE (project code: T1EDK-00094). This work was also financially supported by Associate Laboratory LSRE-LCM—UID/EQU/50020/2019—funded by national funds through FCT/MCTES (PIDDAC). S.A.C.C. acknowledges Fundação para a Ciência e a Tecnologia (Portugal) for Investigador FCT program (IF/01381/2013/CP1160/CT0007), with financing from the European Social Fund and the Human Potential Operational Program.

Acknowledgments

We are grateful to Pedro Tavares (UTAD) for assistance with the TEM and XRD analyses and to Nuno Costa (Requimte) for assistance with N2 adsorption-desorption measurements.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Olah, G.A.; Prakash, G.K.S.; Goeppert, A. Anthropogenic chemical carbon cycle for a sustainable future. J. Am. Chem. Soc. 2011, 133, 12881–12898. [Google Scholar] [CrossRef] [PubMed]
  2. Wang, W.; Wang, S.; Ma, X.; Gong, J. Recent advances in catalytic hydrogenation of carbon dioxide. Chem. Soc. Rev. 2011, 40, 3703–3727. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Rodriguez, J.A.; Liu, P.; Stacchiola, D.J.; Senanayake, S.D.; White, M.G.; Chen, J.G. Hydrogenation of CO2 to Methanol: Importance of Metal-Oxide and Metal-Carbide Interfaces in the Activation of CO2. ACS Catal. 2015, 5, 6696–6706. [Google Scholar] [CrossRef]
  4. Xu, X.; Moulijn, J.A. Mitigation of CO2 by chemical conversion: Plausible chemical reactions and promising products. Energy Fuels 1996, 10, 305–325. [Google Scholar] [CrossRef]
  5. Centi, G.; Perathoner, S. Opportunities and prospects in the chemical recycling of carbon dioxide to fuels. Catal. Today 2009, 148, 191–205. [Google Scholar] [CrossRef]
  6. Sabatier, P.; Senderens, J.B. New Synthesis of Methane. Comptes Rendus Hebd. Seances Acad. Scrences 1902, 134, 514–516. [Google Scholar]
  7. Bahruji, H.; Bowker, M.; Hutchings, G.; Dimitratos, N.; Wells, P.; Gibson, E.; Jones, W.; Brookes, C.; Morgan, D.; Lalev, G. Pd/ZnO catalysts for direct CO2 hydrogenation to methanol. J. Catal. 2016, 343, 133–146. [Google Scholar] [CrossRef] [Green Version]
  8. Li, M.M.J.; Zeng, Z.; Liao, F.; Hong, X.; Tsang, S.C.E. Enhanced CO2 hydrogenation to methanol over CuZn nanoalloy in Ga modified Cu/ZnO catalysts. J. Catal. 2016, 343, 157–167. [Google Scholar] [CrossRef]
  9. Senanayake, S.D.; Ramírez, P.J.; Waluyo, I.; Kundu, S.; Mudiyanselage, K.; Liu, Z.; Liu, Z.; Axnanda, S.; Stacchiola, D.J.; Evans, J.; et al. Hydrogenation of CO2 to Methanol on CeOx/Cu(111) and ZnO/Cu(111) Catalysts: Role of the Metal-Oxide Interface and Importance of Ce3+ Sites. J. Phys. Chem. C 2016, 120, 1778–1784. [Google Scholar] [CrossRef] [Green Version]
  10. Vourros, A.; Garagounis, I.; Kyriakou, V.; Carabineiro, S.A.C.; Maldonado-Hódar, F.J.; Marnellos, G.E.; Konsolakis, M. Carbon dioxide hydrogenation over supported Au nanoparticles: Effect of the support. J. CO2 Util. 2017, 19, 247–256. [Google Scholar] [CrossRef]
  11. Aguayo, A.T.; Ereña, J.; Sierra, I.; Olazar, M.; Bilbao, J. Deactivation and regeneration of hybrid catalysts in the single-step synthesis of dimethyl ether from syngas and CO2. Catal. Today 2005, 106, 265–270. [Google Scholar] [CrossRef]
  12. Zhang, Y.; Fei, J.; Yu, Y.; Zheng, X. Silica immobilized ruthenium catalyst used for carbon dioxide hydrogenation to formic acid (I): The effect of functionalizing group and additive on the catalyst performance. Catal. Commun. 2004, 5, 643–646. [Google Scholar] [CrossRef]
  13. 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]
  14. Kaiser, P.; Unde, R.B.; Kern, C.; Jess, A. Production of liquid hydrocarbons with CO2 as carbon source based on reverse water-gas shift and Fischer-Tropsch synthesis. Chem. Ing. Tech. 2013, 85, 489–499. [Google Scholar] [CrossRef]
  15. Pastor-Pérez, L.; Baibars, F.; Le Sache, E.; Arellano-García, H.; Gu, S.; Reina, T.R. CO2 valorisation via Reverse Water-Gas Shift reaction using advanced Cs doped Fe-Cu/Al2O3 catalysts. J. CO2 Util. 2017, 21, 423–428. [Google Scholar] [CrossRef] [Green Version]
  16. Daza, Y.A.; Kuhn, J.N. CO2 conversion by reverse water gas shift catalysis: Comparison of catalysts, mechanisms and their consequences for CO2 conversion to liquid fuels. RSC Adv. 2016, 6, 49675–49691. [Google Scholar] [CrossRef]
  17. Rönsch, S.; Schneider, J.; Matthischke, S.; Schlüter, M.; Götz, M.; Lefebvre, J.; Prabhakaran, P.; Bajohr, S. Review on methanation—From fundamentals to current projects. Fuel 2016, 166, 276–296. [Google Scholar] [CrossRef]
  18. 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]
  19. Tada, S.; Ochieng, O.J.; Kikuchi, R.; Haneda, T.; Kameyama, H. Promotion of CO2 methanation activity and CH4 selectivity at low temperatures over Ru/CeO2/Al2O3 catalysts. Int. J. Hydrog. Energy 2014, 39, 10090–10100. [Google Scholar] [CrossRef]
  20. Klankermayer, J.; Wesselbaum, S.; Beydoun, K.; Leitner, W. Selective Catalytic Synthesis Using the Combination of Carbon Dioxide and Hydrogen: Catalytic Chess at the Interface of Energy and Chemistry. Angew. Chem. Int. Ed. 2016, 55, 7296–7343. [Google Scholar] [CrossRef]
  21. Li, W.; Wang, H.; Jiang, X.; Zhu, J.; Liu, Z.; Guo, X.; Song, C. A short review of recent advances in CO2 hydrogenation to hydrocarbons over heterogeneous catalysts. RSC Adv. 2018, 8, 7651–7669. [Google Scholar] [CrossRef] [Green Version]
  22. Mazza, A.; Bompard, E.; Chicco, G. Applications of power to gas technologies in emerging electrical systems. Renew. Sustain. Energy Rev. 2018, 92, 794–806. [Google Scholar] [CrossRef]
  23. Lewandowska-Bernat, A.; Desideri, U. Opportunities of Power-to-Gas technology. Energy Procedia 2017, 105, 4569–4574. [Google Scholar] [CrossRef]
  24. Van Leeuwen, C.; Mulder, M. Power-to-gas in electricity markets dominated by renewables. Appl. Energy 2018, 232, 258–272. [Google Scholar] [CrossRef]
  25. Götz, M.; Lefebvre, J.; Mörs, F.; McDaniel Koch, A.; Graf, F.; Bajohr, S.; Reimert, R.; Kolb, T. Renewable Power-to-Gas: A technological and economic review. Renew. Energy 2016, 85, 1371–1390. [Google Scholar] [CrossRef] [Green Version]
  26. Ghaib, K.; Ben-Fares, F.Z. Power-to-Methane: A state-of-the-art review. Renew. Sustain. Energy Rev. 2018, 81, 433–446. [Google Scholar] [CrossRef]
  27. Zangeneh, F.T.; Sahebdelfar, S.; Ravanchi, M.T. Conversion of carbon dioxide to valuable petrochemicals: An approach to clean development mechanism. J. Nat. Gas Chem. 2011, 20, 219–231. [Google Scholar] [CrossRef]
  28. Ma, J.; Sun, N.; Zhang, X.; Zhao, N.; Xiao, F.; Wei, W.; Sun, Y. A short review of catalysis for CO2 conversion. Catal. Today 2009, 148, 221–231. [Google Scholar] [CrossRef]
  29. 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]
  30. Centi, G.; Perathoner, S. Heterogeneous catalytic reactions with CO2: Status and perspectives. Stud. Surf. Sci. Catal. 2004, 153, 1–8. [Google Scholar] [CrossRef]
  31. 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]
  32. Porosoff, M.D.; Yan, B.; Chen, J.G. Catalytic reduction of CO2 by H2 for synthesis of CO, methanol and hydrocarbons: Challenges and opportunities. Energy Environ. Sci. 2016, 9, 62–73. [Google Scholar] [CrossRef]
  33. Ghaib, K.; Nitz, K.; Ben-Fares, F.Z. Chemical Methanation of CO2: A Review. ChemBioEng Rev. 2016, 3, 266–275. [Google Scholar] [CrossRef]
  34. Frontera, P.; Macario, A.; Ferraro, M.; Antonucci, P. Supported Catalysts for CO2 Methanation: A Review. Catalysts 2017, 7, 59. [Google Scholar] [CrossRef]
  35. Scibioh, M.A.; Viswanathan, B. Chapter 5—Heterogeneous Hydrogenation of CO2. Carbon Dioxide Chem. Fuels 2018, 191–253. [Google Scholar] [CrossRef]
  36. Gao, J.; Liu, Q.; Gu, F.; Liu, B.; Zhong, Z.; Su, F. Recent advances in methanation catalysts for the production of synthetic natural gas. RSC Adv. 2015, 5, 22759–22776. [Google Scholar] [CrossRef]
  37. Winter, L.R.; Chen, R.; Chen, X.; Chang, K.; Liu, Z.; Senanayake, S.D.; Ebrahim, A.M.; Chen, J.G. Elucidating the roles of metallic Ni and oxygen vacancies in CO2 hydrogenation over Ni/CeO2 using isotope exchange and in situ measurements. Appl. Catal. B Environ. 2019, 245, 360–366. [Google Scholar] [CrossRef]
  38. Melchionna, M.; Fornasiero, P. The role of ceria-based nanostructured materials in energy applications. Mater. Today 2014, 17, 349–357. [Google Scholar] [CrossRef]
  39. Montini, T.; Melchionna, M.; Monai, M.; Fornasiero, P. Fundamentals and Catalytic Applications of CeO2-Based Materials. Chem. Rev. 2016, 116, 5987–6041. [Google Scholar] [CrossRef] [PubMed]
  40. Hemalatha, P.; Bhagiyalakshmi, M.; Ganesh, M.; Palanichamy, M.; Murugesan, V.; Jang, H.T. Role of ceria in CO2 adsorption on NaZSM-5 synthesized using rice husk ash. J. Ind. Eng. Chem. 2012, 18, 260–265. [Google Scholar] [CrossRef]
  41. Cargnello, M.; Doan-Nguyen, V.V.T.; Gordon, T.R.; Diaz, R.E.; Stach, E.A.; Gorte, R.J.; Fornasiero, P.; Murray, C.B. Control of metal nanocrystal size reveals metal-support interface role for ceria catalysts. Science 2013, 341, 771–773. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. 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]
  43. Konsolakis, M. The role of Copper–Ceria interactions in catalysis science: Recent theoretical and experimental advances. Appl. Catal. B Environ. 2016, 198, 49–66. [Google Scholar] [CrossRef]
  44. Sharma, S.; Hu, Z.; Zhang, P.; McFarland, E.W.; Metiu, H. CO2 methanation on Ru-doped ceria. J. Catal. 2011, 278, 297–309. [Google Scholar] [CrossRef]
  45. Beuls, A.; Swalus, C.; Jacquemin, M.; Heyen, G.; Karelovic, A.; Ruiz, P. Methanation of CO2: Further insight into the mechanism over Rh/γ-Al2O3 catalyst. Appl. Catal. B Environ. 2012, 113–114, 2–10. [Google Scholar] [CrossRef]
  46. Díez-Ramírez, J.; Valverde, J.L.; Sánchez, P.; Dorado, F. CO2 Hydrogenation to Methanol at Atmospheric Pressure: Influence of the Preparation Method of Pd/ZnO Catalysts. Catal. Lett. 2016, 146, 373–382. [Google Scholar] [CrossRef]
  47. Spezzati, G.; Benavidez, A.D.; DeLaRiva, A.T.; Su, Y.; Hofmann, J.P.; Asahina, S.; Olivier, E.J.; Neethling, J.H.; Miller, J.T.; Datye, A.K.; et al. CO oxidation by Pd supported on CeO2(100) and CeO2(111) facets. Appl. Catal. B Environ. 2019, 243, 36–46. [Google Scholar] [CrossRef]
  48. 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]
  49. Konsolakis, M.; Ioakeimidis, Z. Surface/structure functionalization of copper-based catalysts by metal-support and/or metal-metal interactions. Appl. Surf. Sci. 2014, 320, 244–255. [Google Scholar] [CrossRef]
  50. Park, J.B.; Graciani, J.; Evans, J.; Stacchiola, D.; Senanayake, S.D.; Barrio, L.; Liu, P.; Sanz, J.F.; Hrbek, J.; Rodriguez, J.A. Gold, copper, and platinum nanoparticles dispersed on CeOx/TiO2(110) surfaces: High water-gas shift activity and the nature of the mixed-metal oxide at the nanometer level. J. Am. Chem. Soc. 2010, 132, 356–363. [Google Scholar] [CrossRef]
  51. Kattel, S.; Liu, P.; Chen, J.G. Tuning Selectivity of CO2 Hydrogenation Reactions at the Metal/Oxide Interface. J. Am. Chem. Soc. 2017, 139, 9739–9754. [Google Scholar] [CrossRef] [PubMed]
  52. Owen, R.E.; Plucinski, P.; Mattia, D.; Torrente-Murciano, L.; Ting, V.P.; Jones, M.D. Effect of support of Co-Na-Mo catalysts on the direct conversion of CO2 to hydrocarbons. J. CO2 Util. 2016, 16, 97–103. [Google Scholar] [CrossRef] [Green Version]
  53. Le, T.A.; Kim, T.W.; Lee, S.H.; Park, E.D. Effects of Na content in Na/Ni/SiO2 and Na/Ni/CeO2 catalysts for CO and CO2 methanation. Catal. Today 2018, 303, 159–167. [Google Scholar] [CrossRef]
  54. Sun, C.; Li, H.; Chen, L. Nanostructured ceria-based materials: Synthesis, properties, and applications. Energy Environ. Sci. 2012, 5, 8475–8505. [Google Scholar] [CrossRef]
  55. Yao, X.; Tang, C.; Gao, F.; Dong, L. Research progress on the catalytic elimination of atmospheric molecular contaminants over supported metal-oxide catalysts. Catal. Sci. Technol. 2014, 4, 2814–2829. [Google Scholar] [CrossRef]
  56. Ganduglia-Pirovano, M.V.; Hofmann, A.; Sauer, J. Oxygen vacancies in transition metal and rare earth oxides: Current state of understanding and remaining challenges. Surf. Sci. Rep. 2007, 62, 219–270. [Google Scholar] [CrossRef]
  57. Ouyang, B.; Tan, W.; Liu, B. Morphology effect of nanostructure ceria on the Cu/CeO2 catalysts for synthesis of methanol from CO2 hydrogenation. Catal. Commun. 2017, 95, 36–39. [Google Scholar] [CrossRef]
  58. Si, R.; Flytzani-Stephanopoulos, M. Shape and crystal-plane effects of nanoscale ceria on the activity of Au-CeO2 catalysts for the water-gas shift reaction. Angew. Chem. Int. Ed. 2008, 47, 2884–2887. [Google Scholar] [CrossRef]
  59. Liu, Y.; Li, Z.; Xu, H.; Han, Y. Reverse water-gas shift reaction over ceria nanocube synthesized by hydrothermal method. Catal. Commun. 2016, 76, 1–6. [Google Scholar] [CrossRef]
  60. Yang, S.C.; Pang, S.H.; Sulmonetti, T.P.; Su, W.N.; Lee, J.F.; Hwang, B.J.; Jones, C.W. Synergy between Ceria Oxygen Vacancies and Cu Nanoparticles Facilitates the Catalytic Conversion of CO2 to CO under Mild Conditions. ACS Catal. 2018, 8, 12056–12066. [Google Scholar] [CrossRef]
  61. Muroyama, H.; Tsuda, Y.; Asakoshi, T.; Masitah, H.; Okanishi, T.; Matsui, T.; Eguchi, K. Carbon dioxide methanation over Ni catalysts supported on various metal oxides. J. Catal. 2016, 343, 178–184. [Google Scholar] [CrossRef] [Green Version]
  62. Pan, Q.; Peng, J.; Sun, T.; Wang, S.; Wang, S. Insight into the reaction route of CO2 methanation: Promotion effect of medium basic sites. Catal. Commun. 2014, 45, 74–78. [Google Scholar] [CrossRef]
  63. Le, T.A.; Kim, M.S.; Lee, S.H.; Kim, T.W.; Park, E.D. CO and CO2 methanation over supported Ni catalysts. Catal. Today 2017, 293–294, 89–96. [Google Scholar] [CrossRef]
  64. Aldana, P.A.U.; Ocampo, F.; Kobl, K.; Louis, B.; Thibault-Starzyk, F.; Daturi, M.; Bazin, P.; Thomas, S.; Roger, A.C. Catalytic CO2 valorization into CH4 on Ni-based ceria-zirconia. Reaction mechanism by operando IR spectroscopy. Catal. Today 2013, 215, 201–207. [Google Scholar] [CrossRef]
  65. 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] [Green Version]
  66. 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] [Green Version]
  67. Guo, C.; Wei, S.; Zhou, S.; Zhang, T.; Wang, Z.; Ng, S.P.; Lu, X.; Wu, C.M.L.; Guo, W. Initial Reduction of CO2 on Pd-, Ru-, and Cu-Doped CeO2(111) Surfaces: Effects of Surface Modification on Catalytic Activity and Selectivity. ACS Appl. Mater. Interfaces 2017, 9, 26107–26117. [Google Scholar] [CrossRef]
  68. 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]
  69. Sharma, V.; Eberhardt, K.M.; Sharma, R.; Adams, J.B.; Crozier, P.A. A spray drying system for synthesis of rare-earth doped cerium oxide nanoparticles. Chem. Phys. Lett. 2010, 495, 280–286. [Google Scholar] [CrossRef]
  70. Zhu, P.; Liu, M.; Zhou, R. Effect of interaction between CuO and CeO2 on the performance of CuO-CeO2 catalysts for selective oxidation of CO in H2-rich streams. Indian J. Chem. 2012, 51, 1529–1537. [Google Scholar]
  71. Luo, J.Y.; Meng, M.; Li, X.; Li, X.G.; Zha, Y.Q.; Hu, T.D.; Xie, Y.N.; Zhang, J. Mesoporous Co3O4-CeO2 and Pd/Co3O4-CeO2 catalysts: Synthesis, characterization and mechanistic study of their catalytic properties for low-temperature CO oxidation. J. Catal. 2008, 254, 310–324. [Google Scholar] [CrossRef]
  72. Zabilskiy, M.; Djinović, P.; Tchernychova, E.; Tkachenko, O.P.; Kustov, L.M.; Pintar, A. Nanoshaped CuO/CeO2 Materials: Effect of the Exposed Ceria Surfaces on Catalytic Activity in N2O Decomposition Reaction. ACS Catal. 2015, 5, 5357–5365. [Google Scholar] [CrossRef]
  73. Konsolakis, M.; Ioakimidis, Z.; Kraia, T.; Marnellos, G.E. Hydrogen Production by Ethanol Steam Reforming (ESR) over CeO2 Supported Transition Metal (Fe, Co, Ni, Cu) Catalysts: Insight into the Structure-Activity Relationship. Catalysts 2016, 6, 39. [Google Scholar] [CrossRef] [Green Version]
  74. Gamarra, D.; Lόpez Cámara, A.; Monte, M.; Rasmussen, S.B.; Chinchilla, L.E.; Hungría, A.B.; Munuera, G.; Gyorffy, N.; Schay, Z.; Cortés Corberán, V.; et al. Preferential oxidation of CO in excess H2 over CuO/CeO2 catalysts: Characterization and performance as a function of the exposed face present in the CeO2 support. Appl. Catal. B Environ. 2013, 130–131, 224–238. [Google Scholar] [CrossRef]
  75. Senanayake, S.D.; Sadowski, J.T.; Evans, J.; Kundu, S.; Agnoli, S.; Yang, F.; Stacchiola, D.; Flege, J.I.; Hrbek, J.; Rodriguez, J.A. Nanopattering in CeOx/Cu(111): A New Type of Surface Reconstruction and Enhancement of Catalytic Activity. J. Phys. Chem. Lett. 2012, 3, 839–843. [Google Scholar] [CrossRef]
  76. Lykaki, M.; Pachatouridou, E.; Iliopoulou, E.; Carabineiro, S.A.C.; Konsolakis, M. Impact of the synthesis parameters on the solid state properties and the CO oxidation performance of ceria nanoparticles. RSC Adv. 2017, 7, 6160–6169. [Google Scholar] [CrossRef] [Green Version]
  77. Yu, S.W.; Huang, H.H.; Tang, C.W.; Wang, C.B. The effect of accessible oxygen over Co3O4-CeO2 catalysts on the steam reforming of ethanol. Int. J. Hydrog. Energy 2014, 39, 20700–20711. [Google Scholar] [CrossRef]
  78. 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]
  79. Aboukaïs, A.; Skaf, M.; Hany, S.; Cousin, R.; Aouad, S.; Labaki, M.; Abi-Aad, E. A comparative study of Cu, Ag and Au doped CeO2 in the total oxidation of volatile organic compounds (VOCs). Mater. Chem. Phys. 2016, 177, 570–576. [Google Scholar] [CrossRef]
  80. Jampa, S.; Jamieson, A.M.; Chaisuwan, T.; Luengnaruemitchai, A.; Wongkasemjit, S. Achievement of hydrogen production from autothermal steam reforming of methanol over Cu-loaded mesoporous CeO2 and Cu-loaded mesoporous CeO2–ZrO2 catalysts. Int. J. Hydrog. Energy 2017, 42, 15073–15084. [Google Scholar] [CrossRef]
  81. Kovacevic, M.; Mojet, B.L.; Van Ommen, J.G.; Lefferts, L. Effects of Morphology of Cerium Oxide Catalysts for Reverse Water Gas Shift Reaction. Catal. Lett. 2016, 146, 770–777. [Google Scholar] [CrossRef] [Green Version]
  82. Dai, B.; Zhou, G.; Ge, S.; Xie, H.; Jiao, Z.; Zhang, G.; Xiong, K. CO2 reverse water-gas shift reaction on mesoporous M-CeO2 catalysts. Can. J. Chem. Eng. 2017, 95, 634–642. [Google Scholar] [CrossRef]
  83. Lin, L.; Yao, S.; Liu, Z.; Zhang, F.; Li, N.; Vovchok, D.; Martínez-Arias, A.; Castañeda, R.; Lin, J.; Senanayake, S.D.; et al. In Situ Characterization of Cu/CeO2 Nanocatalysts for CO2 Hydrogenation: Morphological Effects of Nanostructured Ceria on the Catalytic Activity. J. Phys. Chem. C 2018, 122, 12934–12943. [Google Scholar] [CrossRef]
  84. Liu, Y.; Liu, D. Study of bimetallic Cu-Ni/γ-Al2O3 catalysts for carbon dioxide hydrogenation. Int. J. Hydrog. Energy 1999, 24, 351–354. [Google Scholar] [CrossRef]
  85. Chen, C.S.; Cheng, W.H.; Lin, S.S. Study of iron-promoted Cu/SiO2 catalyst on high temperature reverse water gas shift reaction. Appl. Catal. A Gen. 2004, 257, 97–106. [Google Scholar] [CrossRef]
  86. Zhang, X.; Zhu, X.; Lin, L.; Yao, S.; Zhang, M.; Liu, X.; Wang, X.; Li, Y.W.; Shi, C.; Ma, D. Highly Dispersed Copper over β-Mo2C as an Efficient and Stable Catalyst for the Reverse Water Gas Shift (RWGS) Reaction. ACS Catal. 2017, 7, 912–918. [Google Scholar] [CrossRef]
  87. Zhou, G.; Liu, H.; Xing, Y.; Xu, S.; Xie, H.; Xiong, K. CO2 hydrogenation to methane over mesoporous Co/SiO2 catalysts: Effect of structure. J. CO2 Util. 2018, 26, 221–229. [Google Scholar] [CrossRef]
  88. Zhou, G.; Wu, T.; Xie, H.; Zheng, X. Effects of structure on the carbon dioxide methanation performance of Co-based catalysts. Int. J. Hydrog. Energy 2013, 38, 10012–10018. [Google Scholar] [CrossRef]
  89. Zhu, H.; Razzaq, R.; Li, C.; Muhmmad, Y.; Zhang, S. Catalytic Methanation of Carbon Dioxide by Active Oxygen Material CexZr1-xO2 Supported Ni-Co Bimetallic Nanocatalysts. AIChE J. 2013, 59, 2567–2576. [Google Scholar] [CrossRef]
  90. Suslova, E.V.; Chernyak, S.A.; Egorov, A.V.; Savilov, S.V.; Lunin, V.V. CO2 Hydrogenation over Cobalt-Containing Catalysts. Kinet. Catal. 2015, 56, 646–654. [Google Scholar] [CrossRef]
  91. Díez-Ramírez, J.; Sánchez, P.; Kyriakou, V.; Zafeiratos, S.; Marnellos, G.E.; Konsolakis, M.; Dorado, F. Effect of support nature on the cobalt-catalyzed CO2 hydrogenation. J. CO2 Util. 2017, 21, 562–571. [Google Scholar] [CrossRef]
  92. Jwa, E.; Lee, S.B.; Lee, H.W.; Mok, Y.S. Plasma-assisted catalytic methanation of CO and CO2 over Ni-zeolite catalysts. Fuel Process. Technol. 2013, 108, 89–93. [Google Scholar] [CrossRef]
  93. Miao, B.; Ma, S.S.K.; Wang, X.; Su, H.; Chan, S.H. Catalysis mechanisms of CO2 and CO methanation. Catal. Sci. Technol. 2016, 6, 4048–4058. [Google Scholar] [CrossRef]
  94. Zhen, W.; Li, B.; Lu, G.; Ma, J. Enhancing catalytic activity and stability for CO2 methanation on Ni@MOF-5 via control of active species dispersion. Chem. Commun. 2015, 51, 1728–1731. [Google Scholar] [CrossRef] [PubMed]
  95. Weatherbee, G.D.; Bartholomew, C.H. Hydrogenation of CO2 on Group VIII metals: II. Kinetics and Mechanism of CO2 Hydrogenation on Nickel. J. Catal. 1982, 77, 460–472. [Google Scholar] [CrossRef]
  96. Ginés, M.J.L.; Marchi, A.J.; Apesteguía, C.R. Kinetic study of the reverse water-gas shift reaction over CuO/ZnO/Al2O3 catalysts. Appl. Catal. A Gen. 1997, 154, 155–171. [Google Scholar] [CrossRef]
  97. Chen, C.S.; Cheng, W.H. Study on the Mechanism of CO Formation in Reverse Water Gas Shift Reaction Over Cu/SiO2 Catalyst by Pulse Reaction, TPD and TPR. Catal. Lett. 2002, 83, 121–126. [Google Scholar] [CrossRef]
  98. Wang, F.; Li, C.; Zhang, X.; Wei, M.; Evans, D.G.; Duan, X. Catalytic behavior of supported Ru nanoparticles on the {100}, {110}, and {111} facet of CeO2. J. Catal. 2015, 329, 177–186. [Google Scholar] [CrossRef]
  99. Sahebdelfar, S.; Takht Ravanchi, M. Carbon dioxide utilization for methane production: A thermodynamic analysis. J. Pet. Sci. Eng. 2015, 134, 14–22. [Google Scholar] [CrossRef]
  100. Gao, J.; Wang, Y.; Ping, Y.; Hu, D.; Xu, G.; Gu, F.; Su, F. A thermodynamic analysis of methanation reactions of carbon oxides for the production of synthetic natural gas. RSC Adv. 2012, 2, 2358–2368. [Google Scholar] [CrossRef]
  101. Reiter, G.; Lindorfer, J. Global warming potential of hydrogen and methane production from renewable electricity via power-to-gas technology. Int. J. Life Cycle Assess. 2015, 20, 477–489. [Google Scholar] [CrossRef]
  102. Swapnesh, A.; Srivastava, V.C.; Mall, I.D. Comparative study on thermodynamic analysis of CO2 utilization reactions. Chem. Eng. Technol. 2014, 37, 1765–1777. [Google Scholar] [CrossRef]
  103. Chen, C.S.; Cheng, W.H.; Lin, S.S. Study of reverse water gas shift reaction by TPD, TPR and CO2 hydrogenation over potassium-promoted Cu/SiO2 catalyst. Appl. Catal. A Gen. 2003, 238, 55–67. [Google Scholar] [CrossRef]
  104. Porosoff, M.D.; Yang, X.; Boscoboinik, J.A.; Chen, J.G. Molybdenum carbide as alternative catalysts to precious metals for highly selective reduction of CO2 to CO. Angew. Chem. Int. Ed. 2014, 53, 6705–6709. [Google Scholar] [CrossRef] [PubMed]
  105. Yang, L.; Pastor-Pérez, L.; Gu, S.; Sepúlveda-Escribano, A.; Reina, T.R. Highly efficient Ni/CeO2-Al2O3 catalysts for CO2 upgrading via reverse water-gas shift: Effect of selected transition metal promoters. Appl. Catal. B Environ. 2018, 232, 464–471. [Google Scholar] [CrossRef]
  106. Chen, C.S.; Cheng, W.H.; Lin, S.S. Enhanced activity and stability of a Cu/SiO2 catalyst for the reverse water gas shift reaction by an iron promoter. Chem. Commun. 2001, 1770–1771. [Google Scholar] [CrossRef] [PubMed]
  107. Okamoto, Y.; Kubota, T.; Gotoh, H.; Ohto, Y.; Aritani, H.; Tanaka, T.; Yoshida, S. XAFS study of zirconia-supported copper catalysts for the NO-CO reaction: Deactivation, rejuvenation and stabilization of Cu species. J. Chem. Soc. Faraday Trans. 1998, 94, 3743–3752. [Google Scholar] [CrossRef]
  108. Li, J.; Zhou, L.; Zhu, Q.; Li, H. Enhanced methanation over aerogel NiCo/Al2O3 catalyst in a magnetic fluidized bed. Ind. Eng. Chem. Res. 2013, 52, 6647–6654. [Google Scholar] [CrossRef]
  109. Li, J.; Zhou, L.; Li, P.; Zhu, Q.; Gao, J.; Gu, F.; Su, F. Enhanced fluidized bed methanation over a Ni/Al2O3 catalyst for production of synthetic natural gas. Chem. Eng. J. 2013, 219, 183–189. [Google Scholar] [CrossRef]
Figure 1. Transmission electron microscope (TEM) images of the samples: (a) CeO2-NR, (b) Cu/CeO2-NR, (c) Co/CeO2-NR, (d) CeO2-NC, (e) Cu/CeO2-NC, (f) Co/CeO2-NC.
Figure 1. Transmission electron microscope (TEM) images of the samples: (a) CeO2-NR, (b) Cu/CeO2-NR, (c) Co/CeO2-NR, (d) CeO2-NC, (e) Cu/CeO2-NC, (f) Co/CeO2-NC.
Nanomaterials 09 01739 g001
Figure 2. X-ray diffraction (XRD) patterns of the as-prepared samples.
Figure 2. X-ray diffraction (XRD) patterns of the as-prepared samples.
Nanomaterials 09 01739 g002
Figure 3. Hydrogen temperature-programmed reduction (H2-TPR) profiles of the as-prepared samples; (a) CeO2-NX, (b) Co/CeO2-NX, (c) Cu/CeO2-NX. H2-TPR profiles of bare Co3O4 (b) and CuO (c) oxides are also depicted for comparison purposes.
Figure 3. Hydrogen temperature-programmed reduction (H2-TPR) profiles of the as-prepared samples; (a) CeO2-NX, (b) Co/CeO2-NX, (c) Cu/CeO2-NX. H2-TPR profiles of bare Co3O4 (b) and CuO (c) oxides are also depicted for comparison purposes.
Nanomaterials 09 01739 g003
Figure 4. Experimental (solid lines) and theoretical (green lines) CO2 conversion profiles for commercial CeO2, bare ceria-NX and M/CeO2-NX samples in CO2 hydrogenation reaction. The dotted and dashed green line correspond to reverse water-gas shift (rWGS) and methanation reactions equilibrium, respectively. (F = 100 cm3/min, gas hourly space velocity (GHSV) = 20,000 h−1, H2:CO2 = 9:1, P = 1 atm).
Figure 4. Experimental (solid lines) and theoretical (green lines) CO2 conversion profiles for commercial CeO2, bare ceria-NX and M/CeO2-NX samples in CO2 hydrogenation reaction. The dotted and dashed green line correspond to reverse water-gas shift (rWGS) and methanation reactions equilibrium, respectively. (F = 100 cm3/min, gas hourly space velocity (GHSV) = 20,000 h−1, H2:CO2 = 9:1, P = 1 atm).
Nanomaterials 09 01739 g004
Figure 5. Selectivity to (a) CO and (b) CH4 as a function of temperature for all samples in the CO2 hydrogenation reaction. (F = 100 cm3/min, GHSV = 20,000 h−1, H2:CO2 = 9:1, P = 1 atm).
Figure 5. Selectivity to (a) CO and (b) CH4 as a function of temperature for all samples in the CO2 hydrogenation reaction. (F = 100 cm3/min, GHSV = 20,000 h−1, H2:CO2 = 9:1, P = 1 atm).
Nanomaterials 09 01739 g005
Figure 6. Effect of H2:CO2 ratio on theoretical and experimental CO2 conversion and selectivity to CH4 (triangle) and CO (square) values for (a,b) Co/CeO2-NR and for (c,d) Cu/CeO2-NR. Dotted and dashed lines represent equilibrium CO2 conversion values for the methanation and rWGS reactions, respectively.
Figure 6. Effect of H2:CO2 ratio on theoretical and experimental CO2 conversion and selectivity to CH4 (triangle) and CO (square) values for (a,b) Co/CeO2-NR and for (c,d) Cu/CeO2-NR. Dotted and dashed lines represent equilibrium CO2 conversion values for the methanation and rWGS reactions, respectively.
Nanomaterials 09 01739 g006aNanomaterials 09 01739 g006b
Figure 7. Dependence of (a) CO2 conversion and (b) selectivity to CO (square) and CH4 (triangle) on time on stream for Cu/CeO2-NR (red) and Co/CeO2-NR (black). Dotted lines refer to the corresponding equilibrium conversions. Experimental conditions: T = 450 °C, H2:CO2 = 9:1, P = 1 atm, GHSV = 20,000 h−1.
Figure 7. Dependence of (a) CO2 conversion and (b) selectivity to CO (square) and CH4 (triangle) on time on stream for Cu/CeO2-NR (red) and Co/CeO2-NR (black). Dotted lines refer to the corresponding equilibrium conversions. Experimental conditions: T = 450 °C, H2:CO2 = 9:1, P = 1 atm, GHSV = 20,000 h−1.
Nanomaterials 09 01739 g007
Table 1. Textural, structural and redox properties of the as-prepared samples.
Table 1. Textural, structural and redox properties of the as-prepared samples.
SampleBET AnalysisXRD AnalysisH2-TPR Analysis
SBET (m2/g)Average Crystallite Diameter, DXRD (nm)H2 Consumption
(mmol H2 g−1) 1
Theoretical H2
(mmol H2 g−1) 2
Peak Temperature (°C)
CeO2Co3O4/CuO
CeO2-NC3727-0.41-589809
CeO2-NR7915-0.59-545788
Co/CeO2-NC2824192.051.76335405
Co/CeO2-NR7214162.371.76318388
Cu/CeO2-NC3419521.501.34194228
Cu/CeO2-NR7512431.801.34181217
1 Estimated by the quantification of H2 uptake in the low temperature range (100–600 °C) of the TPR profiles. 2 Calculated as the amount of H2 required for the complete reduction of Co3O4 to Co and CuO to Cu.
Table 2. Conversion (XCO2), selectivity (S) and specific activity (rs, rm) of investigated samples at 400 °C.
Table 2. Conversion (XCO2), selectivity (S) and specific activity (rs, rm) of investigated samples at 400 °C.
Sample% XCO2% SCO% SCH4Reaction Rates
rs (µmol CO2·m−2·s−1)rm (µmol CO2·g−1·s−1)
CeO2-NR21.188.511.50.097.2
CeO2-NC19.389.810.20.186.6
Cu/CeO2-NR55.097.03.00.2518.8
Cu/CeO2-NC50.197.52.50.5117.1
Co/CeO2-NR84.95.594.50.4028.9
Co/CeO2-NC87.73.796.31.0729.9
Table 3. Comparison of the CO2 reduction performance of Cu- and Co-based catalysts for the rWGS and CO2 methanation reactions, respectively, at atmospheric pressure.
Table 3. Comparison of the CO2 reduction performance of Cu- and Co-based catalysts for the rWGS and CO2 methanation reactions, respectively, at atmospheric pressure.
SampleT (°C)% XCO2% SCO % SCH4 H2:CO2%wt. Cu or CoRef.
Cu-Catalyzed rWGS Reaction
Cu/CeO2-NR4001999.6 18.5This work
3899.0 4
Fe-Cu/Al2O34003689 48.2[15]
Cu/CeO240031.3100 413[82]
Cu/CeO2-NR45049N/A 55[83]
Cu/CeO2300~18100 39[60]
Cu-Ni/γ-Al2O350023.275.5 115[84]
Cu-Fe/SiO260015N/A 110[85]
Cu/β-Mo2C4001697.6 21.3[86]
Co-Catalyzed CO2 Methanation
Co/CeO2-NR40062.8 91.147.9This work
84.9 94.59
Co/CeO240034.9 37410[82]
Co/SiO236044.3 86.5410[87]
Co/KIT-630051 98.94.620[88]
Ni-Co/Ce0.25Zr0.75O228085 9845[89]
Co/Al2O330038 100415[90]
Co/CeO230097 ~96942.3[91]

Share and Cite

MDPI and ACS Style

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. https://doi.org/10.3390/nano9121739

AMA Style

Konsolakis M, Lykaki M, Stefa S, Carabineiro SAC, Varvoutis G, Papista E, Marnellos GE. 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(12):1739. https://doi.org/10.3390/nano9121739

Chicago/Turabian Style

Konsolakis, Michalis, Maria Lykaki, Sofia Stefa, Sόnia A. C. Carabineiro, Georgios Varvoutis, Eleni Papista, and Georgios E. Marnellos. 2019. "CO2 Hydrogenation over Nanoceria-Supported Transition Metal Catalysts: Role of Ceria Morphology (Nanorods versus Nanocubes) and Active Phase Nature (Co versus Cu)" Nanomaterials 9, no. 12: 1739. https://doi.org/10.3390/nano9121739

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