Next Article in Journal
H2 Photoproduction Efficiency: Implications of the Reaction Mechanism as a Function of the Methanol/Water Mixture
Previous Article in Journal
Sulfur Vacancies Enriched 2D ZnIn2S4 Nanosheets for Improving Photoelectrochemical Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thermocatalytic Hydrogenation of CO2 to Methanol Using Cu-ZnO Bimetallic Catalysts Supported on Metal–Organic Frameworks

1
Centre for Nanostructures and Advanced Materials (CeNAM), Chemicals Cluster, Council for Scientific and Industrial Research (CSIR), Meiring Naudé Road, Brummeria, Pretoria 0184, South Africa
2
Molecular Sciences Institute, School of Chemistry, University of the Witwatersrand, Private Bag 2001, Johannesburg 2050, South Africa
3
Department of Chemistry, University of Pretoria, Private Bag X20, Hatfield 0028, South Africa
4
Energy and Fuels for a Sustainable Environment Team, Institut de Chimie et Procédés pour l’Energie, l’Environnement et la Santé, UMR 7515 CNRS—ECPM, Université de Strasbourg, 25 Rue Becquerel, CEDEX 2, 67087 Strasbourg, France
*
Authors to whom correspondence should be addressed.
Submission received: 4 March 2022 / Revised: 28 March 2022 / Accepted: 30 March 2022 / Published: 5 April 2022
(This article belongs to the Topic Catalysis for Sustainable Chemistry and Energy)

Abstract

:
The thermocatalytic hydrogenation of carbon dioxide (CO2) to methanol is considered as a potential route for green hydrogen storage as well as a mean for utilizing captured CO2, owing to the many established applications of methanol. Copper–zinc bimetallic catalysts supported on a zirconium-based UiO-66 metal–organic framework (MOF) were prepared via slurry phase impregnation and benchmarked against the promoted, co-precipitated, conventional ternary CuO/ZnO/Al2O3 (CZA) catalyst for the thermocatalytic hydrogenation of CO2 to methanol. A decrease in crystallinity and specific surface area of the UiO-66 support was observed using X-ray diffraction and N2-sorption isotherms, whereas hydrogen-temperature-programmed reduction and X-ray photoelectron spectroscopy revealed the presence of copper active sites after impregnation and thermal activation. Other characterisation techniques such as scanning electron microscopy, transmission electron microscopy, and thermogravimetric analysis were employed to assess the physicochemical properties of the resulting catalysts. The UiO-66 (Zr) MOF-supported catalyst exhibited a good CO2 conversion of 27 and 16% selectivity towards methanol, whereas the magnesium-promoted CZA catalyst had a CO2 conversion of 31% and methanol selectivity of 24%. The prepared catalysts performed similarly to a CZA commercial catalyst which exhibited a CO2 conversion and methanol selectivity of 30 and 15%. The study demonstrates the prospective use of Cu-Zn bimetallic catalysts supported on MOFs for direct CO2 hydrogenation to produce green methanol.

Graphical Abstract

1. Introduction

The continuous increase in atmospheric CO2 levels propelled by fossil fuel combustion has caused an unprecedented rise in mean global temperatures [1,2,3,4]. The increase in the atmospheric concentration of anthropogenic CO2 contributes vastly to the greenhouse effect, which culminates in heat being trapped in the Earth’s atmosphere [5]. In recent years, there has been a collective effort to research ways to capture, store, and utilise CO2. A possible route for curbing the deleterious effects of global warming caused by greenhouse gas emissions is the conversion of CO2 into value-added chemicals such as methanol, dimethyl ether, formaldehyde, and acetic acid [5,6,7]. The thermocatalytic hydrogenation of CO2 is a proven approach towards the production of renewable methanol in substantial quantities [8,9,10,11,12,13]. Indeed, “The Methanol Economy” outlined by Olah et al. cites the advantages of CO2 conversion to methanol where each molecule produced consists of four hydrogen atoms and is, therefore, a suitable energy storage molecule which is a liquid at ambient conditions. Furthermore, methanol may be used directly as a fuel in internal combustion engines to further curb the deleterious effects of anthropogenic CO2 emissions [14]. Since the late 1960s, methanol has been produced from syngas (CO, CO2, and H2) over a coprecipitated ternary catalyst, Cu/ZnO/Al2O3, at pressures and temperatures ranging between 50 and 80 bar and 210 and 290 °C [15,16,17].
In recent years, Cu/ZnO/Al2O3 catalysts have been studied for the thermocatalytic hydrogenation of CO2 to methanol [18,19,20,21,22]. Though the catalysts exhibit CO2 conversion and methanol selectivity, the promoter effects conferred by ZnOx, which induces morphological changes, and its function as a spacer for Cu active sites are diminished by the temperature and pressure conditions in the reactor with time [23,24,25]. The phase separation of Cu/ZnOx results in the agglomeration, sintering, and poor dispersion of Cu nanoparticles, which consequently decreases activity towards methanol production due to a decrease in the metallic surface area of the Cu active sites. In addition, the deactivation attributed to the phase separation and sintering of Cu nanoparticles can catalyse the reverse water–gas shift reaction (RWGS), which produces carbon monoxide (CO) and water (H2O) from the CO2/H2 inlet gas mixture, thereby decreasing the catalysts’ selectivity to methanol [23,26,27]. The phase separation of Cu-based catalysts can be circumvented by the deposition and/or encapsulation of Cu/ZnOx nanoparticles into metal–organic frameworks (MOFs) [28,29]. MOFs are a class of highly crystalline, nanoporous materials prepared via the self-assembly of metal ion clusters and multidentate organic linkers [30,31]. Recently, MOFs have been identified as prospective supports for Cu-based CO2 hydrogenation catalysts. Rungtaweevoranit et al., for instance, encapsulated and deposited Cu nanoparticles on a Zr-based MOF (UiO-66) and observed that the latter exhibited an 8-fold increase in turnover frequency when compared to the conventional Cu/ZnO/Al2O3 catalyst [4]. In addition to the confinement of Cu nanoparticles by strong support–metal interactions (SMSI), MOFs such as UiO-66 exhibit chemical promotion by charge transfer from Cu to the MOF, which significantly enhances the activity of the hybrid catalysts towards CO2 hydrogenation [32]. Therefore, the purpose of this study is to elucidate the performance of novel Cu-Zn bimetallic catalysts supported on a zirconium-based MOF, UiO-66(Zr), prepared via slurry-phase impregnation compared with conventional ternary promoted Cu/ZnO/Al2O3 catalysts.

2. Results and Discussion

2.1. Characterisation

2.1.1. X-ray Diffraction and (XRD) Thermogravimetric Analysis (TGA)

The XRD analytical results of the co-precipitated commercial and CuO/ZnO/Al2O3/MgO catalysts are shown in Figure 1a, where the crystalline phase of CuO is observed at 35° < 2θ < 39° whilst that of ZnO can be seen at 31, 33, and 37° [5,18,33,34,35]. The average crystallite sizes of CuO at 2θ = 38.7° in the commercial and CuO/ZnO/Al2O3/MgO catalysts, calculated using the Debye–Scherrer equation, were found to be 5.2 and 7.8 nm, respectively. The smaller CuO crystallite sizes in the commercial catalyst can be attributed to a high Cu dispersion [36,37]. The XRD results of the pristine UiO-66 support and Cu/ZnO/UiO-66 also shown in Figure 1a are typical of the crystalline phase with peaks at 2θ = 7.4, 8.5 and 14.1, 14.7, 17, 18.6, and 19.1° [38,39]. Neither the copper nor zinc phases (Cu0, CuO, Cu2O, or ZnO) could be observed in Cu/ZnO/UiO-66, potentially attributed to the high Cu and Zn dispersion over the framework or low metal loading [3].
The thermal stabilities of the commercial and Cu/ZnO/Al2O3/MgO catalysts as well as the MOF-supported catalyst were determined using thermogravimetric analysis in comparison to the commercial catalyst shown in Figure 1b. The commercial and Cu/ZnO/Al2O3/MgO catalysts were observed to exhibit thermal stability up to 900 °C. The mass loss observed in UiO-66 and Cu/ZnO/UiO-66 from 100 to 350 °C is attributed to the removal of dimethylformamide (DMF), ethanol, H2O, and residual carboxylic acid groups (Figure 1b) [38,40]. The almost linear weight loss of UiO-66 and Cu/Zn/UiO-66 between 300 and 500 °C is indicative of the high thermal stability of the Zr-based MOF, after which ZrO2 is formed due to the degradation of the organic linkers [41]. Furthermore, the thermal stability of UiO-66 decreases after impregnation and thermal activation at 350 °C. The thermal stability of the catalysts is essential for the stability under reactor conditions at 230 °C and 50 bar to prevent catalyst deactivation due to thermal degradation and MOF-support structural collapse. Therefore, all the as-prepared catalysts are anticipated to be relatively stable over time on stream while the reactor operates continuously for CO2 hydrogenation. Furthermore, it can be seen that the commercial and Cu/ZnO/Al2O3/MgO catalysts are relatively more thermally stable than UiO-66 and the Cu/ZnO/UiO-66 catalyst, respectively.

2.1.2. N2 Physisorption, Fourier Transform Infrared Spectroscopy (FTIR), and Metal Loading

The surface area and pore volume of the catalysts and pristine UiO-66 were characterised using nitrogen adsorption–desorption isotherms where the results are listed in Table 1. The Brunauer–Emmett–Teller (BET) surface area of UiO-66 was 1238 m2/g with a pore volume of 0.47 cm3/g. The reported values are consistent with data previously reported for UiO-66 [38,40]. The nitrogen sorption isotherms of pristine UiO-66, Cu/ZnO/UiO-66, Cu/ZnO/Al2O3/MgO, and commercial catalysts are shown in Figure 2. It is clear from Figure 2b that both pristine and impregnated MOFs exhibit type-I sorption isotherms owing to their microporous structure. Furthermore, the decrease in specific surface area (SSA) observed in Cu/ZnO/UiO-66 can be attributed to the presence of Cu and Zn nanoparticles which occupy available surface sites of the UiO-66 support [3,32]. The commercial catalyst had a higher SSA compared to the CuO/ZnO/Al2O3/MgO catalyst. The effect of larger crystallite sizes, corroborated by the XRD derived data in Figure 1a, culminates in a relatively low SSA when compared to the commercial catalyst [33,42]. The functional groups of the organic linkers present in the samples were studied using FTIR (Figure 2c). The peak at 656 cm−1 is attributed to the vibration of µ3-O in the Zr6O4(OH)4(CO2)12 nodes of UiO-66, whereas the peak at 550 cm−1 is due to asymmetric Zr-(OC) vibrations which further indicate successful coordination of the Zr secondary building units (SBUs) to the organic benzene dicarboxylate (BDC) linkers [3,40,43]. The stretching symmetric and asymmetric vibrations of the carboxylate group can be observed at 1572 and 1392 cm−1, whereas the π C = C vibration in the aromatic benzene ring in the linker can be observed at 1508 cm−1 [40,44]. The decrease in the intensity of peaks can be observed in the Cu/ZnO/UiO-66 catalyst, which may be attributed to the decrease in the amount of the functional groups, due to partial degradation, associated with UiO-66 after impregnation and thermal treatment, which is corroborated by the decrease in crystallinity and SSA seen in Figure 1a and Figure 2b [45].
The inductively coupled plasma-optical emission spectroscopy (ICP-OES) results showed that the concentration of Cu in the Cu/ZnO/UiO-66 catalyst was 12 wt%, which confirms successful Cu impregnation although it was slightly higher than the nominal loading of 7 wt%. The amount of Zn was also found to be 4.2% compared to a loading of 3%. The discrepancy in nominal loading may be due to inhomogeneous dispersion over the UiO-66 framework, albeit scanning electron microscopy-electron dispersive X-ray spectroscopy (SEM-EDS) elemental maps revealed a good dispersion of both Cu and Zn (Figure S1). The amount of Cu, Zn, Al, and Mg in the Cu/ZnO/Al2O3/MgO catalysts was found to be 61.7/27.7/10.6/0.1 weight% (wt%) compared with the nominal loading of 64.2/24.5/9.8/1.3 wt%. In addition, the elemental map shows good dispersion of all the elements, and Mg was not observed in the analysis, most likely due to its low loading (Figure S2). Similarly, the EDS results of the commercial catalyst showed a good dispersion as seen in Figure S3 and a loading of 59.9, 27.5, 10.1, and 2.54 wt%, which corresponds to the certificate of analysis issued by the supplier, reporting a loading of 64.2/24.5/9.8/1.3 wt%. The elemental loading in each catalyst is summarised in Table 1.

2.1.3. Electron Microscopies

The scanning and transmission electron microscopy (SEM and TEM) images of the samples are shown in Figure 3. The morphology of pristine UiO-66 (Figure 3a,b) is typical of octahedral-shaped crystals with a homogeneous particle size distribution [15,21,25]. The octahedral shape of UiO-66 is also preserved after impregnation with Cu and Zn (Figure 3c,d). The commercial catalyst (Figure 3e,f) exhibited irregular shapes which are typically observed when the ageing time in the mother liquor after co-precipitation is too short [46]. The CuO/ZnO/Al2O3/MgO catalyst (Figure 3g,h) exhibited a needle-like morphology similar to those observed by Mota, et al., being characteristic of catalysts prepared via the zincian malachite precursor as were prepared in this study [34].
The TEM images of the pristine UiO-66 and Cu/ZnO/UiO-66 catalyst shown in Figure 3b,d corroborate the octahedral shape of UiO-66 particles seen in Figure 3a,b. Furthermore, on the surface of the UiO-66 particles in the Cu/ZnO/UiO-66 catalyst, small spherical nanoparticles can be seen which are indicative of CuO/ZnO nanoparticles [4,28]. These nanoparticles are well dispersed on the surface of UiO-66 particles as indicated by SEM-EDS (Figure S2). The micrographs confirm the successful dispersion of Cu and Zn in the Cu/ZnO/UiO-66 catalyst.

2.1.4. X-ray Photoelectron Spectroscopy (XPS)

The XPS results revealed that Cu was indeed successfully loaded on the surface of the UiO-66 support (Figure 4a,b). Cu was present as cupric hydroxide (Cu(OH)2) after drying at 160 °C in air. Furthermore, the 2p1/2 and 2p3/2 peaks at 934.5 and 953.6 3eV, and the satellite peak at 943 ± 0.2 eV, corroborate the presence of Cu2+ species (Figure 4b) [47,48]. Therefore, to ensure the preferential formation of the Cu2+ precursor to the copper active phase precursor, Cu0, the dried pre-catalyst was activated by thermal treatment under a constant flow of argon at 350 °C. CuO is desired as it is the precursor phase to the metallic Cu active phase formed by reduction with H2 at temperatures between 200 and 300 °C [15]. Furthermore, Zr was present as ZrO2 as revealed by the peak at Zr 3d5/2 at 182.1 eV and the Zr 3d3/2 peak at 185.2 eV, respectively (Figure 4c) [45]. XPS analysis did not detect any Zn, probably due to its low loading or high dispersion [3]. It is also likely that the analysis depth of the X-rays used in the XPS analysis are beyond the penetration depth of the Zn nanoparticles [4]. Furthermore, near ambient pressure XPS would be required to ascertain the surface and subsurface oxidation states of nanoparticles smaller than 4 nm [49]. Zn was, however, observed using SEM-EDS with a good dispersion due to the relatively high analysis depth of SEM-EDS which extends to 1–3 μM, whereas XPS is limited to up to 10 nm of the surface (Figure S1) [50,51]. The XPS results of the pristine UiO-66 support are shown in Figure S4.

2.1.5. Hydrogen Temperature-Programmed Reduction (H2-TPR)

The reducibility of the catalysts was studied using hydrogen temperature-programmed reduction (H2-TPR) where the reduction profiles are shown in Figure 5. All the Cu-loaded catalysts exhibited hydrogen uptake. The commercial and Cu/ZnO/Al2O3/MgO catalysts showed reduction peaks at Tmax = 195 °C and 223 °C (Figure 5a). The reduction peaks are indicative of CuO crystallite formed after calcination. The broader curve of the Cu/ZnO/Al2O3/MgO may be due to the larger CuO crystallite size of 7.8 nm compared to 5.2 nm in the commercial catalyst [5,34]. The H2-TPR profiles of UiO-66 and Cu/ZnO/UiO-66 are in Figure 5b, where no Cu reduction peak was seen for pristine UiO-66. H2 uptake, conversely, was observed in Cu/ZnO/UiO-66 at Tmax = 209 °C; the reduction peak further corroborates the presence of H2-active copper species on the surface of UiO-66 [45]. The peaks at Tmax = 446 °C in Figure 5b indicate the reduction of Zr species in the metal centre or clusters of the UiO-66 framework. It can also be seen that the loading of Cu and Zn on UiO-66 shifts the Tmax of the ZrO2 reduction peak to 446 °C, which may be due to the ease of reduction facilitated by H2 spillover from Cu and Zn, respectively [45,52].

2.2. Catalyst Testing and Evaluation

Conversion, Selectivity and Productivity

The CO2 conversion results of all catalysts are shown in Figure 6. In the study, it was predominantly CO, MeOH, and methane (CH4) that were formed as products in addition to H2O. The reaction progresses through the exothermic CO2 hydrogenation reaction shown in Equation (1). CO is formed via the endothermic reverse water–gas shift reaction in which CO2 is reduced to CO and H2O, Equation (2) [4,10,15,26]. In addition, CH4 was presumably formed via the Sabatier reaction, in which CO2 methanation takes place to form CH4 and H2O, Equation (3) [53]. The highest average CO2 conversion after 24 h online was exhibited by the Cu/ZnO/Al2O3/MgO catalyst (30.6%) followed by the commercial catalyst (29.8%) and the Cu/ZnO/UiO-66 catalyst (26.7%). Furthermore, all the catalysts tested exhibited a good stability within the 24 h evaluation period.
C O 2 + 3 H 2     C H 3 O H + H 2 O                                               Δ H = 49.47   K J / m o l
C O 2 + H 2     C O + H 2 O                                                               Δ H = 41.14   K J / m o l Δ
C O 2 + 4 H 2     C H 4 + 2 H 2 O                                                     Δ H = 165   K J / m o l
The selectivity of the catalysts towards methanol was also evaluated, where the Cu/ZnO/Al2O3/MgO catalyst exhibited the highest selectivity towards methanol (24%) followed by the Cu/ZnO/UiO-66 catalyst (16%), whereas the commercial catalyst had the lowest selectivity of 15.0% (Figure 7a). Furthermore, owing to the competing RWGS in thermocatalytic CO2 hydrogenation, the selectivity of the catalysts towards CO was evaluated [54]. As can be seen in Figure 7b, the selectivity of the catalysts towards CO was very high for all catalysts with the highest, at 84%, exhibited by the commercial catalyst, despite small CuO crystallites and a high SSA, which are anticipated to suppress CO selectivity [23,26]. In contrast, the Cu/ZnO/Al2O3/MgO and Cu/ZnO/UiO-66 catalysts had relatively lower CO selectivities of 74 and 83%, respectively. In addition, CH4 was also observed in the product streams of all the catalysts, which is presumably from the Sabatier reaction in which CO2 methanation occurs [53]. As such, the Cu/ZnO/UiO-66 and Cu/ZnO/Al2O3/MgO catalysts had the most selectivity towards CH4 at 1.6 and 1.7%, respectively, with the lowest observed in the commercial catalyst (0.9%). The methanol productivity, space–time yield (STY), of the Cu/ZnO/UiO-66 catalyst commercial catalyst was 128 gMeOH/Kgcat/h, followed by the commercial catalyst at 52 gMeOH/Kgcat/h (Figure 7d). The Cu/ZnO/Al2O3/MgO catalyst had the lowest methanol productivity at 37 gMeOH/Kgcat/h, which may be attributed to the relatively large CuO crystallite size and consequently low SSA [54].
In comparison to An et al., who observed a 100% methanol selectivity, the Cu/ZnO/UiO-66 catalyst showed a relatively low methanol selectivity of 15.7%. However, they also reported relatively low CO2 conversions of 3.3% at a GHSV of 18,000 h−1 compared to 10,000 h−1, as high GHSV results in low per-pass conversion [55]. Furthermore, the commercial and Cu/ZnO/Al2O3/MgO co-precipitated catalysts also exhibit superior CO2 conversion when compared to the 11% observed by Portha et al., although their catalyst, with a copper–zinc–alumina composition, exhibited a methanol selectively of 43% compared to the 24% of the Cu/ZnO/Al2O3/MgO catalyst in this work [56]. These results highlight the promise of using Cu-based MOF-supported catalysts for CO2 hydrogenation to methanol due to the relatively good catalytic performance of Cu/ZnO/UiO-66 despite having ten times less the Cu loading compared to the Cu/ZnO/Al2O3/MgO and commercial catalysts. The high CO2 conversion and relative methanol STY may be attributed to the good dispersion of Cu over UiO-66, the high SSA, and the enhancement of the active sites by the ZrO2 secondary building units [4,57].

3. Materials and Methods

3.1. Reagents and Chemicals

Ethanol (Associated Chemical Enterprises, 95%), N,N-dimethyl formamide (Associated Chemical Enterprises, 99.5%), cupric nitrate trihydrate (Cu(NO3)·3H2O, Associated Chemical Enterprises, 99.9%), magnesium nitrate hexahydrate (Mg(NO3)2·6H2O, Associated Chemical Enterprises, 98%), zinc nitrate hexahydrate (Zn(NO3)2·6H2O, Associated Chemical Enterprises, 99.9%), zirconium tetrachloride (ZrCl4, Sigma-Aldrich, 99.5%), terephthalic acid (Sigma-Aldrich, 98%), and aluminium nitrate nonahydrate (Al(NO3)·9H2O, Sigma-Aldrich, ≥98%) were commercially sourced from Associated Chemical Enterprises Pty Ltd. (Johannesburg, South Africa) and Merck KGaA (Darmstadt, Germany),and used without further purification. De-ionised water was sourced from a water demineralisation system (Instrubal, Zeneer Power II) located on-site. A copper-based methanol synthesis catalyst was procured from Alfa Aesar, Thermo Fischer Scientific (Kandel, Germany). The pelletised catalyst with a percentage composition of 64.2/24.5/9.8/1.3 (Cu/Zn/Al/Mg) was ground to a powder with a pestle and mortar before any use.

3.2. Catalysts’ Synthesis

3.2.1. Cu/ZnO/Al2O3/MgO Catalyst

The Cu/ZnO/Al2O2/MgO catalyst was prepared via a co-precipitation method. In the procedure, a catalyst with a mass percentage loading of 64.2/24.5/9.8/1.3 (Cu/Zn/Al/Mg) was prepared by dissolving 50.5 mmol of Cu(NO3)·3H2O, 11.9 mmol of Zn(NO3)2·6H2O, 10.3 mmol of Al(NO3)·9H2O, and 1.55 mmol of Mg(NO3)2·6H2O in demineralised water (90 mL) followed by acidification with 20 mL of 65% HNO3 to give a 1 M metal nitrates’ solution. The solution, maintained at 60 °C, was dosed with 1 M Na2CO3 solution at a rate of 3 mL/min whilst stirring at 400 rpm. Once the solution reached a pH of 6.5 ± 0.5, it was left to age for 1 h. The solution was then filtered, and the precipitates were washed several times with demineralised water and then dried overnight at 90 °C. The dried precursors were then calcined in open air at 330 °C for 3 h.

3.2.2. UiO-66 (Zr) MOF Synthesis

UiO-66 MOFs were prepared via a modified solvothermal method reported by Ren et al. (2014) [38]. In the procedure, terephthalic acid (4.54 mmol) and ZrCl4 (4.54 mmol) were dissolved in 500 mL of dimethylformamide (DMF) and dissolved by magnetic stirring in a beaker at room temperature. After dissolution, 171.3 mL of formic acid (0.453 mmol) was added as a modulator. The crystallisation of UiO-66 was carried out at a constant reaction temperature of 120 °C under static reflux for 4 h. The white precipitate was recovered by vacuum filtration, washed by immersing in a round-bottom flask with hot ethanol at 60 °C overnight, vacuum filtered, and dried at 90 °C for 24 h.

3.2.3. UiO-66 (Zr) MOF-Supported Catalyst Synthesis

Cu((NO3)2·3H2O (9.97 mmol) and Zn(NO3)2.6H2O (4.06 mmol) were dissolved in 14 mL of ethanol. Thereafter, UiO-66 (8.005 g) was added slowly to the metal nitrate ethanolic solution with continuous stirring. The mixture was then dried overnight at ambient temperature. The dried, light blue, catalyst precursor was thermally treated at 350 °C for 4 h under a constant flow of argon and a heating rate of 5 °C/min to furnish a 7/3/90 wt% Cu/ZnO/UiO-66 catalyst.

3.3. Characterisation

Powder X-ray diffraction (PXRD), for phase identification, was conducted with a Rigaku Ultima IV diffractometer with a Ni-filtered radiation of 0.154 nm, a voltage of 40 kV, and a current of 30 mA at room temperature (Akishima-shi, Tokyo, Japan). A scanning range of 2θ = 3–90° at a rate of 0.01° s−1 was used for all the samples.
The thermal stability of the precursors and catalysts was measured using thermogravimetric analysis (TGA) using the Mettler, Toledo TGA/SDTA 851e instrument (Mettler Toledo, Greifensee, Switzerland). In the procedure, 10 mg of the sample was heated to 1000 °C at a rate of 10 °C min−1 in nitrogen.
Surface area and porosity measurements were performed with an ASAP 2020 HD analyser (Micromeritics Instrument Corporation, Georgia, United States of America), and high purity (99.999%) grade nitrogen (N2) gas. The Brunauer–Emmett–Teller (BET) surface area and pore volume were determined using N2 gas sorption isotherms at 77 K and relative partial pressures (p/po) of up to 1. Prior to each analysis, the samples (>0.2 g) were degassed under vacuum using a Micromeritics SmartVac with heating to no more than 200 °C.
Fourier transform infrared spectroscopy (FTIR) was recorded on a PerkinElmer Spectrum 100 FTIR spectrometer at a mid-IR range of 4000–550 cm−1 (PerkinElmer, Waltham, United States of America).
The powder skeletal density of the samples was determined using a Micromeritics AccuPyc II 1340 pycnometer and helium gas at 135 kPa (Micromeritics Instrument Corporation, Georgia, United States of America).
The morphology of the samples was imaged using an Auriga Cobra focused-ion beam scanning electron microscope coupled with energy-dispersive X-ray spectroscopy (Zeiss, Oberkochen, Germany). Prior to imaging and elemental mapping, the samples were mounted on an adhesive carbon tape loaded on a sample holder and coated with carbon to prevent charging. An FEI Tecnai T12 transmission electron microscope (Thermo Fischer Scientific, Kandel, Germany) was used to image the Cu/ZnO/UiO-66 catalyst. The sample was crushed to a fine powder, dispersed in ethanol by ultrasonication, and a drop was placed on a carbon-coated copper grid which was loaded into the microscope. Elemental analyses were done using EDS for the commercial and Cu/ZnO/Al2O3/MgO catalysts, whereas inductively coupled plasma-optical emission spectroscopy was used for the MOF-supported catalyst. The latter was digested in aqua regia using a microwave, the digest made up to 50 mL with deionised water and analysed using ICP-OES PlasmaQuant 9100 (Analytik Jena GmbH, Jena, Germany).
Temperature-programmed reduction was used for reducibility testing using a Micromeritics Autochem II analyser (Micromeritics Instrument Corporation, Georgia, United States of America). Each sample, placed between two pieces of glass wool in a U-tube, was degassed at 150 °C at a heating rate of 10 °C/min under a constant flow of helium for 30 min and then cooled to 50 °C. Thereafter, the sample was heated to 900 °C at 10 °C/min under a 50 mL/min flow of 5% H2/Ar. The hydrogen consumption was measured using a thermal conductivity detector.
The surface oxidation state of Cu, chemical composition of the SBUs, and metal oxide bonds were characterised using X-ray photoelectron spectroscopy (XPS). A Thermo Scientific ESCALAB 250Xi instrument was used with a monochromatic X-ray Al source with an energy of 1486.7 eV, which was operated at 300 W and a pass energy of 100 eV. The analysis was conducted under an ultrahigh vacuum of <10−8 mbar (Thermo Fischer Scientific, Kandel, Germany).

3.4. Catalyst Testing

Synthesised catalysts were tested in a fixed-bed reactor. The reactor was 21 cm in length with an internal and external diameter of 1.01 cm and 1.27 cm. Typically, the catalyst was sieved to a 50–125 µM size fraction, weighed and packed between two pieces of inert glass wool, and positioned at the centre of the reactor. Before each run, the catalyst was reduced under a stream of pure hydrogen for approximately 2 h at 40 mL/min, 250 °C, and 20 bar. Thereafter, the reactor was cooled to 50 °C, and a pre-mixed gas consisting of 3:1:0.2 H2/CO2/Ar was introduced into the reactor and left for 12 h to stabilise. A CO2 hydrogenation reaction was then carried out at 230 °C, 50 bar, and a flow rate of 40 mL/min with a constant GHSV of 10,000 h−1 for 24 h. The analysis of the outgoing products was carried out using an Agilent 7890B gas chromatograph instrument (Agilent Technologies Inc, California, United States of America) equipped with an FID (CH3OH, CH4) and two TCDs (Ar, CO, CO2, CH4, and H2). Online analysis was carried out for all the outlet gases, whereas the liquid products were condensed in a trap maintained at 0 °C and injected into the GC. The conversions of CO2 were calculated according to Equation (4).
X C O 2   = 1 C O 2 ,   o u t c o 2 ,   i n × A r i n A r o u t
The selectivity for methanol, CO, and CH4 were calculated using Equations (5)–(7):
S C H 3 O H = n C H 3 O H n C H 3 O H + n C H 4 + n C O  
S C O = n C O n C H 3 O H + n C H 4 + n C O  
S C H 4 = 1 S C O + S C H 3 O H

4. Conclusions

In this study, a Zr-based MOF, UiO-66, was used as a support to prepare catalysts via slurry phase impregnation with Cu as well as Zn and tested for the thermocatalytic conversion of CO2 to methanol. The catalyst was benchmarked against the conventional, co-precipitated, MgO-promoted ternary catalyst, and a CZA commercial catalyst. Hydrogen uptake shown by TPR revealed the presence of active Cu species for H2 uptake and a good dispersion of Cu and Zn over the UiO-66 support was observed from SEM-EDS elemental maps. Furthermore, the performance of Cu/ZnO/UiO-66 displayed a good CO2 conversion, selectivity towards methanol, and the highest methanol space–time yield, making MOF-supported Cu-based catalysts promising for their utilisation in the thermocatalytic hydrogenation of CO2 to methanol.

Supplementary Materials

The following supporting information can be downloaded at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/catal12040401/s1, Figure S1: Elemental maps of Cu/ZnO/UiO-66; Figure S2: Elemental maps of Cu/ZnO/Al2O3/MgO catalyst; Figure S3: Elemental maps of commercial catalyst; Figure S4: XPS results of UiO-66: (a) full survey, (b) Zr3d, and O1s scans.

Author Contributions

Conceptualization, Z.G.D., B.L., K.P., and N.M.M.; methodology, Z.G.D., B.L., and K.P.; investigation, Z.G.D.; resources, N.M.M., H.W.L., and J.M.; writing—original draft preparation, Z.G.D.; writing—review and editing, Z.G.D., H.W.L., J.M., N.M.M., X.D., and B.L.; supervision, N.M.M., H.W.L., and J.M.; formal analysis, X.D.; funding acquisition, N.M.M., B.L., and H.W.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Royal Society Foreign, Commonwealth & Development Office (FCDO) Africa Capacity Building Initiative (ACBI) Programme (grant AQ150029) and the South African Department of Science and Innovation (DSI) for research activities under HySA Infrastructure (grant No. CNMH20X).

Acknowledgments

The authors would like to acknowledge financial support from the Royal Society Foreign, Commonwealth & Development Office (FCDO) Africa Capacity Building Initiative (ACBI) Programme (grant AQ150029), the Council for Scientific and Industrial Research (CSIR), and the South African Department of Science and Innovation (DSI) for research activities under HySA Infrastructure (grant No. CNMH20X) and the South Africa–France PROTEA Programme, which is a bilateral incentive programme dedicated to strengthening research collaborations between the two countries (project No. CNMH02X).

Conflicts of Interest

The authors declare no personal or financial conflict of interest.

References

  1. Tursunov, O.; Kustov, L.; Kustov, A. A Brief Review of Carbon Dioxide Hydrogenation to Methanol Over Copper and Iron Based Catalysts. Oil Gas Sci. Technol. Rev. IFP Energ. Nouv. 2017, 72, 30. [Google Scholar] [CrossRef] [Green Version]
  2. Fischer, H.; Wahlen, M.; Smith, J.; Mastroianni, D.; Deck, B. Ice core records of atmospheric CO2 around the last three glacial terminations. Science 1999, 283, 1712–1714. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Stawowy, M.; Ciesielski, R.; Maniecki, T.; Matus, K.; Łużny, R.; Trawczynski, J.; Silvestre-Albero, J.; Łamacz, A. CO2 Hydrogenation to Methanol over Ce and Zr Containing UiO-66 and Cu/UiO-66. Catalysts 2020, 10, 39. [Google Scholar] [CrossRef] [Green Version]
  4. Rungtaweevoranit, B.; Baek, J.; Araujo, J.R.; Archanjo, B.S.; Choi, K.M.; Yaghi, O.M.; Somorjai, G.A. Copper Nanocrystals Encapsulated in Zr-based Metal-Organic Frameworks for Highly Selective CO2 Hydrogenation to Methanol. Nano Lett. 2016, 16, 7645–7649. [Google Scholar] [CrossRef] [PubMed]
  5. Gesmanee, S.; Koo-Amornpattana, W. Catalytic hydrogenation of CO2 for methanol production in fixed-bed reactor using Cu-Zn supported on gamma-Al2O3. Energy Procedia 2017, 138, 739–744. [Google Scholar] [CrossRef]
  6. Bowker, M. Methanol Synthesis from CO2 Hydrogenation. ChemCatChem 2019, 11, 4238–4246. [Google Scholar] [CrossRef] [Green Version]
  7. Fichtl, M.B.; Schlereth, D.; Jacobsen, N.; Kasatkin, I.; Schumann, J.; Behrens, M.; Schlögl, R.; Hinrichsen, O. Kinetics of deactivation on Cu/ZnO/Al2O3 methanol synthesis catalysts. Appl. Catal. A Gen. 2015, 502, 262–270. [Google Scholar] [CrossRef]
  8. Markewitz, P.; Kuckshinrichs, W.; Leitner, W.; Linssen, J.; Zapp, P.; Bongartz, R.; Schreiber, A.; Müller, T.E. Worldwide innovations in the development of carbon capture technologies and the utilization of CO2. Energy Environ. Sci. 2012, 5, 7281–7305. [Google Scholar] [CrossRef] [Green Version]
  9. 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] [Green Version]
  10. Shi, Z.; Tan, Q.; Tian, C.; Pan, Y.; Sun, X.; Zhang, J.; Wu, D. CO2 hydrogenation to methanol over Cu-In intermetallic catalysts: Effect of reduction temperature. J. Catal. 2019, 379, 78–89. [Google Scholar] [CrossRef]
  11. Centi, G.; Perathoner, S.; Rak, Z.S. Reduction of greenhouse gas emissions by catalytic processes. Appl. Catal. B Environ. 2003, 41, 143–155. [Google Scholar] [CrossRef]
  12. Song, C. Global challenges and strategies for control, conversion and utilization of CO2 for sustainable development involving energy, catalysis, adsorption and chemical processing. Catal. Today 2006, 115, 2–32. [Google Scholar] [CrossRef]
  13. Dorner, R.W.; Hardy, D.R.; Williams, F.W.; Willauer, H.D. Heterogeneous catalytic CO2 conversion to value-added hydrocarbons. Energy Environ. Sci. 2010, 3, 884–890. [Google Scholar] [CrossRef]
  14. Olah, G.A.; Goeppert, A.; Prakash, G.S. Beyond Oil and Gas: The Methanol Economy. John Wiley & Sons: Hoboken, NJ, USA, 2018. [Google Scholar]
  15. Natesakhawat, S.; Lekse, J.W.; Baltrus, J.P.; Ohodnicki, P.R.; Howard, B.H.; Deng, X.; Matranga, C. Active Sites and Structure–Activity Relationships of Copper-Based Catalysts for Carbon Dioxide Hydrogenation to Methanol. ACS Catal. 2012, 2, 1667–1676. [Google Scholar] [CrossRef]
  16. Agarwal, A.S.; Rode, E.; Sridhar, N.; Hill, D. Conversion of CO2 to Value-Added Chemicals: Opportunities and Challenges. In Handbook of Climate Change Mitigation and Adaptation; Chen, W.-Y., Suzuki, T., Lackner, M., Eds.; Springer: New York, NY, USA, 2014; pp. 1–40. [Google Scholar]
  17. Etim, U.J.; Song, Y.; Zhong, Z. Improving the Cu/ZnO-Based Catalysts for Carbon Dioxide Hydrogenation to Methanol, and the Use of Methanol As a Renewable Energy Storage Media. Front. Earth Sci. 2020, 8. [Google Scholar] [CrossRef]
  18. Ahouari, H.; Soualah, A.; Le Valant, A.; Pinard, L.; Magnoux, P.; Pouilloux, Y. Methanol synthesis from CO2 hydrogenation over copper based catalysts. React. Kinet. Mech. Catal. 2013, 110, 131–145. [Google Scholar] [CrossRef]
  19. Qaderi, J. A brief review on the reaction mechanisms of CO2 hydrogenation into methanol. Int. J. Innov. Res. Sci. Stud. 2020, 3, 53–63. [Google Scholar] [CrossRef]
  20. Melián-Cabrera, I.; López Granados, M.; Fierro, J.L.G. Reverse Topotactic Transformation of a Cu–Zn–Al Catalyst during Wet Pd Impregnation: Relevance for the Performance in Methanol Synthesis from CO2/H2 Mixtures. J. Catal. 2002, 210, 273–284. [Google Scholar] [CrossRef]
  21. Wang, J.; Funk, S.; Burghaus, U. Indications for Metal-support Interactions: The Case of CO2Adsorption on Cu/ZnO(0001). Catal. Lett. 2005, 103, 219–223. [Google Scholar] [CrossRef]
  22. Waugh, K.C. Methanol Synthesis. Catal. Today 1992, 15, 51–75. [Google Scholar] [CrossRef]
  23. Kuld, S.; Thorhauge, M.; Falsig, H.; Elkjær, C.F.; Helveg, S.; Chorkendorff, I.; Sehested, J. Quantifying the promotion of Cu catalysts by ZnO for methanol synthesis. Science 2016, 352, 969–974. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Spencer, M.S. The role of zinc oxide in Cu/ZnO catalysts for methanol synthesis and the water–gas shift reaction. Top. Catal. 1999, 8, 259. [Google Scholar] [CrossRef]
  25. Prieto, G.; Meeldijk, J.D.; de Jong, K.P.; de Jongh, P.E. Interplay between pore size and nanoparticle spatial distribution: Consequences for the stability of CuZn/SiO2 methanol synthesis catalysts. J. Catal. 2013, 303, 31–40. [Google Scholar] [CrossRef]
  26. An, X.; Li, J.; Zuo, Y.; Zhang, Q.; Wang, D.; Wang, J. A Cu/Zn/Al/Zr Fibrous Catalyst that is an Improved CO2 Hydrogenation to Methanol Catalyst. Catal. Lett. 2007, 118, 264–269. [Google Scholar] [CrossRef]
  27. Dalebout, R.; Visser, N.L.; Pompe, C.E.L.; de Jong, K.P.; de Jongh, P.E. Interplay between carbon dioxide enrichment and zinc oxide promotion of copper catalysts in methanol synthesis. J. Catal. 2020, 392, 150–158. [Google Scholar] [CrossRef]
  28. Lozano, L.A.; Faroldi, B.M.C.; Ulla, M.A.; Zamaro, J.M. Metal–Organic Framework-Based Sustainable Nanocatalysts for CO Oxidation. Nanomaterials 2020, 10, 165. [Google Scholar] [CrossRef] [Green Version]
  29. Falcaro, P.; Ricco, R.; Yazdi, A.; Imaz, I.; Furukawa, S.; Maspoch, D.; Ameloot, R.; Evans, J.D.; Doonan, C.J. Application of metal and metal oxide nanoparticles@MOFs. Coord. Chem. Rev. 2016, 307, 237–254. [Google Scholar] [CrossRef]
  30. Jiao, Y.; Liu, Y.; Zhu, G.; Hungerford, J.T.; Bhattacharyya, S.; Lively, R.P.; Sholl, D.S.; Walton, K.S. Heat-Treatment of Defective UiO-66 from Modulated Synthesis: Adsorption and Stability Studies. J. Phys. Chem. C 2017, 121, 23471–23479. [Google Scholar] [CrossRef] [Green Version]
  31. Farrusseng, D.; Aguado, S.; Pinel, C. Metal–organic frameworks: Opportunities for catalysis. Angew. Chem. Int. Ed. 2009, 48, 7502–7513. [Google Scholar] [CrossRef]
  32. Kobayashi, H.; Taylor, J.M.; Mitsuka, Y.; Ogiwara, N.; Yamamoto, T.; Toriyama, T.; Matsumura, S.; Kitagawa, H. Charge transfer dependence on CO2 hydrogenation activity to methanol in Cu nanoparticles covered with metal–organic framework systems. Chem. Sci. 2019, 10, 3289–3294. [Google Scholar] [CrossRef] [Green Version]
  33. Kang, S.-H.; Bae, J.W.; Prasad, P.S.S.; Oh, J.-H.; Jun, K.-W.; Song, S.-L.; Min, K.-S. Influence of Ga addition on the methanol synthesis activity of Cu/ZnO catalyst in the presence and absence of alumina. J. Ind. Eng. Chem. 2009, 15, 665–669. [Google Scholar] [CrossRef]
  34. Mota, N.; Guil-Lopez, R.; Pawelec, B.G.; Fierro, J.L.G.; Navarro, R.M. Highly active Cu/ZnO–Al catalyst for methanol synthesis: Eeffect of aging on its structure and activity. RSC Adv. 2018, 8, 20619–20629. [Google Scholar] [CrossRef] [Green Version]
  35. Arena, F.; Barbera, K.; Italiano, G.; Bonura, G.; Spadaro, L.; Frusteri, F. Synthesis, characterization and activity pattern of Cu–ZnO/ZrO2 catalysts in the hydrogenation of carbon dioxide to methanol. J. Catal. 2007, 249, 185–194. [Google Scholar] [CrossRef]
  36. Fei, J.-H.; Yang, M.-X.; Hou, Z.-Y.; Zheng, X.-M. Effect of the Addition of Manganese and Zinc on the Properties of Copper-Based Catalyst for the Synthesis of Syngas to Dimethyl Ether. Energy Fuels 2004, 18, 1584–1587. [Google Scholar] [CrossRef]
  37. Wu, J.; Luo, S.; Toyir, J.; Saito, M.; Takeuchi, M.; Watanabe, T. Optimization of preparation conditions and improvement of stability of Cu/ZnO-based multicomponent catalysts for methanol synthesis from CO2 and H2. Catal. Today 1998, 45, 215–220. [Google Scholar] [CrossRef]
  38. Ren, J.; Langmi, H.W.; North, B.C.; Mathe, M.; Bessarabov, D. Modulated synthesis of zirconium-metal organic framework (Zr-MOF) for hydrogen storage applications. Int. J. Hydrogen Energy 2014, 39, 890–895. [Google Scholar] [CrossRef]
  39. Cavka, J.H.; Jakobsen, S.; Olsbye, U.; Guillou, N.; Lamberti, C.; Bordiga, S.; Lillerud, K.P. A new zirconium inorganic building brick forming metal organic frameworks with exceptional stability. J. Am. Chem. Soc. 2008, 130, 13850–13851. [Google Scholar] [CrossRef]
  40. Dyosiba, X.; Ren, J.; Musyoka, N.M.; Langmi, H.W.; Mathe, M.; Onyango, M.S. Feasibility of Varied Polyethylene Terephthalate Wastes as a Linker Source in Metal–Organic Framework UiO-66(Zr) Synthesis. Ind. Eng. Chem. Res. 2019, 58, 17010–17016. [Google Scholar] [CrossRef]
  41. Li, Z.; Zhao, S.; Wang, H.; Peng, Y.; Tan, Z.; Tang, B. Functional groups influence and mechanism research of UiO-66-type metal-organic frameworks for ketoprofen delivery. Colloids Surf. B Biointerfaces 2019, 178, 1–7. [Google Scholar] [CrossRef]
  42. Li, C.; Yuan, X.; Fujimoto, K. Development of highly stable catalyst for methanol synthesis from carbon dioxide. Appl. Catal. A Gen. 2014, 469, 306–311. [Google Scholar] [CrossRef]
  43. Vellingiri, K.; Kumar, P.; Deep, A.; Kim, K.-H. Metal-organic frameworks for the adsorption of gaseous toluene under ambient temperature and pressure. Chem. Eng. J. 2017, 307, 1116–1126. [Google Scholar] [CrossRef]
  44. Valenzano, L.; Civalleri, B.; Chavan, S.; Bordiga, S.; Nilsen, M.H.; Jakobsen, S.; Lillerud, K.P.; Lamberti, C. Disclosing the Complex Structure of UiO-66 Metal Organic Framework: A Synergic Combination of Experiment and Theory. Chem. Mater. 2011, 23, 1700–1718. [Google Scholar] [CrossRef]
  45. Pan, Y.; Jiang, S.; Xiong, W.; Liu, D.; Li, M.; He, B.; Fan, X.; Luo, D. Supported CuO catalysts on metal-organic framework (Cu-UiO-66) for efficient catalytic wet peroxide oxidation of 4-chlorophenol in wastewater. Microporous Mesoporous Mater. 2020, 291, 109703. [Google Scholar] [CrossRef]
  46. Behrens, M. Meso-and nano-structuring of industrial Cu/ZnO/(Al2O3) catalysts. J. Catal. 2009, 267, 24–29. [Google Scholar] [CrossRef] [Green Version]
  47. Peng, B.; Feng, C.; Liu, S.; Zhang, R. Synthesis of CuO catalyst derived from HKUST-1 temple for the low-temperature NH3-SCR process. Catal. Today 2018, 314, 122–128. [Google Scholar] [CrossRef]
  48. Ye, Q.; Wang, L.; Yang, R.T. Activity, propene poisoning resistance and hydrothermal stability of copper exchanged chabazite-like zeolite catalysts for SCR of NO with ammonia in comparison to Cu/ZSM-5. Appl. Catal. A Gen. 2012, 427–428, 24–34. [Google Scholar] [CrossRef]
  49. Alayoglu, S.; Beaumont, S.K.; Zheng, F.; Pushkarev, V.V.; Zheng, H.; Iablokov, V.; Liu, Z.; Guo, J.; Kruse, N.; Somorjai, G.A. CO2 hydrogenation studies on Co and CoPt bimetallic nanoparticles under reaction conditions using TEM, XPS and NEXAFS. Top. Catal. 2011, 54, 778–785. [Google Scholar] [CrossRef] [Green Version]
  50. Zhong, L.; Chen, D.; Zafeiratos, S. A mini review of in situ near-ambient pressure XPS studies on non-noble, late transition metal catalysts. Catal. Sci. Technol. 2019, 9, 3851–3867. [Google Scholar] [CrossRef]
  51. Titus, D.; James Jebaseelan Samuel, E.; Roopan, S.M. Chapter 12—Nanoparticle characterization techniques. In Green Synthesis, Characterization and Applications of Nanoparticles; Shukla, A.K., Iravani, S., Eds.; Elsevier: Amsterdam, The Netherlands, 2019; pp. 303–319. [Google Scholar]
  52. Yang, Y.; Xu, Y.; Ding, H.; Yang, D.; Cheng, E.; Hao, Y.; Wang, H.; Hong, Y.; Su, Y.; Wang, Y.; et al. Cu/ZnOx@UiO-66 synthesized from a double solvent method as an efficient catalyst for CO2 hydrogenation to methanol. Catal. Sci. Technol. 2021, 11, 4367–4375. [Google Scholar] [CrossRef]
  53. Stangeland, K.; Kalai, D.; Li, H.; Yu, Z. CO2 Methanation: The Effect of Catalysts and Reaction Conditions. Energy Procedia 2017, 105, 2022–2027. [Google Scholar] [CrossRef]
  54. Joo, O.-S.; Jung, K.-D.; Moon, I.; Rozovskii, A.Y.; Lin, G.I.; Han, S.-H.; Uhm, S.-J. Carbon Dioxide Hydrogenation To Form Methanol via a Reverse-Water-Gas-Shift Reaction (the CAMERE Process). Ind. Eng. Chem. Res. 1999, 38, 1808–1812. [Google Scholar] [CrossRef]
  55. An, B.; Zhang, J.; Cheng, K.; Ji, P.; Wang, C.; Lin, W. Confinement of Ultrasmall Cu/ZnOx Nanoparticles in Metal–Organic Frameworks for Selective Methanol Synthesis from Catalytic Hydrogenation of CO2. J. Am. Chem. Soc. 2017, 139, 3834–3840. [Google Scholar] [CrossRef] [PubMed]
  56. Portha, J.-F.; Parkhomenko, K.; Kobl, K.; Roger, A.-C.; Arab, S.; Commenge, J.-M.; Falk, L. Kinetics of Methanol Synthesis from Carbon Dioxide Hydrogenation over Copper–Zinc Oxide Catalysts. Ind. Eng. Chem. Res. 2017, 56, 13133–13145. [Google Scholar] [CrossRef]
  57. Joo, S.H.; Park, J.Y.; Tsung, C.-K.; Yamada, Y.; Yang, P.; Somorjai, G.A. Thermally stable Pt/mesoporous silica core–shell nanocatalysts for high-temperature reactions. Nat. Mater. 2009, 8, 126–131. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) XRD patterns and (b) TGA profiles of commercial catalyst, Cu/ZnO/Al2O3/MgO catalyst, UiO-66, and Cu/ZnO/UiO-66. ♦ ZnO, ▼ CuO.
Figure 1. (a) XRD patterns and (b) TGA profiles of commercial catalyst, Cu/ZnO/Al2O3/MgO catalyst, UiO-66, and Cu/ZnO/UiO-66. ♦ ZnO, ▼ CuO.
Catalysts 12 00401 g001
Figure 2. Nitrogen sorption isotherms at 77 K of (a) commercial catalyst and Cu/ZnO/Al2O3/MgO catalyst; (b) UiO-66 and Cu/ZnO/UiO-66; and (c) FTIR spectra of UiO-66 and Cu/ZnO/UiO-66.
Figure 2. Nitrogen sorption isotherms at 77 K of (a) commercial catalyst and Cu/ZnO/Al2O3/MgO catalyst; (b) UiO-66 and Cu/ZnO/UiO-66; and (c) FTIR spectra of UiO-66 and Cu/ZnO/UiO-66.
Catalysts 12 00401 g002
Figure 3. SEM and TEM images of (a,b) UiO-66; (c,d) Cu/ZnO/UiO-66; (e,f) commercial catalyst; and (g,h) CuO/ZnO/Al2O3/MgO catalyst. The images were taken at different magnifications.
Figure 3. SEM and TEM images of (a,b) UiO-66; (c,d) Cu/ZnO/UiO-66; (e,f) commercial catalyst; and (g,h) CuO/ZnO/Al2O3/MgO catalyst. The images were taken at different magnifications.
Catalysts 12 00401 g003
Figure 4. XPS spectra of (a) Cu/ZnO/UiO-66 full survey, (b) Cu2p3, and (c) Zr3d.
Figure 4. XPS spectra of (a) Cu/ZnO/UiO-66 full survey, (b) Cu2p3, and (c) Zr3d.
Catalysts 12 00401 g004
Figure 5. H2-TPR profiles of the prepared catalysts: (a) Commercial and Cu/ZnO/Al2O3/MgO catalysts; (b) UiO-66 and Cu/ZnO/UiO-66.
Figure 5. H2-TPR profiles of the prepared catalysts: (a) Commercial and Cu/ZnO/Al2O3/MgO catalysts; (b) UiO-66 and Cu/ZnO/UiO-66.
Catalysts 12 00401 g005
Figure 6. CO2 conversions of catalysts. T = 230 °C; P = 50 bar; Qv, 0 = 40 mL·min−1; GHSV = 10,000 h−1.
Figure 6. CO2 conversions of catalysts. T = 230 °C; P = 50 bar; Qv, 0 = 40 mL·min−1; GHSV = 10,000 h−1.
Catalysts 12 00401 g006
Figure 7. Catalysts’ (a) methanol selectivity; (b) CO selectivity; (c) methane selectivity and (d) space–time yield.
Figure 7. Catalysts’ (a) methanol selectivity; (b) CO selectivity; (c) methane selectivity and (d) space–time yield.
Catalysts 12 00401 g007
Table 1. Specific surface area, pore volume, and elemental loading of prepared materials.
Table 1. Specific surface area, pore volume, and elemental loading of prepared materials.
SampleSSA (m2/g)Pore Volume (cm3/g)Cu wt%Zn wt%
Commercial catalyst970.18859.927.5
Cu/ZnO/Al2O3/MgO380.10361.727.7
UiO-6612380.471--
Cu/ZnO/UiO-665610.200124.2
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Duma, Z.G.; Dyosiba, X.; Moma, J.; Langmi, H.W.; Louis, B.; Parkhomenko, K.; Musyoka, N.M. Thermocatalytic Hydrogenation of CO2 to Methanol Using Cu-ZnO Bimetallic Catalysts Supported on Metal–Organic Frameworks. Catalysts 2022, 12, 401. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12040401

AMA Style

Duma ZG, Dyosiba X, Moma J, Langmi HW, Louis B, Parkhomenko K, Musyoka NM. Thermocatalytic Hydrogenation of CO2 to Methanol Using Cu-ZnO Bimetallic Catalysts Supported on Metal–Organic Frameworks. Catalysts. 2022; 12(4):401. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12040401

Chicago/Turabian Style

Duma, Zama G., Xoliswa Dyosiba, John Moma, Henrietta W. Langmi, Benoit Louis, Ksenia Parkhomenko, and Nicholas M. Musyoka. 2022. "Thermocatalytic Hydrogenation of CO2 to Methanol Using Cu-ZnO Bimetallic Catalysts Supported on Metal–Organic Frameworks" Catalysts 12, no. 4: 401. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12040401

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