Next Article in Journal
Decorating Zirconium on Graphene Oxide to Design a Multifunctional Nanozyme for Eco-Friendly Detection of Hydrogen Peroxide
Next Article in Special Issue
Enhanced CuAl2O4 Catalytic Activity via Alkalinization Treatment toward High CO2 Conversion during Reverse Water Gas Shift Reaction
Previous Article in Journal
Research on Modified Carbon Nanotubes in Wastewater Treatment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Towards High CO2 Conversions Using Cu/Zn Catalysts Supported on Aluminum Fumarate Metal-Organic Framework for Methanol Synthesis

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, 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: 30 August 2022 / Revised: 16 September 2022 / Accepted: 19 September 2022 / Published: 24 September 2022

Abstract

:
Green methanol is a viable alternative for the storage of hydrogen and may be produced from captured anthropogenic sources of carbon dioxide. The latter was hydrogenated over Cu-ZnO catalysts supported on an aluminum fumarate metal-organic framework (AlFum MOF). The catalysts, prepared via slurry phase impregnation, were assessed for thermocatalytic hydrogenation of CO2 to methanol. PXRD, FTIR, and SBET exhibited a decrease in crystallinity of the AlFum MOF support after impregnation with Cu-Zn active sites. SEM, SEM-EDS, and TEM revealed that the morphology of the support is preserved after metal loading, where H2-TPR confirmed the presence of active sites for hydrogen uptake. The catalysts exhibited good activity, with a doubling in Cu and Zn loading over the AlFum MOF, resulting in a 4-fold increase in CO2 conversions from 10.8% to 45.6% and an increase in methanol productivity from 34.4 to 56.5 gMeOH/Kgcat/h. The catalysts exhibited comparatively high CO selectivity and high yields of H2O, thereby favoring the reverse water-gas shift reaction. The selectivity of the catalysts towards methanol was found to be 12.9% and 6.9%. The performance of the catalyst supported on AlFum MOF further highlights the potential use of MOFs as supports in the heterogeneous thermocatalytic conversion of CO2 to value-added products.

Graphical Abstract

1. Introduction

It is evident that the continuous increase in anthropogenic CO2 levels in the atmosphere has resulted in global warming, which leads to higher average temperatures on Earth and an increase in the acidity of oceans [1,2,3]. As such, driven by the large-scale combustion of carbonaceous fossil fuel sources such as coal, natural gas, and oil, CO2 concentrations have increased by 20% in the past four decades [4]. Therefore, to abate climate change, the mitigation of CO2 emissions is required via capture, storage, and utilization. The latter includes the chemical transformation of CO2 to value-added chemicals such as, inter alia, methanol, dimethyl ether, organic acids, higher alcohols, and hydrocarbons [4,5]. The valorized products have a myriad of applications as chemical feedstocks and in direct usage, and in the case of methanol, as fuel sources in internal combustion engines, which makes it an ideal energy carrier [6,7].
The chemical conversion of CO2 to methanol is predominantly carried out over ternary co-precipitated Cu/ZnO/Al2O3 catalysts, which are typically used for methanol production from synthesis gas (CO, CO2, H2), at industrial conditions of 210–300 °C and 50–80 bar [8,9,10,11]. However, the deactivation of the catalysts driven by thermal sintering of the Cu particles, phase segregation of ZnO and Al2O3 from Cu active sites, and decreased dispersion result in poor performance concerning the selectivity to methanol, thereby facilitating CO production [12,13,14]. Furthermore, zinc has been reported to react with the Al2O3 promoter to form zinc aluminates, which further deactivates the catalyst [14]. As such, the preparation of supported catalysts that favor good Cu dispersion, high activity toward CO2 hydrogenation, and methanol selectivity may prevent the aforementioned catalyst deactivation mechanisms [15]. Therefore, the usage of metal-organic frameworks (MOFs), which are materials with exceptional crystallinity, high surface area, and good porosity with a large surface area as supports for Cu-based catalysts promises to circumvent the poor performance inherent to ternary, co-precipitated Cu-based catalysts [16,17].
The application of MOFs, characterized by metal-ion clusters coordinated to multidentate ligands, as catalyst supports has been reported to increase methanol selectivity and turnover frequency [18,19]. Therefore, the ability of MOFs as supports that have high metallic active phase dispersion makes them attractive in heterogeneous catalysis [20]. In this study, we present, for the first time according to our knowledge, the loading of Cu-ZnO on aluminum fumarate MOF (AlFum MOF) for the thermocatalytic hydrogenation of CO2 to methanol.

2. Results

2.1. Characterization

2.1.1. Powder X-ray Diffraction (PXRD) and Fourier Transform Infrared Spectroscopy (FTIR)

The PXRD patterns of the catalysts supported on AlFum MOF are shown in Figure 1a. The 7Cu/3ZnO/AlFum MOF catalyst exhibits a diffraction pattern typical of AlFum MOF, with characteristic peaks at 2θ = 10.2°, 14.9°, 21.1°, 31.8° and 42.7°, with no Cu or Zn peaks observed, which may be due to low metal loading, and good dispersion [16,21]. Figure 1a also shows the diffraction pattern of the pristine AlFum MOF whereas Figure 1b, zooms into the 30° < 2θ < 40° region of Cu/ZnO/AlFum MOF catalysts. In comparison to 7Cu/3ZnO/AlFum MOF, 15Cu/6.4ZnO/AlFum MOF has double the loading of Cu and Zn and, as a result, a CuO peak at 2θ = 38.7° can be observed [7,22]. Furthermore, it can be seen that the crystallinity of the AlFum MOF in the supported catalysts decreases, which may be due to the high thermal treatment temperature of 350 °C under a constant flow of argon during thermal activation.
The FTIR spectra of the pristine AlFum MOF and Cu/ZnO/AlFum MOF catalysts are shown in Figure 1c–e. The broad peak at 3400–3600 cm−1 is ascribed to O-H vibrations from physisorbed or intercrystalline water and/or absorbed moisture [23,24,25]. The symmetric and asymmetric vibrations of the fumarate carboxylate group can be seen at 1403 and 1591 cm−1 [21,23]. Furthermore, it can be seen in the Cu/ZnO/AlFum MOF catalysts, Figure 1c,d, that after impregnation and thermal treatment, the intensity of the peaks decreases, which corroborates the decrease in crystallinity observed in the PXRD diffractograms, as shown in Figure 1a,b.

2.1.2. N2 Sorption, and Elemental Loading

The pristine AlFum MOF exhibited a BET surface area (SBET) of 910 m2/g, which is consistent with previously reported values [26]. The nitrogen sorption isotherms of the pristine and AlFum MOF-supported catalysts are shown in Figure 2a whereas the H-K pore volumes and NLDFT PSDs are summarized in Table 1. It can be observed that the pristine MOF and catalysts exhibit type-I sorption isotherms, which are indicative of their microporous structure [16]. In addition, the PSDs, shown in Figure 2b, show that the majority of the pores in the AlFum MOF-support and catalysts are in the micropore range, with a median pore diameter below 2 nm, as shown in Table 1 [21,23]. Furthermore, the decrease in SBET observed in the impregnated MOFs to 757 and 416 m2/g is due to the loading of Cu and Zn active sites and thermal treatment [16,27]. The decrease in SBET observed in both the AlFum MOF-supported catalysts is consistent with the decrease in the intensity of the PXRD and FTIR peaks in Figure 1. In addition, a decrease in the pore volume of 0.344 cm3/g in pristine AlFum MOF upon impregnation to 0.300 and 0.148 cm3/g in the 7Cu/3ZnO/AlFum MOF and 15Cu/6.4ZnO/AlFum MOF catalysts is shown in Table 1. Furthermore, SEM-EDX elemental analysis revealed a Cu loading of 11.2 wt% and 18.2 wt% compared with a theoretical composition of 7 and 15 wt% shown in parentheses. In addition, the Zn content was 3.12 and 5.77 wt% in the Cu/ZnO/AlFum MOF and 7Cu/3ZnO/AlFum MOF catalysts, respectively. This discrepancy may be attributed to high Cu dispersion in the framework of the AlFum MOF support with regions of surface enrichment, potentially leading to higher Cu content albeit it must be noted that SEM-EDX is a semi-quantitative technique and other techniques such as inductively coupled plasma optical emission spectroscopy may provide more accurate, quantitative, elemental analysis results [28,29].

2.1.3. Electron Microscopies

The morphology images obtained using SEM and TEM of the pristine AlFum MOF support and Cu/ZnO/AlFum MOF catalysts are shown in Figure 3. The AlFum MOF support particles exhibit a lozenge, quadrilateral shape characterized by the geometry of Al-octahedra coordinated by fumarate linkers [23,30]. The AlFum MOF particles exhibit some degree of agglomeration, which is consistent with literature reports [26]. The Cu/ZnO/AlFum MOF catalysts, as shown in Figure 3c–f, have a similar quadrilateral morphology to the pristine AlFum MOF prepared via the solvothermal method, as shown in Figure 3a,b [31]. Therefore, it can be seen that the morphology of the AlFum MOF support is preserved after impregnation, drying, and thermal treatment. This may be attributed to the good relative thermal stability of the support [32]. The SEM-EDX elemental maps of the 7Cu/3ZnO/AlFum MOF and 15Cu/6.4ZnO/AlFum MOF catalysts are shown in Figure 3g,h.
In both catalysts, a uniform, homogenous distribution of Cu and Zn is observed, which is indicative of successful impregnation. In addition, the micrographs show that in contrast to the pristine AlFum MOF support in Figure 3a,b, tiny spherical nanoparticles can be observed on the surface of the MOF support in the catalysts, which may be attributed to Cu and Zn nanoparticles, as shown in Figure 3c–f. The individual SEM-EDX elemental maps of each metal in the 7Cu/3ZnO/AlFum MOF and 15Cu/6.4ZnO/AlFum MOF catalysts are shown in Figures S1 and S2 (Supplementary Material), where the homogenous, uniform dispersion of Cu and Zn over the AlFum MOF support observed in Figure 3g,h is further corroborated.

2.1.4. Thermogravimetric Analysis (TGA) and Hydrogen-Temperature Programmed Reduction (H2-TPR)

The TGA profiles of the pristine AlFum MOFs and Cu/ZnO/AlFum MOF catalysts are shown in Figure 4a whereas the derivative TGA plots are shown in Figures S3–S5 (Supplementary Material). The TGA profile of the pristine AlFum MOF is consistent with previously reported literature results, where the initial mass loss below 100 °C is attributed to physisorbed water, which is in accordance with the -OH bending vibration FTIR peak at 3300–3600 cm−1 in FTIR, as shown in Figure 1c–e [21,24]. The weight loss peak at 172 °C is due to the removal of residual DMF solvent used during the synthesis step, as shown in Figure 4a [24]. The derivative TGA plot of the pristine AlFum MOF is shown in Figure S3 (Supplementary Material), where an additional weight loss peak at 469 °C indicates the thermal degradation of the fumarate organic linkers [21]. In addition, the TGA profiles and derivative TGA plots (Supplementary Material, Figures S3–S5) reveal that the thermal stability in the catalysts decreases after impregnation and thermal activation, where the weight loss of the fumarate organic linker temperatures shifts from 469 °C in the pristine AlFum MOF to 375 and 370 °C in the 7Cu/3ZnO/AlFum MOF and 15Cu/6.4ZnO/AlFum MOF catalysts (Supplementary Material, Figures S3–S5). The rapid weight loss indicates the decomposition of the organic linkers followed by the formation of relatively thermally stable metal oxides [21,33,34].
The H2-TPR profiles of 7Cu/3ZnO/AlFum MOF and 15Cu/6.4ZnO/AlFum MOF catalysts exhibit α reduction peaks indicative of the presence of CuO at Tmax = 213 °C (α1) and 229 °C (α2), as shown in Figure 4b. The β peaks at Tmax = 339 °C (β1) and 340 Tmax = 340 °C (β2) indicate the presence of a second phase of CuO that formed after thermal treatment [35]. The α peaks indicate the reduction in Cu ions or CuO clusters that are in close contact with the AlFum MOF whereas the β peaks indicate the reduction in “bulk-like” CuO particles [36]. Furthermore, the α peaks indicate the reduction in finely dispersed CuO particles over the surface of the AlFum MOF support [37]. The H2-reduction peaks further corroborate the successful loading of Cu and the formation of the desired CuO species after thermal treatment. The pristine peak of AlFum MOF did not show any reduction peaks due to the absence of Cu active sites.

2.1.5. X-ray Photoelectron Spectroscopy (XPS)

The surface oxidation states of the Cu and Zn active sites in the catalysts are shown in Figure 5. Conventionally, the peaks at 932.5 ± 0.2 and 952.3 ± 0.2 eV indicate Cu+ 2p3/2 and 2p1/2 oxidation states whereas the peaks at 933.7 ± 0.2 and 953.6 ± 0.2 eV are ascribed to Cu2+ 2p3/2 and 2p1/2, respectively [38,39]. In addition, a satellite peak at 943.2 eV is indicative of Cu2+ [39]. In both catalysts, the peaks at 935 and 954 eV are indicative of the Cu2+ species, which is further corroborated by the satellite peak at 944 eV, respectively [40]. The XPS spectra confirm the presence of CuO in both catalysts. Figure 5b shows the spectra of Zn 2p, where the peaks at 1022 and 1046 eV indicate the presence of Zn2+ 2p3/2 and 2p1/2 [41].

2.2. Catalysts’ Testing and Evaluation

Conversions, Selectivity, and Productivity

The CO2 conversion results of all tested catalysts, including a commercial catalyst, are shown in Figure 6. The main reaction of interest is the exothermic, direct CO2 hydrogenation to methanol shown in Equation (1). In addition to methanol, the catalysts formed CO, H2O, and CH4, with no other products observed. The competing, endothermic reverse water-gas shift reaction (RWGS) results in the formation of CO, where CO2 is hydrogenated to CO and H2O, as shown in Equation (2) [8,12,17,42,43]. In addition, CH4 was a by-product that was formed in the reactor presumably from the Sabatier reaction, where CO2 methanation occurs to form CH4 and H2O as shown in Equation (3) [44]. The 15Cu/6.4ZnO/AlFum MOF catalyst (45.6%) had the best average CO2 conversion after 24 h onstream in comparison to the commercial catalyst (29.8%) and the 7Cu/3ZnO/AlFum MOF catalyst (10.8%), respectively. In addition, the catalysts showed relatively good stability over the 24-h testing period. In our previous studies of the thermodynamic equilibrium of the CO2 hydrogenation reaction, it was shown that the best H2 and CO2 conversions and the optimal CH3OH selectivity are obtained at low temperatures. For example, at 50 bars and 230 °C, the conversion of CO2 could achieve 37% and 25% for H2 conversion [45]. At high temperatures, the additional increase in CO2 conversion up to 45% is clearly correlated with CO production with 100% selectivity. Such high conversion (45.6%) observed in the 15Cu/6.4ZnO/AlFum MOF catalyst at a lower temperature of 230 °C is phenomenal and could be correlated to the changes in the selectivity and thus extensive CO production and the presence of CH4. The continuous increase in the slope of the CO2 conversion for 7Cu/3ZnO/AlFum MOF catalyst could be attributed to the reconstruction of the catalytic active sites during reaction; more precisely, the formation of metallic copper nanoparticles stabilized by the concomitant structure. Due to the low copper loading in the 7Cu/3ZnO/AlFum MOF catalyst, the nanoparticles grow, and stabilization may take time:
CO 2 + 3 H 2     CH 3 OH +   H 2 O   Δ H = 49.47   KJ / mol
CO 2 +   H 2     CO +   H 2 O   Δ H = 41.14   KJ / mol  
CO 2 + 4 H 2     CH 4 + 2 H 2 O   Δ H = 165   KJ / mol
The methanol selectivity of the catalysts was also investigated. The commercial catalyst exhibited the highest selectivity (15.0%) whereas 7Cu/3ZnO/AlFum MOF and 15Cu/6.4ZnO/AlFum MOF exhibited relatively lower methanol selectivity of 12.9% and 6.90%, as shown in Figure 7a. In addition, CO was produced by all the evaluated catalysts due to the competing RWGS reaction [46]. The evaluated catalysts’ selectivity to CO was significant in all catalysts, with the highest at 88% displayed by the 15Cu/6.4ZnO/AlFum MOF catalyst followed by the 7Cu/3ZnO/AlFum MOF catalyst (85.2%) whereas the selectivity of the commercial catalyst was the lowest at 84%, as shown in Figure 7a. In addition, the catalysts produced CH4, which is attributed to the CO2 methanation reaction [44]. Therefore, the 15Cu/6.4ZnO/AlFum MOF and 7Cu/3ZnO/AlFum MOF catalysts had the highest selectivity to CH4 at 5.17% and 1.82%, respectively, whereas the commercial catalyst had the lowest CH4 selectivity at 0.93%. Furthermore, the mole product distribution, as shown in Table 2, indicates that the catalysts form a significantly larger amount of CO and H2O, which further corroborates the high selectivity towards CO, which is attributed to the RWGS reaction. Furthermore, all the catalysts produced 48–50% of H2O due to both the RWGS and CO2 to methanol reaction shown in Equations (1) and (2), respectively.
In contrast, Bonura et al. observed a methanol selectivity of 49% using copper-zinc-zirconia catalysts at 240 °C, 30 bar, and GHSV of 10,000 h−1. However, the catalysts had a relatively low degree of CO2 conversion of 16.8% compared with 15Cu/6.4ZnO/AlFum MOF catalyst (45.6%) [47]. Furthermore, An et al. also reported a comparatively high methanol selectivity of 100% albeit the CO2 conversion was 3.3% at 250 °C, 40 bar, and GHSV of 6000 h−1 [40]. Though the 15Cu/6.4ZnO/AlFum MOF catalyst showed relatively low selectivity toward methanol, it exhibited a superior degree of CO2 conversion. As such, the dispersion of Cu and Zn active sites on AlFum MOF demonstrates the advantages of using highly crystalline, high surface area, porous materials, where 10–15% of the active sites may be loaded on the catalysts as opposed to 70% in conventional, co-precipitated catalysts. Furthermore, AlFum MOF is commercially sold by BASF as Basolite A520TM, and a water-based green synthesis procedure has been adopted for the production of green AlFum MOF at a ton-level. Therefore, the scalability of the AlFum MOF via a green, hydrothermal process makes the MOF, Basolite A520TM, a sustainable and alternative support for Cu-based catalysts to drive sustainable CO2 capture and utilization [26,30,31,48].
The methanol space-time yield (STY) obtained over 15Cu/6.4ZnO/AlFum MOF catalyst was found to be 56.5 gMeOH/kgcat/h followed by the commercial catalyst at 51.8 gMeOH/kgcat/h, whereas 7Cu/3ZnO/AlFum MOF catalyst led to the lowest STY of 34.4 gMeOH/kgcat/h (Figure 7b). The methanol STY increased with an increase in the Cu loading: 11.2% (7Cu/3ZnO/AlFum MOF), 18.2% (15Cu/6.4ZnO/AlFum MOF), and 64.2% (commercial catalyst). It appears that increasing the loading of Cu on high surface area microporous materials such as MOFs enhances the catalytic activity with respect to CO2 conversion and methanol productivity due to the high dispersion conferred by the porous support [17,49].

3. Methods and Materials

3.1. Reagents

Aluminum chloride hexahydrate (AlCl3·6H2O, Associated Chemical Enterprises, Theta, Johannesburg, South Africa), cupric nitrate trihydrate (Cu(NO3)2·3H2O, Associated Chemical Enterprises, 99.9%, Theta, Johannesburg, South Africa), N,N-dimethyl formamide (DMF, Associated Chemical Enterprises, 99.5%, Johannesburg, Theta, South Africa), sodium hydroxide (NaOH, Associated Chemical Enterprises, Theta, Johannesburg, South Africa), and fumaric acid (C4H4O4, Sigma-Aldrich, 99.9%, Louis, MO, United States) were procured and used directly without any purification. De-ionized water from a water demineralization system, Thermo Fischer Barnstead Smart2Pure, located in the laboratory was used. The commercial catalyst was bought from Alfa Aesar. The pelletized catalyst with a composition of 64.2/24.5/9.8/1.3 wt% (Cu/Zn/Al/Mg) was crushed to a powdery texture prior to usage.

3.2. Synthesis of AlFum MOF

The solvothermal synthesis of AlFum MOF was carried out at ambient pressure. In the procedure, 19.62 g of AlCl3·6H2O and 14.76 g of fumaric acid were simultaneously added to 400 mL of DMF in a round-bottom flask. Thereafter, the resultant mixture was stirred at 250 rpm and 130 °C for 4 days. The suspension was then washed twice with acetone and methanol, collected using centrifugation, and dried at 80 °C in an oven. The samples were activated at 90 °C, under vacuum using a rotary evaporator. Before impregnation with Cu and Zn, the AlFum MOF support was sieved to a size fraction of 50–125 µm.

3.3. Synthesis of AlFum MOF-Supported Catalysts

Cu(NO3)2·3H2O (5.50 mmol) and Zn(NO3)2·6H2O (2.33 mmol) were dissolved using a magnetic stirrer in deionized water (30 mL) in a round-bottom flask. Thereafter, the Cu and Zn-nitrate solution was added slowly to 4.50 g of AlFum MOF whilst stirring continuously. The obtained slurry was continuously mixed at 90 °C, 50 rpm, at ambient pressure for 2 h in a rotary evaporator. Thereafter, the temperature was decreased to 65 °C and the vacuum was activated and maintained for 4 h to furnish a dry, free-flowing powder. The catalyst was labeled 7Cu/3ZnO/AlFum MOF. A similar procedure was followed to produce a catalyst with double the loading of Cu and Zn by mixing Cu(NO3)2·3H2O (9.44 mmol) and Zn(NO3)2·6H2O (4.00 mmol) with AlFum MOF (3.16 g). The catalyst was labeled 15Cu/6.4ZnO/AlFum MOF.
The dried catalyst precursors were heated at 350 °C under a constant flow of argon. The precursors were heated at a rate of 5 °C/min with a hold time of 4 h to furnish catalysts with a theoretical weight percent (wt%) composition of 7/3/90 (7Cu/3ZnO/AlFum MOF) and 15/6.4/78.6 (15Cu/6.4ZnO/AlFum MOF), respectively.

3.4. Characterization

Powder X-ray diffraction (PXRD) was used for crystallinity and the identification of different phases. A Rigaku Ultima IV diffractometer was equipped with Ni-filtered radiation (0.154 nm), 40 kV voltage, and 30 mA current at ambient temperature (Akishima-shi, Tokyo, Japan). All analyses were carried out at 2θ = 3–90° with steps of 0.01° s−1.
The Mettler Toledo, TGA/SDTA 851e instrument (Mettler Toledo, Greifensee, Switzerland) at a heating rate of 10 °C min−1 to 900 °C under a constant flow of nitrogen was used for thermogravimetric analysis (TGA). In the analysis, 10 mg of sample was used.
The ASAP 2020 HD analyser (Micromeritics Instrument Corporation, Norcross, GA, USA) was employed to determine the porosity and surface area. The Brunauer-Emmett-Teller (BET) surface areas were measured with nitrogen gas adsorption-desorption isotherms at −196 °C and relative partial pressures (p/po) of up to 1. Non-linear density functional theory (NLDFT) was used for the estimation of the pore size distributions (PSDs). H-K isotherms were used to determine the pore volumes. Prior to analysis, all samples (>0.2 g) were outgassed under vacuum at temperatures between 150 and 200 °C.
A PerkinElmer Spectrum 100 FTIR spectrometer (PerkinElmer, Waltham, MA, USA) was used for Fourier transform infrared spectroscopy (FTIR) analysis. The spectra were collected at 4000–550 cm−1.
Helium pycnometry was used to determine the apparent density of the samples. A Micromeritics AccuPyc II 1340 pycnometer (Micromeritics Instrument Corporation, Norcross, GA, USA) was employed with the analysis pressure of helium gas maintained at 135 kPa.
An Auriga Cobra focused-ion beam scanning electron microscope (SEM) coupled with energy-dispersive X-ray spectroscopy (EDS) (Zeiss, Oberkochen, Germany) was used for imaging, elemental mapping, and elemental loading determination in the pristine AlFum MOF and Cu/ZnO/AlFum MOF catalysts. All samples were coated with carbon prior to analysis.
Transmission electron microscope (TEM) imaging was carried out using a JEOL-Jem 2100 model (Joel, Tokyo, Japan), where, prior to analyses, each sample was dispersed in ethanol via ultrasonication for 20 min, transferred to a carbon-coated copper grid, and left to dry overnight. The samples were then mounted onto the stage and analyzed under vacuum.
A Micromeritics Autochem II 2920 analyser (Micromeritics Instrument Corporation, Norcross, GA, USA) was used to determine the reduction profiles of the catalysts via hydrogen temperature-programmed reduction (H2-TPR). A heating rate of 10 °C/min to 150 °C under constant helium flow was used to degas the samples, which were positioned between glass wool in a quartz U-tube. After cooling to 50 °C, each sample was heated at 10 °C/min to 900 °C whilst a constant flow (50 mL/min) of 5% H2/Ar was maintained. A thermal conductivity detector (TCD) was used to measure the consumption of hydrogen.
X-ray photoelectron spectroscopy (XPS) was used to determine the chemical state of the Cu and Zn active sites. A Thermo Scientific ESCALAB 250Xi instrument (Thermo Fischer Scientific, Kandel, Germany) at 300 W, 100 eV, and X-ray aluminum source at 1486.7 eV was used. All analyses were performed in ultrahigh vacuum of <10−8 mbar.

3.5. Catalyst Testing

A fixed-bed tubular reactor with dimensions of 27 cm length, internal diameter of 1.01 cm, and external diameter of 1.27 cm was used to evaluate the performance of the catalysts. Each catalyst was sieved to 50–125 µm followed by loading in a catalyst bed with glass wool and placed in the middle of the reactor. After loading in the reactor, the catalysts were reduced with hydrogen at a flow rate of 40 mL/min for 2 h, temperature of 250 °C, and pressure of 20 bar. Subsequently, the temperature of the tubular reactor was decreased to 50 °C, and the gas flow changed to 72% H2, 24% CO2, and 4% Ar from a gas bottle mixture. The reactor was then left for 12 h for the gas composition to stabilize. The hydrogenation of CO2 was performed for 24 h under reactor conditions: temperature = 230 °C; pressure = 50 bar; flow rate = 40 mL/min; and GHSV = 10,000 h−1. An Agilent gas chromatograph instrument: Model 7890B (Agilent Technologies Inc., Santa Clara, CA, USA) was employed for analysis. The hydrocarbons, CH4 and CH3OH, were analyzed using a flame ionization detector (FID) and the other gases, CO, CH4, CO2, and Ar, were measured with two TCDs. The concentration of the gases was analyzed online and the condensate (CH3OH and H2O) was collected in a steel trap immersed in an ice-water bath and analyzed offline. The conversions of CO2, X, 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 ) ] × 100
Selectivity to methanol, CO, and CH4 was calculated using Equations (5)–(7):
S CH 3 OH = n CH 3 OH n CH 3 OH +   n CH 4 +   n CO   × 100
S CO = n CO n CH 3 OH +   n CH 4 +   n CO   × 100
S CH 4 = [ 1 ( S CO +   S CH 3 OH ) ] × 100
where CO2 (in/out) and Ar (in/out) refer to the initial and outgoing concentrations of CO2 and Ar, respectively; and S and n refer to the selectivity and moles of each compound, respectively.

4. Conclusions

In this work, Cu-based catalysts supported on AlFum MOF were, for the first time, successfully prepared and evaluated for thermocatalytic hydrogenation of CO2. The catalyst preparation method resulted in a homogenous, uniform dispersion of Cu and Zn active sites. Doubling the copper loading resulted in a 4-fold increase in CO2 hydrogenation from 10.8% to 45.6% and an enhancement in methanol productivity. The catalysts exhibited relatively high selectivity toward CO and low selectivity toward methanol when benchmarked with a commercial-grade catalyst. Furthermore, the evaluated catalysts favored the reverse water-gas shift reaction and thereby formed considerable amounts of CO and H2O under the catalyst testing conditions employed in this study.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/catal12101104/s1, Figure S1: Individual elemental maps of 7Cu/3Zn/AlFum MOF. (a) Al Kα1; (b) Cu Kα1; and (c) Zn Kα1; Figure S2: Individual elemental maps of 15Cu/6.4Zn/AlFum MOF. (a) Al Kα1; (b) Cu Kα1; and (c) Zn Kα1; Figure S3: Derivative TGA plot of AlFum MOF. Figure S4: Derivative TGA plot of 7Cu/3ZnO/AlFum MOF; Figure S5: Derivative TGA plot of 15Cu/6.4ZnO/AlFum MOF.

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. and B.L.; supervision, N.M.M., H.W.L. and J.M.; Formal analysis, Z.G.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).

Data Availability Statement

Not applicable.

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 the strengthening of research collaborations between the two countries (Project No. CNMH02X).

Conflicts of Interest

The authors declare no personal and financial conflict of interest.

References

  1. Jiang, X.; Nie, X.; Guo, X.; Song, C.; Chen, J.G. Recent Advances in Carbon Dioxide Hydrogenation to Methanol via Heterogeneous Catalysis. Chem. Rev. 2020, 120, 7984–8034. [Google Scholar] [CrossRef] [PubMed]
  2. 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]
  3. 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]
  4. 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]
  5. Xiaoding, 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]
  6. Bozzano, G.; Manenti, F. Efficient methanol synthesis: Perspectives, technologies and optimization strategies. Prog. Energy Combust. Sci. 2016, 56, 71–105. [Google Scholar] [CrossRef]
  7. 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]
  8. 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]
  9. 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, NY, USA, 2014; pp. 1–40. [Google Scholar] [CrossRef]
  10. 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, 545431. [Google Scholar] [CrossRef]
  11. 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]
  12. 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]
  13. 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]
  14. 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]
  15. Fisher, I.A.; Bell, A.T. In Situ Infrared Study of Methanol Synthesis from H2/CO over Cu/SiO2 and Cu/ZrO2/SiO2. J. Catal. 1998, 178, 153–173. [Google Scholar] [CrossRef]
  16. 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]
  17. 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 CO(2) Hydrogenation to Methanol. Nano Lett. 2016, 16, 7645–7649. [Google Scholar] [CrossRef] [PubMed]
  18. 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]
  19. Farrusseng, D.; Aguado, S.; Pinel, C. Metal-organic frameworks: Opportunities for catalysis. Angew. Chem. Int. Ed. 2009, 48, 7502–7513. [Google Scholar] [CrossRef]
  20. Liang, J.; Liang, Z.; Zou, R.; Zhao, Y. Heterogeneous catalysis in zeolites, mesoporous silica, and metal-organic frameworks. Adv. Mater. 2017, 29, 1701139. [Google Scholar] [CrossRef] [PubMed]
  21. Rambau, K.M.; Musyoka, N.M.; Panek, R.; Franus, W.; Wdowin, M.; Manyala, N. Preparation of coal fly ash derived metal organic frameworks and their carbon derivatives. Mater. Today Commun. 2021, 27, 102433. [Google Scholar] [CrossRef]
  22. 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]
  23. Karmakar, S.; Dechnik, J.; Janiak, C.; De, S. Aluminium fumarate metal-organic framework: A super adsorbent for fluoride from water. J. Hazard Mater. 2016, 303, 10–20. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, Y.; Qu, Q.; Liu, G.; Battaglia, V.S.; Zheng, H. Aluminum fumarate-based metal organic frameworks with tremella-like structure as ultrafast and stable anode for lithium-ion batteries. Nano Energy 2017, 39, 200–210. [Google Scholar] [CrossRef]
  25. 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]
  26. Teo, H.W.B.; Chakraborty, A.; Kitagawa, Y.; Kayal, S. Experimental study of isotherms and kinetics for adsorption of water on Aluminium Fumarate. Int. J. Heat Mass Transf. 2017, 114, 621–627. [Google Scholar] [CrossRef]
  27. 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]
  28. Smyrnioti, M.; Ioannides, T. Dimethyl Ether Hydrolysis over WO3/γ-Al2O3 Supported Catalysts. Catalysts 2022, 12, 396. [Google Scholar] [CrossRef]
  29. Felgueiras, M.B.S.; Restivo, J.; Sousa, J.P.S.; Pereira, M.F.R.; Soares, O.S.G.P. Copper Supported on Mesoporous Structured Catalysts for NO Reduction. Catalysts 2022, 12, 170. [Google Scholar] [CrossRef]
  30. Alvarez, E.; Guillou, N.; Martineau, C.; Bueken, B.; Van de Voorde, B.; Le Guillouzer, C.; Fabry, P.; Nouar, F.; Taulelle, F.; De Vos, D. The structure of the aluminum fumarate metal-organic framework A520. Angew. Chem. 2015, 127, 3735–3739. [Google Scholar] [CrossRef]
  31. Kayal, S.; Chakraborty, A.; Teo, H.W.B. Green synthesis and characterization of aluminium fumarate metal-organic framework for heat transformation applications. Mater. Lett. 2018, 221, 165–167. [Google Scholar] [CrossRef]
  32. Jeremias, F.; Fröhlich, D.; Janiak, C.; Henninger, S.K. Advancement of sorption-based heat transformation by a metal coating of highly-stable, hydrophilic aluminium fumarate MOF. RSC Adv. 2014, 4, 24073–24082. [Google Scholar] [CrossRef]
  33. Coelho, J.A.; Ribeiro, A.M.; Ferreira, A.F.P.; Lucena, S.M.P.; Rodrigues, A.E.; Azevedo, D.C.S.d. Stability of an Al-Fumarate MOF and Its Potential for CO2 Capture from Wet Stream. Ind. Eng. Chem. Res. 2016, 55, 2134–2143. [Google Scholar] [CrossRef]
  34. 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. [Google Scholar] [CrossRef]
  35. Kamyar, N.; Khani, Y.; Amini, M.M.; Bahadoran, F.; Safari, N. Copper-based catalysts over A520-MOF derived aluminum spinels for hydrogen production by methanol steam reforming: The role of spinal support on the performance. Int. J. Hydrog Energy 2020, 45, 21341–21353. [Google Scholar] [CrossRef]
  36. Avgouropoulos, G.; Ioannides, T. Effect of synthesis parameters on catalytic properties of CuO-CeO2. Appl. Catal. B Environ. 2006, 67, 1–11. [Google Scholar] [CrossRef]
  37. Djinovic, P.; Batista, J.; Levec, J.; Pintar, A. Influence of morphological, redox and surface acidity properties on WGS activity of CuO–CeO2 catalysts. J. Chem. Eng. Jpn. 2009, 42, s3–s9. [Google Scholar] [CrossRef]
  38. 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]
  39. 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]
  40. 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]
  41. Farrusseng, D.; Daniel, C.; Hamill, C.; Casaban, J.; Didriksen, T.; Blom, R.; Velte, A.; Fueldner, G.; Gantenbein, P.; Persdorf, P.; et al. Adsorber heat exchanger using Al-fumarate beads for heat-pump applications—A transport study. Faraday Discuss. 2021, 225, 384–402. [Google Scholar] [CrossRef]
  42. 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]
  43. García, A.C.; Moral-Vico, J.; Abo Markeb, A.; Sánchez, A. Conversion of Carbon Dioxide into Methanol Using Cu–Zn Nanostructured Materials as Catalysts. Nanomaterials 2022, 12, 999. [Google Scholar] [CrossRef] [PubMed]
  44. 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]
  45. Angelo, L.; Kobl, K.; Tejada, L.M.M.; Zimmermann, Y.; Parkhomenko, K.; Roger, A.-C. Study of CuZnMOx oxides (M=Al, Zr, Ce, CeZr) for the catalytic hydrogenation of CO2 into methanol. Comptes Rendus Chim. 2015, 18, 250–260. [Google Scholar] [CrossRef]
  46. 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]
  47. Bonura, G.; Cordaro, M.; Cannilla, C.; Arena, F.; Frusteri, F. The changing nature of the active site of Cu-Zn-Zr catalysts for the CO2 hydrogenation reaction to methanol. Appl. Catal. B Environ. 2014, 152–153, 152–161. [Google Scholar] [CrossRef]
  48. Gaab, M.; Trukhan, N.; Maurer, S.; Gummaraju, R.; Müller, U. The progression of Al-based metal-organic frameworks—From academic research to industrial production and applications. Microporous Mesoporous Mater. 2012, 157, 131–136. [Google Scholar] [CrossRef]
  49. 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]
Figure 1. PXRD patterns of (a) AlFum MOF, 7Cu/3ZnO/AlFum MOF, 15Cu/6.4ZnO/AlFum MOF; (b) 7Cu/3ZnO/AlFum MOF, 15Cu/6.4ZnO/AlFum MOF at 30° < 2θ < 40°; and (ce) FTIR spectra of AlFum MOF, 7Cu/3ZnO/AlFum MOF, and 15Cu/6.4ZnO/AlFum MOF. CuO (♦).
Figure 1. PXRD patterns of (a) AlFum MOF, 7Cu/3ZnO/AlFum MOF, 15Cu/6.4ZnO/AlFum MOF; (b) 7Cu/3ZnO/AlFum MOF, 15Cu/6.4ZnO/AlFum MOF at 30° < 2θ < 40°; and (ce) FTIR spectra of AlFum MOF, 7Cu/3ZnO/AlFum MOF, and 15Cu/6.4ZnO/AlFum MOF. CuO (♦).
Catalysts 12 01104 g001
Figure 2. N2 sorption isotherms at 77 K of: (a) AlFum MOF, 7Cu/3ZnO/AlFum MOF, 15Cu/6.4ZnO/AlFum MOF; and (b) NLDFT pore size distributions of AlFum MOF, 7Cu/6.4ZnO/AlFum MOF, and 15Cu/6.4ZnO/AlFum MOF.
Figure 2. N2 sorption isotherms at 77 K of: (a) AlFum MOF, 7Cu/3ZnO/AlFum MOF, 15Cu/6.4ZnO/AlFum MOF; and (b) NLDFT pore size distributions of AlFum MOF, 7Cu/6.4ZnO/AlFum MOF, and 15Cu/6.4ZnO/AlFum MOF.
Catalysts 12 01104 g002
Figure 3. SEM and TEM images of: (a,b) AlFum MOF; (c,d) 7Cu/3ZnO/AlFum MOF; (e,f) 15Cu/6.4ZnO/AlFum MOF; EDX elemental maps of (g) 7Cu/3ZnO/AlFum MOF; and (h) 15Cu/6.4ZnO/AlFum MOF.
Figure 3. SEM and TEM images of: (a,b) AlFum MOF; (c,d) 7Cu/3ZnO/AlFum MOF; (e,f) 15Cu/6.4ZnO/AlFum MOF; EDX elemental maps of (g) 7Cu/3ZnO/AlFum MOF; and (h) 15Cu/6.4ZnO/AlFum MOF.
Catalysts 12 01104 g003
Figure 4. (a) TGA profiles of AlFum MOF, 7Cu/3ZnO/AlFum MOF, 15Cu/ZnO/6.4AlFum MOF; (b) H2-TPR profiles of: AlFum MOF, 7Cu/3ZnO/AlFum MOF, 15Cu/6.4ZnO/AlFum MOF.
Figure 4. (a) TGA profiles of AlFum MOF, 7Cu/3ZnO/AlFum MOF, 15Cu/ZnO/6.4AlFum MOF; (b) H2-TPR profiles of: AlFum MOF, 7Cu/3ZnO/AlFum MOF, 15Cu/6.4ZnO/AlFum MOF.
Catalysts 12 01104 g004
Figure 5. XPS spectra (a) Cu 2p and (b) Zn 2p of 7Cu/3ZnO/AlFum MOF and 15Cu/6.4ZnO/AlFum MOF catalysts.
Figure 5. XPS spectra (a) Cu 2p and (b) Zn 2p of 7Cu/3ZnO/AlFum MOF and 15Cu/6.4ZnO/AlFum MOF catalysts.
Catalysts 12 01104 g005
Figure 6. CO2 conversions of the evaluated catalysts. Temperature = 230 °C; pressure = 50 bar; Qv,0 = 40 mL min−1; gas hourly space velocity = 10,000 h−1; H2/CO2 = 3:1.
Figure 6. CO2 conversions of the evaluated catalysts. Temperature = 230 °C; pressure = 50 bar; Qv,0 = 40 mL min−1; gas hourly space velocity = 10,000 h−1; H2/CO2 = 3:1.
Catalysts 12 01104 g006
Figure 7. (a) MeOH, CO, and CH4 selectivity and (b) MeOH space-time yield of all the tested catalysts.
Figure 7. (a) MeOH, CO, and CH4 selectivity and (b) MeOH space-time yield of all the tested catalysts.
Catalysts 12 01104 g007
Table 1. SBET, pore volume, mean pore size diameter, and elemental composition of pristine AlFum MOF and Cu/ZnO/AlFum MOF catalysts.
Table 1. SBET, pore volume, mean pore size diameter, and elemental composition of pristine AlFum MOF and Cu/ZnO/AlFum MOF catalysts.
SampleSBET (m2/g)Pore Volume (cm3/g)d (nm)Elemental Loading (Weight %)
CuZn
AlFum MOF9100.3441.27--
7Cu/3ZnO/AlFum MOF7570.3001.2311.2 (7)3.12 (3)
15.4Cu/6.4ZnO/AlFum MOF4160.1481.3018.2 (15)5.77 (6.4)
Table 2. Product distribution for all the evaluated catalysts.
Table 2. Product distribution for all the evaluated catalysts.
Product Distribution (mol%)
CatalystMeOHCOCH4H2O
Commercial12.836.02.5048.7
7Cu/3ZnO/AlFum MOF6.5042.80.949.8
15Cu/6.4ZnO/AlFum MOF3.4844.62.6249.3
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.; Moma, J.; Langmi, H.W.; Louis, B.; Parkhomenko, K.; Musyoka, N.M. Towards High CO2 Conversions Using Cu/Zn Catalysts Supported on Aluminum Fumarate Metal-Organic Framework for Methanol Synthesis. Catalysts 2022, 12, 1104. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12101104

AMA Style

Duma ZG, Moma J, Langmi HW, Louis B, Parkhomenko K, Musyoka NM. Towards High CO2 Conversions Using Cu/Zn Catalysts Supported on Aluminum Fumarate Metal-Organic Framework for Methanol Synthesis. Catalysts. 2022; 12(10):1104. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12101104

Chicago/Turabian Style

Duma, Zama G., John Moma, Henrietta W. Langmi, Benoit Louis, Ksenia Parkhomenko, and Nicholas M. Musyoka. 2022. "Towards High CO2 Conversions Using Cu/Zn Catalysts Supported on Aluminum Fumarate Metal-Organic Framework for Methanol Synthesis" Catalysts 12, no. 10: 1104. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12101104

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