Next Article in Journal / Special Issue
Synthesis and Structure of COE-11, a New Borosilicate Zeolite with a Two-Dimensional Pore System of 12-Ring Channels
Previous Article in Journal
Molecular Modeling and Potential Ca2+ Channel Blocker Activity of Diphenylmethoxypiperidine Derivatives
Previous Article in Special Issue
Cesium Heteropolyacid Salts: Synthesis, Characterization and Activity of the Solid and Versatile Heterogeneous Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

UiO-66 MOF-Derived Ru@ZrO2 Catalysts for Photo-Thermal CO2 Hydrogenation

1
Institute for Advanced Materials and Mathematics (InaMat2), Universidad Pública de Navarra (UPNA), Campus de Arrosadia, 31006 Pamplona-Iruña, Spain
2
Universidad de Zaragoza (UNIZAR), Campus San Francisco, 50014 Zaragoza, Spain
*
Authors to whom correspondence should be addressed.
Chemistry 2023, 5(2), 720-729; https://doi.org/10.3390/chemistry5020051
Submission received: 2 March 2023 / Revised: 20 March 2023 / Accepted: 23 March 2023 / Published: 25 March 2023

Abstract

:
The use of metal–organic frameworks (MOFs) as templates or precursors in the manufacture of heterogeneous catalysts is highly attractive due to the transfer of MOFs’ inherent porosity and homogeneous metallic distribution to the derived structure. Herein, we report on the preparation of MOF-derived Ru@ZrO2 catalysts by controlled thermal treatment of zirconium-based MOF UiO-66 with ruthenium moieties. Ru3+ (3 or 10 mol%) precursor was added to UiO-66 synthesis and, subsequently, the as-synthesized hybrid structure was calcined in flowing air at different temperatures (400–600 °C) to obtain ZrO2-derived oxides doped with highly dispersed Ru metallic clusters. The materials were tested for the catalytic photo-thermal conversion of CO2 to CH4. Methanation experiments were conducted in a continuous flow (feed flow rate of 5 sccm and 1:4 CO2 to H2 molar ratio) reactor at temperatures from 80 to 300 °C. Ru0.10@ZrO2 catalyst calcined at 600 °C was able to hydrogenate CO2 to CH4 with production rates up to 65 mmolCH4·gcat.–1·h–1, CH4 yield of 80% and nearly 100% selectivity at 300 °C. The effect of the illumination was investigated with this catalyst using a high-power visible LED. A CO2 conversion enhancement from 18% to 38% was measured when 24 sun of visible LED radiation was applied, mainly due to the increase in the temperature as a result of the efficient absorption of the radiation received. MOF-derived Ru@ZrO2 catalysts have resulted to be noticeably active materials for the photo-thermal hydrogenation of CO2 for the purpose of the production of carbon-neutral methane. A remarkable effect of the ZrO2 crystalline phase on the CH4 selectivity has been found, with monoclinic zirconia being much more selective to CH4 than its cubic allotrope.

Graphical Abstract

1. Introduction

Anthropogenic emission of greenhouse gases is causing an extraordinary increment of atmospheric CO2 levels, leading to the rise of global temperature in a very short period of time [1]. During the last two decades, enormous efforts have been carried out for decarbonizing transportation and industrial processes in order to mitigate carbon dioxide emissions, so the development of technologies able of recycling CO2 into value-added products is highly desired [2,3,4,5]. Carbon-neutral fuels obtained from CO2 directly captured from air or other sources, such as biomass or organic wastes, is considered a very promising way of mitigating overall carbon emissions, while at the same time, taking advantage of the use of existing infrastructures and technologies [6]. For this purpose, one of the most interesting processes for recycling anthropogenic CO2 is the Sabatier reaction, first reported in 1902 [7,8], which consists of the exothermic hydrogenation of carbon dioxide to produce methane and water.
Photo-thermal catalysis is the synergistic combination of photo- and thermochemical processes that occur under broad solar spectrum illumination of the catalyst, resulting in the promotion of the chemical reaction [9,10,11,12]. The photo-thermal effect holds promising features for the valorization of CO2 into solar fuels using green hydrogen as a chemical vector and solar light as the sole energy input [13,14]. The photo-thermal CO2 methanation can be conducted under light using noble metals nanoparticles (NPs) based mainly on Pd, Rh and Ru as active catalysts, though the more abundant and less costly transition metals such Ni or Co [11,15,16,17,18] can be also used. The good performance of the catalytic system depends not only on the properties of the metallic nanoparticle, but also on the physical and chemical characteristics of the support (large light absorption band, high surface area, thermal conductivity, etc.) that plays key roles in the CO2 adsorption and activation processes. The distribution and size of the metallic nanoparticles on the support and the interface between them are also crucial for proper activity and selectivity [19,20,21,22]. Looking for these convenient properties, in recent years, the use of metal–organic frameworks (MOFs) as templates or precursors for the manufacture of heterogeneous catalysts has gained attention due to the MOFs’ inherent porosity and homogeneous metallic distribution, which is also transferred to the derived structure [23,24,25]. MOFs-derived materials can be obtained upon thermal treatment under inert gas atmosphere to produce carbonaceous solids [26], or under oxidizing atmospheres to obtain metallic oxides [27,28,29]. In both cases, the selection of the conditions, i.e., temperature and time of the thermal treatment, is crucial to obtain active materials. It is necessary to guarantee the complete degradation/removal of the parent organic ligands while avoiding large particle growth to minimize aggregation and loss of surface area.
Herein, we present the synthesis of a CO2 methanation catalyst from a well known Zr(IV)-based MOF UiO-66 (Zr-BDC (Benzene-1,4-dicarboxylic acid)) [30] through the addition of controlled amounts of Ru3+. It is possible in this way to keep the MOF porous structure while containing the active metal fully dispersed within the framework [31,32]. The incorporation of addenda atoms to a stable MOF structure provokes defects in the crystalline network, resulting in unstable structures [33]. Subsequent controlled calcination yields zirconium oxide doped with highly dispersed Ru clusters. To achieve this non-trivial target (high metal dispersion and controlled support properties), the use of MOFs as catalyst precursors constitutes a very appealing approach.

2. Results and Discussion

2.1. Synthesis and Characterization of Ru-Loaded UiO-66

Ru-loaded UiO-66 (RuxUiO-66, where x is the nominal Ru-to-Zr molar ratio) was prepared following the procedure described in the literature [31], substituting part of the bulk Zr by Ru in the MOF synthesis gel. The materials will be referred to as Ru0.03UiO-66 (0.3 Ru to 9.7 Zr mol ratio) and Ru0.10UiO-66 (1 Ru to 9 Zr mol ratio). After the hydrothermal synthesis at 100 °C, a nanoparticulated (mean size of 590 ± 20 nm) solid was obtained, as revealed by scanning electron microscopy (SEM) (Figure 1a). Powder X-ray diffraction (XRD) (Figure 1b) patterns of the as-synthesized MOFs revealed that the addition of Ru does not affect the structure of the parent UiO-66, as evidenced by the coincidence with the simulated characteristics diffraction peaks. In addition, no diffraction peaks corresponding to Ru species can be observed, though the low Ru loading can explain their absence. On the other hand, the Fourier-transform infrared (FTIR) spectra of the RuxUiO-66 nanocomposites (Figure S1) showed peaks for the Zr−O mode and Zr−O−C symmetric stretching, proving that Zr is the coordination center in the organic framework. Ru vibrational peaks were not detected by FTIR, suggesting that Ru is not coordinated to the UiO-66 skeleton [32,34]. Thermo-gravimetric analyses (TGA) (Figure 1c and Figure S4) were conducted in air to evaluate the thermal stability of RuxUiO-66 samples. The weight loss at temperatures below 100 °C is attributed to water desorption, and the loss at around 220 °C is attributed to the presence of residual solvent. Meanwhile, decomposition of the framework organic linker (corresponding to a total weight loss of around 65%) took place at 420 °C for Ru0.10UiO-66, at 510 °C for Ru0.03UiO-66 and 520 °C for UiO-66. This result can be attributed to a weakening of the interation between the organic linker and Zr4+ due to the presence of Ru moieties, thus facilitating the decompositon of the hybrid material [32], though a catalytic effect of Ru on the linkers degradation cannot be ruled out. N2 adsorption-desorption analyses provided very high specific surface areas of 1220 m2·g−1 and 1247 m2·g−1 for Ru0.10UiO-66 and Ru0.03UiO-66, respectively, which are only slightly lower than that of pristine UiO-66 (1380 m2·g−1). This reduction of superficial area for the hybrid materials suggests the presence of Ru clusters inside the porous network without seriously affecting the textural features of the pure UiO-66.

2.2. Preparation and Characterization of MOF-Derived Ru@ZrO2 Catalysts

500 mg of as-synthesized RuxUiO-66 materials were treated under air flow at three different calcination temperatures (400, 500 and 600 °C) to obtain the catalysts. Calcined MOF-derived composites (denoted as Ru@ZrO2-X, where X indicates the calcination temperature, 400 °C, 500 °C, or 600 °C, were characterized by TEM, XRD, TGA, and N2 adsorption–desorption. As expected, after calcination, the samples presented a large loss of weight and bulk density due to the elimination of the organic ligand. Their color turned dark; the darker they were, the more Ru in the framework, and they also possessed improved light absorption in the visible range, as evidenced by UV-Vis reflectance spectroscopy (Figure 2c) [35]. Ru content of the materials was determined by inductively coupled plasma atomic emission spectrometry (ICP-AES). The Ru incorporation to the MOF during the synthesis was much lower than expected (0.37 wt.% for the as-synthesized Ru10UiO-66 vs. the nominal value of 5.84 wt.%), resulting in catalysts with very low Ru loads. The ruthenium incorporation could be favored by increasing the synthesis temperature. Results also showed that the calcination treatment caused a slight change in the ruthenium content from 0.34 wt.% in Ru0.10@ZrO2-400 °C to 0.26 wt.% and 0.24 wt.% for samples Ru0.10@ZrO2-500 °C and Ru0.10@ZrO2-600 °C, respectively. In the case of the lowest Ru to Zr molar ratio, the Ru content diminished to 0.04 wt.% for Ru0.03@ZrO2-600 °C.
As it can be seen in the TEM images depicted in Figure 2a and Figure S2, polycrystalline nanoparticulated solids were obtained. The crystallinity of the particles (encircled in the images) increased with the calcination temperature. No Ru nanoclusters could be observed in the micrographs, probably due to the low incorporation of Ru and its high dispersion, as confirmed by EDS (Energy Dispersive Spectroscopy) analysis (Figure S2). X-ray patterns in Figure 2b and Figure S3 reveal the presence of monoclinic ZrO2 in the Ru@ZrO2-600 °C samples. Meanwhile, Ru@ZrO2-400 °C and Ru@ZrO2-500 °C contain cubic ZrO2 [36,37,38]. Lippi et al. [31] demonstrated, by operating XRD, the importance of the monoclinic ZrO2 phase transitions during thermal treatment to stabilize Ru and obtain active CO2 methanation catalysts. In our case, Ru species could not be detected by XRD, mainly due to the low Ru content of the bulk materials. Although all treated samples exhibited crystallinity, materials calcined at 400 and 500 °C still contained low amounts of organic residue or non-degraded MOF, as observed by TGA (Figure S4) and FTIR (Figure S1), confirming an incomplete removal of the organic linker.

2.3. Catalytic Tests

Firstly, all the UiO-66 MOF-derived catalysts were tested for the CO2 methanation in a temperature-controlled Harrick HVC-MRA-5 reaction chamber. An amount of 30 mg of the catalysts were deposited on the chamber mesh placing a thermocouple directly in contact with the catalyst sample. Prior to the methanation tests, an in situ reduction pre-treatment with pure H2 at 150 °C for 3 h was carried out to assure the presence of Ru in the metallic state. A stoichiometric mixture of H2 and CO2, according to a 4:1 molar ratio, was fed to the reactor at a space velocity of 10,000 N mL gcat.−1 h−1, after which the temperature was increased at a controlled rate of 3 °C·min−1 from 80 to 300 °C (see Figure 3a). As pointed out before, the temperature of the thermal treatment was crucial to obtain active materials and achieve good CH4 yields.
The CH4 yields and selectivities (see Figure 3b) achieved by the different catalysts demonstrated that the materials treated at 600 °C present excellent methanation performance, meanwhile those treated at lower temperatures exhibit clearly lower activity and higher selectivity to CO. These results can be attributed to the presence of monoclinic ZrO2 in the catalyst obtained after calcination at 600°C of the RuxUiO-66 precursors, while the samples calcined at lower temperatures contain cubic ZrO2. It is remarkable that, whereas the Ru0.03@ZrO2-600 °C shows only improved CH4 selectivity, Ru0.10@ZrO2-600 °C exhibits both improved activity and selectivity, thus suggesting a cooperative effect of both Ru and monoclinic ZrO2. Indeed, Ru0.10@ZrO2-600 °C catalyst was able to hydrogenate CO2 to CH4 with a production rate at 300 °C as high as 65 mmolCH4·gcat.−1·h−1, CH4 yield of about 80% and nearly 100% selectivity. The CH4 production rate achieved with Ru0.03@ZrO2-600 °C decreased to 20 mmolCH4·gcat.−1·h−1 due to the low amount of Ru, and the selectivity to methane was 90%. In spite of their comparatively low activity, both Ru0.03@ZrO2 and Ru0.10@ZrO2 calcined at 400 °C and 500 °C exhibited high selectivities to CO, similar (low activity and high selectivity to CO) to those achieved with bare ZrO2 supports for calcined UiO-66 without Ru (Figure S5).
The higher activity and selectivity of the monoclinic ZrO2 materials obtained after calcination at 600 °C in comparison with their cubic ZrO2 counterparts reveals a key role of the allotropic form of zirconia on the CO2 methanation selectivity, indicating that the support is involved in the reaction pathway. Different studies reported in the literature also highlight the importance of the crystalline phase [39], the presence of dopants [40], and the nature of the metallic phase [41] on the CO2 methanation over Ru/ZrO2 systems. Ilsemann et al. [42] performed in operando DRIFTS studies with a Ru/ZrO2 catalysts and found that the formate pathway is predominant at temperatures below 300 °C, at which this species would be formed after C–O bond cleavage of a bidentate carbonate adsorbed on the ZrO2 surface. These authors also suggested that the formation of formate species is related with the presence of Lewis basic oxygen vacancies that could be a possible reason for the higher activity of monoclinic zirconia.
The most active material, Ru0.10@ZrO2-600 °C, was selected to carry out the photo-thermal CO2 hydrogenation (Figure 4). The temperature was increased with the internal heater up to 230 °C, and then, the reaction chamber was illuminated with an ultra-high power visible-range LED. Under an irradiance of 8 Sun, the temperature raised to 250 °C, provoking an increment of the production of CH4 from 17 mmolCH4·gcat.−1·h−1 in dark conditions to 23 mmolCH4·gcat.−1·h−1, which corresponded to an increase in the CO2 conversion from about 15% to almost 25%. Then, the lamp was switched off for 70 min, and the CH4 production and temperature values came back to those of the dark conditions. The procedure was repeated at higher irradiance (24 Sun) that led to a temperature rise up to 310 °C, as well as an increase in the CO2 conversion up to about 38%. CH4 production jumped to 32 mmolCH4·gcat.−1·h−1. Methane selectivity during the whole process was above 99% under dark or illuminated conditions. The conversion enhancement is mainly attributed to the increase in the catalyst temperature, resulting from the absorption of the radiation received. More specific conversion vs. light intensity experiments should be carried out to separately determine both thermal and photocatalytic effects of light as pointed out by Mateo et al. [43,44]. In addition, the temperature increase resulting from light absorption occurs only in the exposed part of the catalytic bed, so a relatively large temperature gradient can be established [45], causing the actual mean temperature, and hence the CO2 conversion, to be lower than that expected from the whole bed operating isothermally at the temperature read by the thermocouple. Specifically designed catalytic chambers have to be developed for better control of the reaction temperature.

3. Materials and Methods

3.1. Materials

Terephthalic acid (98%), anhydrous zirconium (IV) chloride (ZrCl4), ruthenium (III) chloride hydrate (RuCl3·xH2O), and hydrochloric acid (HCl, 37%) were supplied by Sigma Aldrich. N,N-dimethylformamide (DMF), ethanol (96%), and methanol were purchased from Scharlab. Calibrated reaction gas mixture 17.9 vol. % CO2, 71.9 vol. % H2 and 10.1 vol. % N2 was provided by Nippon Gases. All reactants were used without further purification.

3.2. MOF Synthesis

Ru-loaded MOF RuxUiO-66 (x stands for the nominal Ru/Zr molar ratio) were synthesized following well established procedures for its pristine counterpart UiO-66 (ZrBDC) with a metal to organic ligand molar ratio of 1:2 [30]. Ru was introduced by replacing 3% or 10% of the initial Zr precursor by the Ru salt. A typical synthesis procedure was as follows: for the synthesis of Ru0.10UiO-66, ZrCl4 (9 mmol) and RuCl3·xH2O (1 mmol) were added to a mixture of DMF (60 mL) and 37% HCl (2.4 mL) [31]. Afterwards, H2BDC (20 mmol) was added to the solution and stirred at room temperature until complete dissolution. The mixture was heated at autogenous pressure at 100 °C for 20 h. The product was recovered by centrifugation and washed with distilled water, ethanol, and methanol. Finally, the Ru-loaded MOF was dried at 70 °C overnight.

3.3. Thermal Treatment

The MOF-derived catalysts were obtained by controlled calcination of the Ru-loaded MOF precursors in a flow-through tubular oven (Hobersal) for 3 h under 20 N mL·min−1 air flow (Nippon Gases, 99.9%). The influence of the temperature of calcination on the catalytic properties of the final product was studied by calcination at 400, 500, and 600 °C. The heating ramp was kept, in all cases, at 10 °C·min−1.

3.4. Characterization Techniques

Scanning electron microscopy (SEM) analyses (CSEM-FEG Inspect) were performed on Ru-loaded MOF precursors to assess crystal morphology and chemical composition. For the MOF-derived oxides, transmission electron microscopy with energy-dispersive X-ray spectroscopy (TEM-EDX) analyses (FEI TECNAI F30) were performed for the same purposes. Purity and crystallinity of both the MOF-precursors and final oxides were evaluated by powder X-ray diffraction (XRD, RIGAKU Ru2500). The XRD patterns were compared to simulated patterns calculated using the software VESTA. Thermal stability of the synthesized compounds was evaluated by differential thermogravimetric analysis (DTGA) over 6 mg of sample with a heating rate of 5 °C·min−1 from room temperature to 700 °C under air flow (TGA/DSC 3+ Mettler Toledo). Textural properties were assessed by N2 physisorption at 77 K (Micromeritics GEMINI V 2380), and surface area was evaluated applying the Langmuir model. Degree of oxidation after calcination of the MOF precursor was further assessed by Fourier-transform infrared spectroscopy (FTIR) with a high throughput monolithic diamond attenuated total reflectance (ATR) (Jasco FT/IR-4700 Spectrometer), accumulation of 50 scans, resolution of 1 cm−1. Light absorption properties were analyzed with a flame spectrometer (Ocean Optics) in the 200–800 nm range with a UV-visible reflection probe. Finally, the Ru content of the catalysts precursors was analyzed by inductively coupled plasma atomic emission spectroscopy (ICP-AES) (Varian Vista MPX Radial). The digestion of the ruthenium containing materials was carried out following the fusion method reported by T. Suoranta et al. [46].

3.5. Catalytic Tests

The CO2 hydrogenation tests were carried out in temperature-controlled Harrick HVC-MRA-5 reaction chamber. The reactor temperature was measured by means of a type K thermocouple placed on the catalyst surface [45] Reaction gases were analyzed with an online gas chromatograph (Agilent 990 Micro GC System provided with a molecular sieve 5A column for permanent gases and PPU column for CO2 determination). Stoichiometric mixture of H2 and CO2 according to a 4:1 molar ratio (17.9 vol. % CO2, 71.9 vol. % H2 and 10.1 vol. % of N2 as internal standard) was fed to the reactor at a space velocity of 10,000 mL g−1 h−1. Prior to any experiments, the Ru-based catalysts (30 mg) were subjected to an in situ reduction pre-treatment with pure H2 for 2 h at 150 °C. Then, for the direct experiments, the reaction mixture was introduced into the reactor, and the temperature was increased with a rate of 3 °C·min−1 from 80 to 300 °C, and it was kept constant for 50 min at 300 °C.
To evaluate the photo-thermal activity, the most active catalysts were placed into the reactor, and the temperature was increased up to 230 °C and kept at this temperature for 60 min. Subsequently, the catalyst was illuminated with an ultra-high power (30 W) visible-range LED (Figure S6) (LEDM-5500 White HP—Pyroistech S.L.) placed 1 cm above the catalyst surface. Position of the LED-lamp and its luminous flux can be controlled to regulate the irradiance received by de sample. Maximum irradiance received corresponds to 2.4 W·cm−2 (1 Sun = 0.1 W·cm−2). The temperature on the surface of the catalyst was continuously registered during the process.
The CO2 conversion (X, %), CH4 selectivity (SCH4, %), and yield (YCH4, %) were calculated as follows:
X C O 2   = F C O 2   ( i n ) F C O 2   ( o u t ) F C O 2   ( i n ) × 100 %
S C H 4   = F C H 4   ( o u t ) F C H 4   ( o u t ) + F C O   ( o u t ) × 100 %
Y C H 4   = X C O 2   · S C H 4   × 100 %
where FCO2 (in) and FCO2 (out) are the molar flows of CO2 at the reactor inlet and outlet, respectively. Similarly, FCH4(out) and FCO(out) are the molar flows of CH4 and CO at the reactor outlet.

4. Conclusions

The use of MOFs as catalyst precursors has resulted in an interesting approach to achieve high metallic dispersion all over the catalyst surface, as well as good control of the support properties. Small quantities of Ru were incorporated during the hydrothermal synthesis of the Zr-based UiO-66 MOF, and subsequently, MOF-derived Ru@ZrO2 catalysts were obtained via controlled calcination. Calcination at 600 °C was crucial to tune the crystalline phase (monoclinic) of ZrO2 that provided the most outstanding methanation results (in terms of conversion and selectivity). The most active material, Ru0.10@ZrO2-600 °C, with a final amount of Ru around 0.24 wt. %, was tested under high power visible LED illumination. A CO2 conversion enhancement from 18% to 38% was observed under 24 Sun irradiance as a result of the increase in the catalyst temperature due to the absorption of the radiation received.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/chemistry5020051/s1, Figure S1: FTIR spectra, Figure S2: TEM images and EDX analysis, Figure S3: Simulated XRD for ZrO2, Figure S4: TGA curves of RuxUiO-66 and derived-Rux@ZrO2 catalysts, Figure S5: CO2 methanation performance of MOF-derived ZrO2 without Ru content, Figure S6: Spectral power distribution of ultra-high power LEDM-5500 White HP—Pyroistech S.L.

Author Contributions

Conceptualization, I.P. and L.M.G.; methodology, M.L., F.A., M.I., and I.P.; investigation and data curation, M.L., F.A., A.E., and M.I.; writing and editing, M.L., F.A., I.P., and L.M.G.; funding acquisition, I.P. and L.M.G. All authors have read and agreed to the published version of the manuscript.

Funding

Financial support was obtained from Spanish Agencia Estatal de Investigación and Spanish Ministerio de Ciencia e Innovación MCIN/AEI/10.13039/501100011033/ and FEDER “Una manera de hacer Europa” (grants PID2019-106687RJ-I00/AEI/ 10.13039/501100011033 and PID2021-127265OB-C21), as well as from Plan de Recuperación, Transformación y Resiliencia and NextGenerationEU (grants PLEC2022-009221 and TED2021-130846B-100), which is gratefully acknowledged. M.L. and M.I. thank Spanish Ministerio de Universidades and Unión Europea-NextGenerationEU for the postdoctoral ¨Margarita Salas¨ and FPU 18/01877 grants, respectively. A.E. thanks Universidad Pública de Navarra for the predoctoral fellowship. L.M.G. thanks the Banco de Santander and Universidad Pública de Navarra for their financial support under “Programa de Intensificación de la Investigación 2018” initiative.

Data Availability Statement

The data presented in this study are available on request from the corresponding authors.

Acknowledgments

Authors acknowledge the use of XRD analysis service of Servicio General de Apoyo a la Investigación-SAI, Universidad de Zaragoza and the instrumentation as well as the technical advice provided by the National Facility ELECMI ICTS, node Laboratorio de Microscopias Avanzadas (LMA) at Universidad de Zaragoza.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Intergovernmental Panel on Climate Change, IPCC. Climate Change 2014: Mitigation of Climate Change; Cambridge University Press: Cambridge, UK, 2015; ISBN 9781107058217. [Google Scholar]
  2. Álvarez, A.; Bansode, A.; Urakawa, A.; Bavykina, A.V.; Wezendonk, T.A.; Makkee, M.; Gascon, J.; Kapteijn, F. Challenges in the Greener Production of Formates/Formic Acid, Methanol, and DME by Heterogeneously Catalyzed CO2 Hydrogenation Processes. Chem. Rev. 2017, 117, 9804–9838. [Google Scholar] [CrossRef] [PubMed]
  3. Goud, D.; Gupta, R.; Maligal-Ganesh, R.; Peter, S.C. Review of Catalyst Design and Mechanistic Studies for the Production of Olefins from Anthropogenic CO2. ACS Catal. 2020, 10, 14258–14282. [Google Scholar] [CrossRef]
  4. Ye, R.P.; Ding, J.; Gong, W.; Argyle, M.D.; Zhong, Q.; Wang, Y.; Russell, C.K.; Xu, Z.; Russell, A.G.; Li, Q.; et al. CO2 Hydrogenation to High-Value Products via Heterogeneous Catalysis. Nat. Commun. 2019, 10, 5698. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Corma, A.; Garcia, H. Photocatalytic Reduction of CO2 for Fuel Production: Possibilities and Challenges. J. Catal. 2013, 308, 168–175. [Google Scholar] [CrossRef]
  6. Ramirez, A.; Sarathy, S.M.; Gascon, J. CO2 Derived E-Fuels: Research Trends, Misconceptions, and Future Directions. Trends Chem. 2020, 2, 785–795. [Google Scholar] [CrossRef]
  7. Senderens, J.B.; Sabatier, P. Nouvelles Synthèses Du Méthane. Comptes. Rendus. Acad. Sci. 1902, 82, 514–516. [Google Scholar]
  8. Vogt, C.; Monai, M.; Kramer, G.J.; Weckhuysen, B.M. The Renaissance of the Sabatier Reaction and Its Applications on Earth and in Space. Nat. Catal. 2019, 2, 188–197. [Google Scholar] [CrossRef]
  9. Zhang, F.; Li, Y.H.; Qi, M.Y.; Yamada, Y.M.A.; Anpo, M.; Tang, Z.R.; Xu, Y.J. Photothermal Catalytic CO2 Reduction over Nanomaterials. Chem. Catal. 2021, 1, 272–297. [Google Scholar] [CrossRef]
  10. Song, C.; Wang, Z.; Yin, Z.; Xiao, D.; Ma, D. Principles and Applications of Photothermal Catalysis. Chem. Catal. 2022, 2, 52–83. [Google Scholar] [CrossRef]
  11. Meng, X.; Wang, T.; Liu, L.; Ouyang, S.; Li, P.; Hu, H.; Kako, T.; Iwai, H.; Tanaka, A.; Ye, J. Photothermal Conversion of CO2 into CH4 with H2 over Group VIII Nanocatalysts: An Alternative Approach for Solar Fuel Production. Angew. Chem. 2014, 126, 11662–11666. [Google Scholar] [CrossRef]
  12. Ghoussoub, M.; Xia, M.; Duchesne, P.N.; Segal, D.; Ozin, G. Principles of Photothermal Gas-Phase Heterogeneous CO2 Catalysis. Energy Environ. Sci. 2019, 12, 1122–1142. [Google Scholar] [CrossRef]
  13. Fang, S.; Hu, Y.H. Thermo-Photo Catalysis: A Whole Greater than the Sum of Its Parts. Chem. Soc. Rev. 2022, 51, 3609–3647. [Google Scholar] [CrossRef] [PubMed]
  14. Lv, C.; Bai, X.; Ning, S.; Song, C.; Guan, Q.; Liu, B.; Li, Y.; Ye, J. Nanostructured Materials for Photothermal Carbon Dioxide Hydrogenation: Regulating Solar Utilization and Catalytic Performance. ACS Nano 2022, 17, 1725–1738. [Google Scholar] [CrossRef] [PubMed]
  15. Li, Z.; Shi, R.; Zhao, J.; Zhang, T. Ni-Based Catalysts Derived from Layered-Double-Hydroxide Nanosheets for Efficient Photothermal CO2 Reduction under Flow-Type System. Nano Res. 2021, 14, 4828–4832. [Google Scholar] [CrossRef]
  16. Li, X.; Yu, J.; Jaroniec, M.; Chen, X. Cocatalysts for Selective Photoreduction of CO 2 into Solar Fuels. Chem. Rev. 2019, 119, 3962–4179. [Google Scholar] [CrossRef]
  17. Li, Y.; Liu, Z.; Rao, Z.; Yu, F.; Bao, W.; Tang, Y.; Zhao, H.; Zhang, J.; Wang, Z.; Li, J.; et al. Experimental and Theoretical Insights into an Enhanced CO2 Methanation Mechanism over a Ru-Based Catalyst. Appl. Catal. B Environ. 2022, 319, 121903. [Google Scholar] [CrossRef]
  18. Ullah, S.; Lovell, E.C.; Wong, R.J.; Tan, T.H.; Scott, J.; Amal, R. Light-Enhanced CO2 Reduction to CH4 Using Nonprecious Transition-Metal Catalysts. ACS Sustain. Chem. Eng. 2020, 8, 5056–5066. [Google Scholar] [CrossRef]
  19. Ulmer, U.; Dingle, T.; Duchesne, P.N.; Morris, R.H.; Tavasoli, A.; Wood, T.; Ozin, G.A. Fundamentals and Applications of Photocatalytic CO2 Methanation. Nat. Commun. 2019, 10, 3169. [Google Scholar] [CrossRef] [Green Version]
  20. Kho, E.T.; Jantarang, S.; Zheng, Z.; Scott, J.; Amal, R. Harnessing the Beneficial Attributes of Ceria and Titania in a Mixed-Oxide Support for Nickel-Catalyzed Photothermal CO2 Methanation. Engineering 2017, 3, 393–401. [Google Scholar] [CrossRef]
  21. Albero, J.; Garcia, H.; Corma, A. Temperature Dependence of Solar Light Assisted CO2 Reduction on Ni Based Photocatalyst. Top. Catal. 2016, 59, 787–791. [Google Scholar] [CrossRef] [Green Version]
  22. González-Rangulan, V.V.; Reyero, I.; Bimbela, F.; Romero-Sarria, F.; Daturi, M.; Gandía, L.M. CO2 Methanation over Nickel Catalysts: Support Effects Investigated through Specific Activity and Operando IR Spectroscopy Measurements. Catalysts 2023, 13, 448. [Google Scholar] [CrossRef]
  23. Zhang, Y.; Xu, J.; Zhou, J.; Wang, L. Metal-Organic Framework-Derived Multifunctional Photocatalysts. Chin. J. Catal. 2022, 43, 971–1000. [Google Scholar] [CrossRef]
  24. Oar-Arteta, L.; Wezendonk, T.; Sun, X.; Kapteijn, F.; Gascon, J. Metal Organic Frameworks as Precursors for the Manufacture of Advanced Catalytic Materials. Mater. Chem. Front. 2017, 1, 1709–1745. [Google Scholar] [CrossRef] [Green Version]
  25. Khan, I.S.; Garzon-Tovar, L.; Mateo, D.; Gascon, J. Metal-Organic-Frameworks and Their Derived Materials in Photo-Thermal Catalysis. Eur. J. Inorg. Chem. 2022, 2022, e202200316. [Google Scholar] [CrossRef]
  26. Khan, I.S.; Mateo, D.; Shterk, G.; Shoinkhorova, T.; Poloneeva, D.; Garzón-Tovar, L.; Gascon, J. An Efficient Metal–Organic Framework-Derived Nickel Catalyst for the Light Driven Methanation of CO2. Angew. Chem. Int. Ed. 2021, 60, 26476–26482. [Google Scholar] [CrossRef]
  27. Gong, X.; Wang, W.W.; Fu, X.P.; Wei, S.; Yu, W.Z.; Liu, B.; Jia, C.J.; Zhang, J. Metal-Organic-Framework Derived Controllable Synthesis of Mesoporous Copper-Cerium Oxide Composite Catalysts for the Preferential Oxidation of Carbon Monoxide. Fuel 2018, 229, 217–226. [Google Scholar] [CrossRef]
  28. Zheng, L.; Li, X.; Du, W.; Shi, D.; Ning, W.; Lu, X.; Hou, Z. Metal-Organic Framework Derived Cu/ZnO Catalysts for Continuous Hydrogenolysis of Glycerol. Appl. Catal. B Environ. 2017, 203, 146–153. [Google Scholar] [CrossRef]
  29. Zeng, L.; Wang, Y.; Li, Z.; Song, Y.; Zhang, J.; Wang, J.; He, X.; Wang, C.; Lin, W. Highly Dispersed Ni Catalyst on Metal-Organic Framework-Derived Porous Hydrous Zirconia for CO2 Methanation. ACS Appl. Mater. Interfaces 2020, 12, 17436–17442. [Google Scholar] [CrossRef]
  30. Shearer, G.C.; Chavan, S.; Ethiraj, J.; Vitillo, J.G.; Svelle, S.; Olsbye, U.; Lamberti, C.; Bordiga, S.; Lillerud, K.P. Tuned to Perfection: Ironing out the Defects in Metal-Organic Framework UiO-66. Chem. Mater. 2014, 26, 4068–4071. [Google Scholar] [CrossRef]
  31. Lippi, R.; D’Angelo, A.M.; Li, C.; Howard, S.C.; Madsen, I.C.; Wilson, K.; Lee, A.F.; Sumby, C.J.; Doonan, C.J.; Patel, J.; et al. Unveiling the Structural Transitions during Activation of a CO2 Methanation Catalyst Ru0/ZrO2 Synthesised from a MOF Precursor. Catal. Today 2021, 368, 66–77. [Google Scholar] [CrossRef]
  32. Xu, W.; Dong, M.; Di, L.; Zhang, X. A Facile Method for Preparing UiO-66 Encapsulated Ru Catalyst and Its Application in Plasma-Assisted CO2 Methanation. Nanomaterials 2019, 9, 1432. [Google Scholar] [CrossRef] [Green Version]
  33. Fereja, S.L.; Zhang, Z.; Fang, Z.; Guo, J.; Zhang, X.; Liu, K.; Li, Z.; Chen, W. High-Entropy Oxide Derived from Metal-Organic Framework as a Bifunctional Electrocatalyst for Efficient Urea Oxidation and Oxygen Evolution Reactions. ACS Appl. Mater. Interfaces 2022, 14, 38727–38738. [Google Scholar] [CrossRef] [PubMed]
  34. Pourkhosravani, M.; Dehghanpour, S.; Farzaneh, F. Palladium Nanoparticles Supported on Zirconium Metal Organic Framework as an Efficient Heterogeneous Catalyst for the Suzuki-Miyaura Coupling Reaction. Catal. Lett. 2016, 146, 499–508. [Google Scholar] [CrossRef]
  35. Akilandeswari, S.; Rajesh, G.; Govindarajan, D.; Thirumalai, K.; Swaminathan, M. Efficacy of Photoluminescence and Photocatalytic Properties of Mn Doped ZrO2 Nanoparticles by Facile Precipitation Method. J. Mater. Sci. Mater. Electron. 2018, 29, 18258–18270. [Google Scholar] [CrossRef]
  36. Han, Y.; Zhu, J. Surface Science Studies on the Zirconia-Based Model Catalysts. Top. Catal. 2013, 56, 1525–1541. [Google Scholar] [CrossRef]
  37. Sadati, S.M.; Feghhi, S.A.H.; Mohammadi, K. Effect of Time Exposure on Themoluminiscend Glow Curve For UV-Induced ZrO2:Mg Phosphor. Radiat. Prot. Dosim. 2017, 173, 333–337. [Google Scholar]
  38. Prakashbabu, D.; Hari Krishna, R.; Nagabhushana, B.M.; Nagabhushana, H.; Shivakumara, C.; Chakradar, R.P.S.; Ramalingam, H.B.; Sharma, S.C.; Chandramohan, R. Low Temperature Synthesis of Pure Cubic ZrO2 Nanopowder: Structural and Luminescence Studies. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2014, 122, 216–222. [Google Scholar] [CrossRef]
  39. Nagase, H.; Naito, R.; Tada, S.; Kikuchi, R.; Fujiwara, K.; Nishijima, M.; Honma, T. Ru Nanoparticles Supported on Amorphous ZrO2 for CO2 Methanation. Catal. Sci. Technol. 2020, 10, 4522–4531. [Google Scholar] [CrossRef]
  40. Gao, M.; Zhang, J.; Zhu, P.; Liu, X.; Zheng, Z. Unveiling the Origin of Alkali Metal Promotion in CO2 Methanation over Ru/ZrO2. Appl. Catal. B Environ. 2022, 314, 121476. [Google Scholar] [CrossRef]
  41. Alves, L.M.N.C.; Almeida, M.P.; Ayala, M.; Watson, C.D.; Jacobs, G.; Rabelo-Neto, R.C.; Noronha, F.B.; Mattos, L.V. CO2 Methanation over Metal Catalysts Supported on ZrO2: Effect of the Nature of the Metallic Phase on Catalytic Performance. Chem. Eng. Sci. 2021, 239, 116604. [Google Scholar] [CrossRef]
  42. Ilsemann, J.; Murshed, M.M.; Gesing, T.M.; Kopyscinski, J.; Bäumer, M. On the Support Dependency of the CO2 Methanation—Decoupling Size and Support Effects. Catal. Sci. Technol. 2021, 11, 4098–4114. [Google Scholar] [CrossRef]
  43. Mateo, D.; Cerrillo, J.L.; Durini, S.; Gascon, J. Fundamentals and Applications of Photo-Thermal Catalysis. Chem. Soc. Rev. 2021, 50, 2173–2210. [Google Scholar] [CrossRef]
  44. Mateo, D.; Maity, P.; Shterk, G.; Mohammed, O.F.; Gascon, J. Tunable Selectivity in CO2 Photo-Thermal Reduction by Perovskite-Supported Pd Nanoparticles. ChemSusChem 2021, 14, 5525–5533. [Google Scholar] [CrossRef] [PubMed]
  45. Li, X.; Liu, J.; Everitt, H.O. Untangling Thermal and Nonthermal Effects in Plasmonic Photocatalysis. In Plasmonic Catalysis: From Fundamentals to Applications; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2021; pp. 191–230. ISBN 9783527347506. [Google Scholar]
  46. Suoranta, T.; Niemelä, M.; Perämäki, P. Comparison of Digestion Methods for the Determination of Ruthenium in Catalyst Materials. Talanta 2014, 119, 425–429. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Characterization of RuxUiO-66 samples: (A) SEM images, (B) XRD patterns, and (C) Differential TGA curves.
Figure 1. Characterization of RuxUiO-66 samples: (A) SEM images, (B) XRD patterns, and (C) Differential TGA curves.
Chemistry 05 00051 g001
Figure 2. Characterization of the MOF-derived catalysts obtained via controlled calcination at 400, 500, and 600 °C of RuxUiO-66 materials. (A) TEM images; (B) XRD patterns of the Ru0.10@ZrO2 catalysts; and (C) UV-Vis reflectance measurements.
Figure 2. Characterization of the MOF-derived catalysts obtained via controlled calcination at 400, 500, and 600 °C of RuxUiO-66 materials. (A) TEM images; (B) XRD patterns of the Ru0.10@ZrO2 catalysts; and (C) UV-Vis reflectance measurements.
Chemistry 05 00051 g002
Figure 3. CO2 methanation performance of MOF-derived materials. (A) Evolution of the CO2 conversion with the reaction temperature for the catalysts calcined at 600 °C. (B) Effect of the calcination temperature on the reaction yield and selectivity to CH4 (selected data recorded at 300 °C).
Figure 3. CO2 methanation performance of MOF-derived materials. (A) Evolution of the CO2 conversion with the reaction temperature for the catalysts calcined at 600 °C. (B) Effect of the calcination temperature on the reaction yield and selectivity to CH4 (selected data recorded at 300 °C).
Chemistry 05 00051 g003
Figure 4. Effect of the visible LED irradiation (8 and 24 Sun) on the photo-thermal hydrogenation of CO2 with the catalyst Ru0.10@ZrO2-600 °C.
Figure 4. Effect of the visible LED irradiation (8 and 24 Sun) on the photo-thermal hydrogenation of CO2 with the catalyst Ru0.10@ZrO2-600 °C.
Chemistry 05 00051 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Almazán, F.; Lafuente, M.; Echarte, A.; Imizcoz, M.; Pellejero, I.; Gandía, L.M. UiO-66 MOF-Derived Ru@ZrO2 Catalysts for Photo-Thermal CO2 Hydrogenation. Chemistry 2023, 5, 720-729. https://doi.org/10.3390/chemistry5020051

AMA Style

Almazán F, Lafuente M, Echarte A, Imizcoz M, Pellejero I, Gandía LM. UiO-66 MOF-Derived Ru@ZrO2 Catalysts for Photo-Thermal CO2 Hydrogenation. Chemistry. 2023; 5(2):720-729. https://doi.org/10.3390/chemistry5020051

Chicago/Turabian Style

Almazán, Fernando, Marta Lafuente, Amaya Echarte, Mikel Imizcoz, Ismael Pellejero, and Luis M. Gandía. 2023. "UiO-66 MOF-Derived Ru@ZrO2 Catalysts for Photo-Thermal CO2 Hydrogenation" Chemistry 5, no. 2: 720-729. https://doi.org/10.3390/chemistry5020051

Article Metrics

Back to TopTop