Nickel Catalyst for the Light Driven Methanation of CO2
Item Type Article
Authors Khan, Il Son;Mateo, Diego;Shterk, Genrikh;Shoinkhorova,
Tuiana;Poloneeva, Daria;Garzon Tovar, Luis Carlos;Gascon, Jorge Citation Khan, I. S., Mateo, D., Shterk, G., Shoinkhorova, T., Poloneeva,
D., Garzón-Tovar, L., & Gascon, J. (2021). An Efficient Metal- Organic Framework - Derived Nickel Catalyst for the Light Driven Methanation of CO2. Angewandte Chemie. doi:10.1002/
ange.202111854 Eprint version Post-print
DOI 10.1002/ange.202111854
Publisher Wiley
Journal Angewandte Chemie
Rights Archived with thanks to Angewandte Chemie Download date 2023-12-04 17:02:00
Link to Item http://hdl.handle.net/10754/672880
COMMUNICATION
An Efficient Metal-Organic Framework - Derived Nickel Catalyst for the Light Driven Methanation of CO 2
Il Son Khan,
[a],+Diego Mateo,
[a],+Genrikh Shterk,
[a]Tuiana Shoinkhorova,
[a]Daria Poloneeva,
[a], Luis Garzón-Tovar
[a],* and Jorge Gascon
[a],*
[a] I. Khan+, Dr. D. Mateo+, G. Shterk, T. Shoinkhorova, D. Poloneeva, Dr. L. Garzón-Tovar* and Prof. Dr. J. Gascon
Advanced Catalytic Materials, KAUST Catalysis Center (KCC), King Abdullah University of Science and Technology (KAUST), Thuwal 23955, Saudi Arabia E-mail: [email protected], [email protected]
+Theseauthors contributed equally
Supporting information for this article is given via a link at the end of the document.
Abstract: We report the synthesis of a highly active and stable metal- organic framework derived Ni-based catalyst for the photo-thermal reduction of CO2 to CH4. Through the controlled pyrolysis of MOF-74 (Ni), the nature of the carbonaceous species and therefore photo- thermal performance can be tuned. CH4 production rates of 488 mmol g-1 h-1 under UV–visible-IR irradiation are achieved when the catalyst is prepared under optimized conditions. No particle aggregation or significant loss of activity were observed after ten consecutive reaction cycles or more than 12 hours under continuous flow configuration. Finally, as a proof-of-concept, we performed an outdoor experiment under ambient solar irradiation, demonstrating the potential of our catalyst to reduce CO2 to CH4 using only solar energy.
Anthropogenic CO2 emissions have increased over the past decade to levels as high as 34.2 Gt per annum in 2019.[1] In the quest for strategies to mitigate the impact of CO2 emissions, the development of technologies able to recycle CO2 are expected to play a large role.[2] Among the different products attainable through CO2 conversion, the production of methane has attracted considerable attention.[3] Methanation is the reaction by which carbon oxides and hydrogen are converted to methane and water and is industrially used to purify synthesis gas over Ni based catalysts.[4] Methanation catalysts are also required for the Power- to-methane process where gaseous methane is used as chemical storage materials of surplus renewable energy.[5] Additionally, methanation catalysts are also used as side catalysts in the production of ammonia to transform poisonous for most of the catalysts CO and CO2.[6]
Photo-thermal catalysis results from the combination of photo- and thermo-chemical processes.[7] Upon illumination, hot charge carriers generated through the decay of localized surface plasmon resonance (LSPR) in metallic nanoparticles may dissipate energy through local heating, resulting in an increase in temperature of the catalyst.[8] This photo-thermal effect has been used to run different catalytic processes and holds great promise for the distributed valorization of different feedstocks such as carbon dioxide.
In case of CO2 methanation, noble metals nanoparticles (NPs), specially based on Pd and Ru have been widely explored.[9] As it is the case for more traditional heterogeneous catalysts, the use of more abundant transition metals, such Ni, Co and Fe has also been studied.[10] The nature of the metal nanoparticle, however, is not the only factor determining catalytic performance. Indeed the support, and, more specifically its thermal and optical properties, play a very important role too.[7b]
For example, carbon-based materials (i.e. graphene, carbon nanotubes, etc) hold great promise as a photo-thermal supports due to their broadband light absorption, high efficiency in the conversion of photo-energy into thermal energy and high surface areas.[11] Furthermore, their high thermal conductivity offers additional advantages for CO2 methanation.[12]
We envisaged the synthesis of inexpensive Ni on carbon based catalysts for the photo-thermal methanation of CO2. Given the abundance and low price of Ni, effective catalysts should contain a high metal loading in the form of small nanoparticles.
When it comes to the support, control over the nature of carbonaceous species and a good nanoparticle dispersion are also very important. In order to achieve this non trivial target (high Ni loading, dispersion and controlled support properties) the use of Metal-Organic Frameworks (MOFs) as catalyst precursors seemed like the most appropriate approach. Indeed, the last decade has seen a number of examples in which MOFs have been used as sacrificial templates for the synthesis of heterogeneous catalysts. This can be achieved through routes such as thermal treatment under inert or oxidative atmospheres[13]
or even electrochemically.[14] Upon thermal treatment under inert atmosphere, MOFs decompose to form highly dispersed metal nanoparticles encapsulated in a porous graphitic carbon matrix,[15]
while oxidative treatments are used to generate in situ metal oxide supports.[16] In fact, in many cases, the properties of the resulting solids far exceed the ones of similar materials prepared through more traditional synthetic approaches due to the intrinsic features of the parent MOF.[17] In the field of photothermal CO2
methanation, Chen et al. have recently reported a Co/Al2O3
composite, which was obtained by the calcination of the Al2O3@ZIF-67 composite. The catalyst exhibited CH4 production rate of 6036 μmol g−1 h−1and selectivity of 97.7% after UV−vis−IR light irradiation.[18]
Herein, we explore the MOF-mediated approach to synthesize highly dispersed Ni NPs supported on a carbon matrix derived from Ni-MOF-74. We demonstrate the outstanding catalytic performance of the Ni@C materials in the methanation reaction, achieving high conversion rates, selectivity and recyclability without any external heating.
Ni-MOF-74 was synthesized according to the reported method,[19] and obtained as a pure phase as confirmed by X-ray powder diffraction (PXRD) and N2 adsorption (SBET = 1299 m² g-
1; Figures S1-S3). Once synthesized, Ni-MOF-74 was pyrolyzed in a tubular oven under a continuous flow of N2 for 6 hours at temperatures of 500, 600 and 700 ºC. PXRD patterns and High‐
angle annular dark‐field scanning transmission electron
Accepted Manuscript
COMMUNICATION
2
microscopy (HAADF‐STEM) images of the resulting materials (denoted as Ni@C-X, where X = 500, 600, 700) confirmed the formation of Ni nanoparticles (fcc form) confined within a graphitic carbon matrix (2θ = 27°) with an average particle size (Ni@C-500, 600, 700) of 10.1 ± 4.3, 9.5 ± 5.3, 9.4 ± 5.1 nm, respectively (Figures 1, S4-S7). Analysis of the surface of the Ni@C-X catalysts by X-ray photoelectron spectroscopy (XPS) revealed that the Ni surface exists in form of Ni0 (Ni 2p3/2, 852.7 eV) and Ni2+ (Ni 2p3/2, 855.9 eV) (Figures S8-S10).[20] The presence of Ni2+ is attributed to the partial oxidation of Ni0 during the passivation step after the pyrolysis of Ni-MOF-74. In fact, the O 1s core level spectra were deconvoluted into three peaks located at 529.7, 532.0, and 533.6 eV corresponding to the M–O–M, M–
OH, M–H2O/C=O bonds, respectively. The C 1s spectra were fitted with five components located at 284.3, 285.3, 286.8, 288.3 and 289.6 eV corresponding to the C=C(sp2), C–C (sp3), C–O, C=O, and O–C=O bonds, respectively.[21] Additionally, STEM- EELS images of Ni@C-X materials revealed the homogeneous distribution of the Ni NPs through the carbon matrix (Figures 1, S11
(
The content of Ni was estimated by thermogravimetric analysis (TGA) under air atmosphere, from which Ni content of 69.9 %, 72.5 % and 71.4 % for Ni@C-500, Ni@C-600 and Ni@C- 700 were determined (Figures S12-S14).The graphitization degree of the carbon matrix in the Ni@C- X catalysts was evaluated by Raman spectroscopy. As shown in Figure S15, the Raman spectra of the catalysts exhibit the characteristic vibration bands of disordered graphite (D band) at 1360 cm-1 and the E2g mode of graphite (G band) at 1586 cm−1. The intensity ratio of the two bands (IG/ID) was used to evaluate the graphitization degree of the different catalysts. Thus, we found that higher pyrolysis temperatures lead to a higher degree of graphitization (Ni@C-600: 1.19 and Ni@C-700: 1.50). Particularly, different IG/ID values were obtained when the Raman spectrum was collected at different spots of the Ni@C-500 sample, revealing the heterogeneous nature of samples pyrolyzed at the lowest temperature. This can be attributed to the fact that the Ni- MOF-74 was not fully pyrolized at 500 °C. Indeed, the TGA curve
of the pristine MOF under N2 atmosphere (Figure S2) showed a mass loss of ~ 6 % from 500 °C to 600 °C.
UV-Vis absorption spectra of Ni@C-X catalysts revealed a strong light absorption over a broad photo-irradiation range including UV and visible light (Figure S16), which is characteristic of carbon based-materials. This strong absorption suggests that a high utilization efficiency of solar energy could be feasible by the Ni@C-X catalysts.[22] Finally, as expected, pyrolysis also led a significant reduction of the porosity compared to the pristine MOF (Figures S17-S19). The BET surface areas were reduced from 1299 m² g-1 to 149-171 m² g-1 (Table S1).
The photo-thermal catalytic activity of Ni@C-X materials was tested using a quartz photoreactor and a 300 W Xe lamp as UV-visible-IR light source (see Supplementary Information for further experimental details). As it can be seen in Figure 2a, all samples showed a progressive increase in the temperature of the catalyst bed upon light irradiation, along with a drop in the reactor pressure due to the consumption of gas reactants. However, Ni@C-600 displayed the highest temperature and the most significant pressure decrease after 10 min of reaction thus leading to very high CH4 production rates: 488 mmol g-1 h-1 (Figure 2b, Figure S23-S25). To the best of our knowledge, this activity is among the highest reported for the CO2 photo-methanation reaction (Table S5). To provide a more accurate comparison of the catalytic activities of the pyrolized samples used in this study, we also calculated the turnover frequency (TOF) values for CH4
production (Table S5). As it can be seen, Ni@C-600 rendered a 2-fold and 5-fold higher TOF values compared to Ni@C-500 and Ni@C-700, respectively. Because of the higher activity of Ni@C- 600, further experiments were performed with this sample, unless otherwise indicated.
The differences in performance between the catalysts can be attributed to the inaccessibility of the nickel nanoparticles, which are stuck in the carbon shell of the Ni@C-700 catalyst,[23]
and to the heterogeneous structure of the carbon matrix in the Ni@C-500. In this regard, it has been demonstrated that a higher degree of graphitization of carbon-based materials is related with a higher absorption of light in the NIR region together with an Figure 1. High-resolution dark field STEM images of Ni@C-600 and STEM-EELS mapping showing the distribution of Ni throughout the solid.
Accepted Manuscript
enhanced thermal conductivity that ultimately contribute to a better photo-thermal performance.[24] This would explain the better photo-thermal performance of Ni@C-600 compared to Ni@C-500. However, in the case of Ni@C-700, Raman spectroscopy confirmed a strong degree of graphitization that led to an excessive coverage of Ni NPs (vide supra). XPS data revealed that the difference in the performance between Ni@C- 700 and Ni@C-600 can also be attributed to the lower surface accessibility of the nickel nanoparticles in the Ni@C-700 (Table S3). In the case of Ni@C-500 the accessible metallic content was found to be similarly lower than the content of Ni@C-600.
However, the lower pyrolysis temperature results not only in a more heterogeneous carbon matrix, but also in the higher content of surface nickel oxide (Table S3). Moreover, the accessible nickel surface areas were determined by H2 chemisorption analysis. As expected, the metallic surface area and metal dispersion for Ni@C-600 was almost 4-fold higher compared with Ni@C-500 and Ni@C-700 (Figure S20, Table S2). In order to investigate the structural changes of the Ni@C-X (X = 500, 600, 700) after the photo-thermal reaction, HAADF-STEM, XRD and XPS were performed on the spent catalysts. Remarkable, HAADF-STEM images showed that the catalysts remain intact without any significant agglomeration of the Ni nanoparticles (Figures S27-S29, S34). This can be attributed to the confinement of the nanoparticles into the porous carbon matrix. Although PXRD patterns show only the Ni nanoparticles phase (Figure S26), Ni core-level XPS spectra revealed that after reaction slightly more Ni0 has been oxidized (Figures S31-S33, Table S3).
These findings confirm the stability of the catalysts after reaction.
Once the pyrolysis temperature was optimized, we performed additional experiments in order to disentangle the operating mechanisms in the photo-thermal CO2 conversion by Ni@C-600. As it has been previously mentioned, photo-thermal catalysis arises from the synergistic combination of photochemical and thermochemical contributions. In order to determine which of the aforementioned pathways is predominant in our system, we studied the relationship between reaction rate and light intensity. Interestingly, a distinct exponential correlation between the light intensity and the CH4 production was observed (Figure S35a). This observation is a clear signature of thermal- driven reactions, thus suggesting that this system is mainly operated by an effective light-to-heat conversion.[25] To provide
further evidence of this thermal enhancement, we performed a blank experiment under dark conditions using external heating to achieve similar temperature as in the light-driven tests (Figure S35b). The fact that both CO2 conversion and CH4 production rate under dark thermal conditions were comparable to the ones observed under light radiation, confirm that the thermal enhancement is preeminent.
Figure 2. a) Temperature (turquoise) and pressure (orange) profiles of Ni@C-500 (squares), Ni@C-600 (circles) and Ni@C-700 (triangles) under photo- thermal CO2 conversion. b) CO2 conversion (turquoise bars) and CH4 production rates (orange bars) of Ni@C-500, Ni@C-600 and Ni@C-700 under photo- thermal CO2 conversion. Reaction conditions: 25 mg of photocatalyst, 5 bar (H2/CO2=4) and 4.3 W cm-2 UV-visible-IR light intensity.
Accepted Manuscript
COMMUNICATION
4
One of the main advantages of photo-thermal catalysis compared to traditional semiconductor-based photocatalysis is the possibility of taking advantage of low energy photons from the visible-IR region that are usually unable to trigger photochemical reactions. Therefore, we compared the catalytic performance of Ni@C-600 under full spectrum radiation and using a L42 (λ>420 nm) cut-off filter at a constant power density. Noticeably, we found that both reaction kinetics and temperature increase were considerably enhanced under visible-IR radiation, while the experiment under full spectrum displayed a slower reaction rate and a lower temperature (Figure 3). Indeed, full spectrum test rendered 90 % CO2 conversion and a CH4 production rate of 154 mmol g-1 h-1 after 35 min of reaction, whereas visible-IR radiation displayed higher conversion and a CH4 production rate of 226 mmol g-1 h-1 in just 25 min of reaction. These observations are in line with related precedents that have reported a superior photo- thermal conversion efficiency of supported Ni NPs under visible- IR radiation than under UV-visible-IR illumination.[26] This photo-
thermal response under low-energy photons implies that the Ni@C-600 photocatalyst converts more efficiently visible-IR light into heat, which ultimately provides the driving force to hydrogenate CO2 to CH4. In addition, the individual contributions of visible and IR light were also investigated. To this end, we used 420 and 800 nm cut-off filters to irradiate with visible-IR or sole IR radiation, respectively. As it can be seen in Figure S36, under the same light intensity, IR radiation displays a better light-to-heat conversion which translates into a three-fold higher catalytic performance compared to visible-IR radiation. Indeed, the maximum temperature reached in these experiments was 170 °C for visible-IR and 181 °C for IR, thus again evidencing the more effective light-to-heat conversion of long-wavelength radiation.
This enhanced photo-thermal conversion can be explained thanks to the broad absorption of light of Ni@C in the NIR region (inset in Figure S36). Nonetheless, this light-to-heat conversion derives not only from the well-known photo-thermal activity of Ni NPs but also from the carbonaceous matrix present in the Ni@C photocatalyst.[27] In spite that the operating mechanisms of the photo-thermal effect in carbon materials are not completely understood, especially under UV and visible radiation, a generally accepted theory is that photo-excited electrons can transfer their excess of energy in the form of lattice phonons that eventually produce heat.[11b, 24b] If we take into account that carbon materials commonly display a wide absorption band and a considerable thermal conductivity, these carbonaceous structures are considered perfect host materials for photo-thermal applications.[28] Altogether, the noteworthy photo-thermal performances of both Ni NPs and the carbon matrix make the Ni@C composite here presented a promising photocatalyst to perform the conversion of CO2 into CH4 using sunlight as unique energy source.
In order to demonstrate the advantages of our MOF-derived Ni@C photocatalyst, we performed additional experiments using Co NPs supported on a carbon matrix derived from Co-MOF-74, and Ni NPs supported on different commercial substrates. To this end, we prepared a Co@C-600 catalyst under similar conditions of Ni@C-600 (Figures S37-S38), and Ni NPs on carbon (Ni/C) and inert SiO2 (Ni/SiO2) supports with 22% and 21% metal content respectively. Co NPs were selected due to their well-known high Figure 3. Temperature (turquoise lines) and pressure (orange lines)
profiles of Ni@C-600 under full spectrum radiation (squares) or using a L42 (λ>420 nm) cut-off filter (circles) under constant power density.
Reaction conditions: 25 mg of photocatalyst, 5 bar (H2/CO2=4) and 3.3 W cm-2 light intensity.
Figure 4. a) CO2 conversion values of Ni@C-600 upon consecutive catalytic cycles. Reaction conditions: 30 mg of photo-catalyst, 5 bar (H2/CO2=4) and 4.3 W cm-2 UV-visible-IR light intensity. b) CO2 conversion values of Ni@C-600 under continuous flow configuration. Reaction conditions: 45 mg of photocatalyst, 15 mL min-1 total flow (H2/CO2=4), 1 bar and 4.3 W cm-2 UV-visible-IR light intensity * Ni@C-600 was re-activated under H2 and light for 40 min.
Accepted Manuscript
CO2 methanation activity. On the other hand, carbon was used to compare the catalytic performance of Ni NPs supported on an alternative carbon matrix, while SiO2 was selected as a good photo-thermal support for metallic NPs according to its low thermal conductivity that maximizes temperature at the active sites.[29] Figure S39 shows the CO2 conversion values of these experiments along with our Ni@C-600 photocatalyst. It should be mentioned that in these experiments, the total amount of Ni was constant in order to establish a proper comparison. As expected, Co@C-600 showed high catalytic activity, demonstrating that Co is very effective methanation catalyst. Importantly, we found that Ni NPs supported on commercial carbon (Ni/C) displayed a very low catalytic activity whereas Ni NPs supported on SiO2 exhibited a catalytic performance comparable to Ni@C-600. These observations confirmed that the excellent photo-thermal performance of the MOF-derived photocatalysts stems from the pyrolysis step of the MOF structure. In fact, this preparation procedure favors the encapsulation of Ni NPs in the carbon matrix in a way that prevents structural changes of the metallic active sites, compared to the case of Ni NPs just supported on carbon substrate. Indeed, previous reports have described the enhanced photo-thermal activity of metallic NPs coated by graphene- derived layers, in contrast to the uncoated counterparts.[30] In a similar way, our photocatalyst benefits from the cooperation between Ni NPs and the carbon layer that collectively maximizes the photo-thermal performance of the metal-carbon composite.
On the other hand, according to the inertness of the SiO2 support, the notable catalytic activity exhibited by Ni/SiO2 demonstrated that Ni NPs are the active phase and that this activity mainly derives from a thermal enhancement, in this case favored by the heat insulating properties of SiO2.
Stability issues represent one of the major drawbacks for the practical application of many catalysts.[31] For this reason, we studied the stability of our catalyst upon ten consecutive cycles, as depicted in Figure 4. As per these results, the catalytic activity exhibited a slight decrease due to the re-oxidation of the Ni active sites by water inside the reactor. Indeed, after seven uses the catalyst was re-activated under H2 and light for 40 minutes to in situ reduce the Ni active sites and this treatment translated into an improved catalytic activity for at least three more catalytic cycles. All in all, these results demonstrate that, in spite that the catalyst suffers from a certain deactivation upon reuses due to the oxidation associated with the inherent presence of water as by- product, it is possible to completely restore its catalytic activity by activating the material using H2 and light. Oppositely, Ni/SiO2
photocatalyst displayed a drastic deactivation upon reuses (Figure S40b). This deactivation has been already described in other Ni-based photocatalysts and it has been attributed to the partial oxidation of Ni NPs by H2O formed as byproduct in the methanation reaction under batch conditions.[32] In contrast, the here presented Ni@C-600 photocatalyst better preserves the activity of Ni NPs against the oxidation by H2O, probably due to the presence of a carbonaceous protective shell resulting from the pyrolysis of the MOF precursor. Indeed, from the deconvoluted Ni 2p XPS spectra of the Ni@C-600 (fresh and spent), the calculated ratio of Ni2+/Ni0 slightly increased from 0.33 to 0.46 after the photo-thermal reaction, whereas from the XPS spectra of the Ni@SiO2, the ratio Ni2+/Ni0 increased significantly from 1.88 to 10.7, consistent with the fast deactivation of Ni@SiO2 (Figure S41, Table S3). This features an additional advantage of the synthesis method of the catalyst, which provides not only an active photo-
thermal carbonaceous matrix for the embedded Ni active sites, but also a preservative layer that stabilizes metallic NPs.
In an effort to prove even more the robustness of our catalyst and given the limited duration of the above-mentioned stability tests (10 minutes), we performed an additional catalytic test using a flow-type reactor (see Supplementary information for further details). This approach allowed us to better study the stability of the catalyst under long-term reaction conditions. As it can be seen in Figure 4b, the catalytic activity for Ni@C-600 exhibited a notable CO2 conversion ranging from 42 to 57 % after more than 12 consecutive hours of reaction without evident deactivation.
This key experiment under flow conditions demonstrated the remarkable stability of our photocatalyst and further confirmed that the source of catalyst deactivation under batch configuration was none other than the surface oxidation of Ni active sites by water by-product. In fact, here we have demonstrated that one of the benefits from the continuous flow configuration is the possibility to remove all the products from the reaction chamber, thus avoiding the presence of any potential poison for the catalyst.
One of the drawbacks of the transition metal (i.e. Ni)/carbon- based catalyst during the conventional thermal methanation reaction is coke formation. Therefore, in order to investigate the coke formation during the reaction, we performed thermogravimetric analysis (TGA) for the fresh and spent Ni@C- 600 catalysts (Figure S30). Interestingly, TGA curves showed that there is a negligible coke formation on our Ni-carbon catalyst, demonstrating the advantage of photo-thermal approach over the thermal methanation reaction.
In the particular case of carbon-based photocatalysts for CO2 reduction, it is from pivotal importance to demonstrate that reaction products actually come from the fed CO2 rather than from adventitious carbon sources present in the sample. With this purpose, we performed a blank experiment loading the reactor with the photocatalyst and H2, but in the absence of CO2. Under these conditions, negligible amounts of CH4 below 0.08 % were detected (Table S4). Furthermore, additional experiments using isotopically labelled 13CO2 showed the presence of 13CH4 (m/z = 17), H2O (m/z = 18) and 13CO2 (m/z = 45) when the reaction mixture was analyzed by GC-MS (Figure S32). Altogether, these results provided convincing evidences of the origin of the observed CH4 product, which was none other than the reactant CO2 gas.
In order to demonstrate the practical application of the here presented Ni@C-600 photocatalyst in the light-induced CO2
methanation reaction, we performed an additional experiment under outdoor sunlight. The experiment was carried out on March 2nd, 2021 from 10:00 to 16:00 under natural sunlight in Thuwal, Saudi Arabia (Figure S43). Under these conditions, a productivity rate of 1.1 µmol CH4 g-1 h-1 was achieved. As expected, the catalytic performance of this outdoor experiment was considerably low compared to the one achieved under concentrated light from the Xe lamp (∼40 sun). However, this experiment provides a valuable proof-of-concept to demonstrate that the light-promoted CO2 methanation reaction under ambient solar light is a feasible process. In this case, reactor engineering to maximize light harvesting or the use of condenser lens to concentrate sunlight could be possible strategies to further enhance the catalytic activity under outdoor conditions.
In summary, we demonstrated the effectiveness of MOF- mediated approach for the synthesis of metal nanoparticles embedded on carbon matrix catalysts. Our catalyst showed high
Accepted Manuscript
COMMUNICATION
6
performance for the photo-thermal methanation reaction under visible and IR irradiation. Additional outdoor experiments showed that the catalyst is able to produce methane under ambient solar irradiation. We have also demonstrated that our catalyst displayed excellent recyclability without loss of the catalytic activity compared with the catalysts of Ni NPs supported on conventional substrates. We found that the carbon matrix played a beneficial role in the stability of our system preventing the deactivation of the nickel NPs by the water formed as a byproduct during the reaction.
Acknowledgements
Funding for this work was provided by King Abdullah University of Science and Technology (KAUST). The authors wish to thank and acknowledge to Dr. Serhii Vasylevskyi from KAUST-Core labs for his support during the Raman spectroscopy measurements, Dr.
Natalia Morlanes for her assistance for the flow-reactor configuration, and Sandra Ramirez for the art work.
Keywords: MOF, Photo-thermal catalysis, Methanation, Pyrolysis, Carbon dioxide
[1] B. Petroleum, 69 ed., London, 2020.
[2] a) R.-P. Ye, J. Ding, W. Gong, M. D. Argyle, Q. Zhong, Y.
Wang, C. K. Russell, Z. Xu, A. G. Russell, Q. Li, M. Fan, Y.-G. Yao, Nat. Commun. 2019, 10, 5698; b) D. Goud, R.
Gupta, R. Maligal-Ganesh, S. C. Peter, ACS Catal. 2020, 10, 14258-14282; c) A. Álvarez, A. Bansode, A. Urakawa, A. V. Bavykina, T. A. Wezendonk, M. Makkee, J. Gascon, F. Kapteijn, Chem. Rev. 2017, 117, 9804-9838; d) Y. Ni, Z. Chen, Y. Fu, Y. Liu, W. Zhu, Z. Liu, Nat. Commun.
2018, 9, 3457; e) S. De, A. Dokania, A. Ramirez, J.
Gascon, ACS Catal. 2020, 10, 14147-14185.
[3] a) M. Jacquemin, A. Beuls, P. Ruiz, Catal. Today 2010, 157, 462-466; b) M. D. Burkart, N. Hazari, C. L. Tway, E.
L. Zeitler, ACS Catal. 2019, 9, 7937-7956.
[4] a) I. Sreedhar, Y. Varun, S. A. Singh, A. Venugopal, B. M.
Reddy, Catalysis Science & Technology 2019, 9, 4478- 4504; b) M. A. A. Aziz, A. A. Jalil, S. Triwahyono, A.
Ahmad, Green Chem. 2015, 17, 2647-2663; c) S. Ronsch, J. Schneider, S. Matthischke, M. Schluter, M. Gotz, J.
Lefebvre, P. Prabhakaran, S. Bajohr, Fuel 2016, 166, 276- 296.
[5] K. Ghaib, F.-Z. Ben-Fares, Renewable and Sustainable Energy Reviews 2018, 81, 433-446.
[6] A. S.Abdullah, A. A.Abdulabbas, A. A.Al-Asadi, International Journal of Advanced Research in Chemical Science 2015, 2, 80-84.
[7] a) M. Ghoussoub, M. Xia, P. N. Duchesne, D. Segal, G.
Ozin, Energy Environ. Sci. 2019, 12, 1122-1142; b) U.
Ulmer, T. Dingle, P. N. Duchesne, R. H. Morris, A.
Tavasoli, T. Wood, G. A. Ozin, Nat. Commun. 2019, 10, 3169; c) L. Zhu, M. Gao, C. K. N. Peh, G. W. Ho, Mater.
Horiz. 2018, 5, 323-343.
[8] D. Mateo, J. L. Cerrillo, S. Durini, J. Gascon, Chem. Soc.
Rev. 2021.
[9] a) J. Jia, H. Wang, Z. Lu, P. G. O'Brien, M. Ghoussoub, P.
Duchesne, Z. Zheng, P. Li, Q. Qiao, L. Wang, A. Gu, A. A.
Jelle, Y. Dong, Q. Wang, K. K. Ghuman, T. Wood, C.
Qian, Y. Shao, C. Qiu, M. Ye, Y. Zhu, Z.-H. Lu, P. Zhang, A. S. Helmy, C. V. Singh, N. P. Kherani, D. D. Perovic, G.
A. Ozin, Advanced Science 2017, 4, 1700252; b) D.
Mateo, J. Albero, H. García, Joule 2019, 3, 1949-1962.
[10] a) X. Meng, T. Wang, L. Liu, S. Ouyang, P. Li, H. Hu, T.
Kako, H. Iwai, A. Tanaka, J. Ye, Angew. Chem. Int. Ed.
2014, 53, 11478-11482; b) G. Chen, R. Gao, Y. Zhao, Z.
Li, G. I. N. Waterhouse, R. Shi, J. Zhao, M. Zhang, L.
Shang, G. Sheng, X. Zhang, X. Wen, L.-Z. Wu, C.-H.
Tung, T. Zhang, Adv. Mater. 2018, 30, 1704663.
[11] a) G. M. Neelgund, A. Oki, Mater. Chem. Front. 2018, 2, 64-75; b) J. Wang, G. Zhang, P. Zhang, Appl. Catal., B 2018, 239, 77-85.
[12] a) D. Mateo, J. Albero, H. García, Energy Environ. Sci.
2017, 10, 2392-2400; b) D. Mateo, N. Morlanes, P. Maity, G. Shterk, O. F. Mohammed, J. Gascon, Adv. Funct.
Mater. 2021, 31, 2008244.
[13] A. Bavykina, N. Kolobov, I. S. Khan, J. A. Bau, A.
Ramirez, J. Gascon, Chem. Rev. 2020, 120, 8468-8535.
[14] R. Wang, H. Haspel, A. Pustovarenko, A. Dikhtiarenko, A.
Russkikh, G. Shterk, D. Osadchii, S. Ould-Chikh, M. Ma, W. A. Smith, K. Takanabe, F. Kapteijn, J. Gascon, ACS Energy Letters 2019, 4, 2024-2031.
[15] a) S. Dang, Q.-L. Zhu, Q. Xu, Nat. Rev. Mater. 2017, 3, 17075; b) L. Oar-Arteta, T. Wezendonk, X. Sun, F.
Kapteijn, J. Gascon, Mater. Chem. Front. 2017, 1, 1709- 1745.
[16] X. Sun, A. I. O. Suarez, M. Meijerink, T. van Deelen, S.
Ould-Chikh, J. Zečević, K. P. de Jong, F. Kapteijn, J.
Gascon, Nat. Commun. 2017, 8, 1680.
[17] a) J. Hwang, A. Ejsmont, R. Freund, J. Goscianska, B. V.
K. J. Schmidt, S. Wuttke, Chem. Soc. Rev. 2020, 49, 3348-3422; b) J. Castells-Gil, S. Ould-Chikh, A. Ramírez, R. Ahmad, G. Prieto, A. R. Gómez, L. Garzón-Tovar, S.
Telalovic, L. Liu, A. Genovese, N. M. Padial, A. Aguilar- Tapia, P. Bordet, L. Cavallo, C. Martí-Gastaldo, J.
Gascon, Chem Catalysis 2021, 1, 364-382.
[18] X. Chen, Q. Li, M. Zhang, J. Li, S. Cai, J. Chen, H. Jia, ACS Appl. Mater. Interfaces 2020, 12, 39304-39317.
[19] L. Garzón-Tovar, A. Carné-Sánchez, C. Carbonell, I. Imaz, D. Maspoch, J. Mater. Chem. A 2015, 3, 20819-20826.
[20] a) A. P. Grosvenor, M. C. Biesinger, R. S. C. Smart, N. S.
McIntyre, Surf. Sci. 2006, 600, 1771-1779; b) M. C.
Biesinger, B. P. Payne, L. W. M. Lau, A. Gerson, R. S. C.
Smart, Surf. Interface Anal. 2009, 41, 324-332.
[21] L. Zhang, G. Chen, M. N. Hedhili, H. Zhang, P. Wang, Nanoscale 2012, 4, 7038-7045.
[22] H. Zhang, T. Wang, J. Wang, H. Liu, T. D. Dao, M. Li, G.
Liu, X. Meng, K. Chang, L. Shi, T. Nagao, J. Ye, Adv.
Mater. 2016, 28, 3703-3710.
[23] I. S. Khan, A. Ramirez, G. Shterk, L. Garzón-Tovar, J.
Gascon, Catalysts 2020, 10.
[24] a) G. M. Neelgund, A. R. Oki, J. Colloid Interface Sci.
2016, 484, 135-145; b) B. Han, Y.-L. Zhang, Q.-D. Chen, H.-B. Sun, Adv. Funct. Mater. 2018, 28, 1802235.
[25] M. J. Kale, T. Avanesian, P. Christopher, ACS Catal.
2014, 4, 116-128.
[26] Q. Zhang, M. Mao, Y. Li, Y. Yang, H. Huang, Z. Jiang, Q.
Hu, S. Wu, X. Zhao, Appl. Catal., B 2018, 239, 555-564.
[27] a) X. Meng, T. Wang, L. Liu, S. Ouyang, P. Li, H. Hu, T.
Kako, H. Iwai, A. Tanaka, J. Ye, Angew. Chem. Int. Ed.
2014, 53, 11478-11482; b) Y. Zhao, A. Dunn, J. Lin, D.
Shi, in Novel Nanomater. Biomed., Environ. Energy Appl., Elsevier, 2019, pp. 415-434.
[28] X. Wu, G. Y. Chen, G. Owens, D. Chu, H. Xu, Materials Today Energy 2019, 12, 277-296.
[29] T. Gao, B. P. Jelle, J. Nanopart. Res. 2019, 21, 108.
[30] D.-K. Lim, A. Barhoumi, R. G. Wylie, G. Reznor, R. S.
Langer, D. S. Kohane, Nano Lett. 2013, 13, 4075-4079.
[31] a) D. Wierzbicki, R. Baran, R. Dębek, M. Motak, T.
Grzybek, M. E. Gálvez, P. Da Costa, Int. J. Hydrogen Energy 2017, 42, 23548-23555; b) Z. Zhang, Y. Tian, L.
Zhang, S. Hu, J. Xiang, Y. Wang, L. Xu, Q. Liu, S. Zhang, X. Hu, Int. J. Hydrogen Energy 2019, 44, 9291-9306.
[32] a) D. Mateo, J. Albero, H. García, Appl. Catal., B 2018, 224, 563-571; b) D. Mateo, N. Morlanes, P. Maity, G.
Shterk, O. F. Mohammed, J. Gascon, Adv. Funct. Mater., 2008244.
…
Accepted Manuscript
COMMUNICATION
Entry for the Table of Contents
A Ni-based catalyst derived from Ni-MOF-74 was synthesized and studied for the photo-thermal methanation reaction. The catalyst shows high performance for CH4 production under visible-IR irradiation. The catalysts also shows high stability and excellent recyclability after several consecutive reaction cycles.
Institute and/or researcher Twitter usernames: @jorgascon, @adv_cat_mat, @KCCkaust