Next Article in Journal
Progress & Prospect of Enzyme-Mediated Structured Phospholipids Preparation
Next Article in Special Issue
Effect of Modulation and Functionalization of UiO-66 Type MOFs on Their Surface Thermodynamic Properties and Lewis Acid–Base Behavior
Previous Article in Journal
Recent Breakthrough in Layered Double Hydroxides and Their Applications in Petroleum, Green Energy, and Environmental Remediation
Previous Article in Special Issue
Copper and Iron Cooperation on Micro-Spherical Silica during Methanol Synthesis via CO2 Hydrogenation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Differentiating the Reactivity of ZrO2-Bound Formates Formed on Cu/ZrO2 during CO2 Hydrogenation

Institut de Recherches sur la Catalyse et l’Environnement de Lyon, IRCELYON, Université Claude Bernard Lyon 1, Univ Lyon, CNRS, 2 Av. Albert Einstein, 69626 Villeurbanne, France
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(7), 793; https://doi.org/10.3390/catal12070793
Submission received: 29 June 2022 / Revised: 6 July 2022 / Accepted: 13 July 2022 / Published: 19 July 2022

Abstract

:
The surface species formed during the hydrogenation of CO2 with H2 over a ZrO2-supported Cu catalyst were investigated by operando diffuse reflectance FT-IR spectroscopy at 220 °C and 3 bar. The reactivity of two different formates located on zirconia could be unraveled. The data pointed to ZrO2 hydroxyl groups at 3755 cm−1 as the sites on which carbonates and then formates were hydrogenated to methoxy species. Formate hydrogenation appeared as the slowest step. The most reactive ZrO2-bound formates exhibited a rate constant of reaction about 65 times higher than that of the slower formate.

1. Introduction

The hydrogenation of CO2 to produce methanol enables recycling CO2 into a base chemical that can also be directly used as a fuel [1,2,3,4]. Cu/ZnO/Al2O3 is the reference industrial catalyst for methanol synthesis and has been developed primarily for CO-containing feed [3]. A major difference between using CO (Equation (1)) and using CO2 (Equation (2)) for manufacturing methanol is the formation of water, which can also be obtained through the reverse water gas-shift reaction (RWGS, Equation (3)):
CO + 2H2→CH3OH
CO2 + 3H2→CH3OH + 2H2O
CO2 + H2→CO + H2O
The increased presence of water in the reactor will impact the long-term stability of Cu/ZnO/Al2O3 catalysts due to the poor hydrothermal stability of alumina. In fact, Prašnikar et al. showed that a correlation existed between alumina surface area and Cu particle sizes during accelerated aging tests [5]. Golunski and Burch [3]. stressed the need to stabilize alumina against densification or find a more durable support.
ZrO2 is a strong candidate to be used as a support for methanol synthesis catalysts because of its well-known hydrothermal stability. The melting point of zirconia is 2715 °C, in comparison to 2072 °C for alumina. ZrO2-based catalysts have been successfully used for many demanding reactions such as the dry [6] and steam [7] reforming of methane, steam-reforming of methanol [8] (the reverse reaction of that of interest here) and aqueous-phase reactions in acidic media [9].
Unpromoted ZrO2 present some activity by itself for CO2 hydrogenation, its combination with ZnO yield even greater interesting activity, although less than those obtained when combined with Cu [10]. In fact, the combination of Cu and ZrO2 is one of the most efficient to obtain high reaction rates to methanol with CO/CO2/H2 [3,11] or CO2/H2 mixture [12,13,14,15] Cu supported on amorphous zirconia was found to be more active than the catalysts made from monoclinic or tetragonal polymorphs, though ageing was not considered [16].
Numerous IR studies have been reported on the Cu-catalyzed hydrogenation of CO [17,18] and CO2 [19,20,21,22,23] to methanol, with formates often proposed as being a main reaction intermediate. Fischer and Bell investigated the hydrogenation of CO2 over Cu/ZrO2/SiO2 and proposed that Cu favored the activation of H2 that spilled over to ZrO2 to promote carbonate/hydrogenocarbonate hydrogenation into formates and then to methoxide species [21].
The role of formates in the hydrogenation of CO2 over ZrO2-supported Cu has actually been the subject of a controversy. Two distinct pathways have been proposed (Figure 1). Kattel et al. [22] proposed a formate-free pathway, starting with the reverse WGS (RWGS) and the formation of a CO-OH hydroxycarbonyl intermediate. In contrast, Larmier et al. [24] proposed a formate-based route, in which a ZrO2-bound formate located at the interface with Cu nanoparticles is hydrogenated by H2 activated on Cu to methoxy, still on the support.
We have previously investigated formate decomposition over alumina-supported cobalt used for CO hydrogenation [25]. The reactivity of formates was determined by inducing a chemical transient, i.e., by removing CO from the feed. Interestingly, two types of formate reactivity were observed, despite the fact that both formate species exhibited identical DRIFTS spectra. These facts were rationalized by proposing that the reactivity of these alumina-bound formates differed depending on their proximity to the alumina-metal interface.
A similar experimental approach has been used here to investigate the hydrogenation of CO2 over a zirconia-supported Cu catalyst. The present work aims at highlighting different types of formate species present at the surface of the catalyst and their relative reactivity.

2. Experimental Section

A 6 wt.% Cu/ZrO2 catalyst was prepared by incipient wetness impregnation of zirconia (from MEL Chemicals, monoclinic, 131 m2 g−1) using Cu(NO3)2. 6 H2O. The impregnated sample was then dried at room temperature for 24 h before being oven-dried at 110 °C for 12 h. The sample was finally calcined at 450 °C for 12 h in synthetic flowing air. The sample was reduced in situ at 350 °C under flowing 80% H2/He for 1 h before the CO2 hydrogenation experiments.
Copper dispersion was measured using N2O titration (frontal chromatography) using a 10 cm-pathlength FT-IR gas cell as detector. The sample was reduced in situ at 300 °C in 50% H2/He for 1 h, purge in He and brought to 75 °C to be exposed to 2000 ppm of N2O in a flow rate of 36 mL min−1. It was assumed that each O deposited on the surface titrated two Cu surface atoms [26,27].
Powder X-ray diffraction patterns (XRD) were recorded to assess the crystallinity of the samples. Diffractograms were collected between 10 and 80° (2Θ) with steps of 0.02° and 1 s per step with a Bruker D8-Advance diffractometer using CuKα radiation at λ = 1.5418 Å. Nitrogen adsorption isotherms were measured at 77 K on an ASAP 2020 from Micromeritics. Samples were first outgassed under vacuum at 300 °C for 3 h.
High-purity gases He, CO, CO2 and H2 from Air Liquid were used for the operando catalytic tests. The experiments were carried out at 3 bars using a mixture of 20% CO2 + 60% H2 in He at a total flowrate of 75 mL min−1, unless otherwise stated. Operando DRIFTS experiments were performed on a modified high-temperature DRIFT cell (from Spectra-Tech, Hong Kong) fitted with CaF2 windows using a Collector II assembly. A description and properties of the cell can be found in earlier references [28,29,30]. The spectrophotometer used was a Nicolet 8700 (ThermoFischer Scientific, Waltham, MA, USA) fitted with a liquid-N2 cooled MCT detector. The DRIFT spectra were recorded at a resolution of 4 cm−1 and 8 scans were averaged. The DRIFTS spectra are reported as log (1/R), where R is the sample reflectance. This pseudo-absorbance gives a better linear representation of the band intensity against surface coverage than that given by the Kubelka–Munk function for strongly absorbing media such as those based on metals supported on oxides [31]. The contribution of gas-phase CO was subtracted using a CO(g) spectrum collected under the same experimental conditions over KBr powder [32,33].
The reaction products were analyzed using a transmission FT-IR gas-cell (200 mL dead-volume) with a 2 m pathlength kept at 60 °C to prevent product condensation. The pressure in the system was controlled by a back-pressure regulator located after the DRIFTS cell and before the transmission FT-IR gas cell. The pressure in the line was precisely measured using an electronic gauge.

3. Results and Discussion

The surface area of the calcined 6 wt.% Cu/ZrO2 catalyst was 123 m2 g−1. The sample exhibited a Cu surface area measured by N2O frontal chromatography of about 8.3 m2 g−1, corresponding to Cu spherical particles with an average diameter of 4.9 nm. In view of these two surface measurements, it can be concluded that the ZrO2 surface represented the largest fraction (ca. 93%) of the surface area of the sample.
The diffraction patterns of the ZrO2 support and Cu/ZrO2 sample showed the presence of mostly monoclinic zirconia with traces of a tetragonal phase (Figure 2A). Very weak and broad peaks associated with CuO were also observed in the case of the calcined Cu/ZrO2 sample, which are consistent with the high dispersion of the Cu phase determined through N2O titration.
The in situ spectrum at 220 °C of the Cu/ZrO2 sample reduced at 350 °C prior to exposure to the feed (Figure 2B, Red) showed two main broad bands at about 3755 and 3665 cm−1 typical of zirconia surface hydroxyl groups [34]. A group of bands near 1505 cm−1 was also observed, indicated the residual presence of strongly bound carbonates that were not decomposed during the reduction step.
Methanol and CO were the only C-containing products observed at 220 °C and 3 bar when the 20% CO2 + 60% H2 feed was introduced. A steady state was achieved within one hour on stream in terms of methanol and CO concentrations measured at the exit of the DRIFTS cell and in terms of DRIFTS signals of the various adsorbates observed. Differential conditions were obtained with a CO2 conversion of about 0.8%. The selectivity to methanol was about 47%, and the productivity was 1.0 µmol(CH3OH) s−1 gCatalyst−1. These values are similar to those reported by Fischer and Bell [21] at 250 °C and 6.5 bar for a Cu/ZrO2/SiO2, that were 1.3 µmol(CH3OH) s−1 gCatalyst−1 and a selectivity to methanol of 43%. The operando spectrum obtained at steady state under the CO2 and H2 feed at 220 °C is shown in Figure 3A (top). The set of bands at 2873, 1575 and 1380 cm−1 corresponded to formate species [35,36], mostly, if not wholly, adsorbed on the zirconia support (Table 1). The band around 1468 cm−1 can be assigned to polydendate carbonates formed from CO2 adsorption on zirconia basic sites [21]. The bands at 2927, 2822, 1150 and 1049 cm−1 were assigned to methoxy groups [34,37]. The methoxy species with the band at 1150 cm−1 (corresponding to the stretching C-O vibration ν(O-C)) was proposed to be formed over zirconia hydroxyl groups located near 3755 cm−1 and be a monodendate species. In contrast, the methoxy with a band at 1049 cm−1 was proposed to derive from the adsorption on the site associated with hydroxyl groups at 3665 cm−1 and be a bridged methoxy [34,37].
Interestingly, no evidence of CO adsorbed on Cu was apparent. To assess whether or not adsorbed CO could ever be observed, an experiment was carried out replacing CO2 with CO. A large band centered around 2010 cm−1 assigned to CO adsorbed on metallic copper [38] was observed when CO was used (Figure 3A (bottom)). This indicates that the coverage of the Cu nanoparticles is very different in the presence of CO2, as compared to the case of CO. More work would be needed to determine the nature of the main species covering the Cu surface in this case (e.g., H, O). It is interesting to note that, apart from the CO(ads) signal, the rest of the spectra was almost identical, pointing to similar coverages of methoxy, formates and carbonates over the ZrO2 support whether CO or CO2 were used.
Table 1. Tentative IR band assignments.
Table 1. Tentative IR band assignments.
Wavenumber (cm−1) References
3755 and 3665Hydroxyls on zirconia[34]
2873, 1575 and 1380Formate on zirconia[35,36]
1468Multi-bonded carbonates[21]
2927, 2822, 1150Methoxy type I[34,37]
2927, 2822, 1049Methoxy type II[34,37]
2010carbonyl on Cu0[38]
CO2 was removed from the feed, and the spectrum obtained after 60 min in H2 at 220 °C is shown in Figure 3B and compared to the steady-state spectrum. The carbonate and formate signals decreased significantly, while that of the methoxy at 1150 cm−1 increased. In contrast, the methoxy at 1049 cm−1 remained essentially unchanged. These observations show first that all the surface species were bound rather strongly to the surface and were poorly reactive under H2 at 220 °C, all likely bound to ZrO2 sites. The data also suggest that a fraction of carbonates and formates had been converted into the methoxy at 1150 cm−1. These results would suggest that zirconia hydroxyl groups at 3755 cm−1, associated with the 1150 cm−1 methoxy, are the location at which carbonate and formate species are hydrogenated to methoxy. Bensitel et al. [34] had actually reported that only the 3755 cm−1 hydroxyl species reacted with CO2 to form hydrogenocarbonate species, stressing the unique reactivity of this site.
The decay of the signal in the formate region was investigated by integrating over the range 1415.5–1313.1 cm−1, using a single point baseline located at 1900 cm−1 (see inset in Figure 3B). The signal decay in the first minute was quite complex, consisting of a rapid evolution of overlapping increasing and decreasing contributions, and will be dealt with in a subsequent contribution.
The signal decay in the time range 1–60 min is presented in Figure 4. The total signal could be decomposed into a slow linear contribution and a faster decay (Figure 4A). A semi-logarithmic plot of the faster signal decay showed a linear behavior, indicating a uniform reactivity (Figure 4B). The difference DRIFTS spectrum obtained by subtracting the spectra collected after having removed CO2 for 2 min and 20 min, when most of the fast signal species had gone, is shown in Figure 5. It shows that the lost signal corresponding to this fast-removed species was essentially a formate similar to that prevailing at steady state. The larger band width observed was possibly due to a larger distribution of heterogeneous sites. The slow remaining species left after more than 30 min under H2 was already discussed above (Figure 3B) and also corresponded to ZrO2-bound formates.
The rate constant of reaction/decomposition of the two normalized signals of the so-called “slow” and “fast” formates on ZrO2 were approximately equal to 2 × 10−3 min−1 and 1.3 × 10−1 min−1, respectively. The fast formate thus exhibited a ca. 65-fold higher reactivity than the slow formate. It must be stressed here that the products (adsorbed or gas-phase) to which these formates decomposed into could not be directly measured.
The evolution of the methoxy DRIFTS band area at 1150 cm−1 was plotted as a function of those of the total formate band (Figure 6A) and fast formate-only band (Figure 6B) following the removal of CO2. These plots show that a strong correlation existed between these quantities, primarily the fast formate band (since the slow formate hardly changed over the duration of the experiment) and the methoxy at 1150−1. This quantitative correlation supports the afore-mentioned model that the ZrO2 hydroxyl groups at 3755 cm−1, on which the 1150 cm−1 methoxy is formed, were the sites where carbonates and then formates were hydrogenated to methoxy species. It is interesting to note that the 3755 cm−1 hydroxyl group was still totally missing after 60 min exposure to H2 at 220 °C (Figure 2B, Black), indicating that those were still involved in bonding some adsorbates (formates or methoxy).
The findings here are consistent with a reaction model in which carbonates formed from CO2 adsorption on zirconia are readily reduced to formate species. The hydrogenation of formates to methoxy is then kinetically limiting, explaining the large signal of formate species. These steps are all favored by H2 activation over Cu followed by spillover of H onto the support, as proposed earlier by Fisher and Bell [21]. The accumulation of methoxy species may also occur, if methanol readsorption is significant and if the reductive elimination or hydrolysis of methoxy groups is slow. Fisher and Bell have actually observed of a similar sample that methoxy hydrolysis (by water formed in the reaction) was significantly faster than reductive elimination [21]. This may explain the accumulation of methoxy species in conditions under which water production is limited or totally absent, as in the present case under the H2/He stream.
Some of the main findings obtained here are summarized in Figure 7. Formate species are observed under reactions conditions that are mostly ZrO2-bound. The reactivity of ZrO2-bound formates is two-fold, with a fast formate exhibiting a rate constant of decomposition 65 times higher than that of a slower formate species. The fast formate species appears to be hydrogenated into methoxy groups that are associated with ZrO2 hydroxyl groups at 3755 cm−1. It is not yet clear if the hydrogenation ability if solely related to the nature of the zirconia sites or if the distance to the copper–zirconia interface matters, for instance, as a result of H-spillover. It is possible that the slow formates were located on domains or ZrO2 crystallites on which no Cu nanoparticles were present and thus almost no spillover H would be available. On the contrary, the fast formates could be located on ZrO2 crystals on which Cu nanoparticles would be present and spillover H would be readily available, enabling a uniform reactivity of formates on such domains.

4. Conclusions

This contribution reveals that on a zirconia-supported copper catalyst used for CO2 hydrogenation, two main types of zirconia-bound formates are present. A significant difference in reactivity is observed (ca. 65-fold), partly related to the nature of the hydroxyl group on which these formates are adsorbed. More work is under way to quantitatively relate the rate of formation of methanol to the reaction rate of the formate species present over zirconia.

Author Contributions

F.C.M. conceptualized the work. F.C.M. and K.E. carried out the IR analyses. I.D. carried out the N2O-TPD and XRD analyses. F.C.M. prepared the draft of the paper and all authors contributed to the final version. Validation: Y.S.; Writing–review & editing: Y.S. The manuscript was written through contributions of all authors. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Meunier, F.C. Mixing Copper Nanoparticles and ZnO Nanocrystals: A Route towards Understanding the Hydrogenation of CO2 to Methanol? Angew. Chem. Int. Ed. 2011, 50, 4053–4054. [Google Scholar] [CrossRef] [PubMed]
  2. Bowker, M. Methanol Synthesis from CO2 Hydrogenation. Methanol Synthesis from CO2 Hydrogenation. ChemCatChem 2019, 11, 4238–4246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Golunski, S.; Burch, R. CO2 Hydrogenation to Methanol over Copper Catalysts: Learning from Syngas Conversion. Top. Catal. 2021, 64, 974–983. [Google Scholar] [CrossRef]
  4. 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]
  5. Pavlišič, P.A.; Ruiz–Zepeda, F.; Kovač, J.; Likozar, B. Mechanisms of Copper-Based Catalyst Deactivation during CO2 Reduction to Methanol. Ind. Eng. Chem. Res. 2019, 58, 13021–13029. [Google Scholar]
  6. O’Connor, A.M.; Meunier, F.C.; Ross, J.R.H. An In-situ DRIFTS Study of the Mechanism of the CO2 Reforming of CH4 over a Pt/ZrO2 Catalyst. Stud. Surf. Sci. Catal. 1998, 119, 819–824. [Google Scholar]
  7. Lim, Z.-Y.; Wu, C.Z.; Wang, W.G.; Choy, K.-L.; Yin, H. Porosity effect on ZrO2 hollow shells and hydrothermal stability for catalytic steam reforming of methane. J. Mater. Chem. A 2016, 4, 153–159. [Google Scholar] [CrossRef]
  8. Breen, J.P.; Meunier, F.C.; Ross, J.R.H. Mechanistic aspects of the steam reforming of methanol over a CuO/ZnO/ZrO2/Al2O3 catalyst. Chem. Commun. 1999, 22, 2247–2248. [Google Scholar] [CrossRef]
  9. Al-Naji, M.; Popova, M.; Chen, Z.; Wilde, N.; Glaser, R. Aqueous-Phase Hydrogenation of Levulinic Acid Using Formic Acid as a Sustainable Reducing Agent Over Pt Catalysts Supported on Mesoporous Zirconia. ACS Sust. Chem. Eng. 2020, 8, 393–402. [Google Scholar] [CrossRef]
  10. Wang, J.; Li, G.; Li, Z.; Tang, C.; Feng, Z.; An, H.; Liu, H.; Liu, T.; Li, C. A highly selective and stable ZnO-ZrO2 solid solution catalyst for CO2 hydrogenation to methanol. Sci. Adv. 2017, 3, e1701290. [Google Scholar] [CrossRef] [Green Version]
  11. Burch, R.; Golunski, S.E.; Spencer, M.S. The role of copper and zinc oxide in methanol synthesis catalysts. J. Chem. Soc. Faraday Trans. 1990, 86, 2683–2691. [Google Scholar] [CrossRef]
  12. Arena, F.; Barbera, K.; Italiano, G.; Bonura, G.; Spadaro, L.; Frusteri, F. Synthesis, characterization and activity pattern of Cu–ZnO/ZrO2 catalysts in the hydrogenation of carbon dioxide to methanol. J. Catal. 2007, 249, 185–194. [Google Scholar] [CrossRef]
  13. Arena, F.; Italiano, G.; Barbera, K.; Bordiga, S.; Bonura, G.; Spadaro, L.; Frusteri, F. Solid-state interactions, adsorption sites and functionality of Cu-ZnO/ZrO2 catalysts in the CO2 hydrogenation to CH3OH. Appl. Catal. A Gen. 2008, 350, 16–23. [Google Scholar] [CrossRef]
  14. Amenomiya, Y. Methanol synthesis from CO2 + H2 II. Copper-based binary and ternary catalysts. Appl. Catal. 1987, 30, 57–68. [Google Scholar] [CrossRef]
  15. Li, K.; Chen, J.G. CO2 Hydrogenation to Methanol over ZrO2-Containing Catalysts: Insights into ZrO2 Induced Synergy. ACS Catal. 2019, 9, 7840–7861. [Google Scholar] [CrossRef]
  16. Tada, S.; Kayamori, S.; Honma, T.; Kamei, H.; Nariyuki, A.; Kon, K.; Toyao, T.; Shimizu, K.-I.; Satokawa, S. Design of Interfacial Sites between Cu and Amorphous ZrO2 Dedicated to CO2-to-Methanol Hydrogenation. ACS Catal. 2018, 8, 7809–7819. [Google Scholar] [CrossRef]
  17. Fisher, I.A.; Bell, A.T. In Situ Infrared Study of Methanol Synthesis from H2/CO over Cu/SiO2and Cu/ZrO2/SiO2. J. Catal. 1998, 178, 153–173. [Google Scholar] [CrossRef]
  18. Rhodes, M.D.; Pokrovski, K.A.; Bell, A.T. The effects of zirconia morphology on methanol synthesis from CO and H2 over Cu/ZrO2 catalysts: Part II. Transient-response infrared studies. J. Catal. 2005, 233, 210–220. [Google Scholar] [CrossRef] [Green Version]
  19. Fujita, S.-I.; Usui, M.; Ohara, E.; Takezawa, N. Methanol synthesis from carbon dioxide at atmospheric pressure over Cu/ZnO catalyst. Role of methoxide species formed on ZnO support. Catal. Lett. 1992, 13, 349–358. [Google Scholar] [CrossRef]
  20. Millar, G.J.; Rochester, C.H.; Waugh, K.C. An in situ high pressure FT-IR study of CO2/H2 interactions with model ZnO/SiO2, Cu/SiO2 and Cu/ZnO/SiO2 methanol synthesis catalysts. Catal. Lett. 1992, 14, 289–295. [Google Scholar] [CrossRef]
  21. Fisher, I.A.; Bell, A.T. In Situ Infrared Study of Methanol Synthesis from H2/CO2 over Cu/SiO2and Cu/ZrO2/SiO2. J. Catal. 1997, 172, 222–237. [Google Scholar] [CrossRef]
  22. Kattel, S.; Yan, B.; Yang, Y.; Chen, J.G.; Liu, P. Optimizing Binding Energies of Key Intermediates for CO2 Hydrogenation to Methanol over Oxide-Supported Copper. J. Am. Chem. Soc. 2016, 138, 12440–12450. [Google Scholar] [CrossRef] [PubMed]
  23. Lam, E.; Corral-Pérez, J.J.; Larmier, K.; Noh, G.; Wolf, P.; Comas-Vives, A.; Urakawa, A.; Copéret, C. CO2 Hydrogenation on Cu/Al2O3: Role of the Metal/Support Interface in Driving Activity and Selectivity of a Bifunctional Catalyst. Angew. Chem. Int. Ed. 2019, 58, 13989–13996. [Google Scholar] [CrossRef] [PubMed]
  24. Larmier, K.; Liao, W.C.; Tada, S.; Lam, E.; Verel, R.; Bansode, A.; Urakawa, A.; Comas–Vives, A.; Copéret, C. CO2-to-Methanol Hydrogenation on Zirconia-Supported Copper Nanoparticles: Reaction Intermediates and the Role of the Metal–Support Interface. Angew. Chem. Int. Ed. 2017, 56, 2318–2323. [Google Scholar] [CrossRef]
  25. Paredes-Nunez, L.A.; Mirodatos, C.; Schuurman, Y.; Meunier, F.C. Determination of formate decomposition rates and relation to product formation during CO hydrogenation over supported cobalt. Catal. Today 2015, 259, 192–196. [Google Scholar]
  26. Chinchen, G.C.; Hay, C.M.; Vandervell, H.D.; Waugh, K.C. The measurement of copper surface areas by reactive frontal chromatography. J. Catal. 1987, 103, 79–86. [Google Scholar] [CrossRef]
  27. Bennici, G.S. Dispersion and surface states of copper catalysts by temperature-programmed-reduction of oxidized surfaces (s-TPR). Appl. Catal. A Gen. 2005, 281, 199–205. [Google Scholar]
  28. Meunier, F.C.; Goguet, A.; Shekhtman, S.; Rooney, D.; Daly, H. A modified commercial DRIFTS cell for kinetically relevant operando studies of heterogeneous catalytic reactions. Appl. Catal. A Gen. 2008, 340, 196–202. [Google Scholar] [CrossRef]
  29. Li, H.; Rivallan, M.; Thibault–Starzyk, F.; Travert, A.; Meunier, F.C. Effective bulk and surface temperatures of the catalyst bed of FT-IR cells used for in situ and operando studies. Phys. Chem. Chem. Phys. 2013, 5, 7321–7327. [Google Scholar] [CrossRef]
  30. Bredy, P.; Farrusseng, D.; Schuurman, Y.; Meunier, F.C. On the link between CO surface coverage and selectivity to CH4 during CO2 hydrogenation over supported cobalt catalysts. J. Catal. 2022, 411, 93–96. [Google Scholar] [CrossRef]
  31. Sirita, J.; Phanichphant, S.; Meunier, F.C. Quantitative Analysis of Adsorbate Concentrations by Diffuse Reflectance FT-IR. Anal. Chem. 2007, 79, 3912–3918. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Jbir, P.I.; Bianchi, D.; Meunier, F.C. Spectrum baseline artefacts and correction of gas-phase species signal during diffuse reflectance FT-IR analyses of catalysts at variable temperatures. Appl. Catal. A Gen. 2015, 495, 17–22. [Google Scholar]
  33. Meunier, F.C. Pitfalls and benefits of in situ and operando diffuse reflectance FT-IR spectroscopy (DRIFTS) applied to catalytic reactions. React. Chem. Eng. 2016, 1, 134–141. [Google Scholar] [CrossRef]
  34. Bensitel, M.; Moravek, V.; Lamotte, J.; Saur, O.; Lavalley, J.C. Infrared study of alcohols adsorption on zirconium oxide: Reactivity of alkoxy species towards CO2. Spectrochim. Acta 1987, 43, 1487–1491. [Google Scholar] [CrossRef]
  35. Busca, G.; Lamotte, J.; Lavalley, J.C.; Lorenzelli, V. FT-IR study of the adsorption and transformation of formaldehyde on oxide surfaces. J. Am. Chem. Soc. 1987, 109, 5197–5202. [Google Scholar] [CrossRef]
  36. Meunier, T.F.C.; Goguet, A.; Reid, D.; Burch, R.; Boaro, M.; Vicario, M.; Trovarelli, A. An investigation of possible mechanisms for the water–gas shift reaction over a ZrO2-supported Pt catalyst. J. Catal. 2006, 244, 183–191. [Google Scholar]
  37. Binet, C.; Daturi, M. Methanol as an IR probe to study the reduction process in ceria–zirconia mixed compounds. Catal. Today 2001, 70, 155–167. [Google Scholar] [CrossRef]
  38. Courtois, D.X.; Perrichon, V.; Bianchi, D. Heats of Adsorption of CO on a Cu/Al2O3 Catalyst Using FTIR Spectroscopy at High Temperatures and under Adsorption Equilibrium Conditions. J. Phys. Chem. B 2000, 104, 6001–6011. [Google Scholar]
Figure 1. Reaction pathways proposed on Cu/ZrO2. (Adapted from reference [22], Kattel et al. J. Am. Chem. Soc. 138 (2016) 12440–12450 with permission, © 2016 American Chemical Society and from reference [24]. Larmier et al., Angew Chem. Int. Ed. Angew. Chem. Int. Ed. 2017, 56, 2318–2323 © 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim). The * symbol indicates an adsorbed species.
Figure 1. Reaction pathways proposed on Cu/ZrO2. (Adapted from reference [22], Kattel et al. J. Am. Chem. Soc. 138 (2016) 12440–12450 with permission, © 2016 American Chemical Society and from reference [24]. Larmier et al., Angew Chem. Int. Ed. Angew. Chem. Int. Ed. 2017, 56, 2318–2323 © 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim). The * symbol indicates an adsorbed species.
Catalysts 12 00793 g001
Figure 2. (A) XRD patterns of the zirconia support and the calcined zirconia-supported Cu. The main peaks of PDF reference files are given underneath. (B) In situ DRIFTS spectrum at 220 °C of the ZrO2-supported Cu sample under 80% H2/He (Red) after reduction at 350 °C and (Black) after reaction and then exposed to H2-only feed for 60 min at 220 °C.
Figure 2. (A) XRD patterns of the zirconia support and the calcined zirconia-supported Cu. The main peaks of PDF reference files are given underneath. (B) In situ DRIFTS spectrum at 220 °C of the ZrO2-supported Cu sample under 80% H2/He (Red) after reduction at 350 °C and (Black) after reaction and then exposed to H2-only feed for 60 min at 220 °C.
Catalysts 12 00793 g002
Figure 3. (A) Operando DRIFTS spectra collected at steady state 220 °C and 3 bar under (black, top) 20% CO2 + 60% H2 and (red, bottom) 25% CO + 75% H2. (B) (Black) same as (A, black, top) and (Red) in situ DRIFTS spectra collected at 220 °C and 3 bar under H2/He after having removed CO2 from the feed for 60 min. Inset: region used to integrate the formate signal.
Figure 3. (A) Operando DRIFTS spectra collected at steady state 220 °C and 3 bar under (black, top) 20% CO2 + 60% H2 and (red, bottom) 25% CO + 75% H2. (B) (Black) same as (A, black, top) and (Red) in situ DRIFTS spectra collected at 220 °C and 3 bar under H2/He after having removed CO2 from the feed for 60 min. Inset: region used to integrate the formate signal.
Catalysts 12 00793 g003
Figure 4. (A) Evolution of DRIFTS band area recorded over Cu/ZrO2 at 220 °C as a function of time under 85% H2 in He. The time range 1–60 min is presented. The total signal (open square) could be decomposed into a slow linear contribution (red dotted line) and a faster decay (green circles). Total flow = 100 mL min−1. (B) Semi-logarithmic plot of the normalized signal associated with the faster decay showing a linear behavior.
Figure 4. (A) Evolution of DRIFTS band area recorded over Cu/ZrO2 at 220 °C as a function of time under 85% H2 in He. The time range 1–60 min is presented. The total signal (open square) could be decomposed into a slow linear contribution (red dotted line) and a faster decay (green circles). Total flow = 100 mL min−1. (B) Semi-logarithmic plot of the normalized signal associated with the faster decay showing a linear behavior.
Catalysts 12 00793 g004
Figure 5. Operando DRIFTS spectra collected (Red) at steady state under CO2 + H2 at 220 °C and (Black) difference between the spectra collected after 2 and 20 min after having removed CO2 from the feed.
Figure 5. Operando DRIFTS spectra collected (Red) at steady state under CO2 + H2 at 220 °C and (Black) difference between the spectra collected after 2 and 20 min after having removed CO2 from the feed.
Catalysts 12 00793 g005
Figure 6. Variation of the methoxy DRIFTS band area at 1150 cm−1 as a function of that of the formate band following the removal of CO2. Feed: 100 mL min−1 of 85% H2/He.
Figure 6. Variation of the methoxy DRIFTS band area at 1150 cm−1 as a function of that of the formate band following the removal of CO2. Feed: 100 mL min−1 of 85% H2/He.
Catalysts 12 00793 g006
Figure 7. Schematic representation of the evolution of the ZrO2-bound formates under H2. See text for more details.
Figure 7. Schematic representation of the evolution of the ZrO2-bound formates under H2. See text for more details.
Catalysts 12 00793 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Meunier, F.C.; Dansette, I.; Eng, K.; Schuurman, Y. Differentiating the Reactivity of ZrO2-Bound Formates Formed on Cu/ZrO2 during CO2 Hydrogenation. Catalysts 2022, 12, 793. https://doi.org/10.3390/catal12070793

AMA Style

Meunier FC, Dansette I, Eng K, Schuurman Y. Differentiating the Reactivity of ZrO2-Bound Formates Formed on Cu/ZrO2 during CO2 Hydrogenation. Catalysts. 2022; 12(7):793. https://doi.org/10.3390/catal12070793

Chicago/Turabian Style

Meunier, Frederic C., Isaac Dansette, Kimleang Eng, and Yves Schuurman. 2022. "Differentiating the Reactivity of ZrO2-Bound Formates Formed on Cu/ZrO2 during CO2 Hydrogenation" Catalysts 12, no. 7: 793. https://doi.org/10.3390/catal12070793

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