Next Article in Journal
Use of Radioisotopes to Produce High Yielding Crops in Order to Increase Agricultural Production
Previous Article in Journal
Synthesis of 2-aminopyridine Lactones and Studies of Their Antioxidant, Antibacterial and Antifungal Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Proceeding Paper

Comparative Investigation of (10%Co+0.5%Pd)/TiO2(Al2O3) Catalysts in CO Hydrogenation at Low and High Pressure †

1
Institute of Catalysis, Bulgarian Academy of Sciences, 1113 Sofia, Bulgaria
2
Inorganic Chemistry Department, Materials Science Institute (CSIC-US), 41092 Seville, Spain
3
Institute of Geotechnics, Slovak Academy of Sciences, 04001 Kosice, Slovakia
*
Author to whom correspondence should be addressed.
Presented at the 2nd International Electronic Conference on Catalysis Sciences—A Celebration of Catalysts 10th Anniversary, 15–30 October 2021; Available online: https://eccs2021.sciforum.net/.
Chem. Proc. 2022, 6(1), 11; https://doi.org/10.3390/ECCS2021-11105
Published: 14 October 2021

Abstract

:
Surface properties of prereduced (Co+Pd)/Al2O3 and (Co+Pd)/TiO2 catalysts were studied. Metal dispersion was 1–3%. CoPdA demonstrated high temperature H2 desorption and firmly held CO and carbonate species on the surface. SMSI operated on CoPdT even after contact with H2O and air. Metal surface reconstruction and increased formation of CH2 groups occurred during catalysis. At low pressure, CoPdT was more active, whereas CoPdA had higher CH4 selectivity. At high pressure, catalysis on CoPdA revealed dependence on Tred, synthesis of C2+ hydrocarbons, decreased CO2 production, and higher CH4/CO2 ratio. CO conversion decreased with time due to difficulties in the surface diffusion of reagents, intermediates, and products, and metal particle agglomeration.

1. Introduction

CO hydrogenation in synthesis gas is an environmentally friendly process which offers an alternative to oil refinement [1]. The products obtained have low or no nitrogen and sulphur [1,2] content. The main products of the CO hydrogenation process are CH4, CO2, and light and heavy hydrocarbons. CH4 and CO2 are considered unwanted products [3,4].
Co-Pd catalysts are active in the process of CO hydrogenation [5,6,7,8]. Many factors affect their activity and selectivity [1,2,9,10,11]. Product distribution is influenced directly by some process parameters, but others affect it indirectly through their effect on the conversion [10]. The CO hydrogenation reaction is thermodynamically favoured by increasing pressure. Since the reaction mechanism is very complex, involving separate hydrogenation and polymerization routes [12], investigations at low and high pressure were carried out to obtain detailed information about the reasons for and ways of product distribution change. Generally, the effect of increased pressure results in enhanced CO conversion, a decrease in light hydrocarbon synthesis, and an increase in the C5+ compound quota [2,6,9,10,11,12,13,14,15,16,17,18]. Comparative analyses at different pressures in intervals of 0.33–40 atm have been done [2,6,9,10,12,13,14,15,16,17,18]. A 10 atm pressure has been accepted as an optimum pressure as a result of investigations on combined influence of process parameters as pressure, temperature, H2/CO ratio, flow rate, and conversion [12,14].
The present study discusses the surface properties of alumina- and titania-supported (10%Co+0.5%Pd) catalysts and selectivity in CO hydrogenation at low and high pressure. The aim is a better understanding of the specific role of the support.

2. Materials and Methods

Bimetallic catalysts with 10%Co and 0.5%Pd were prepared by deposition of metal nitrate salts from aqueous solution on non-porous TiO2 and Al2O3 supports. The precursors were reduced in H2 flow, applying two modes: (i) for studies at P = 1 atm—heating at 100 and 200 °C for 1 h, and 2 h at 300 °C; (ii) for studies at P = 10 atm—heating at 260 or 400 °C for 15 h. Catalytic activity measurements were carried out at P = 1 atm and Treac = 150–365 °C or P = 10 atm and 260 °C. Samples of both groups of catalysts were studied using a number of methods. Chemisorption of H2 (at 100 °C) and CO (at 25 °C) was measured by volumetric method [19,20,21]. Irreversibly adsorbed CO was determined as a difference between total and reversible adsorption [22]. Particle size distribution was derived by photon cross-correlation spectroscopy (PCCS). Electron paramagnetic resonance (EPR) spectra were recorded at 25 °C in X-band. X-ray photoelectron spectra (XPS) were recorded using AlKα X-ray source. The spectra were processed according to Refs. [23,24] Temperature-programmed desorption (TPD) of H2 (Tads = 100 °C) and CO (Tads = 25 and 200 °C) was studied using a differential scanning calorimeter. CO hydrogenation at 1 atm was studied in situ by diffuse-reflectance infrared spectroscopy (DRIFTS) in a high temperature vacuum chamber. PCCS, EPR, XPS, TPD, and DRIFTS measurements were made with catalyst samples after a CO hydrogenation test at 1 atm.

3. Results and Discussion

Two main reactions are running in our investigations at the chosen reaction conditions: (i) CO + 3H2 = CH4 + H2O; and (ii) CO + H2O = CO2 + H2 [25]. Tred and Treac effect on the catalytic behaviour of the synthesized materials was examined at 1 atm in temperature intervals of 300–450 °C and 150–375 °C, respectively. A Tred over 300 °C led to a decrease in CO conversion and CH4 and CO2 yields. A definitely sharp decrease was found in the case of Al2O3-supported system. The dependences observed were due to metal particle agglomeration. The increase in Treac resulted in an increase in CO conversion and CH4 and CO2 yields [6] and a decrease in the CH4/CO2 selectivity ratio. Values within 2–19 of the CH4/CO2 ratio were registered on TiO2-supported samples in the interval Treac = 285–335 °C and 3–24 for Al2O3-supported samples in the range 315–365 °C. These results showed that water gas-shift reaction (WGSR) was favoured to a significant extent by the temperature. (Co+Pd)/TiO2 samples were more active and the highest activity exhibited that one reduced at 300 °C (CoPdT). (Co+Pd)/Al2O3 samples had lower activity but demonstrated higher selectivity to CH4 if particularly reduced at 400 °C (CoPdA).
The properties of two samples were compared, namely the most active CoPdT one reduced at 300 °C and the most selective CoPdA entity reduced at 400 °C. At the initial stage of the catalytic test, H/CO ratio values of 2.9 and 2.8, respectively, were determined. Metal dispersion was calculated based on H2 chemisorption and it was estimated to be very low: 3.61 and 1% for CoPdT and CoPdA, respectively. The data showed presence of large metal particles and a very small part of the supported metal was accessible to contact with reagents. The reasons for this can be found in both the low BET area of the titania, as well as to the processes which take place during the decomposition of the cobalt nitrate in a reductive medium, followed by metal particle agglomeration facilitated by the subsequent reduction at 400 °C. In the case of CoPdA, the surface Pd atoms are highly diluted in the bimetallic particles.
In order to reveal why titania-supported catalysts were more active and those on alumina produced more CH4, samples of both were studied after the catalytic tests. We consider that samples selected after catalysis would manifest surface properties that do not change, or change insignificantly, during subsequent studies.
A different number of peaks corresponding to particle hydrodynamic radii and representing particle size distribution in both materials characterised correlation functions obtained by PCCS. Values of the hydrodynamic radii showed that 100% of the CoPdT catalyst particles were of 40–120-nm. For the CoPdA catalyst, the result was indicative about bimodal particle size distribution of 90–102 nm (58%) and 7.5–10 μm (42%). Thus, in both cases the catalyst particles were found to be agglomerates.
EPR registered spectra with a g factor of 2.255 ± 0.005, which was representative of tetrahedrally coordinated Co2+ ions [26]. Concerning their amount, it was found that CoPdT > CoPdA is valid. Perhaps, the pretreatment mode and Al2O3 support gave rise to the large amount of cobalt in the diamagnetic state (Co, CoPd alloy particles and/or Co3+) after reduction at 400 °C. XPS also revealed Co3+, metallic Pd, and Pd2+. Most probably ion presence was due to the ex-situ measurements, where oxygen adsorption and oxidation of the particle surface layer proceeded without penetration into the bulk while being exposed to air [27,28], and/or owing to the presence of unreduced phases. The (Co+Pd)/support ratio of CoPdA catalyst was lower than that for CoPdT. This peculiarity was attributed to a lower TiO2 surface area, which presumed that all the metal was on the carrier grains but not in the bulk. The EPR and XPS data could be attributed to the higher extent of metal particle agglomeration and alloying in the CoPdA sample. On the surface of this sample, smaller amounts of carbon were registered. The deconvolution of the C1s peaks revealed that about 20% and 50% carbon with CoPdT and CoPdA, respectively, on the surface was in the form of carbonates, indicating that alumina exposed stronger adsorption sites [29]. The deconvolution of the O1s spectra of CoPdT showed a composition of three sub-peaks [30] and Ti/O > 0.5, which is below the stoichiometry and presupposes oxygen deficit (TiO2−x). Thus, strong metal-support interaction (SMSI) has been invoked to occur during sample reduction, which is preserved after catalytic runs and exposure to air.
In situ DRIFTS studies of CO hydrogenation were performed at Treac = 50–250 °C. Registered bands, band maxima, and shoulders of the adsorbed species were ascribed as follows: 1767 cm−1–CO multiple (bridge) bonded on Co0; 1864 cm−1–CO multiple bonded on Pd0; 1934–90 cm−1–CO bridge bonded on Pd0 and/or linear bonded on Co0; 2000–36 cm−1–CO linear bonded on Co0; 2040–60 cm−1–CO linear bonded on Coδ+; 2050/60 cm−1–hydrocarbonyl (H-Co-CO); and 2073–100 cm−1–CO linear bonded on Pd0 [31,32,33,34,35,36]. The registration of many bands due to one type of adsorbed CO species was attributed to the existence of various sites on the metal particles with differences in nature, coordination of atoms, electronic state, and bond energy with adsorbates. A comparative analysis of the spectra during the reaction showed the modification of CO adsorption forms/sites with temperature change [15]. The spectral changes indicated: (i) facile CO interaction with hydrogen even at room temperature, which mostly concerned linear species; (ii) formation of new sites on CoPdT due to SMSI destruction/reduction because of synthesized H2O reaction product [37] and because of partial oxidation of surface Co atoms by water molecules and/or reduction of residual oxide phases with the formation of new adsorption centres in close contact with the support (CoPdT, CoPdA) [38,39]; and (iii) cobalt hydrocarbonyl species could contribute to some band broadening and intensity increase. When the Treac was around 250 °C, the bands in the CoPdT spectra decreased significantly in intensity to (almost) complete disappearance, particularly those of bridge-bonded CO on Pd, thus displaying acceleration of the hydrogenation process [34] as a result of the predominant dissociative adsorption of both CO and H2 favouring formation of CHx intermediates and CH4 [5]. The spectra of CoPdA demonstrated significant stability of the CO adsorption on the surface via irreversibly adsorbed CO that is considered a precondition which further favours the hydrogenation process to methane [34,38].
Bands characteristic of hydrocarbon species were distinguished by vibrations of a CH3 group at 1370–93, 1420–52, and 2957 cm−1, a CH2 group at 1461, 2851, and 2926 cm−1, and a CH group at 1356 cm−1 (Figure 1a,b) [40,41]. With all bands, an increase of intensity was registered at Treac ≥ 175 °C for CoPdA and 150 °C for CoPdT samples. This mostly concerned the band of CH2 groups at 2926 cm−1. With CoPdT, the ICH2/ICH3 intensity ratio was 1.2–2 in the range 175–250 °C, which showed the availability of sites for steady adsorption of CHx intermediates [42,43]. Synthesized CH4 (1303–1305, 3013–3016, 3099 cm−1) [40,41] became perceptible at 200 and 225 °C in the spectra of CoPdT and CoPdA, respectively. Band intensities increased with Treac much more with the CoPdT sample in the range 225–250 °C, where a 3.7-fold increase in the 3015-cm−1 band intensity was detected. This particularity shows that both catalysts have potential to produce higher hydrocarbons than CH4, even at 1 atm.
Bidentate (1245–1268, 1566–1580, 1616–1618, 1640 cm−1) and monodentate (1320–1375, 1472, 1521–1522 cm−1) carbonate and formate species (1341–1390, 1566–1580, 1616–1618 cm−1) were registered on the surface of catalysts [41]. Bidentate carbonates are classified as weakly held carbonates [44], whereas desorption of formate and monodentate carbonate species takes place at higher temperatures. It can be supposed that the release of small amounts of CO2 registered with CoPdA at Treac ≥ 125 °C was due to desorption of bidentate carbonates. Thus, the results suppose the existence of sites for strong adsorption of formate and carbonate species on the surface. The formation of CO2 was suppressed and the catalyst became more selective to CH4. In the case of CoPdT the increase of Treac over 150 °C facilitated the transformation of carbonate(-like) intermediate species to increase the amount of CO2 in the gaseous phase. The catalyst manifested high activity but poor selectivity.
TPD studies showed two regions of desorption of H2 and CO. From the CoPdT sample, the main amount of H2 was desorbed at low temperatures, while the situation was opposite with CoPdA. High temperature desorption of H2 (T > 360 °C) from CoPdA displayed the presence of low reactive and low mobile adsorption form, which could not participate in the formation of formates and CH4. The presence of such H2 species can cause diminished contact between the CO and catalytically active sites, a lower number of adsorbed CO species, and increased H/CO ratio on sites generating CH4 formation. On comparing CO species adsorbed at 25 °C and at 200 °C by TPD, we observed that with CoPdT the amount of CO species desorbed at a low temperature (below ~200 °C, two types) increased with Tads, at the expense of those desorbed at a high temperature. Considering CoPdA, the CO species desorbing at a low temperature changed only their share upon Tads increase. High temperature species represented a significant part of the adsorbed CO on CoPdT at both used Tads. Based on DRIFTS results, the CO desorbed at a low temperature occurs from linear and bridge-bonded species, whereas CO desorbed at a high temperature originates from carbonate(-like) species. Depending on Tads and support, the linear CO species prevailed over the bridge forms to a different extent. All desorption peaks independent of adsorbed gas and Tads fell into common temperature intervals that supposes surface species of close bonding energies and a similar structure of adsorption sites. This facilitates interaction between the species to form hydrocarbon(s) and CO2, the latter being result of carbonate(-like) intermediate decomposition. Comparative analyses of CO25C-TPD and CO200C-TPD together with DRIFTS data supposed facilitated partial destruction of carbonate species, most probably those of bidentate type. Since the desorption from CoPdA catalyst in the high temperature region was negligible compared to that from CoPdT, it could be assumed that the surface possesses a peculiar set of homogeneity of intermediates that contributes to a higher selectivity of the CoPdA catalyst.
Catalytic tests at 10 atm showed that the CO conversion and selectivity depended on Tred (Figure 2). Methane was still the main hydrocarbon product, and the CH4/CO2 selectivity ratio was CoPdA > CoPdT, taking into account that the CO2 amount produced by the CoPdT catalyst was 1.5–1.6 times higher. The CO conversion over CoPdT decreased faster with time on stream (1.4 times) after both Tred (260 and 400 °C), in contrast to CoPdA, which was obviously more stable, especially after reduction at 400 °C.
Selectivity was generally preserved with some changes after Tred = 400 °C: C5+ and CO2 amounts decreased, while the share of CH4 and C2-C4 compounds slightly increased. The decrease in CO conversion was attributed to the higher carbon deposition on the surface; difficult diffusion of reagents, intermediates, and products due to synthesis of C5+ hydrocarbons [6,18]; consumption of reagents in metal oxide reduction; and metal particle agglomeration. Reduction at 400 °C supposed a higher extent of metal reduction favouring a decrease in CO2 formation through WGSR, since the number of active sites for this case as Con+ decreased. CoPdT preserved higher CO2 production most probably because of better dispersion and SMSI providing active sites. The agglomeration renders different effects on the conversion, depending on carrier material. CO conversion over CoPdA was more stable due to improved resistance of the metal particles to agglomeration. In the case of CoPdT, the agglomeration and SMSI (see XPS data) led to diminished CO conversion by lowering dispersion and the number of active sites. By increasing metal particle size, the selectivity of both catalysts to different hydrocarbons is changed, because the formation and homogenization of the bimetallic particle surface is facilitated. Modified properties of the surface cobalt atoms and larger particle size favoured CO dissociation over cobalt [15,36]. Together with the effect of increased pressure, the probability for polymerization and carbon chain growth was augmented. However, increased PH2 enriched the surface in CHx species [1,9,11] and helped to divide bigger intermediates into smaller fragments and thus decreased chain growth to some extent in situation of slightly enhanced CO dissociation [9]. In the case of CoPdA catalyst, the increased PH2 could contribute to stable CO conversion, avoiding deactivation of metal by oxidation during the process [15].

4. Conclusions

Co-Pd catalysts with alumina support pretreated in H2 were more selective toward methane and produced a smaller amount of CO2 during CO hydrogenation at 1 atm compared to catalysts with titania. The properties were preserved in the reaction at 10 atm due to the strong metal-support interaction in CoPdT, which was not reduced by contact with H2O and air, thus determining sites for carbonate species formation and CO2 production. CoPdT produced a smaller C5+ amount because of stable adsorption of CHx intermediates.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ECCS2021-11105/s1.

Author Contributions

Conceptualisation, writing—original draft preparation, visualisation, project administration, funding acquisition, M.S.; methodology, M.S., S.T., A.C. and G.K.; validation, M.S., S.T., A.C., H.K., M.F., K.A. and K.T.; investigation, M.S., H.K., M.F., K.A., K.T., S.T., A.C. and G.K.; resources, M.S., S.T. and A.C.; data curation, M.S., S.T., A.C., H.K., M.F., K.A. and K.T.; writing—review and editing, G.K.; supervision, S.T., A.C. and G.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Bulgarian National Science Fund, contract number KP-06-H29-9/2018.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

IC, BAS, Sofia, Bulgaria; ICD, MSI (CSIC-US), Seville, Spain; IG, SAS, Kosice, Slovakia.

Acknowledgments

English language editing by Ch. Bonev is appreciated.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, Q.; Kang, J.; Wang, Y. Development of novel catalysts for Fischer–Tropsch synthesis: Tuning the product selectivity. ChemCatChem 2010, 2, 1030–1058. [Google Scholar] [CrossRef]
  2. Riyahin, M.; Atashi, H.; Mohebbi-Kalhori, D. Optimization of reaction condition on the product selectivity of Fischer-Tropsch synthesis over a Co-SiO2/SiC catalyst using a fixed bed reaction. Petroleum Sci. Technol. 2017, 35, 1078–1084. [Google Scholar] [CrossRef]
  3. Schmidt-Rohr, K. Why combustions are always exothermic, yielding about 418 kJ per mole of O2. J. Chem. Educ. 2015, 92, 2094–2099. [Google Scholar] [CrossRef] [Green Version]
  4. Murdoch, A. Structural and Compositional Analysis of Co-Pd Model Catalyst Surfaces. Ph.D. Thesis, University of St. Andrews, St. Andrews, UK, 2012. [Google Scholar]
  5. Davis, B.H.; Iglesia, E. Technology Development for Iron and Cobalt Fischer-Tropsch Catalysts, Final Technical Report DE-FC26-98FT40308; University of California: Berkeley, CA, USA; University of Kentucky Research Foundation: Lexington, KY, USA, 2002. [Google Scholar]
  6. Arsalanfar, M.; Mirzaei, A.A.; Bozorgzadeh, H.R.; Samimi, A. A review of Fischer–Tropsch synthesis on the cobalt based catalysts. Phys. Chem. Res. 2014, 2, 179–201. [Google Scholar] [CrossRef]
  7. Rabo, J.A.; Risch, A.P.; Poutsma, M.L. Reactions of carbon monoxide and hydrogen on Co, Ni, Ru, and Pd metals. J. Catal. 1978, 53, 295–311. [Google Scholar] [CrossRef]
  8. Dinse, A.; Aigner, M.; Ulbrich, M.; Johnson, G.R.; Bell, A.T. Effects of Mn promotion on the activity and selectivity of Co/SiO2 for Fischer-Tropsch synthesis. J. Catal. 2012, 288, 104–114. [Google Scholar] [CrossRef]
  9. Sari, A.; Zamani, Y.; Taheri, S.A. Intrinsic kinetics of Fischer-Tropsch reactions over an industrial Co-Ru/γ-Al2O3 catalyst in slurry phase reactor. Fuel Proc. Technol. 2009, 90, 1305–1313. [Google Scholar] [CrossRef]
  10. Gibson, E.J.; Hall, C.C. The Fischer-Tropsch synthesis with cobalt catalysts: The effect of process conditions on the composition of the reaction products. J. Appl. Chem. 1954, 4, 49–61. [Google Scholar] [CrossRef]
  11. Zhou, W.; Chen, J.-G.; Fang, K.-G.; Sun, Y.-H. The deactivation of Co/SiO2 catalyst for Fischer-Tropsch synthesis at different ratios of H2 to CO. Fuel Proc. Thechnol. 2006, 87, 609–616. [Google Scholar] [CrossRef]
  12. Mirzaei, A.A.; Shirzadi, B.; Atashi, H.; Mansouri, M. Modeling and operating conditions optimization of Fischer-Tropsch synthesis in a fixed-bed reactor. J. Ind. Eng. Chem. 2012, 18, 1515–1521. [Google Scholar] [CrossRef]
  13. Kwack, S.-H.; Park, M.-J.; Bae, J.W.; Ha, K.-S.; Jun, K.-W. Development of a kinetic model of the Fischer-Tropsch synthesis reaction with a cobalt-based catalyst. React. Kinet. Mech. Catal. 2011, 104, 483–502. [Google Scholar] [CrossRef]
  14. Jalama, K. Chapter 39—Effect of operating pressure on Fischer-Tropsch synthesis kinetics over titania-supported cobalt catalyst. In Proceedings of the Transactions on Engineering Technologies, World Congress on Engineering and Computer Science, San Francisco, CA, USA, 21–23 October 2015; Ao, S.-I., Kim, H.K., Amouzegar, M.A., Eds.; Springer: Singapore, 2017; pp. 555–562. [Google Scholar] [CrossRef]
  15. de la Pena O’Shea, V.A.; Alvarez-Galavan, M.C.; Campos-Martin, J.M.; Fierro, J.L.G. Strong dependence on pressure of the performance of a Co/SiO2 catalyst in Fischer-Tropsch slurry reactor synthesis. Catal. Lett. 2005, 100, 105–110. [Google Scholar] [CrossRef]
  16. Geerlings, J.J.C.; Wilson, J.H.; Kramer, G.J.; Kuipers, H.P.C.E.; Hoek, A.; Huisman, H.M. Fischer-Tropsch technology—From active site to commercial process. Appl. Catal. A Gen. 1999, 186, 27–40. [Google Scholar] [CrossRef]
  17. Yates, I.C.; Sattarfield, C.N. Hydrocarbon selectivity from cobalt Fischer-Tropsch catalysts. Energy Fuels 1992, 6, 308–314. [Google Scholar] [CrossRef]
  18. Yan, Z.; Wang, Z.; Bukur, D.B.; Goodman, D.W. Fischer-Tropsch synthesis on a model Co/SiO2 catalyst. J. Catal. 2009, 268, 196–200. [Google Scholar] [CrossRef]
  19. Aben, P.C. Palladium areas in supported catalysts: Determination of palladium surface areas in supported catalysts by means of hydrogen chemisorption. J. Catal. 1968, 10, 224. [Google Scholar] [CrossRef]
  20. Reuel, R.C.; Bartholomew, C.H. The stoichiometries of H2 and CO adsorptions on cobalt: Effects of support and preparation. J. Catal. 1984, 85, 63. [Google Scholar] [CrossRef]
  21. Zowtiak, J.M.; Bartholomew, C.H. The kinetics of H2 adsorption on and desorption from cobalt and the effects of support thereon. J. Catal. 1983, 83, 107. [Google Scholar] [CrossRef]
  22. Anderson, J.R. Structure of Metallic Catalysts; Mir: Moscow, Russia, 1978. (In Russian) [Google Scholar]
  23. Shirley, D.A. High-resolution X-ray photoemission spectrum of the valence bands of gold. Phys. Rev. B 1972, 5, 4709–4714. [Google Scholar] [CrossRef] [Green Version]
  24. Scofield, J.H. Hertree-Slater subshell photoionization cross-sections at 1254 and 1487 eV. J. Electron Spectrosc. Relat. Phenom. 1976, 8, 129–137. [Google Scholar] [CrossRef]
  25. Zheng, S.; Liu, Y.; Li, J.; Shi, B. Deuterium tracer study of pressure effect on product distribution in the cobalt-catalyzed Fischer-Tropsch synthesis. Appl. Catal. A Gen. 2007, 330, 63–68. [Google Scholar] [CrossRef]
  26. Guskos, N.; Typek, J.; Maryniak, M.; Żolnierkiewicz, G.; Podsiadly, M.; Arabczyk, W.; Lendzion-Bielun, Z.; Narkiewicz, U. Effect of calcination and structural additives on the EPR spectra of nanocrystalline cobalt oxides. Mater. Sci.-Pol. 2006, 24, 4. [Google Scholar]
  27. Popova, N.M.; Babenkova, L.V.; Savel’eva, G.A. Adsorption and Interaction of the Simplest Gases with Metals from VIII Group; Nauka: Alma-Ata, Kazakhstan, 1979. (In Russian) [Google Scholar]
  28. Potoczna-Petru, D.; Jablonski, J.M.; Okal, J.; Krajczyk, L. Influence of oxidation-reduction treatment on the microstructure of Co/SiO2 catalyst. Appl. Catal. A Gen. 1998, 175, 113–120. [Google Scholar] [CrossRef]
  29. Singh, B.; Patial, J.; Sharma, P.; Chandra, S.; Maity, S.; Lingaiah, N. A comparative study on basicity based on supported K-salt catalysts for isomerization of 1-methoxy-4-(2-propene-1-yl) benzene. Indian J. Chem. Technol. 2010, 17, 446–450. [Google Scholar]
  30. Pan, X.; Yang, M.-Q.; Fu, X.; Zhang, N.; Xu, Y.-J. Defective TiO2 with oxygen vacancies: Synthesis, properties and photocatalytic applications. Nanoscale 2013, 5, 3601–3614. [Google Scholar] [CrossRef] [PubMed]
  31. Zhang, J.; Chen, J.; Ren, J.; Sun, Y. Chemical treatment of γ-Al2O3 and its influence on the properties of Co-based catalysts for Fischer–Tropsch synthesis. Appl. Catal. A 2003, 243, 121. [Google Scholar] [CrossRef]
  32. Tsubaki, N.; Sun, S.; Fujimoto, K. Different functions of the noble metals added to cobalt catalysts for Fischer–Tropsch synthesis. J. Catal. 2001, 199, 236. [Google Scholar] [CrossRef]
  33. Kadinov, G.; Bonev, C.; Todorova, S.; Palazov, A. IR spectroscopy study of CO adsorption and of the interaction between CO and hydrogen on alumina supported cobalt. J. Chem. Soc. Faraday Trans. 1998, 94, 3027–3031. [Google Scholar] [CrossRef]
  34. Singh, J.A.; Yang, N.; Liu, X.; Tsai, C.; Stone, K.H.; Johnson, B.; Koh, A.L.; Bent, S.F. Understanding the active sites of CO hydrogenation on Pt-Co catalysts prepared using atomic layer deposition. J. Phys. Chem. C 2018, 122, 2184–2194. [Google Scholar] [CrossRef]
  35. Xiong, H.; Zhang, Y.; Liew, K.; Li, J. Ruthenium promotion of Co/SBA-15 catalysts with high cobalt loading for Fischer-Tropsch synthesis. Fuel Proc. Technol. 2009, 90, 237–246. [Google Scholar] [CrossRef]
  36. Tuxen, A.; Carenco, S.; Chintapalli, M.; Chuang, C.-H.; Escudero, C.; Pach, E.; Jiang, P.; Borondics, F.; Beberwyck, B.; Alivisatos, A.P.; et al. Size-dependent dissociation of carbon monoxide on cobalt nanoparticles. J. Am. Chem. Soc. 2013, 135, 2273–2278. [Google Scholar] [CrossRef]
  37. Li, J.; Xu, L.; Keogh, R.A.; Davis, B.H. Effect of boron and ruthenium on the catalytic properties of Co/TiO2 Fischer-Tropsch catalysts. Prepr.-Am. Chem. Society. Div. Pet. Chem. 2000, 45, 253. [Google Scholar]
  38. Sun, X.; Sartipi, S.; Kapteijn, F.; Gascon, J. Effect of pretreatment atmosphere on the activity and selectivity of Co/mesoHZSM-5 for Fischer–Tropsch synthesis. New J. Chem. 2016, 40, 4167–4177. [Google Scholar] [CrossRef]
  39. Chen, T.-Y.; Su, J.; Zhang, Z.; Cao, C.; Wang, X.; Si, R.; Liu, X.; Shi, B.; Xu, J.; Han, Y.-F. Structure evolution of Co-CoOx interface for higher alcohol synthesis from syngas over Co/CeO2 catalysts. ACS Catal. 2018, 8, 8606–8617. [Google Scholar] [CrossRef]
  40. Sanchez-Escribano, V.; Larrubio-Vargas, M.A.; Finocchio, E.; Busca, G. On the mechanisms and the selectivity determining steps in syngas conversion over supported metal catalysts: An IR study. Appl. Catal. A Gen. 2007, 316, 68–74. [Google Scholar] [CrossRef]
  41. Little, L.H. Infrared Spectra of Adsorbed Species; Academic Press Inc.: London, UK; New York, NY, USA, 1966. [Google Scholar]
  42. Geerlings, J.J.C.; Zonnevylle, M.C.; de Groot, C.P.M. Studies of the Fischer-Tropsch reaction on Co(0001). Surf. Sci. 1991, 241, 302–314. [Google Scholar] [CrossRef]
  43. Geerlings, J.J.C.; Zonnevylle, M.C.; de Groot, C.P.M. Structure sensitivity of the Fischer-Tropsch rection on cobalt single crystals. Surf. Sci. 1991, 241, 315–324. [Google Scholar] [CrossRef]
  44. Szanyi, J.; Kwak, J.H. Dissecting the steps of CO2 reduction: 1. The interaction of CO and CO2 with γ-Al2O3: An in situ FTIR study. Phys. Chem. Chem. Phys. 2014, 16, 15117–15125. [Google Scholar] [CrossRef]
Figure 1. In situ DRIFTS spectra recorded at different temperatures during CO hydrogenation at 1 atm in the presence of: (a) CoPdA sample; (b) CoPdT sample.
Figure 1. In situ DRIFTS spectra recorded at different temperatures during CO hydrogenation at 1 atm in the presence of: (a) CoPdA sample; (b) CoPdT sample.
Chemproc 06 00011 g001
Figure 2. CO conversion at P = 10 atm over the studied samples after reduction at different temperatures. Inset: Product distribution.
Figure 2. CO conversion at P = 10 atm over the studied samples after reduction at different temperatures. Inset: Product distribution.
Chemproc 06 00011 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shopska, M.; Caballero, A.; Todorova, S.; Aleksieva, K.; Tenchev, K.; Kolev, H.; Fabian, M.; Kadinov, G. Comparative Investigation of (10%Co+0.5%Pd)/TiO2(Al2O3) Catalysts in CO Hydrogenation at Low and High Pressure. Chem. Proc. 2022, 6, 11. https://doi.org/10.3390/ECCS2021-11105

AMA Style

Shopska M, Caballero A, Todorova S, Aleksieva K, Tenchev K, Kolev H, Fabian M, Kadinov G. Comparative Investigation of (10%Co+0.5%Pd)/TiO2(Al2O3) Catalysts in CO Hydrogenation at Low and High Pressure. Chemistry Proceedings. 2022; 6(1):11. https://doi.org/10.3390/ECCS2021-11105

Chicago/Turabian Style

Shopska, Maya, Alfonso Caballero, Silviya Todorova, Katerina Aleksieva, Krassimir Tenchev, Hristo Kolev, Martin Fabian, and Georgi Kadinov. 2022. "Comparative Investigation of (10%Co+0.5%Pd)/TiO2(Al2O3) Catalysts in CO Hydrogenation at Low and High Pressure" Chemistry Proceedings 6, no. 1: 11. https://doi.org/10.3390/ECCS2021-11105

Article Metrics

Back to TopTop