Next Article in Journal
Effects of Cu Species on Liquid-Phase Partial Oxidation of Methane with H2O2 over Cu-Fe/ZSM-5 Catalysts
Next Article in Special Issue
Methane-Assisted Iron Oxides Chemical Looping in a Solar Concentrator: A Real Case Study
Previous Article in Journal
Single-Atom Co-Catalysts Employed in Titanium Dioxide Photocatalysis
Previous Article in Special Issue
Efficacy of the Immobilized Kocuria flava Lipase on Fe3O4/Cellulose Nanocomposite for Biodiesel Production from Cooking Oil Wastes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Low-Pt-Based Sn Alloy for the Dehydrogenation of Methylcyclohexane to Toluene: A Density Functional Theory Study

by
Kingsley Onyebuchi Obodo
*,
Cecil Naphtaly Moro Ouma
and
Dmitri Bessarabov
HySA-Infrastructure, Faculty of Engineering, North-West University, Private Bag X6001, Potchefstroom 2520, South Africa
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(10), 1221; https://doi.org/10.3390/catal12101221
Submission received: 12 September 2022 / Revised: 4 October 2022 / Accepted: 10 October 2022 / Published: 12 October 2022
(This article belongs to the Special Issue Advanced Catalysis for Green Fuel Synthesis and Energy Conversion)

Abstract

:
Spin-polarized van der Waals corrected density functional theory calculations were applied to Sn–Pt alloys with Pt content ≤ 50% (referred to as low Pt alloys) to evaluate their catalytic activity towards the dehydrogenation of methylcyclohexane (MCH), with the formation of toluene as product. The calculated adsorption energies of MCH, its intermediates and toluene showed that these molecules bind on the considered Sn–Pt alloys. Sn–Pt alloys had the lowest dehydrogenation energetics, indicating that the activity of this catalytic material is superior to that of a pristine Pt catalyst. Desorption of the intermediate species was feasible for all Sn–Pt alloy configurations considered. The catalytic dehydrogenation reaction energetics for the various Sn–Pt alloy configurations were more favourable than that achieved with pristine Pt surfaces. The current study should motivate experimental realization of Sn–Pt alloys for the catalytic dehydrogenation reaction of MCH.

1. Introduction

The development of green energy technology is attracting wide attention due to the need to replace health- and environment-damaging energy technologies. In this regard, the application of catalytic materials is rapidly expanding; hence, it is necessary to address energy challenges and cost implications for mass deployment. Several catalytic materials require the use of expensive noble metals for their catalytic reactions [1]. Various means of energy storage and deployment are available, all at various stages of development, deployment, and application.
The focus on energy storage and deployment via the use of liquid organic hydrogen carrier (LOHC) materials and, in particular, the development of more efficient and less costly catalytic materials for the dehydrogenation of a LOHC molecule are gaining significant traction [2,3,4,5,6,7,8,9]. LOHC molecules are organic compounds that consist of a carbon backbone, and, in some cases, various functional groups attached to it. LOHC molecules offer a very safe means for hydrogen storage and delivery because they are non-volatile organic liquids. LOHC molecules undergo reversible removal of hydrogen gas (dehydrogenation), resulting in a fully dehydrogenated aromatic compound that is hydrogen-lean, as well as the addition of hydrogen (hydrogenation), resulting in a fully hydrogenated aliphatic compound that is hydrogen-rich and hence, considered renewable [10,11,12,13,14].
The dehydrogenation reaction of the MCH to toluene molecule, for hydrogen storage and production, has been (and is being) extensively investigated [15,16,17,18,19,20]. The reaction of MCH to toluene is of interest, particularly because of the low cost of these molecules, high hydrogen storage density (6.2 wt%) and low toxicity of the dehydrogenated product, toluene [17,20].
Currently, Pt is the most widely used dehydrogenation catalyst for various LOHC molecules [4,20,21,22,23,24,25,26,27,28,29,30]. A reduction in the catalyst’s Pt content, for the dehydrogenation reaction, is essential in efforts to reduce costs associated with LOHCs as energy storage and transport materials. To this end, the use of Sn–Pt-based alloys in various catalytic reactions has been considered, and investigations initiated. The Sn–Pt-based alloys have been found to be useful as electrocatalysts in polymer electrolyte membrane fuel cells [31] and for the dehydrogenation of n-butane to n-butenes [32], selective hydrogenation of crotonaldehyde [33] and dehydrogenation of cyclohexene on Sn/Pt (111) surface alloys [34], and also for the hydrogenation of cyclohexanone on Pt–Sn surface alloys [35].
In this study, we consider application of van der Waals (vdW) spin-polarized density functional theory (DFT) calculations to different Sn–Pt alloys, where emphasis is on a low Pt content in the alloys. Specifically, we considered Sn–Pt alloys with Pt concentration ≤ 50%, i.e., SnPt, Sn2Pt, Sn3Pt2 and Sn4Pt, for the catalytic dehydrogenation of MCH (C7H14) to toluene (C7H8). The considered Sn-Pt alloys were obtained using the convex hull procedure as implemented in the Materials Project (https://materialsproject.org/, accessed on 6 June 2021). For the various Sn–Pt alloy configurations considered, the most dominant surface, as evaluated using the Wulff construction, was taken as a representative surface for the catalytic reaction. The adsorption of the C7H14 and dehydrogenation intermediates on the various surfaces of the Sn–Pt alloy configurations was considered. Stable surface–adsorbate complexes were observed for the various surface configurations. Overall, low Pt content of Sn–Pt alloys gave improved catalytic activity (better than with a pristine Pt catalyst) due to the lower dehydrogenation energy.

2. Results and Discussion

The results obtained for the catalytic properties of the various Sn–Pt alloy surfaces considered are now reported and discussed. First, the lowest energy absorption site for the C7H14 molecule on the Sn–Pt alloy surfaces was evaluated, followed by the removal of each hydrogen molecule, as presented in Figure 1 and Table 1. This was carried out to obtain information on the dehydrogenation of the C7H14 molecule (MCH) to the C7H8 molecule (toluene).
Different approximations play a role in the calculated energetics of a material. The full dehydrogenation of the dominant Pt surface, the Pt (111) surface, was determined using the same approximations as presented in Section 2 to benchmark the various Sn–Pt alloys considered. The calculated dehydrogenation energy of the reaction from C7H14 to C7H8 on the Pt (111) surface was 89.28 kJ/mol. This is higher than recorded in previous theoretical studies [18] and an experimental value of 68.3 kJ/mol; it is attributed to the DFT approximations used in the calculations [16,36]. The application of DFT calculations is known to predict and describe various properties of different materials [1,30,37,38]. As such, our findings here show that a reduction in Pt concentration and alloying with Sn results in overall qualitative and quantitative improvement of the catalytic properties towards the dehydrogenation reaction of C7H14 to C7H8.

2.1. Adsorption Energies of C7H14(g) and Its Intermediates on the Surfaces of Sn–Pt Alloys

The adsorption energies of C7H14 and its intermediate products, as well as C7H8, obtained because of 1/2H2(g) removal on the Sn–Pt surfaces, were evaluated using Equation (2), see Section 3. We wished to determine whether these molecules can stably adsorb on the Sn–Pt surfaces considered and effectively partake in the dehydrogenation reaction. The adsorption locator module in BIOVIA Materials Studio was used to determine the optimal adsorption configurations for the C7H14 molecule, as described in Section 3.
The calculated adsorption energies for the C7H14 molecule and its intermediate products, as well as the C7H8 molecule, were negative on the Sn4Pt, Sn2Pt, Sn3Pt2 and SnPt alloy surfaces considered—as also determined in our previous studies [17]. The implication is that C7H14 and its intermediates products, as well as C7H8, would favourably be formed on the considered surfaces and C7H14 would be dehydrogenated to form C7H8. Furthermore, the adsorption of C7H14 on the Sn–Pt alloy surface had the lowest adsorption energy, compared to other intermediates and the C7H8. The most strongly adsorbed molecules were C7H9, C7H11, C7H13 and C7H11 for the Sn4Pt, Sn2Pt, Sn3Pt2 and SnPt alloy surfaces, respectively. According to results presented in Table 1, there was no clear trend for the adsorption C7H14, its intermediates and the C7H8 molecule.

2.2. Pathways and Energetics of the Dehydrogenation Reaction

2.2.1. Monoatomic Removal of Hydrogen (1/2H2(g))

The dehydrogenation reaction of C7H14 to the C7H8 molecule on the Sn–Pt-based alloy surfaces was undertaken via the removal of a single hydrogen atom from the C7H14 until C7H8 was obtained. To carry out the removal of a single hydrogen atom, a choice should be made on the order of removing the 1/2H2(g). The procedure used in previous studies, which entails removal of a single hydrogen atom, was applied [12,18,28,39,40]. For the Sn–Pt-based alloy surfaces, the 1/2H2(g) removal energies were obtained using Equation (1) for the removal, where E RxN is
E RxN = ( E Surf + Mol - H + E Surf   +   H ) ( E Surf + Mol + E Surf )
Here, E RxN , E Surf + Mol - H , E Surf   +   H , E Surf + Mol , and E Surf are the reaction energies for H removal, the total energy of the surface with the molecules and removed hydrogen, the total energy of the surface with hydrogen, the total energy of the surface with the molecule without removed hydrogen, and the total energy of the surface alone. The surface here refers to the Sn–Pt-based alloy surfaces considered.
As an illustration, the first 1/2H2(g) removal from the MCH molecule was considered. The 1/2H2(g) removal energy was obtained using E RxN = ( E Surf + C 7 H 13 + E Surf +   H ) ( E Surf + C 7 H 14 + E Surf ) , where E S urf   +   C 7 H 14 , E S urf + C 7 H 13 , E Surf and E Surf + H are the calculated total energies of the surface–C7H14 system, surface–C7H13 (MCH with a removed 1/2H2(g)atom) system, surface alone, and a dissociated H adsorbed on the surface, respectively. The calculated removal energies for the single hydrogen atom C7H14, leading to the different intermediates and formation of C7H8, are tabulated in Table 2 for the Sn–Pt-based alloy surfaces considered.
Figure 1 shows a schematic of the C atoms that form the benzene ring of C7H14(g). The procedure was followed for the removal of half hydrogen molecule (1/2H2(g)) up to the sixth hydrogen atom (3 H2(g)) on the different surfaces considered for the Sn–Pt-based alloys. This is an elaborate process, as described above. The calculated energy for the removal of each 1/2H2(g) on the Sn–Pt-based alloy surfaces is presented in Table 3. It is observed that the highest removal energy of H atom on the Sn–Pt-based alloy surfaces does not follow any specific trend for the different surfaces and alloys considered.
The 1/2H2(g) removal pathways from the C7H14 molecule on the SnPt, Sn2Pt, Sn3Pt2 and Sn4Pt surfaces were 541263, 543216, 546123, and 432165, respectively (numbers refer to positions of the aromatic ring: see Figure 1). For the different alloy surfaces, the pathways were not the same. The removal of a single hydrogen atom from the different alloys followed different pathways. However, we did observe that the next nearest 1/2H2(g) to the first abstracted 1/2H2(g) was always preferred, irrespective of the Sn–Pt-based alloy surface. This is a result of the ease of formation of a double bond due to the formation of a free radical initially. The above observation would subsequently reduce the number of calculations required to be carried out in future investigations.

2.2.2. Possible Pathways for H2(g) Removal

The different possible pathways for H2(g) abstraction resulting in the dehydrogenation reaction of C7H14 to the C7H8 molecule on the Sn–Pt-based alloy surfaces are presented in Table 3. This has been shown in previous studies [12,18]. The reaction proceeds stepwise via four different pathways, as presented in Table 3. These pathways entail: (i) the removal of each H2(g), as presented in step A; (ii) the removal of H2(g) followed by 2H2(g), as presented in step B; (iii) the removal of 2H2(g) followed by H2(g), as presented in step C; and (iv) the removal of all 3H2(g) in a single step. The removal of all 3H2(g) is the full dehydrogenation reaction of MCH to toluene, as presented in pathway D. The calculated reaction energies for these pathways and each step were obtained using a modified version of Equation (1).
Taking pathways A, B and C into consideration, the highest energy step for the removal of the first H2(g) on the different Sn–Pt-based alloy surfaces does not follow the similar or equivalent steps or a specific trend. The difference in pathways is attributed to the nature of each of the Sn–Pt-based alloy surfaces considered here and their electronic properties. Thus, different steps are the reaction-limiting cases for the pathways A, B and C. However, no specific trend in terms of reaction pathway was observed. The implication is that Sn–Pt-based catalysts do not follow any specific trend in terms of the dehydrogenation pathways. For the removal of 3H2, which results in the overall dehydrogenation of C7H14 to C7H8, which is pathway D, the pristine Pt has the highest dehydrogenation energy (compared to the various alloys considered). The implication is that alloying Pt with Sn leads to a reduction in the overall activation energy and hence an improvement in the catalytic properties of the material.
The Brønsted–Evans–Polanyi linear scaling relationship uses both the reaction energies and activation energies of the dehydrogenation reactions to determine the activation energy, as reported earlier [41], as has been applied to some studies [15,42,43]. The surface with the highest reaction energy also has the highest activation energy, as determined in an earlier investigation by Obodo et al. [15]. Thus, the catalytic activation energy can be described using only the reaction energies.

3. Computational Method

The spin-polarized vdW DFT [44,45,46] calculations were carried as implemented in the CASTEP code [47,48] within the Materials Studio package. The Perdew–Burke–Ernzerhof form [49] of the generalized gradient approximation [50,51] exchange-correlation functional and ultra-soft pseudopotential were applied to describe the electron–ion interactions. The Tkatcheno–Scheffler method [52] for semi-empirical dispersion correction was applied to account for the vdW interactions.
The SnPt, Sn2Pt, Sn3Pt2 and Sn4Pt in the P63/MMC, P63/MMC, P63/MMC and CCCA space groups were considered. These configurations were obtained using the convex hull from the Materials Project database as the most optimal structures. The various surface configurations for the Sn–Pt-based alloys were determined. We found that the (1 −1 0), (1 1 1), (0 0 2) and (1 −1 0) were the most dominant surface configurations, using the Wulff construction as implemented in the morphology module of the Materials Studio 2020 suite for the SnPt, Sn2Pt, Sn3Pt2 and Sn4Pt, respectively, as presented in Figure 2. These configurations were used to create surface slabs with 4/5 atomic layers in the c-axes. The k-point mesh separation of 0.07 Å−1, an adequately converged kinetic energy cut-off of 408.20 eV, and 10−6 eV/atom convergence criteria for the calculated total energies were applied to optimize the various alloys of the Sn–Pt unit cell considered. The forces in the various alloys of the Sn–Pt unit cell considered were converged up to 0.03 eV/Å for the full geometric optimization.
Various super-cell slabs of the Sn–Pt alloys with vacuum distance 15 Å, and consisting of five-layer slabs, were used for all of the calculations in this study. The vacuum distance applied in the z-direction was adequate to prevent periodic image interactions [53]. The convergence criteria enforced for the unit cell and the super-cell slab calculations were the same. For the super-cell slab calculations, the volume was fixed with the atomic positions optimized. The size of the molecular adsorbates considered, and tractability of the calculations, influenced the choice for the area of the surface slab (length × width). To avoid, first, the straining of the relaxed molecular adsorbates on the surface and, second, the interaction of mirror images, >5 Å was added to the length and width of the longest molecular species to form the length and width of the surface slab, as consistent with previous studies [12].
The lowest energy adsorption sites for the molecular adsorbates (C7H14, C7H13, C7H12, C7H11, C7H10, C7H9 and C7H8) on the Sn–Pt alloy surfaces were evaluated using the adsorption locator module. The latter has been applied in various studies to evaluate the lowest energy surface–adsorbate configuration (SAC) [4,12,18]. The molecular adsorbates were pre-adsorbed to determine the optimal SAC, which is the lowest energy configuration. The determined optimal surface–adsorbate configurations were subsequently relaxed using vdW corrected spin-polarized DFT calculations to determine the final configurations and total energies [54]. Next, the adsorption energy (EAds) of each SAC on the Sn–Pt surfaces was evaluated using Equation (2) [18,55]
E Ads =   E SAC   E Surf   E mol
where E SAC , E Surf   , and Emol are the following, respectively: the total energy of the Sn–Pt surface slabs with the C7H14 molecule and its intermediates, the total energy of the Sn–Pt surface slabs without the C7H14 molecule and its intermediates, and the total energy of the C7H14 molecule and its intermediates. The activation energies were not considered in the current investigation (as in our previous studies) [15].

4. Conclusions

DFT with spin-polarized vdW correction was applied to investigate Sn–Pt-based alloy surfaces, specifically, low-concentration Sn–Pt-based alloy surfaces. The convex hull and Wulff construction were used to determine the dominant surfaces. The calculated adsorption energies of C7H14, its intermediates and C7H8 were determined—they absorb exothermically on the considered Sn–Pt-based alloy surfaces.
The determined reaction energies for the various surfaces of Sn–Pt-based alloys demonstrated that the dehydrogenation pathways differ. The Sn4Pt (101) surface had the lowest overall dehydrogenation energy of all the considered surfaces. Hence, among the Sn–Pt-based alloys, the Sn4Pt alloy is the best dehydrogenation catalyst. Furthermore, all the considered Sn–Pt-based alloys considered in this study had lower dehydrogenation energies compared to Pt (Pt being the optimal dehydrogenation catalyst to date).

Author Contributions

Conceptualization, K.O.O.; Methodology, K.O.O.; Writing—original draft, K.O.O.; Writing—review and editing, K.O.O., C.N.M.O. and D.B.; Funding acquisition, D.B. All authors have read and agreed to the published version of the manuscript.

Funding

This project has received funding from the Fuel Cells and Hydrogen 2 Joint Undertaking under grant agreement No 101007223. This Joint Undertaking receives support from the European Union’s Horizon 2020 Research and Innovation programme, Hydrogen Europe, and Hydrogen Europe research. The project also received funding from DSI South Africa and HySA Infrastructure Centre of Competence, Faculty of Engineering, North-West University via KP5 grant.

Data Availability Statement

Data used in the study are included in the manuscript.

Acknowledgments

The authors thank the Centre for High Performance Computing (CHPC), Cape Town, South Africa for computational resources used in this study and the HySA-Infrastructure Centre of Competence, Faculty of Engineering, North-West University for financial support. KOO thanks National Institute of Theoretical and Computational Sciences, Johannesburg, South Africa.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Obodo, K.O.; Ouma, C.N.M.; Bessarabov, D. Low-Temperature Water Electrolysis. In Power to Fuel: How to Speed Up a Hydrogen Economy; Spazzafumo, G., Ed.; Elsevier: Amsterdam, The Netherlands, 2021; pp. 1–287. [Google Scholar]
  2. Bessarabov, D.; Human, G.; Kruger, A.J.; Chiuta, S.; Modisha, P.M.; du Preez, S.P.; Oelofse, S.P.; Vincent, I.; Van Der Merwe, J.; Langmi, H.W.; et al. South African Hydrogen Infrastructure (HySA Infrastructure) for Fuel Cells and Energy Storage: Overview of a Projects Portfolio. Int. J. Hydrogen Energy 2017, 42, 13568–13588. [Google Scholar] [CrossRef]
  3. Wulf, C.; Zapp, P. Assessment of System Variations for Hydrogen Transport by Liquid Organic Hydrogen Carriers. Int. J. Hydrogen Energy 2018, 43, 11884–11895. [Google Scholar] [CrossRef]
  4. Modisha, P.M.; Ouma, C.N.M.; Garidzirai, R.; Wasserscheid, P.; Bessarabov, D. The Prospect of Hydrogen Storage Using Liquid Organic Hydrogen Carriers. Energy Fuels 2019, 33, 2778–2796. [Google Scholar] [CrossRef]
  5. Jo, Y.; Oh, J.; Kim, D.; Park, J.H.; Baik, H.; Suh, Y.-W. Recent Progress in Dehydrogenation Catalysts for Heterocyclic and Homocyclic Liquid Organic Hydrogen Carriers. Korean J. Chem. Eng. 2022, 39, 20–37. [Google Scholar] [CrossRef]
  6. Ye, H.-L.; Liu, S.-X.; Zhang, C.; Cai, Y.-Q.; Shi, Y.-F. Dehydrogenation of Methylcyclohexane over Pt-Based Catalysts Supported on Functional Granular Activated Carbon. RSC Adv. 2021, 11, 29287. [Google Scholar] [CrossRef]
  7. Wijayanta, A.T.; Oda, T.; Purnomo, C.W.; Kashiwagi, T.; Aziz, M. Liquid Hydrogen, Methylcyclohexane, and Ammonia as Potential Hydrogen Storage: Comparison Review. Int. J. Hydrogen Energy 2019, 44, 15026–15044. [Google Scholar] [CrossRef]
  8. Niermann, M.; Timmerberg, S.; Drünert, S.; Kaltschmitt, M. Liquid Organic Hydrogen Carriers and Alternatives for International Transport of Renewable Hydrogen. Renew. Sustain. Energy Rev. 2021, 135, 110171. [Google Scholar] [CrossRef]
  9. Kwak, Y.; Kirk, J.; Moon, S.; Ohm, T.; Lee, Y.J.; Jang, M.; Park, L.H.; Ahn, C.; Jeong, H.; Sohn, H.; et al. Hydrogen Production from Homocyclic Liquid Organic Hydrogen Carriers (LOHCs): Benchmarking Studies and Energy-Economic Analyses. Energy Convers. Manag. 2021, 239, 114124. [Google Scholar] [CrossRef]
  10. Do, G.; Preuster, P.; Aslam, R.; Bösmann, A.; Müller, K.; Arlt, W.; Wasserscheid, P. Hydrogenation of the Liquid Organic Hydrogen Carrier Compound Dibenzyltoluene—Reaction Pathway Determination by 1H NMR Spectroscopy. React. Chem. Eng. 2016, 1, 313–320. [Google Scholar] [CrossRef] [Green Version]
  11. Leinweber, A.; Müller, K. Hydrogenation of the LOHC Compound Monobenzyl Toluene—Reaction Pathway and Kinetic Effects. Energy Technol. 2017, 6, 513–520. [Google Scholar] [CrossRef]
  12. Ouma, C.N.M.; Modisha, P.; Bessarabov, D. Insight into the Adsorption of a Liquid Organic Hydrogen Carrier, Perhydro- i -Dibenzyltoluene (i = m, o, p), on Pt, Pd and PtPd Planar Surfaces. RSC Adv. 2018, 8, 31895–31904. [Google Scholar] [CrossRef] [Green Version]
  13. Singh, E.; Kim, K.S.; Yeom, G.Y.; Nalwa, H.S. Two-Dimensional Transition Metal Dichalcogenide-Based Counter Electrodes for Dye-Sensitized Solar Cells. RSC Adv. 2017, 7, 28234–28290. [Google Scholar] [CrossRef] [Green Version]
  14. Ouma, C.N.M.; Obodo, K.O.; Modisha, P.M.; Rhyman, L.; Ramasami, P.; Bessarabov, D. Effect of Chalcogen (S, Se and Te) Surface Additives on the Dehydrogenation of a Liquid Organic Hydrogen Carrier System, Octahydroindole–Indole, on a Pt (1 1 1) Surface. Appl. Surf. Sci. 2021, 566, 150636. [Google Scholar] [CrossRef]
  15. Obodo, K.O.; Ouma, C.N.M.; Bessarabov, D. Modified Pt (2 1 1) and (3 1 1) Surfaces towards the Dehydrogenation of Methylcyclohexane to Toluene: A Density Functional Theory Study. Appl. Surf. Sci. 2022, 584, 152590. [Google Scholar] [CrossRef]
  16. Chen, F.; Huang, Y.; Mi, C.; Wu, K.; Wang, W.; Li, W.; Yang, Y. Density Functional Theory Study on Catalytic Dehydrogenation of Methylcyclohexane on Pt(111). Int. J. Hydrogen Energy 2020, 45, 6727–6737. [Google Scholar] [CrossRef]
  17. Alhumaidan, F.; Cresswell, D.; Garforth, A. Hydrogen Storage in Liquid Organic Hydride: Producing Hydrogen Catalytically from Methylcyclohexane. Energy Fuels 2011, 25, 4217–4234. [Google Scholar] [CrossRef]
  18. Obodo, K.O.; Ouma, C.N.M.; Modisha, P.M.; Bessarabov, D. Density Functional Theory Calculation of Ti3C2 MXene Monolayer as Catalytic Support for Platinum towards the Dehydrogenation of Methylcyclohexane. Appl. Surf. Sci. 2020, 529, 147186. [Google Scholar] [CrossRef]
  19. Yolcular, S.; Olgun, Ö. Hydrogen Storage in the Form of Methylcyclohexane. Energy Sources Part A Recover. Util. Environ. Eff. 2007, 30, 149–256. [Google Scholar] [CrossRef]
  20. Boufaden, N.; Akkari, R.; Pawelec, B.; Fierro, J.L.G.; Zina, M.S.; Ghorbel, A. Dehydrogenation of Methylcyclohexane to Toluene over Partially Reduced Silica-Supported Pt-Mo Catalysts. J. Mol. Catal. A Chem. 2016, 420, 96–106. [Google Scholar] [CrossRef]
  21. Sun, G.; Zhao, Z.J.; Mu, R.; Zha, S.; Li, L.; Chen, S.; Zang, K.; Luo, J.; Li, Z.; Purdy, S.C.; et al. Breaking the Scaling Relationship via Thermally Stable Pt/Cu Single Atom Alloys for Catalytic Dehydrogenation. Nat. Commun. 2018, 9, 4454. [Google Scholar] [CrossRef]
  22. Maatman, R.W.; Mahaffy, P.; Hoekstra, P.; Addink, C. The Preparation of Pt-Alumina Catalyst and Its Role in Cyclohexane Dehydrogenation. J. Catal. 1971, 23, 105–118. [Google Scholar] [CrossRef]
  23. Sugiura, Y.; Nagatsuka, T.; Kubo, K.; Hirano, Y.; Nakamura, A.; Miyazawa, K.; Iizuka, Y.; Furuta, S.; Iki, H.; Higo, T.; et al. Dehydrogenation of Methylcyclohexane over Pt/TiO2-Al2O3 Catalysts. Chem. Lett. 2017, 46, 1601–1604. [Google Scholar] [CrossRef]
  24. Lup, A.N.K.; Abnisa, F.; Daud, W.M.A.W.; Aroua, M.K. A Review on Reactivity and Stability of Heterogeneous Metal Catalysts for Deoxygenation of Bio-Oil Model Compounds. J. Ind. Eng. Chem. 2017, 56, 1–34. [Google Scholar] [CrossRef]
  25. Antolini, E. Pt-Ni and Pt-M-Ni (M = Ru, Sn) Anode Catalysts for Low-Temperature Acidic Direct Alcohol Fuel Cells: A Review. Energies 2017, 10, 42. [Google Scholar] [CrossRef] [Green Version]
  26. Frontera, P.; Macario, A.; Ferraro, M.; Antonucci, P. Supported Catalysts for CO2 Methanation: A Review. Catalysts 2017, 7, 59. [Google Scholar] [CrossRef]
  27. Auer, F.; Blaumeiser, D.; Bauer, T.; Bösmann, A.; Szesni, N.; Libuda, J.; Wasserscheid, P. Boosting the Activity of Hydrogen Release from Liquid Organic Hydrogen Carrier Systems by Sulfur-Additives to Pt on Alumina Catalysts. Catal. Sci. Technol. 2019, 9, 3537–3547. [Google Scholar] [CrossRef]
  28. Bachmann, P.; Schwarz, M.; Steinhauer, J.; Späth, F.; Düll, F.; Bauer, U.; Silva, T.N.; Mohr, S.; Hohner, C.; Scheuermeyer, M.; et al. Dehydrogenation of the Liquid Organic Hydrogen Carrier System Indole/Indoline/Octahydroindole on Pt(111). J. Phys. Chem. C 2018, 122, 4470–4479. [Google Scholar] [CrossRef]
  29. Modisha, P.M.; Jordaan, J.H.L.; Bösmann, A.; Wasserscheid, P.; Bessarabov, D. Analysis of Reaction Mixtures of Perhydro-Dibenzyltoluene Using Two-Dimensional Gas Chromatography and Single Quadrupole Gas Chromatography. Int. J. Hydrogen Energy 2018, 43, 5620–5636. [Google Scholar] [CrossRef]
  30. Wu, X.; Kang, F.; Duan, W.; Li, J. Density Functional Theory Calculations: A Powerful Tool to Simulate and Design High-Performance Energy Storage and Conversion Materials. Prog. Nat. Sci. Mater. Int. 2019, 29, 247–255. [Google Scholar] [CrossRef]
  31. Borbáth, I.; Bakos, I.; Pászti, Z.; Szijjártó, G.P.; Tompos, A. Design of SnPt/C Cathode Electrocatalysts with Optimized Sn/Pt Surface Composition for Potential Use in Polymer Electrolyte Membrane Fuel Cells. Catal. Today 2021, 366, 20–30. [Google Scholar] [CrossRef]
  32. Nagaraja, B.M.; Shin, C.H.; Jung, K.D. Selective and Stable Bimetallic PtSn/θ-Al2O3 Catalyst for Dehydrogenation of n-Butane to n-Butenes. Appl. Catal. A Gen. 2013, 467, 211–223. [Google Scholar] [CrossRef]
  33. Taniya, K.; Yu, C.H.; Takado, H.; Hara, T.; Okemoto, A.; Horie, T.; Ichihashi, Y.; Tsang, S.C.; Nishiyama, S. Synthesis of Bimetallic SnPt-Nanoparticle Catalysts for Chemoselective Hydrogenation of Crotonaldehyde: Relationship between SnxPty Alloy Phase and Catalytic Performance. Catal. Today 2018, 303, 241–248. [Google Scholar] [CrossRef]
  34. Xu, C.; Koel, B.E. Dehydrogenation of Cyclohexene on Ordered Sn/Pt(111) Surface Alloys. Surf. Sci. 1994, 304, 249–266. [Google Scholar] [CrossRef]
  35. Olivas, A.; Jerdev, D.I.; Koel, B.E. Hydrogenation of Cyclohexanone on Pt–Sn Surface Alloys. J. Catal. 2004, 222, 285–292. [Google Scholar] [CrossRef]
  36. Preuster, P.; Papp, C.; Wasserscheid, P. Liquid Organic Hydrogen Carriers (LOHCs): Toward a Hydrogen-Free Hydrogen Economy. Acc. Chem. Res. 2017, 50, 74–85. [Google Scholar] [CrossRef]
  37. Obodo, K.O.; Ouma, C.N.M.; Obodo, J.T.; Gebreyesus, G.; Rai, D.P.; Ukpong, A.M.; Bouhafs, B. Sn3C2 Monolayer with Transition Metal Adatom for Gas Sensing: A Density Functional Theory Studies. Nanotechnology 2021, 32, 355502. [Google Scholar] [CrossRef]
  38. Obodo, K.O.; Ouma, C.N.M.; Bessarabov, D. First Principles Evaluation of the OER Properties of TM−X (TM = Cr, Mn, Fe, Mo, Ru, W and Os, and X = F and S) Doped IrO2 (110) Surface. Electrochim. Acta 2021, 403, 139562. [Google Scholar] [CrossRef]
  39. Ouma, C.N.M.; Modisha, P.M.; Bessarabov, D. Catalytic Dehydrogenation Onset of Liquid Organic Hydrogen Carrier, Perhydro-Dibenzyltoluene: The Effect of Pd and Pt Subsurface Configurations. Comput. Mater. Sci. 2020, 172, 109332. [Google Scholar] [CrossRef]
  40. Ouma, C.N.M.; Obodo, K.O.; Modisha, P.M.; Bessarabov, D. Si, P, S and Se Surface Additives as Catalytic Activity Boosters for Dehydrogenation of Methylcyclohexane to Toluene—A Liquid Organic Hydrogen Carrier System: Density Functional Theory Insights. Mater. Chem. Phys. 2022, 279, 125728. [Google Scholar] [CrossRef]
  41. Wang, S.; Petzold, V.; Tripkovic, V.; Kleis, J.; Howalt, J.G.; Skulason, E.; Fernández, E.M.; Hvolbæk, B.; Jones, G.; Toftelund, A.; et al. Universal Transition State Scaling Relations for (de)Hydrogenation over Transition Metals. Phys. Chem. Chem. Phys. 2011, 13, 20760–20765. [Google Scholar] [CrossRef]
  42. Yook, H.; Kim, K.; Park, J.H.; Suh, Y.W.; Han, J.W. Density Functional Theory Study on the Dehydrogenation of 1,2-Dimethyl Cyclohexane and 2-Methyl Piperidine on Pd and Pt Catalysts. Catal. Today 2020, 352, 345–353. [Google Scholar] [CrossRef]
  43. Kim, K.; Oh, J.; Kim, T.W.; Park, J.H.; Han, J.W.; Suh, Y.W. Different Catalytic Behaviors of Pd and Pt Metals in Decalin Dehydrogenation to Naphthalene. Catal. Sci. Technol. 2017, 7, 3728–3735. [Google Scholar] [CrossRef]
  44. Hohenberg, P.; Kohn, W. Inhomogeneous Electron Gas. Phys. Rev. 1964, 136, B864–B871. [Google Scholar] [CrossRef] [Green Version]
  45. Kohn, W.; Sham, L.J. Self-Consistent Equations Including Exchange and Correlation Effects. Phys. Rev. 1965, 140, A1133–A1138. [Google Scholar] [CrossRef] [Green Version]
  46. McNellis, E.R.; Meyer, J.; Reuter, K. Azobenzene at Coinage Metal Surfaces: Role of Dispersive van Der Waals Interactions. Phys. Rev. B Condens. Matter Mater. Phys. 2009, 80, 205414. [Google Scholar] [CrossRef] [Green Version]
  47. Segall, M.D.; Lindan, J.; Probert, M.J.; Pickard, C.J.; Hasnip, P.J.; Clark, S.J.; Payne, M.C. First-Principles Simulation: Ideas, Illustrations and the CASTEP Code. J. Phys. Condens. Matter 2002, 14, 2717–2744. [Google Scholar] [CrossRef]
  48. Clark, S.J.; Segall, M.D.; Pickard, C.J.; Hasnip, P.J.; Probert, M.I.J.; Refson, K.; Payne, M.C. First Principles Methods Using CASTEP. Z. Krist. 2005, 220, 567–570. [Google Scholar] [CrossRef] [Green Version]
  49. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Perdew, J.P.; Zunger, A. Self-Interaction Correction to Density-Functional Approximations for Many-Electron Systems. Phys. Rev. B 1981, 23, 5048–5049. [Google Scholar] [CrossRef] [Green Version]
  51. Perdew, J.P.; Wang, Y. Accurate and Simple Analytic Representation of the Electron-Gas Correlation Energy. Phys. Rev. B 1992, 45, 13244–13249. [Google Scholar] [CrossRef]
  52. Tkatchenko, A.; Scheffler, M. Accurate Molecular van Der Waals Interactions from Ground-State Electron Density and Free-Atom Reference Data. Phys. Rev. Lett. 2009, 102, 073005. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Obodo, K.O.; Ouma, C.N.M.; Obodo, J.T.; Braun, M.; Bessarabov, D. First Principles Study of Single and Multi-Site Transition Metal Dopant Ions in MoS2 Monolayer. Comput. Condens. Matter 2019, 21, e00419. [Google Scholar] [CrossRef]
  54. Hamada, I. Van Der Waals Density Functional Made Accurate. Phys. Rev. B 2014, 89, 121103. [Google Scholar] [CrossRef]
  55. Ouma, C.N.M.; Obodo, K.O.; Braun, M.; Amolo, G.O.; Bessarabov, D. Insights on Hydrogen Evolution Reaction in Transition Metal Doped Monolayer TcS2 from Density Functional Theory Calculations. Appl. Surf. Sci. 2019, 470, 107–113. [Google Scholar] [CrossRef]
Figure 1. Full and stepwise dehydrogenation of methylcyclohexane to toluene molecule with the suggestive position of the single hydrogen atom (1/2 H2(g)) removal site in the molecules shown. The schematic shows that the 1/2H2(g) removal proceeds in the order 1–6, which is an idealised configuration. The * represents removal of one 1/2 H2(g) followed by the formation of a double C–C bond due to H2(g) removal. The other pathways show the removal of 3H2(g), resulting in the full dehydrogenation of methylcyclohexane to toluene.
Figure 1. Full and stepwise dehydrogenation of methylcyclohexane to toluene molecule with the suggestive position of the single hydrogen atom (1/2 H2(g)) removal site in the molecules shown. The schematic shows that the 1/2H2(g) removal proceeds in the order 1–6, which is an idealised configuration. The * represents removal of one 1/2 H2(g) followed by the formation of a double C–C bond due to H2(g) removal. The other pathways show the removal of 3H2(g), resulting in the full dehydrogenation of methylcyclohexane to toluene.
Catalysts 12 01221 g001
Figure 2. Schematic ball and stick model of the Sn–Pt alloy surfaces showing the top and side view, where (a) Sn4Pt, (b) Sn3Pt2 (c) Sn2Pt, and (d) SnPt. The blue and grey balls are the Pt and Sn atoms, respectively.
Figure 2. Schematic ball and stick model of the Sn–Pt alloy surfaces showing the top and side view, where (a) Sn4Pt, (b) Sn3Pt2 (c) Sn2Pt, and (d) SnPt. The blue and grey balls are the Pt and Sn atoms, respectively.
Catalysts 12 01221 g002
Table 1. Methylcyclohexane (C7H14), the dehydrogenation intermediates and toluene (C7H8), and the calculated adsorption energies (in eV) for Sn–Pt surfaces considered.
Table 1. Methylcyclohexane (C7H14), the dehydrogenation intermediates and toluene (C7H8), and the calculated adsorption energies (in eV) for Sn–Pt surfaces considered.
MoleculesSn4PtSn2PtSn3Pt2SnPt
C7H14−0.87−0.91−0.73−2.97
C7H13−3.76−2.11−3.51−3.52
C7H12−2.79−1.31−2.31−3.05
C7H11−3.68−2.07−3.32−4.57
C7H10−3.77−1.32−2.34−3.91
C7H9−4.24−1.91−3.11−4.52
C7H8−3.55−1.03−0.75−3.29
Table 2. Calculated energy for the removal of a half hydrogen molecule (1/2H2(g)) up to the sixth hydrogen atom (3 H2(g)) on the different surfaces considered for the Sn–Pt alloys.
Table 2. Calculated energy for the removal of a half hydrogen molecule (1/2H2(g)) up to the sixth hydrogen atom (3 H2(g)) on the different surfaces considered for the Sn–Pt alloys.
Sn–Pt AlloysSurface1H2H3H4H5H6H
Sn3Pt2(0 0 2)−69.7662.0567.8158.5945.83104.48
Sn4Pt(1 0 1)−78.4939.1179.04−43.4574.82−56.41
Sn2Pt(1 1 1)81.8323.3691.5236.1363.78−38.08
SnPt(1−1 0)143.75−8.0318.0829.0360.803.97
Table 3. Calculated reaction energies on the Sn–Pt-based alloy surfaces (in eV) for the pathways A–D.
Table 3. Calculated reaction energies on the Sn–Pt-based alloy surfaces (in eV) for the pathways A–D.
PathwaysABCD
Sn–Pt AlloysSurfacesH2H2H2H22H22H2H23H2
Sn3Pt2(0 0 2)−7.71126.41150.31−7.71276.72118.70150.3183.33
Sn4Pt(1 0 1)−39.3935.5918.41−39.3954.00−3.8018.414.41
Sn2Pt(1 1 1)105.19127.6525.70105.19153.36232.8425.7085.72
SnPt(1 −1 0)135.7247.1164.77135.72111.88182.8364.7779.45
Pt(1 1 1) 89.28
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Obodo, K.O.; Ouma, C.N.M.; Bessarabov, D. Low-Pt-Based Sn Alloy for the Dehydrogenation of Methylcyclohexane to Toluene: A Density Functional Theory Study. Catalysts 2022, 12, 1221. https://doi.org/10.3390/catal12101221

AMA Style

Obodo KO, Ouma CNM, Bessarabov D. Low-Pt-Based Sn Alloy for the Dehydrogenation of Methylcyclohexane to Toluene: A Density Functional Theory Study. Catalysts. 2022; 12(10):1221. https://doi.org/10.3390/catal12101221

Chicago/Turabian Style

Obodo, Kingsley Onyebuchi, Cecil Naphtaly Moro Ouma, and Dmitri Bessarabov. 2022. "Low-Pt-Based Sn Alloy for the Dehydrogenation of Methylcyclohexane to Toluene: A Density Functional Theory Study" Catalysts 12, no. 10: 1221. https://doi.org/10.3390/catal12101221

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