Acessibilidade / Reportar erro

Theoretical Calculations on the Mechanism of Hydrogenation of Diphenylacetylene over Pdn (n = 1-4) Clusters

Abstract

Diphenylacetylene (DPA) is a precursor of stilbene and benzil, and reduction of DPA or its derivatives with metallic reagents is both an old and contemporary topic of research. By means of density function theory (DFT) calculations, a detailed investigation of the mechanism of the hydrogenation of DPA over Pd clusters was carried out at the molecular level. The various species structures in the hydrogenation of DPA over Pd clusters were optimized and analyzed. The calculations indicate that the reactions over different Pd clusters share similar reaction mechanisms, and the entire reaction path could be divided into approximately two stages: stage 1: the hydrogenation of DPA to stilbene by the addition of one hydrogen molecule; and stage 2: the hydrogenation of stilbene to the final product diphenylethane (DPE) with the recovery of the catalyst. The Pd2- and Pd3-catalyzed systems exhibit the smallest rate-determining step (RDS) energy barrier, and these systems might be the most active and effective catalytic species among the Pd clusters. Although the Pd clusters used in the current work are simple systems, these clusters could eventually provide insights into the specific structure of the Pd catalyst, since Pd and/or clusters and/or nanoparticles could be envisioned as active catalysts in experiments.

Keywords:
palladium-catalyzed; diphenylacetylene; DFT; reaction mechanisms


Introduction

The carbon-carbon triple bond of alkynes is one of the basic functional groups in chemistry, and its reaction belongs to the foundations of organic chemistry.11 Jun, C. H.; Chem. Soc. Rev. 2004, 33, 610.

2 Tobisu, M.; Chatani, N.; Chem. Soc. Rev. 2008, 37, 300.
-33 Wang, A.; Jiang, H.; J. Am. Chem. Soc. 2008, 130, 5030. In recent years, the physical chemistry and theory of few organic molecules have been studied more deeply than those of acetylene. Acetylene chemistry has undergone a renaissance because acetylene not only is present in molecules on the frontiers of organic chemistry, such as pharmaceutical chemistry and biochemistry, but also serves as a building block or a general intermediate for the synthesis of large numbers of chemicals.44 Chinchilla, R.; Najera, C.; Chem. Rev. 2014, 114, 1783.

5 Stang, P. J.; Diederich, F.; Modern Acetylene Chemistry; VCH: Weinheim, 1995.
-66 Diederich, F.; Stang, P. J.; Tykwinski, R. R.; Acetylene Chemistry; Wiley-VCH: Weinheim, 2005. The development of new synthetic methodologies based on transition metal catalysis has provided impetus to alkyne chemistry.77 Schmidt, E. Y.; Tatarinova, I. V.; Semenova, N. V.; Protsuk, N. I.; Ushakov, I. A.; Trofimov, B. A.; J. Org. Chem. 2018, 83, 10272. The partial hydrogenation catalytic reduction of internal alkynes is an efficient method for preparing olefins and alkanes, and metal complex catalysts have been the most effective catalysts for achieving this transformation.88 Borodziński, A.; Bond, G. C.; Catal. Rev. 2008, 50, 379.,99 Molnár, Á.; Sárkány, A.; Varga, M.; J. Mol. Catal. A: Chem. 2001, 173, 185. The methods for the hydrogenation of alkynes provide exciting strategies for the synthesis of complex organic building blocks and attract extensive attention in pharmaceuticals and other valuable chemicals. This type of hydrogenation reaction utilizes homogeneous Ni, Cr, Ru, Rh, Pd, Cu and borate organic catalysts and is known for the reduction of olefins, alkynes, nitriles and ketones. Although hydrogenation reagents are hydrogen activated by transition metal-based catalysts, Pd catalysts require the mildest of conditions and are the most widely used.1010 Barrios-Francisco, R.; García, J. J.; Appl. Catal., A 2010, 385, 108.

11 Espinal-Viguri, M.; Neale, S. E.; Coles, N. T.; Macgregor, S. A.; Webster, R. L.; J. Am. Chem. Soc. 2019, 141, 572.

12 Shao, Z.; Fu, S.; Wei, M.; Zhou, S.; Liu, Q.; Angew. Chem., Int. Ed. 2016, 55, 14653.

13 Zhou, Q.; Zhang, L.; Meng, W.; Feng, X.; Yang, J.; Du, H.; Org. Lett. 2016, 18, 5189.
-1414 Sperger, T.; Sanhueza, I. A.; Kalvet, I.; Schoenebeck, F.; Chem. Rev. 2015, 115, 9532.

Diphenylacetylene (DPA) is particularly representative of the alkynes, and it can act as a Lewis acid and ligand in organometallic chemistry because of its symmetry and high planarity. Additionally, DPA is a precursor of stilbene and benzil and reduction of DPA or its derivatives with metallic reagents is old and contemporary topic of research; much attention has been focused on metal complexes, such as Pd, Ir, Ag, Rh complexes, etc.,1515 Kojima, Y.; Matsuoka, T.; Takahashi, H.; J. Mater. Sci. Lett. 1996, 15, 1543.

16 Li, S. H.; Li, L. C.; Xu, C. H.; Chin. Chem. Lett. 2012, 23, 69.

17 Manna, S.; Antonchick, A. P.; ChemSusChem 2019, 12, 3094.

18 Domínguez-Domínguez, S.; Berenguer-Murcia, Á.; Linares-Solano, Á.; Cazorla-Amorós, D.; J. Catal. 2008, 257, 87.

19 Nishio, R.; Sugiura, M.; Kobayashi, S.; Org. Biomol. Chem. 2006, 4, 992.
-2020 Maazaoui, R.; Abderrahim, R.; Chemla, F.; Ferreira, F.; Perez-Luna, A.; Jackowski, O.; Org. Lett. 2018, 20, 7544. as reductants. In the case of the hydrogenation reaction of diphenylacetylene, however, diphenylethylene is formed as a primary product, but 1,2-diphenylethane (DPE) is absent. In addition, with the development of chemistry, some different synthetic methods, such as the reduction of diphenylacetylene to 1,2-diphenylethane, have been discovered. Recent methods for the synthesis of 1,2-diphenylethane have concentrated on the direct reduction of diphenylacetylene, and most of the reported methods require metal complexes. For example, Cravotto and co-workers2121 Wu, Z.; Cherkasov, N.; Cravotto, G.; Borretto, E.; Ibhadon, A. O.; Medlock, J.; Bonrath, W.; ChemCatChem 2015, 7, 952. reported that the Pd complex can act as a catalyst to generate 1,2-diphenylethane by the reduction of diphenylacetylene. Additionally, Webster and co-workers2222 King, A. K.; Buchard, A.; Mahon M. F.; Webster, R. L.; Chem. - Eur. J. 2015, 21, 15960. reported that using the FeII complex as a catalyst, in 1 equivalent of nBuNH2 and HBpin, the triple bond was reduced, and the 1,2-diphenylethane was obtained. However, low yield and many byproducts affect the efficient utilization of catalysts in the hydrogenation of diphenylacetylene. A new hydrogenation reaction with complex metal catalysts was reported by Jackowski and co-workers.2020 Maazaoui, R.; Abderrahim, R.; Chemla, F.; Ferreira, F.; Perez-Luna, A.; Jackowski, O.; Org. Lett. 2018, 20, 7544. When Cl2Pd(PPh3)2/Me2Zn was used as the catalyst in the reduction of diphenylacetylene, 1,2-diphenylethane was the only product and was produced in 99% yield. Despite the use of a bimetallic complex in hydrogenation reactions, Cl2Pd(PPh3)2 as precatalyst and Me2Zn as a reducing agent, Me2Zn interacts with the PdII precatalyst to deliver Cl2Pd(PPh3)2, which is reduced to Pd0.2020 Maazaoui, R.; Abderrahim, R.; Chemla, F.; Ferreira, F.; Perez-Luna, A.; Jackowski, O.; Org. Lett. 2018, 20, 7544. Therefore, Pd0 is an efficient catalyst in the hydrogenation of diphenylacetylene to 1,2-diphenylethane. Although the authors deduced the reaction mechanism by experiments, the details of the transformations and, more importantly, the origins of the observed selectivity still need to be understood at the molecular level, without which the mechanism of the calculations is incomplete. In challenging situations, the preferred mechanism may depend on the reaction conditions and ligands. While experiments can provide sufficient insights, the interpretation of these results is often inconclusive. In such cases, moving to computational chemistry may attain a deeper understanding of specific chemical problems.

In many cases, Pd catalysis has provided a new world of transformations and has made numerous structural parts more accessible. Although the role of Pd may be apparent, there are usually multiple pathways that need to be considered.1414 Sperger, T.; Sanhueza, I. A.; Kalvet, I.; Schoenebeck, F.; Chem. Rev. 2015, 115, 9532.,2323 Sameera, W. M. C.; Maseras, F.; Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2012, 2, 375.

24 Bonney, K. J.; Schoenebeck, F.; Chem. Soc. Rev. 2014, 43, 6609.

25 Holder, J. C.; Zou, L.; Marziale, A. N.; Liu, P.; Lan, Y.; Gatti, M.; Kikushima, K.; Houk, K. N.; Stoltz, B. M.; J. Am. Chem. Soc. 2013, 135, 14996.

26 Lan, Y.; Houk, K. N.; J. Org. Chem. 2011, 76, 4905.
-2727 Musaev, D. G.; Figg, T. M.; Kaledin, A. L.; Chem. Soc. Rev. 2014, 43, 5009. With the progress of theoretical methods in recent years, the combination of quantum chemical computations and experiments has made great progress in the development of chemistry. Particularly, the density functional theory (DFT) method2828 Kohn, W.; Sham, L. J.; Phys. Rev. 1965, 140, 1133.

29 Dreizler, R. M.; Gross, E. K. U.; Density Function Theory; Springer: Berlin, 1990.

30 Steinmetz, M.; Grimme, S.; ChemistryOpen 2013, 2, 115.

31 Koch, W.; Holthausen, M. C.; Kaupp, M.; Angew. Chem. 2001, 113, 989.

32 Dreizler, R. M.; Gross, E. K. U.; Density Functional Approach to Time-Dependent and to Relativistic Systems; Springer: Boston, 1985.
-3333 Lipshutz, B. H.; Chrisman, W.; Noson, K.; Papa, P.; Sclafani, J. A.; Vivian, R. W.; Keith, J. M.; Tetrahedron 2000, 56, 2779. has been widely used in the study of the mechanisms, molecular interactions, and origins of selectivity in the reactions.

Here, we have employed the DFT method to investigate the mechanism of the reduction reaction at a molecular level. The aim of the present paper is focused on the following aspects: (i) the most feasible step of the whole reaction, including the rate-determining step (RDS); II the effect of the states of the palladium catalysts (Pd-Pd4) on the hydrogenation reaction (Scheme 1); and (iii) the configuration of the stilbenes (semihydrogenation products). This computational study contributes to the understanding of these hydrogenation reactions at the molecular level and can also provide some important suggestions for new hydrogenation reactions.

Scheme 1
Pd-catalyzed hydrogenation of diphenylacetylene to 1,2-diphenylethane.

Methodology

The previous computational literatures demonstrated that the density functional Minnesota 06 (M06)3434 Zhao, Y.; Truhlar, D.; Theor. Chem. Acc. 2008, 120, 215.

35 Zhao, Y.; Truhlar, D.; Acc. Chem. Res. 2008, 41, 157.

36 Bryantsev, V. S.; Diallo, M. S.; van Duin, A. C. T.; Goddard III, W. A.; J. Chem. Theory Comput. 2009, 5, 1016.
-3737 Jacquemin, D.; Perpète, E. A.; Ciofini, I.; Adamo, C.; Valero, R.; Zhao, Y.; Truhlar, D. G.; J. Chem. Theory Comput. 2010, 6, 2071. and Becke, 3-parameter, Lee-Yang-Parr (B3LYP)3838 Becke, A. D.; J. Chem. Phys. 1993, 98, 5648.,3939 Lee, C. T.; Yang, W. T.; Parr, R. G.; Phys. Rev. B: Condens. Matter Mater. Phys. 1988, 37, 785. methods performed well for catalytic reactions based on the transition metal system.3333 Lipshutz, B. H.; Chrisman, W.; Noson, K.; Papa, P.; Sclafani, J. A.; Vivian, R. W.; Keith, J. M.; Tetrahedron 2000, 56, 2779.,4040 Shi, S. L.; Xu, L. W.; Oisaki, K.; Kanai, M.; Shibasaki, M.; J. Am. Chem. Soc. 2010, 132, 6638.

41 Liu, H. Y.; Zhang, W.; He, L.; Luo, M. L.; Qin, S.; RSC Adv. 2014, 4, 5726.

42 Zhang, W.; Li, W. Y.; Qin, S.; Org. Biomol. Chem. 2012, 10, 597.

43 Hong, S.; Huber, S. M.; Gagliardi, L.; Cramer, C. C.; Tolman, W. B.; J. Am. Chem. Soc. 2007, 129, 14190.

44 Bar-Nahum, I.; Gupta, A. K.; Huber, S. M.; Ertem, M. Z.; Cramer, C. J.; Tolman, W. B.; J. Am. Chem. Soc. 2009, 131, 2812.
-4545 Cramer, C. J.; Gour, J. R.; Kinal, A.; Wloch, M.; Piecuch, P.; Shahi, A. R. M.; Gagliardi, L.; J. Chem. Phys. A 2008, 112, 3754. In the present investigations, the geometrical optimizations of all the intermediates (IM) and transition states (TS) were performed using the M06 and B3LYP methods, respectively. The palladium clusters were optimized before the coordination of DPA, the initial Pd−Pd distances of 1.5 and 3 Å were used in the optimization of Pd agglomerates and the coordinates of palladium atoms of the clusters were relaxed in following optimization. The electronic spin state of the considered Pdn clusters is the singlet state. The SDD basis set4646 Igel-Mann, G.; Stoll, H.; Preuss, H.; Mol. Phys. 1988, 65, 1321. was used for Pd atom and the 6-311+G(d,p) basis set4747 Frisch, M. J.; Pople, J. A.; Binkley, J. S.; J. Chem. Phys. 1984, 80, 3265. was employed for the rest atoms. The solvation effect was considered in the geometry optimization by a self-consistent reaction field (SCRF), in which the polarizable continuum model (PCM) implicit solvent model was employed with tetrahydrofuran (THF) as a solvent.4848 Tomasi, J.; Mennucci, B.; Cancès, E.; J. Mol. Struct.: THEOCHEM 1999, 464, 211.

49 Barone, V.; Cossi, M.; J. Phys. Chem. A 1998, 102, 1995.
-5050 Cossi, M.; Rega, N.; Scalmani, G.; Barone, V.; J. Comput. Chem. 2003, 24, 669. All geometries were optimized in vacuum by using PCM model. This model was used for single-point energy calculations based on all of the solvent-phase optimized geometries at a larger basis set (SDD for the Pd atom and 6-311+G(d,p) for other atoms). All DFT calculations were performed with the Gaussian 095151 Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery Jr., J. A.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Keith, T.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, O.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J.; Gaussian 09, Revision D.01; Gaussian, Inc., Wallingford, CT, USA, 2013. series of electronic structure programs. The vibrational frequencies were assessed at the same level of theory as that for the geometrical optimizations, to verify the stationary points as local minima (no imaginary frequencies) or transition states (unique imaginary frequency). Necessarily, using the intrinsic reaction coordinate (IRC) method,5252 Gonzalez, C.; Schlegel, H. B.; J. Chem. Phys. 1989, 90, 2154. the key transition states were confirmed to be the connection between the corresponding reactant and product. The total cartesian coordinates for all minima point and transition states in the gas-phase by B3LYP and M06 are provided in the Supplementary Information (SI) section.

Results and Discussion

Reaction mechanism

The DFT calculations predicted that the reactions over different Pd clusters share a reaction mechanism similar to that in Scheme 1, and the entire reaction path could be divided into approximately two successive stages: stage 1: the hydrogenation of DPA to stilbene by the addition of one hydrogen molecule; and stage 2: the hydrogenation of stilbene to the final product DPE with the recovery of the catalyst.

Additionally, M06 and B3LYP gave the comparable values with few discrepancies in the optimized geometries, which implies that the present calculation is reasonable for the titled reaction system. To give a concise expression, the following discussion will be based on the Pd3-catalyst system at the B3LYP(PCM)/6-311+G(d,p), SDD level unless otherwise specified. The detailed information about the optimized structures and energy profile of the hydrogenation of DPA over the Pd, Pd2 and Pd4 cluster systems is provided in the SI section.

Stage 1: hydrogenation of DPA to stilbene

The energy profile and the optimized IMs and TSs at the B3LYP/6-311+G(d,p), SDD, are shown in Figure 1.

Figure 1
3D models of various species and the energy profile in reaction stage 1 of Pd3-catalyzed hydrogenation of diphenylacetylene (DPA) system obtained at the M06/6-311+G(d,p), SDD level. Bond lengths are in Å, relative energies are in kcal mol–1 and imaginary frequencies are in cm–1.

As shown in Figure 1, the reaction is triggered by the coordination of the alkyne moiety with the Pd atoms with the formation of the Pd3-DPA complex, and the palladium clusters were optimized before its approximation to DPA. In this complex, the Pd−C distances are calculated to be 1.99-2.14 Å, and the negative Laplacian of electronic densities Ñ22 Tobisu, M.; Chatani, N.; Chem. Soc. Rev. 2008, 37, 300.r at (3, -1) bonding critical points by atoms in molecule (AIM) analysis5353 Matta, Ch. F.; Boyd, R. J.; The Quantum Theory of Atoms in Molecules: from Solid State to DNA and Drug Design; Wiley-VCH GmbH & Co. KGaA: Weinheim, 2007. in Figure 2 indicates the covalent interaction between Pd3 and DPA, demonstrating that there is a strong interaction between the substrate and the Pd3 cluster. This result could be enhanced by the lower relative free energy of 54.7 kcal mol-1 with respect to the separated substance and bare Pd3 cluster.

Figure 2
Laplacian (∇2ρ) and electronic density (ρ) of selected bond critical points (BCP) for Pd3-DPA complex were obtained by atoms in molecule (AIM) analysis.

Next, an external H2 molecule gets close to one Pd-end of the Pd3-DPA complex and leads to the yield of Pd3-IM11 Jun, C. H.; Chem. Soc. Rev. 2004, 33, 610. intermediate. In this IM, the H−H bond is slightly enlarged to 0.81 Å, and the Pd−H distances are predicted to be ca. 1.79 Å. The formation of the Pd3-IM11 Jun, C. H.; Chem. Soc. Rev. 2004, 33, 610. is calculated to be energy-favored by −7.8 kcal mol-1 from Pd3-DPA + H2, indicating that the generation of Pd3-IM11 Jun, C. H.; Chem. Soc. Rev. 2004, 33, 610. from reactants is non-spontaneous.

Then, the cleavage of the H−H bond and the insertion of one hydrogen atom into the DPA takes place through the three-membered ring transition state (TS11 Jun, C. H.; Chem. Soc. Rev. 2004, 33, 610.), leading to the formation of Pd3-IM22 Tobisu, M.; Chatani, N.; Chem. Soc. Rev. 2008, 37, 300.. In this species, the three-membered ring is characterized by an elongated C−H bond of 1.66 Å, Pd−C bond of 2.08 Å and Pd−H bond of 1.56 Å. For Pd3-IM22 Tobisu, M.; Chatani, N.; Chem. Soc. Rev. 2008, 37, 300., the C−H distance is calculated to be 1.08 Å, indicating that the migration of hydrogen atom is complete, and the corresponding C−H bond has been formed. In addition, the rest of the alkyne coordinates with the two Pd ends of the Pd3 cluster with the Pd−C distances of 2.09 Å. The calculations show that the above step has a moderate energy barrier of 26.2 kcal mol-1.

From the Pd3-IM22 Tobisu, M.; Chatani, N.; Chem. Soc. Rev. 2008, 37, 300., the next hydrogen atom transfers to another side of the alkyne moiety of the DPA via transition state TS22 Tobisu, M.; Chatani, N.; Chem. Soc. Rev. 2008, 37, 300., which leads to the formation of Pd3-IM33 Wang, A.; Jiang, H.; J. Am. Chem. Soc. 2008, 130, 5030.. As shown in Figure 1, cis-TS22 Tobisu, M.; Chatani, N.; Chem. Soc. Rev. 2008, 37, 300. and trans-TS22 Tobisu, M.; Chatani, N.; Chem. Soc. Rev. 2008, 37, 300. are located on the energy profile: the former leads to the stilbene moiety in the E-configuration in the cis-IM33 Wang, A.; Jiang, H.; J. Am. Chem. Soc. 2008, 130, 5030. intermediate, and the latter leads to the Z-configuration in the trans-IM33 Wang, A.; Jiang, H.; J. Am. Chem. Soc. 2008, 130, 5030. intermediate. Each TS22 Tobisu, M.; Chatani, N.; Chem. Soc. Rev. 2008, 37, 300. features the four-membered ring transition state with the Pd−Pd bond of ca. 2.70 Å, Pd−C bond of ca. 2.00 Å, C−H bond of ca. 1.50 Å and H−Pd bond of 1.70 Å.

As a result, cis- and trans-IM33 Wang, A.; Jiang, H.; J. Am. Chem. Soc. 2008, 130, 5030. are identified at the end of stage 1 of the reaction mechanism. For each IM33 Wang, A.; Jiang, H.; J. Am. Chem. Soc. 2008, 130, 5030., the C−H distance is calculated to be ca. 1.09 Å, indicating that the migration of the hydrogen atom is finished, and the corresponding C−H bond has been formed. In addition, the alkene that is generated coordinates with the Pd3 cluster with the Pd−C distances in the range of 2.14-2.15 Å, demonstrating the interaction between the carbon of the phenyl ring and the Pd3 cluster. As a result, the stilbene adsorbed on the Pd3 cluster is generated at the end of this reaction state after TS22 Tobisu, M.; Chatani, N.; Chem. Soc. Rev. 2008, 37, 300., and the energy barrier of the step via TS22 Tobisu, M.; Chatani, N.; Chem. Soc. Rev. 2008, 37, 300. is calculated to be 16.3 kcal mol-1 for the cis-path and 12.7 kcal mol-1 for the trans-path.

Here, it should be emphasized that the configuration of the stilbene intermediate is closely dependent on the step via the TS22 Tobisu, M.; Chatani, N.; Chem. Soc. Rev. 2008, 37, 300., and therefore the TS22 Tobisu, M.; Chatani, N.; Chem. Soc. Rev. 2008, 37, 300. appears to be the stereo-controlling transition state in the reaction stage 1. The B3LYP calculation predicts the energy barrier via the trans-TS22 Tobisu, M.; Chatani, N.; Chem. Soc. Rev. 2008, 37, 300. is energy-favored by 3.6 kcal mol-1 over its competing transition state of cis-TS22 Tobisu, M.; Chatani, N.; Chem. Soc. Rev. 2008, 37, 300.. The trans-TS22 Tobisu, M.; Chatani, N.; Chem. Soc. Rev. 2008, 37, 300. links to the formation of stilbene in the Z-configuration and the cis-TS22 Tobisu, M.; Chatani, N.; Chem. Soc. Rev. 2008, 37, 300. to the formation of stilbene in the E-configuration, according to the absolute rate theory and the formula as follows:

(1) trans : cis = k trans : k cis = exp Δ G trans = RT : exp Δ G cis = RT

where k is the rate constant, T is the absolute temperature, R is the gas constant, and DG is the activation Gibbs free energy.

The stereo-selectivity of stilbene is predicted to be Z/ E > 1:99 at 298.15 K indicating that the generation of Z-stilbene might be advantageous in thermodynamics. However, considered the error generated from the DFT calculations, this reason may be not accurate enough to account for the selectivity of cis- and trans-IMs in the experiment.

Stage 2: hydrogenation of stilbene to DPE

In stage 2, the reaction must undergo the hydrogenation from stilbene to DPE over Pd-catalyst with the addition of another H2 molecule after the reaction stage 1 has been completed. The results are shown in Figure 3.

Figure 3
3D models of various species and the energy profile in reaction stage 2 of Pd3-catalyzed hydrogenation of diphenylacetylene (DPA) system obtained at the M06/6-311+G(d,p), SDD level. Bond lengths are in Å, relative energies are in kcal mol–1 and imaginary frequencies are in cm–1.

In this stage, an external H2 molecule coordinates to one Pd-end of two Pd3-IM33 Wang, A.; Jiang, H.; J. Am. Chem. Soc. 2008, 130, 5030. complexes and leads to the yield of cis- or trans-IM44 Chinchilla, R.; Najera, C.; Chem. Rev. 2014, 114, 1783. intermediate. In each IM44 Chinchilla, R.; Najera, C.; Chem. Rev. 2014, 114, 1783., the H−H bond is slightly enlarged to ca. 0.85-0.86 Å, and the Pd−H distances are calculated to be ca. 1.70 Å. The formation of the cis-IM44 Chinchilla, R.; Najera, C.; Chem. Rev. 2014, 114, 1783. is predicted to be energy-favored by 9.7 kcal mol-1 from the separated cis-IM33 Wang, A.; Jiang, H.; J. Am. Chem. Soc. 2008, 130, 5030. and H2, and the formation of the trans-IM44 Chinchilla, R.; Najera, C.; Chem. Rev. 2014, 114, 1783. is calculated to be favored by 12.0 kcal mol-1 from the separated trans-IM33 Wang, A.; Jiang, H.; J. Am. Chem. Soc. 2008, 130, 5030. and H2 in relative energy, indicating that the generation of IM44 Chinchilla, R.; Najera, C.; Chem. Rev. 2014, 114, 1783. is spontaneous.

Next, the breakage of the H−H bond and the insertion of one hydrogen atom into the stilbene moiety occur through the four-membered ring transition state (cis- or trans-TS44 Chinchilla, R.; Najera, C.; Chem. Rev. 2014, 114, 1783.) with the generation of the IM55 Stang, P. J.; Diederich, F.; Modern Acetylene Chemistry; VCH: Weinheim, 1995.. The four-membered ring is characterized by an elongated C−H bond of 1.66-1.68 Å, Pd−H bond of 1.61-1.62 Å, Pd−C bond of ca. 2.10 Å, and C−C bond of 1.43 Å. For IM55 Stang, P. J.; Diederich, F.; Modern Acetylene Chemistry; VCH: Weinheim, 1995., the calculations determine the optimized C−H distance to be 1.09 Å, which implies that the transfer of the hydrogen atom is completed with the formation of the corresponding sp33 Wang, A.; Jiang, H.; J. Am. Chem. Soc. 2008, 130, 5030. carbon. In addition, the rest of the alkene part coordinates the two Pd atom Pd3 cluster with the Pd−C distances of ca. 2.09 Å. The calculations predict that the above step has a moderate energy barrier of 12.5 kcal mol-1 in relative energy.

Then, from the IM55 Stang, P. J.; Diederich, F.; Modern Acetylene Chemistry; VCH: Weinheim, 1995., the reaction goes through the four-membered ring transition state (TS55 Stang, P. J.; Diederich, F.; Modern Acetylene Chemistry; VCH: Weinheim, 1995.), leading to the formation of IM66 Diederich, F.; Stang, P. J.; Tykwinski, R. R.; Acetylene Chemistry; Wiley-VCH: Weinheim, 2005.. The three-membered ring is composed of a Pd−C bond of 2.13 Å, a C−H bond of 1.73 Å and a H−Pd bond of 1.61 Å. For IM66 Diederich, F.; Stang, P. J.; Tykwinski, R. R.; Acetylene Chemistry; Wiley-VCH: Weinheim, 2005., the C−H distance is calculated to be 1.09 Å, indicating that the migration of the hydrogen atom is completed, and the corresponding C−H bond is formed. In addition, the rest of the alkyne coordinates with the three Pd atom Pd3 cluster with the Pd−C distances of ca. 2.12 Å. The calculations indicate that the above step has a moderate energy barrier of 13.4 kcal mol-1.

Finally, the release of DPE from IM66 Diederich, F.; Stang, P. J.; Tykwinski, R. R.; Acetylene Chemistry; Wiley-VCH: Weinheim, 2005. takes place with the recovery of the Pd3-cluster. This step is predicted to be endothermic by ca. 21.9 kcal mol-1 without any barrier. As shown in Figures 1 and 3, the entire reaction is clearly favored by 54.7 kcal mol-1 in Gibbs free energy, and the largest barrier of 26.2 kcal mol-1 corresponds to the step from the IM11 Jun, C. H.; Chem. Soc. Rev. 2004, 33, 610. ® TS11 Jun, C. H.; Chem. Soc. Rev. 2004, 33, 610., indicating that the first hydrogenation step should be the rate determining step (RDS).

Comparison of Pd-Pd4 catalyzed reactions

The reaction mechanisms over the Pd-Pd4 cluster are similar, but the properties of the energy profiles vary: for the Pd-Pd3 systems, both the M06 and B3LYP calculations predict that the step of IM11 Jun, C. H.; Chem. Soc. Rev. 2004, 33, 610. ® TS11 Jun, C. H.; Chem. Soc. Rev. 2004, 33, 610. is the RDS with the barrier of ca. 22-29 kcal mol-1 (see SI section), while the RDS of the Pd4 system turns to the final hydrogen migration step in IM55 Stang, P. J.; Diederich, F.; Modern Acetylene Chemistry; VCH: Weinheim, 1995. ® TS55 Stang, P. J.; Diederich, F.; Modern Acetylene Chemistry; VCH: Weinheim, 1995. with the giant barrier larger than 33 kcal mol-1. It is found that B3LYP energies could be considerably underestimated, and this is in accordance with the literature.5454 de Souza, L. A.; Nogueira, C. A. S.; Lopes, J. F.; dos Santos, H. F.; de Almeida, W. B.; J. Phys. Chem. C 2015, 119, 8394.,5555 de Souza, L. A.; da Silva, A. M.; dos Santos, H. F.; de Almeida, W. B.; RSC Adv. 2017, 7, 13212. The 3D models of various species and the energy profile are shown in Figures 4 and 5.

Figure 4
3D models of various species and the energy profile in reaction stage 1 of Pd4-catalyzed hydrogenation of diphenylacetylene (DPA) system obtained at the M06/6-311+G(d,p), SDD level. Bond lengths are in Å, relative energies are in kcal mol–1 and imaginary frequencies are in cm–1.

Figure 5
3D models of various species and the energy profile in reaction stage 2 of Pd4-catalyzed hydrogenation of diphenylacetylene (DPA) system obtained at the M06/6-311+G(d,p), SDD level. Bond lengths are in Å, relative energies are in kcal mol–1 and imaginary frequencies are in cm–1.

Analysis of turnover frequency (TOF) in the catalytic cycle

In addition, based on the transition state theory and energetic span model,5656 Amatore, C.; Jutand, A.; J. Organomet. Chem. 1999, 576, 254. we evaluated the theoretical turnover frequency (TOF) of the catalytic cycle via the intermolecular and H-transfer mechanism catalyzed by the Pdn (n = 1-4) catalyst. In equations 2 and 3,5757 Kozuch, S.; Shaik, S.; Acc. Chem. Res. 2011, 44, 101.

58 Kozuch, S.; Shaik, S.; J. Am. Chem. Soc. 2006, 128, 3355.

59 Kozuch, S.; Shaik, S.; J. Phys. Chem. A 2008, 112, 6032.
-6060 Uhe, A.; Kozuch, S.; Shaik, S.; J. Comput. Chem. 2011, 32, 978. the dE (energy span) is the energy difference between the summit and the through of the catalytic cycle. GTDTS and GTDI are defined as the Gibbs free energy of the TOF-determining transition state (TDTS) and the TOF-determining intermediate (TDI), and DGr is the global free energy of the whole cycle.5656 Amatore, C.; Jutand, A.; J. Organomet. Chem. 1999, 576, 254.,6161 Qi, T.; Yang, H.-Q.; Whitfield, D. M.; Yu, K.; Hu, C.-W.; J. Phys. Chem. A 2016, 120, 918.

(2) TOF K B T h e δ E / RT

where KB is Boltzmann’s constant, h is Planck’s constant, T is the absolute temperature, R is the gas constant, and

(3) δ E = E TDTS E TDI , if TDTS appears after TDI
(4) δ E = E TDTS E TDI + Δ G r , if TDTS appears before TDI

As shown in Table 1, the intermediate IM11 Jun, C. H.; Chem. Soc. Rev. 2004, 33, 610. + H2 and IM55 Stang, P. J.; Diederich, F.; Modern Acetylene Chemistry; VCH: Weinheim, 1995. is predicted to be TDI, and the H-transfer transition states TS11 Jun, C. H.; Chem. Soc. Rev. 2004, 33, 610. and TS55 Stang, P. J.; Diederich, F.; Modern Acetylene Chemistry; VCH: Weinheim, 1995. are TDTS for the whole cycle of the hydrogenation reaction. As expected, the TOFs of the catalytic cycles involving Pdn (n = 1-4) catalyzed H-transfer pathways are significantly higher than that of the intramolecular H-transfer pathways. Moreover, Pd3 exhibits better catalytic performance when the catalytic reaction occurs along the Pd3 reaction pathway, with TOF being 0.25 s-1.

Table 1
Turnover frequency (TOF) of the catalytic cycle for hydrogenation reaction catalyzed by Pdn (n = 1-4) along four paths

Conclusions

The mechanism of the hydrogenation reaction of diphenylacetylene catalyzed by Pd-Pd4 species was investigated by using DFT method. The entire reaction mechanism in the present investigation is predicted to be composed of two processes: stage 1: the hydrogenation of DPA to stilbene with the addition of one hydrogen molecule via TS11 Jun, C. H.; Chem. Soc. Rev. 2004, 33, 610. and TS22 Tobisu, M.; Chatani, N.; Chem. Soc. Rev. 2008, 37, 300.; and stage 2: the hydrogenation of stilbene via TS33 Wang, A.; Jiang, H.; J. Am. Chem. Soc. 2008, 130, 5030. and TS44 Chinchilla, R.; Najera, C.; Chem. Rev. 2014, 114, 1783. to the final product DPE with the recovery of the catalyst.

The calculation on the diphenylacetylene system indicates that the hydrogenation could take place in existence of four catalysis of Pd, Pd2, Pd3 and Pd4. For the Pd-Pd3 systems, both the M06 and B3LYP calculations predict that the step of IM11 Jun, C. H.; Chem. Soc. Rev. 2004, 33, 610. ® TS11 Jun, C. H.; Chem. Soc. Rev. 2004, 33, 610. is the RDS with the barrier of the largest Gibbs energy, and the RDS of the Pd4 system turns to the step of the final hydrogen-migration step in IM55 Stang, P. J.; Diederich, F.; Modern Acetylene Chemistry; VCH: Weinheim, 1995. ® TS55 Stang, P. J.; Diederich, F.; Modern Acetylene Chemistry; VCH: Weinheim, 1995. with the giant barrier larger than 33.1 kcal mol-1. The calculation reproduces the major product E-stilbene intermediate, and exhibits the smallest RDS energy barrier for the Pd2- and Pd3-catalyzed system. Therefore, Pd2- and Pd3-catalysts might be the most active and effective catalysis species among the four clusters. This computational study is expected to provide a full understanding of this hydrogenation reaction at the molecular level and can also provide some important suggestions for the rational design and synthesis of the new Pd-catalytic hydrogenation reactions.

Acknowledgments

This research was funded by Academy-School Cooperation Project S18H321-Q.

Supplementary Information

The total Cartesian coordinates for all minima point and transition states in the gas phase by B3LYP and M06 are available free of charge at http://jbcs.sbq.org.br as PDF file.

References

  • 1
    Jun, C. H.; Chem. Soc. Rev. 2004, 33, 610.
  • 2
    Tobisu, M.; Chatani, N.; Chem. Soc. Rev. 2008, 37, 300.
  • 3
    Wang, A.; Jiang, H.; J. Am. Chem. Soc. 2008, 130, 5030.
  • 4
    Chinchilla, R.; Najera, C.; Chem. Rev. 2014, 114, 1783.
  • 5
    Stang, P. J.; Diederich, F.; Modern Acetylene Chemistry; VCH: Weinheim, 1995
  • 6
    Diederich, F.; Stang, P. J.; Tykwinski, R. R.; Acetylene Chemistry; Wiley-VCH: Weinheim, 2005
  • 7
    Schmidt, E. Y.; Tatarinova, I. V.; Semenova, N. V.; Protsuk, N. I.; Ushakov, I. A.; Trofimov, B. A.; J. Org. Chem 2018, 83, 10272.
  • 8
    Borodziński, A.; Bond, G. C.; Catal. Rev. 2008, 50, 379.
  • 9
    Molnár, Á.; Sárkány, A.; Varga, M.; J. Mol. Catal. A: Chem. 2001, 173, 185.
  • 10
    Barrios-Francisco, R.; García, J. J.; Appl. Catal., A 2010, 385, 108.
  • 11
    Espinal-Viguri, M.; Neale, S. E.; Coles, N. T.; Macgregor, S. A.; Webster, R. L.; J. Am. Chem. Soc. 2019, 141, 572.
  • 12
    Shao, Z.; Fu, S.; Wei, M.; Zhou, S.; Liu, Q.; Angew. Chem., Int. Ed. 2016, 55, 14653.
  • 13
    Zhou, Q.; Zhang, L.; Meng, W.; Feng, X.; Yang, J.; Du, H.; Org. Lett. 2016, 18, 5189.
  • 14
    Sperger, T.; Sanhueza, I. A.; Kalvet, I.; Schoenebeck, F.; Chem. Rev. 2015, 115, 9532.
  • 15
    Kojima, Y.; Matsuoka, T.; Takahashi, H.; J. Mater. Sci. Lett. 1996, 15, 1543.
  • 16
    Li, S. H.; Li, L. C.; Xu, C. H.; Chin. Chem. Lett 2012, 23, 69.
  • 17
    Manna, S.; Antonchick, A. P.; ChemSusChem 2019, 12, 3094.
  • 18
    Domínguez-Domínguez, S.; Berenguer-Murcia, Á.; Linares-Solano, Á.; Cazorla-Amorós, D.; J. Catal. 2008, 257, 87.
  • 19
    Nishio, R.; Sugiura, M.; Kobayashi, S.; Org. Biomol. Chem. 2006, 4, 992.
  • 20
    Maazaoui, R.; Abderrahim, R.; Chemla, F.; Ferreira, F.; Perez-Luna, A.; Jackowski, O.; Org. Lett. 2018, 20, 7544.
  • 21
    Wu, Z.; Cherkasov, N.; Cravotto, G.; Borretto, E.; Ibhadon, A. O.; Medlock, J.; Bonrath, W.; ChemCatChem 2015, 7, 952.
  • 22
    King, A. K.; Buchard, A.; Mahon M. F.; Webster, R. L.; Chem. - Eur. J. 2015, 21, 15960.
  • 23
    Sameera, W. M. C.; Maseras, F.; Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2012, 2, 375.
  • 24
    Bonney, K. J.; Schoenebeck, F.; Chem. Soc. Rev. 2014, 43, 6609.
  • 25
    Holder, J. C.; Zou, L.; Marziale, A. N.; Liu, P.; Lan, Y.; Gatti, M.; Kikushima, K.; Houk, K. N.; Stoltz, B. M.; J. Am. Chem. Soc 2013, 135, 14996.
  • 26
    Lan, Y.; Houk, K. N.; J. Org. Chem. 2011, 76, 4905.
  • 27
    Musaev, D. G.; Figg, T. M.; Kaledin, A. L.; Chem. Soc. Rev. 2014, 43, 5009.
  • 28
    Kohn, W.; Sham, L. J.; Phys. Rev. 1965, 140, 1133.
  • 29
    Dreizler, R. M.; Gross, E. K. U.; Density Function Theory; Springer: Berlin, 1990
  • 30
    Steinmetz, M.; Grimme, S.; ChemistryOpen 2013, 2, 115.
  • 31
    Koch, W.; Holthausen, M. C.; Kaupp, M.; Angew. Chem 2001, 113, 989.
  • 32
    Dreizler, R. M.; Gross, E. K. U.; Density Functional Approach to Time-Dependent and to Relativistic Systems; Springer: Boston, 1985
  • 33
    Lipshutz, B. H.; Chrisman, W.; Noson, K.; Papa, P.; Sclafani, J. A.; Vivian, R. W.; Keith, J. M.; Tetrahedron 2000, 56, 2779.
  • 34
    Zhao, Y.; Truhlar, D.; Theor. Chem. Acc. 2008, 120, 215.
  • 35
    Zhao, Y.; Truhlar, D.; Acc. Chem. Res. 2008, 41, 157.
  • 36
    Bryantsev, V. S.; Diallo, M. S.; van Duin, A. C. T.; Goddard III, W. A.; J. Chem. Theory Comput. 2009, 5, 1016.
  • 37
    Jacquemin, D.; Perpète, E. A.; Ciofini, I.; Adamo, C.; Valero, R.; Zhao, Y.; Truhlar, D. G.; J. Chem. Theory Comput 2010, 6, 2071.
  • 38
    Becke, A. D.; J. Chem. Phys. 1993, 98, 5648.
  • 39
    Lee, C. T.; Yang, W. T.; Parr, R. G.; Phys. Rev. B: Condens. Matter Mater. Phys. 1988, 37, 785.
  • 40
    Shi, S. L.; Xu, L. W.; Oisaki, K.; Kanai, M.; Shibasaki, M.; J. Am. Chem. Soc 2010, 132, 6638.
  • 41
    Liu, H. Y.; Zhang, W.; He, L.; Luo, M. L.; Qin, S.; RSC Adv 2014, 4, 5726.
  • 42
    Zhang, W.; Li, W. Y.; Qin, S.; Org. Biomol. Chem 2012, 10, 597.
  • 43
    Hong, S.; Huber, S. M.; Gagliardi, L.; Cramer, C. C.; Tolman, W. B.; J. Am. Chem. Soc. 2007, 129, 14190.
  • 44
    Bar-Nahum, I.; Gupta, A. K.; Huber, S. M.; Ertem, M. Z.; Cramer, C. J.; Tolman, W. B.; J. Am. Chem. Soc. 2009, 131, 2812.
  • 45
    Cramer, C. J.; Gour, J. R.; Kinal, A.; Wloch, M.; Piecuch, P.; Shahi, A. R. M.; Gagliardi, L.; J. Chem. Phys. A 2008, 112, 3754.
  • 46
    Igel-Mann, G.; Stoll, H.; Preuss, H.; Mol. Phys. 1988, 65, 1321.
  • 47
    Frisch, M. J.; Pople, J. A.; Binkley, J. S.; J. Chem. Phys. 1984, 80, 3265.
  • 48
    Tomasi, J.; Mennucci, B.; Cancès, E.; J. Mol. Struct.: THEOCHEM 1999, 464, 211.
  • 49
    Barone, V.; Cossi, M.; J. Phys. Chem. A 1998, 102, 1995.
  • 50
    Cossi, M.; Rega, N.; Scalmani, G.; Barone, V.; J. Comput. Chem. 2003, 24, 669.
  • 51
    Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery Jr., J. A.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Keith, T.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, O.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J.; Gaussian 09, Revision D.01; Gaussian, Inc., Wallingford, CT, USA, 2013
  • 52
    Gonzalez, C.; Schlegel, H. B.; J. Chem. Phys. 1989, 90, 2154.
  • 53
    Matta, Ch. F.; Boyd, R. J.; The Quantum Theory of Atoms in Molecules: from Solid State to DNA and Drug Design; Wiley-VCH GmbH & Co. KGaA: Weinheim, 2007.
  • 54
    de Souza, L. A.; Nogueira, C. A. S.; Lopes, J. F.; dos Santos, H. F.; de Almeida, W. B.; J. Phys. Chem. C 2015, 119, 8394.
  • 55
    de Souza, L. A.; da Silva, A. M.; dos Santos, H. F.; de Almeida, W. B.; RSC Adv. 2017, 7, 13212.
  • 56
    Amatore, C.; Jutand, A.; J. Organomet. Chem. 1999, 576, 254.
  • 57
    Kozuch, S.; Shaik, S.; Acc. Chem. Res. 2011, 44, 101.
  • 58
    Kozuch, S.; Shaik, S.; J. Am. Chem. Soc. 2006, 128, 3355.
  • 59
    Kozuch, S.; Shaik, S.; J. Phys. Chem. A 2008, 112, 6032.
  • 60
    Uhe, A.; Kozuch, S.; Shaik, S.; J. Comput. Chem. 2011, 32, 978.
  • 61
    Qi, T.; Yang, H.-Q.; Whitfield, D. M.; Yu, K.; Hu, C.-W.; J. Phys. Chem. A 2016, 120, 918.

Publication Dates

  • Publication in this collection
    09 Oct 2020
  • Date of issue
    Oct 2020

History

  • Received
    23 Dec 2019
  • Accepted
    18 June 2020
Sociedade Brasileira de Química Instituto de Química - UNICAMP, Caixa Postal 6154, 13083-970 Campinas SP - Brazil, Tel./FAX.: +55 19 3521-3151 - São Paulo - SP - Brazil
E-mail: office@jbcs.sbq.org.br