Next Article in Journal
Application of POCOP Pincer Nickel Complexes to the Catalytic Hydroboration of Carbon Dioxide
Previous Article in Journal
Catalytic Co-Pyrolysis of Kraft Lignin with Refuse-Derived Fuels Using Ni-Loaded ZSM-5 Type Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Synthesis and Characterization of Cationic Tetramethyl Tantalum(V) Complex

1
King Abdullah University of Science & Technology, KAUST Catalysis Center (KCC), Thuwal 23955-6900, Saudi Arabia
2
Department of Chemistry, National Institute of Technology Calicut, NIT Campus, Kozhikode 673 601, India
3
College of Chemistry & Chemical Engineering, Central South University, Changsha 410083, China
4
Institut de Química Computacional i Catàlisi and Departament de Química, Universitat de Girona, c/ Mª Aurèlia Capmany 69, 17003 Girona, Catalonia, Spain
*
Authors to whom correspondence should be addressed.
Catalysts 2018, 8(11), 507; https://doi.org/10.3390/catal8110507
Submission received: 1 October 2018 / Revised: 26 October 2018 / Accepted: 28 October 2018 / Published: 1 November 2018
(This article belongs to the Section Computational Catalysis)

Abstract

:
A novel method for the synthesis of the homogeneous homoleptic cationic tantalum(V)tetramethyl complex [(TaMe4)+ MeB(C6F5)3−] from neutral tantalumpentamethyl (TaMe5) has been described, by direct demethylation using B(C6F5)3 reagent. The aforesaid higher valent cationic tantalum complex was characterized precisely by liquid state 1H-NMR, 13C-NMR, and 1H-13C-NMR correlation spectroscopy.

Graphical Abstract

1. Introduction

Tantalum alkyl complexes are of vital importance in current organometallic chemistry, from the point of view of stoichiometric reactions and mainly in catalysis [1,2,3,4]. Since 1974, tantalumpentamethyl (TaMe5, 1) has the honour to be the simplest homoleptic complex of this class of compounds [5,6]. 1 was discovered by Richard Schrock and the structure was not unravelled until 1992 by Albright [7] and Haaland et al. [8]. TaMe5 is rather unstable thermally and, consequently, 1 is highly prone to autocatalytically degradation [6]. Recently, in our previous communication [9], we checked that grafting this unstable TaMe5 complex to the silica surface by a Surface Organometallic Chemistry (SOMC) strategy enhances its thermal stability due to the formation of the stable grafted (≡Si-O-)TaMe4. The latter complex proved to be a nice precursor for alkane metathesis, leading to the formation of a surface monopodal tantalum methylidene dimethyl catalyst [9,10].
With the precedents of Buchmeiser and co-workers who developed the first cationic tungsten-oxo-alkylidene-NHC complex (NHC = N-heterocyclic carbene) [11,12], where the NHC was introduced to stabilize the cationic metal centre, together with the also tungsten based cationic complex WMe5+ [13], to improve the catalysts in terms of reactivity and selectivity, and taking into account the idea of increasing the electrophilicity on metal centre, we deepened into the idea of generating the cationic species of TaMe5. This change planned to enter the field of predictive catalysis. Unfortunately, the drawback for tantalum with respect to tungsten is that the similar cationic Ta-complexes are scarce in literature [14,15,16], whereas there are several examples with stable cationic tantalum complexes containing one or more nitrogen [17], phosphorous [18] and cyclopentadienyl containing ligands [19,20,21].

2. Results

As a part of our continuing research program to explore any novel application of tantalumpentamethyl complex, we report here a method for the synthesis of the homogeneous cationic [TaMe4]+ together with [MeB(C6F5)3] anion (2), starting from the neutral TaMe5. The experimental procedure was straightforward: The simple mixture of bulky and non-coordinating boron Lewis acid, B(C6F5)3, with 1, at low temperature (−40 °C) generates the cationic complex 2 (see Scheme 1), since above this temperature the undesired phenomenon of degradation appears in a rather facile way.
Preliminarily, we tried the synthesis of TaMe4+ using an organic Lewis acid, tris(pentafluorophenyl)boron. The latter species is known to form a non-nucleophilic anion after demethylation reaction, a strategy that it is valid also with titanium [22,23], zirconium [24], or hafnium [24,25,26]. When the reaction was monitored by nuclear magnetic resonance (NMR) spectroscopy, at −40 °C, the peak corresponding to TaMe5 at 0.9 ppm in 1H-NMR (Figure 1a) almost completely disappeared in less than 15 min due to the fast reaction between TaMe5 and B(C6F5)3, and two new peaks at 2.6 ppm and 0.5 ppm appeared (Figure 1b) [27].
Based on the above observations one can assume two possibilities (Scheme 2); path I consists of simple demethylation by the B(C6F5)3 from TaMe5 leading to the formation of cationic tantalum complex; whereas in path II there would be a chance for the formation of a neutral tantalum complex by simple ligand exchange. Path II could take place in a one-step mechanism from the original TaMe5, however it seems more plausible after path I.
Looking in detail at the 1H-NMR chemical shift at +0.5 ppm, it is confirmed that it corresponds to [Me-B(C6F5)3], in agreement with past studies [13,25,26]. The peak at 2.6 ppm corresponded to the cationic tantalum complex [TaMe4]+, upfield with respect to the signal at 0.9 ppm assigned to the resonance for the methyl proton of TaMe5.
Switching to 13C-NMR, the spectrum also clearly showed that the peak corresponding to TaMe5 at 82.4 ppm (Figure 1c) was completely replaced by two peaks at 110.9 ppm [TaMe4]+ and 10.8 ppm (Figure 1d) corresponding to the [Me-B(C6F5)3] anion which is in good agreement with the literature value reported for the [Me-B(C6F5)3] anion [26]. We also found in the 1H-13C correlation spectra that the peak at 0.5 ppm in 1H-NMR correlates with the peak at 10.8 ppm in 13C-NMR and the peak at 2.6 ppm in 1H-NMR correlated with the peak at 110.9 ppm in 13C-NMR (Supplementary Materials Figure S1). In separate experiments we synthesised 13C labelled Ta(13CH3)5 and upon treatment with B(C6F5)3 in dichloromethane solution, we identified incorporation of the labelled methyl in [(13CH3)B(C6F5)3] anion in the final product. The above experiment clearly indicates that in [MeB(C6F5)3], ‘Me’ came from TaMe5. These spectroscopic data strongly support the formation of cationic complex 2 and is in favour of path I. The ionic product generated by the reaction was fairly stable at temperatures below −40 °C and in the absence of light. However, upon warming a dichloromethane solution of 2 from −40 °C to 0 °C, a very fast decomposition of the cationic complex was observed with the release of gaseous methane.
Density Functional Theory (DFT) calculations (M06/Def2TZVPP(smd)//PBE0-d3bj/SVP [28], at temperature = 298.15 ºC and pressure = 1354 atm [29,30,31] to remove the overestimation of the entropy [32,33,34]) were envisaged to unravel the mechanistic insights of the first demethylation of TaMe5. The energy barrier to overcome was found to be 6.8 kcal/mol (Figure 2a), releasing 15.4 kcal/mol (see Table 1), thus the favourable kinetic and thermodynamic character of the reaction, respectively, confirmed the observed experimental facile demethylation. Furthermore, the energy barrier for this reaction was even 2.3 kcal/mol lower with respect to the recent homologous process studied for WMe6, together with a release of 1.6 kcal/mol more energy as well, confirms the experimental more facile demethylation by Ta than W based complexes [13]. It is worth pointing out that for TaMe5, the radical demethylation was not in competition with the anion demethylation, because TaMe4·radical was located 44.1 kcal/mol over TaMe5, and thus enormously disfavoured by 25.3 kcal/mol with respect to the homologous WMe5. Surprisingly, TaMe5 is able to allocate easily a sixth anion like chloride, with an energy stabilization of 5.8 kcal/mol, compared with the destabilisation by 3.6 kcal/mol for WMe6, due to steric hindrance in the metal centre of the WMe6Cl anion.
Here, steric effects play a key role, and the more sterically crowded W centre provides less space for a seventh coordination around the metal centre, and at the same time facilitates the loose of a methyl ligand. To evaluate the steric hindrance around the metal, SambVca steric maps were used [35,36,37]. The %Vbur was 74.5% for TaMe5 and 79.9% for WMe6. It is worth pointing out that the quadrants for the latter hexamethylated complex were not equally occupied (see Figure 3), ranging from 75.8% to 84.0%, with 76.1% and 83.8% in between. Anyway, none of the quadrants was less occupied than the equally distributed TaMe5 ones [38].
On the other hand, the substitution of a methyl ligand by a perfluorobenzene ligand from B(C6F5)3, i.e., path II in Scheme 2, requires to overcome an energy barrier that is placed at 26.8 kcal/mol with respect to TaMe5. However, from [MeB(C6F5)3] an unaffordable energy barrier of 41.9 kcal/mol must be overcome (see Figure 2b). The overall reaction pathway following paths I and II in Scheme 2 is displayed in Figure 4. Consequently, the huge kinetic cost of path II confirms that once TaMe5 is demethylated, the next cation intermediate TaMe4+ cannot recover an anionic ligand, i.e., (C6F5), from [MeB(C6F5)3], even though the next intermediate TaMe4(C6F5) is 11.8 kcal/mol lower in energy than TaMe5.
Switching to the grafted system (≡Si-O-)TaMe4, the demethylation process requires to overcome an energy barrier of 9.8 kcal/mol (see Figure 5 for the corresponding transition state) and the next intermediate is 5.1 kcal/mol lower in energy than the initial supported catalyst. Here the difference with tungsten is significant, because apart from facing a more facile barrier by 4.8 kcal/mol, the relative stability of the next intermediate excludes the reversibility of the demethylation observed only for W. Overall, kinetically and thermodynamically, the demethylation process is more prone with Ta based systems. Consequently, we then prepared the surface complex of TaMe5 after grafting on to the dehydroxylated Aerosil SiO2(700) (Figure S2). Synthesized (≡Si-O-)TaMe4 was treated with B(C6F5)3 in order to get a cationic tantalum-methyl complex anchored on the surface of silica, [(≡Si-O-)TaMe3]+. However, after several attempts (varying reaction temperature, reaction time and solvents) we were unable to identify any well-defined heterogeneous cationic tantalum methyl complex [13], except for a mixture of decomposed tantalum methyl complex with a 13C-NMR signal corresponding to the formation of the anionic [MeB(C6F5)3] (see Figures S3 and S4), indicating a probable formation of surface cationic species. The surface cationic complex decomposed much more rapidly, and it was difficult to characterize it by solid state NMR. This may be due to the unstable nature of 8 e [(≡Si-O-)TaMe3]+ complex.

3. Conclusions

In summary, the cationic homoleptic tantalum (V)-methyl complex has been synthesized via a straightforward strategy to address the will to get a more electrophilic metal centre, generating the first tantalum-based species with such a property. The delicate equilibrium between stability and degradation was enforced to work at low temperature, and the high oxidation state cationic tantalum complex 2 was defined by liquid 1H-NMR, 13C-NMR and 1H-13C correlation spectroscopy; together with DFT calculations. Ongoing experimental studies are being undertaken in order to understand the null activity in alkane metathesis of the grafted tantalum system and will be disclosed in due time.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/8/11/507/s1, Figure S1: (a) Two-dimensional (2D) liquid-state 1H-13C heteronuclear single quantum correlation (HSQC) NMR spectrum of [TaMe4+ MeB(C6F5)3] in CD2Cl2 recorded at −40 °C; Figure S2: (a) One-dimensional (1D) 1H MAS NMR spectrum of (≡Si-O-)TaMe4 acquired at 600 MHz with a 22 kHz MAS frequency, a repetition delay of 5 s, and 8 scans. (b) 13C CP MAS NMR spectrum of (≡Si-O-)TaMe4 (acquired at 400 MHz) with a 10 kHz MAS frequency, 10,000 scans, a 4-s repetition delay, and a 2-ms contact time. Exponential line broadening of 80 Hz was applied prior to Fourier transformation. 13C CP MAS spectra were acquired at natural abundance; Figure S3. (c) One-dimensional (1D) 13C CP MAS NMR spectrum of [(≡Si-O-)TaMe4] (acquired at 400 MHz) with a 10 kHz MAS frequency, (d) 13C CP MAS spectra of [(≡Si-O-)TaMe4] when decomposed on the surface at room temperature during the prolonged NMR measurement. 13C CP MAS spectra were acquired at natural abundance; Figure S4. (e) One-dimensional (1D) 1H MAS NMR spectrum of [(≡Si-O-)TaMe3+B(C6F5)3Me] acquired at 400 MHz with a 22 kHz MAS frequency, a repetition delay of 5 s, and 8 scans; (f) 13C CP MAS NMR spectrum of [(≡Si-O-)TaMe3+B(C6F5)3Me].

Author Contributions

Conceptualization, R.D., A.P., and J.-M.B.; experiments, R.D., J.C.M., M.K.S., A.H., S.K., Y.C., E.A.-H.; calculations, L.C. and A.P.; writing R.D., A.P., L.C. and J.-M.B.

Funding

This work was supported by funds from King Abdullah University of Science and Technology (KAUST) office of sponsored research (OSR). R.D. thanks National Institute of Technology, Calicut for financial support. A.P. thanks the Spanish MINECO for a project CTQ2014-59832-JIN.

Acknowledgments

The authors acknowledge the KAUST Nuclear Magnetic Resonance Core Lab and the technical assistance of Xianrong Guo.

Conflicts of Interest

The authors declare no competing financial interests.

References and Notes

  1. Schrock, R.R. Alkylidene Complexes of Niobium and Tantalum. Acc. Chem. Res. 1979, 12, 98–104. [Google Scholar] [CrossRef]
  2. Schrock, R.R. First Isolable Transition Metal Methylene Complex and Analogs. Characterization, Mode of Decomposition, and some Simple Reactions. J. Am. Chem. Soc. 1975, 97, 6577–6578. [Google Scholar] [CrossRef]
  3. Roux, E.L.; Chabanas, M.; Baudouin, A.; de-Mallmann, A.; Copéret, C.; Quadrelli, E.A.; Thivolle-Cazat, J.; Basset, J.-M.; Lukens, W.; Lesage, A.; et al. Detailed Structural Investigation of the Grafting of [Ta(=CHtBu)(CH2tBu)3] and [Cp*TaMe4] on Silica Partially Dehydroxylated at 700 °C and the Activity of the Grafted Complexes toward Alkane Metathesis. J. Am. Chem. Soc. 2004, 126, 13391–13399. [Google Scholar] [CrossRef] [PubMed]
  4. Freundlich, J.S.; Schrock, R.R.; Davis, W.M. Alkyl and Alkylidene Complexes of Tantalum That Contain a Triethylsilyl-Substituted Triamido–Amine Ligand. Organometallics 1996, 15, 2777–2783. [Google Scholar] [CrossRef]
  5. Schrock, R.R.; Meakin, P. Pentamethyl Complexes of Niobium and Tantalum. J. Am. Chem. Soc. 1974, 96, 5288–5290. [Google Scholar] [CrossRef]
  6. Schrock, R.R. Preparation and Characterization of M(CH3)5 (M = Nb or Ta) and Ta(CH2C6H5)5 and Evidence for Decomposition by α-Hydrogen Atom Abstraction. J. Organomet. Chem. 1976, 122, 209–225. [Google Scholar] [CrossRef]
  7. Albright, T.A.; Tang, H. The Structure of Pentamethyltantalum. Angew. Chem. Int. Ed. 1992, 31, 1462–1464. [Google Scholar] [CrossRef]
  8. Pulham, C.; Haaland, A.; Hammel, A.; Rypdal, K.; Verne, H.P.; Volden, H.V. Perfluorotriethylamine: An Amine with Unusual Structure and Reactivity. Angew. Chem. Int. Ed. 1992, 31, 1464–1467. [Google Scholar]
  9. Chen, Y.; Abou-hamad, E.; Hamieh, A.; Hamzaoui, B.; Emsley, L.; Basset, J.-M. Alkane Metathesis with the Tantalum Methylidene [(≡SiO)Ta(=CH2)Me2]/[(≡SiO)2Ta(=CH2)Me] Generated from Well-Defined Surface Organometallic Complex [(≡SiO)TaVMe4]. J. Am. Chem. Soc. 2015, 137, 588–591. [Google Scholar] [CrossRef] [PubMed]
  10. Chen, Y.; Ould-Chikh, S.; Abou-Hamad, E.; Callens, E.; Mohandas, J.C.; Khalid, S.; Basset, J.-M. Facile and Efficient Synthesis of the Surface Tantalum Hydride (≡SiO)2TaIIIH and Tris-Siloxy Tantalum (≡SiO)3TaIII Starting from Novel Tantalum Surface Species (≡SiO)TaMe4 and (≡SiO)2TaMe3. Organometallics 2014, 33, 1205–1211. [Google Scholar] [CrossRef]
  11. Schowner, R.; Frey, W.; Buchmeiser, M.R. Cationic Tungsten-Oxo-Alkylidene-N-Heterocyclic Carbene Complexes: Highly Active Olefin Metathesis Catalysts. J. Am. Chem. Soc. 2015, 137, 6188–6191. [Google Scholar] [CrossRef] [PubMed]
  12. Pucino, M.; Mougel, V.; Schowner, R.; Fedorov, A.; Buchmeiser, M.R.; Copéret, C. Cationic Silica-Supported N-Heterocyclic Carbene Tungsten Oxo Alkylidene Sites: Highly Active and Stable Catalysts for Olefin Metathesis. Angew. Chem. Int. Ed. 2016, 55, 4300–4302. [Google Scholar] [CrossRef] [PubMed]
  13. Dey, R.; Samantaray, M.K.; Poater, A.; Hamieh, A.; Kavitake, S.; Abou-Hamad, E.; Callens, E.; Emwas, A.-H.; Cavallo, L.; Basset, J.-M. Synthesis and Characterization of a Homogeneous and Silica Supported Homoleptic Cationic Tungsten(VI) Methyl Complex: Application in Olefin Metathesis. Chem. Commun. 2016, 52, 11270–11273. [Google Scholar] [CrossRef] [PubMed]
  14. Sanchez-Nieves, J.; Royo, P.; Mosquera, M.E.G. Synthesis of the Cation Complex [TaCp*Me3]+ and a Comparison of Its Reactivity with That of [TaCp*Me4]. Organometallics 2006, 25, 2331–2336. [Google Scholar] [CrossRef]
  15. Sanchez-Nieves, J.; Royo, P. Synthesis of Neutral and Cationic Monocyclopentadienyl Tantalum Alkoxo Complexes and Polymerization of Methyl Methacrylate. Organometallics 2007, 26, 2880–2884. [Google Scholar] [CrossRef]
  16. Mariott, W.R.; Gustafson, L.O.; Chen, E.Y.-X. Activation of Tantalocene(V) Alkyl and Alkylidene Complexes with Strong Organo Lewis Acids and Application to Polymerization Catalysis. Organometallics 2006, 25, 3721–3729. [Google Scholar] [CrossRef]
  17. Fryzuk, M.D.; Johnson, S.A.; Rettig, S.J. Synthesis and Structure of the Tantalum Trimethyl Complex [P2N2]TaMe3 and Its Conversion to the Tantalum Methylidene Species [P2N2]TaCH2(Me) ([P2N2] = PhP(CH2SiMe2NSiMe2CH2)2PPh). Organometallics 1999, 18, 4059–4067. [Google Scholar] [CrossRef]
  18. Guérin, F.; Stephan, D.W. Synthesis and Structure of the Dicationic Bisborate Adduct. Angew. Chem. Int. Ed. 2000, 39, 1298–1300. [Google Scholar] [CrossRef]
  19. Anderson, L.L.; Schmidt, J.A.R.; Arnold, J.; Bergman, R.G. Neutral and Cationic Alkyl Tantalum Imido Complexes:  Synthesis and Migratory Insertion Reactions. Organometallics 2006, 25, 3394–3406. [Google Scholar] [CrossRef] [PubMed]
  20. Anderson, L.L.; Arnold, J.; Bergman, R.G. Catalytic Hydroamination of Alkynes and Norbornene with Neutral and Cationic Tantalum Imido Complexes. Org. Lett. 2004, 6, 2519–2522. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Tomson, N.C.; Arnold, J.; Bergman, R.G. Synthesis and Reactivity of Cationic Niobium and Tantalum Methyl Complexes Supported by Imido and β-diketiminato Ligands. Dalton Trans. 2011, 40, 7718–7729. [Google Scholar] [CrossRef] [PubMed]
  22. Stahl, N.G.; Salata, M.R.; Marks, T.J. B(C6F5)3- vs. Al(C6F5)3-Derived Metallocenium Ion Pairs. Structural, Thermochemical, and Structural Dynamic Divergences. J. Am. Chem. Soc. 2005, 127, 10898–10909. [Google Scholar] [CrossRef] [PubMed]
  23. Mathis, D.; Couzijn, E.P.A.; Chen, P. Structure, Dynamics, and Polymerization Activity of Zirconocenium Ion Pairs Generated with Boron-C6F5 Compounds and Al2R6. Organometallics 2011, 30, 3834–3843. [Google Scholar] [CrossRef]
  24. Alonso-Moreno, C.; Lancaster, S.J.; Wright, J.A.; Hughes, D.L.; Zuccaccia, C.; Correa, A.; Macchioni, A.; Cavallo, L.; Bochmann, M. Ligand Mobility and Solution Structures of the Metallocenium Ion Pairs [Me2C(Cp)(fluorenyl)MCH2SiMe3+···X] (M = Zr, Hf; X = MeB(C6F5)3, B(C6F5)4). Organometallics 2008, 27, 5474–5487. [Google Scholar] [CrossRef]
  25. Spitzmesser, S.K.; Gibson, V.C. Dialkylaluminium Complexes Derived from 1,8-diphenyl-3,6-dimethylcarbazole: a New Sterically Hindered Monodentate Ligand System. J. Organomet. Chem. 2003, 673, 95–101. [Google Scholar] [CrossRef]
  26. Gillis, D.J.; Tudoret, M.-J.; Baird, M.C. Novel Arene Complexes of Titanium(IV), Zirconium(IV), and Hafnium(IV). J. Am. Chem. Soc. 1993, 115, 2543–2545. [Google Scholar] [CrossRef]
  27. Preparation of [TaMe4+ B(C6F5)3Me-](2): A cold solution (−40 °C) of B(C6F5)3 (125 mg, 0.25 mmol) in dichloromethane was added drop wise to the cold (−40 °C) solution of tantalumpentamethyl (50 mg, 0.2 mmol) in dichloromethane. The mixture was stirred for another 15 minutes. Colour of the solution intensified to dark yellowish which indicated the formation of the ionic complex. 1H-NMR(600 MHz) δ (ppm) 0.5(s, 3H, BCH3), 2.6(s, 12H, TaCH3). 13C-NMR(150 MHz) δ (ppm) 10.8(s, 1C, BCH3), 110.9 (s, 4C, TaCH3). (Caution! This 8e- compound is highly unstable and decomposes into black tantalum power while drying.).
  28. All the DFT calculations were performed with the Gaussian09, using the PBE0 functional, corrected by the D3 Grimme pairwise scheme, using the SVP basis set except for the metals the standard SDD. Single-point energy calculations in solution (SMD) were performed with the M06 functional and the Def2TZVPP basis set for main group atoms and again the same SDD pseudopotential for the metals.
  29. Martin, R.L.; Hay, P.J.; Pratt, L.R. Hydrolysis of Ferric Ion in Water and Conformational Equilibrium. J. Phys. Chem. A 1998, 102, 3565–3573. [Google Scholar] [CrossRef] [Green Version]
  30. Poater, A.; Pump, E.; Vummaleti, S.V.C.; Cavallo, L. The Right Computational Recipe for Olefin Metathesis with Ru-Based Catalysts: The Whole Mechanism of Ring-Closing Olefin Metathesis. J. Chem. Theory Comput. 2014, 10, 4442–4448. [Google Scholar] [CrossRef] [PubMed]
  31. Pump, E.; Slugovc, C.; Cavallo, L.; Poater, A. Mechanism of the Ru−allenylidene to Ru−Indenylidene Rearrangement in Ruthenium Precatalysts for Olefin Metathesis. Organometallics 2015, 34, 3107–3111. [Google Scholar] [CrossRef]
  32. Manzini, S.; Poater, A.; Nelson, D.J.; Cavallo, L.; Nolan, S.P. How Phenyl Makes a Difference: Mechanistic Insights into the Ruthenium(II)-Catalysed Isomerisation of Allylic Alcohols. Chem. Sci. 2014, 5, 180–188. [Google Scholar] [CrossRef]
  33. Manzini, S.; Poater, A.; Nelson, D.J.; Cavallo, L.; Slawin, A.M.Z.; Nolan, S.P. Insights into the Decomposition of Olefin Metathesis Precatalysts. Angew. Chem. Int. Ed. 2014, 53, 8995–8999. [Google Scholar] [CrossRef] [PubMed]
  34. Richmond, C.J.; Matheu, R.; Poater, A.; Falivene, L.; Benet-Buchholz, J.; Sala, X.; Cavallo, L.; Llobet, A. Supramolecular Water Oxidation with Ru–bda-based Catalysts. Chem. Eur. J. 2014, 20, 17282–17286. [Google Scholar] [CrossRef] [PubMed]
  35. Falivene, L.; Credendino, R.; Poater, A.; Petta, A.; Serra, L.; Oliva, R.; Scarano, V.; Cavallo, L. SambVca 2. A Web Tool for Analyzing Catalytic Pockets with Topographic Steric Maps. Organometallics 2016, 35, 2286–2293. [Google Scholar] [CrossRef]
  36. Poater, A.; Cosenza, B.; Correa, A.; Giudice, S.; Ragone, F.; Scarano, V.; Cavallo, L. SambVca: A Web Application for the Calculation of Buried Volumes of N-Heterocyclic Carbene Ligands. Eur. J. Inorg. Chem. 2009, 2009, 1759–1766. [Google Scholar] [CrossRef]
  37. Poater, A.; Ragone, F.; Giudice, S.; Costabile, C.; Dorta, R.; Nolan, S.P.; Cavallo, L. Thermodynamics of N-Heterocyclic Carbene Dimerization: The Balance of Sterics and Electronics. Organometallics 2008, 27, 2679–2681. [Google Scholar] [CrossRef]
  38. Poater, A.; Falivene, L.; Urbina-Blanco, C.A.; Manzini, S.; Nolan, S.P.; Cavallo, L. How does the Addition of Steric Hindrance to a Typical N-Heterocyclic Carbene Ligand Affect Catalytic Activity in Olefin Metathesis? Dalton Trans. 2013, 42, 7433–7439. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Demethylation of homoleptic TaMe5 complexes by strong electrophilic B(C6F5)3.
Scheme 1. Demethylation of homoleptic TaMe5 complexes by strong electrophilic B(C6F5)3.
Catalysts 08 00507 sch001
Scheme 2. Possible reaction between TaMe5 and B(C6F5)3.
Scheme 2. Possible reaction between TaMe5 and B(C6F5)3.
Catalysts 08 00507 sch002
Figure 1. (a) 1H-NMR spectra of TaMe5; (b) 1H-NMR spectra of [TaMe4]+[MeB(C6F5)3]; (c) 13C-NMR spectrum of TaMe5; (d) 13C-NMR spectrum of [TaMe4]+[MeB(C6F5)3], (at −40 °C, peak marked by (◊) is due to the slight amounts of pentane vapor which was used as a coolant during the low temperature reaction in the glovebox).
Figure 1. (a) 1H-NMR spectra of TaMe5; (b) 1H-NMR spectra of [TaMe4]+[MeB(C6F5)3]; (c) 13C-NMR spectrum of TaMe5; (d) 13C-NMR spectrum of [TaMe4]+[MeB(C6F5)3], (at −40 °C, peak marked by (◊) is due to the slight amounts of pentane vapor which was used as a coolant during the low temperature reaction in the glovebox).
Catalysts 08 00507 g001
Figure 2. Transition states for (a) the demethylation of TaMe5 by B(C6F5)3 and (b) (C6F5) transfer from B(Me)(C6F5)3 to TaMe4+, main distances are shown in Å.
Figure 2. Transition states for (a) the demethylation of TaMe5 by B(C6F5)3 and (b) (C6F5) transfer from B(Me)(C6F5)3 to TaMe4+, main distances are shown in Å.
Catalysts 08 00507 g002
Figure 3. Topographic steric maps (xy plane) of the metal centres of (a) TaMe5 and (b) WMe6. The corresponding metal is at the origin and one methyl of TaMe5 is on the z axis. The isocontour curves of the steric maps are given in Ǻ.
Figure 3. Topographic steric maps (xy plane) of the metal centres of (a) TaMe5 and (b) WMe6. The corresponding metal is at the origin and one methyl of TaMe5 is on the z axis. The isocontour curves of the steric maps are given in Ǻ.
Catalysts 08 00507 g003
Figure 4. Reaction pathway for the demethylation of TaMe5 with B(C6F5)3 and next C6F5 anionic transfer (Gibbs energies in kcal/mol referred to TaMe5).
Figure 4. Reaction pathway for the demethylation of TaMe5 with B(C6F5)3 and next C6F5 anionic transfer (Gibbs energies in kcal/mol referred to TaMe5).
Catalysts 08 00507 g004
Figure 5. Transition state for the demethylation by B(C6F5)3 of (≡Si-O-)TaMe4, main distances are shown in Å.
Figure 5. Transition state for the demethylation by B(C6F5)3 of (≡Si-O-)TaMe4, main distances are shown in Å.
Catalysts 08 00507 g005
Table 1. Gibbs energy values (in kcal/mol) for (a) the energy barrier of the transition state of the anion demethylation in the reaction [MMex→MMe(x−1)+ + Me, (b) the relative stability of the next demethylated anion MMe(x−1)+, (c) the intermediate MMe(x−1)· after the loose of a radical methyl and (d) the addition of a chloride anion that leads to MMexCl (M = Ta or W).
Table 1. Gibbs energy values (in kcal/mol) for (a) the energy barrier of the transition state of the anion demethylation in the reaction [MMex→MMe(x−1)+ + Me, (b) the relative stability of the next demethylated anion MMe(x−1)+, (c) the intermediate MMe(x−1)· after the loose of a radical methyl and (d) the addition of a chloride anion that leads to MMexCl (M = Ta or W).
Catalyst(a)(b)(c)(d)
TaMe56.8−15.444.1−5.8
WMe69.1−13.818.83.6

Share and Cite

MDPI and ACS Style

Dey, R.; Mohandas, J.C.; Samantaray, M.K.; Hamieh, A.; Kavitake, S.; Chen, Y.; Abou-Hamad, E.; Cavallo, L.; Poater, A.; Basset, J.-M. Synthesis and Characterization of Cationic Tetramethyl Tantalum(V) Complex. Catalysts 2018, 8, 507. https://doi.org/10.3390/catal8110507

AMA Style

Dey R, Mohandas JC, Samantaray MK, Hamieh A, Kavitake S, Chen Y, Abou-Hamad E, Cavallo L, Poater A, Basset J-M. Synthesis and Characterization of Cationic Tetramethyl Tantalum(V) Complex. Catalysts. 2018; 8(11):507. https://doi.org/10.3390/catal8110507

Chicago/Turabian Style

Dey, Raju, Janet C. Mohandas, Manoja K. Samantaray, Ali Hamieh, Santosh Kavitake, Yin Chen, Edy Abou-Hamad, Luigi Cavallo, Albert Poater, and Jean-Marie Basset. 2018. "Synthesis and Characterization of Cationic Tetramethyl Tantalum(V) Complex" Catalysts 8, no. 11: 507. https://doi.org/10.3390/catal8110507

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