Dehydrogenative coupling of silanes

Last updated
The general reaction for dehydrogenative coupling of primary silanes. General primary dehydrogenative coupling of primary polysilanes.png
The general reaction for dehydrogenative coupling of primary silanes.

The dehydrogenative coupling of silanes is a reaction type for the formation of Si-Si bonds. Although never commercialized[ citation needed ], the reaction has been demonstrated for the synthesis of certain disilanes as well as polysilanes. These reactions generally require catalysts.

Contents

Metallocene-based catalysts

Titanocene and related their complexes are typical catalysts. A typical reaction involves phenylsilane: [1] [2]

n PhSiH3 → [PhSiH]n + n H2

Para- and meta-substituted phenylsilanes polymerize readily but ortho-substituted polymers were failed to form. Polymers white/colorless, tacky and soluble in organic solvents. Crosslinking was not observed. [3]

Using Cp 2Ti(OPh)2 as a catalyst, the dehydrogenative coupling of phenylsilane in the presence of vinyltriethoxysilane produces a polysilane terminated with a triethoxysilylvinyl group. [4]

Other catalysts

The nickel(I) complex [(dippe)Ni(μ-H)]2 promotes the dehydrogenative coupling of some silanes. [5] While catalysts for dehydrogenative coupling reactions generally tend to be transition metal complexes, magnesium oxide and calcium oxide promote the dehydrogenation of phenylsilane. Being a heterogeneous process, the products are easily separated from the catalyst. [6] Dehydrogenative coupling of primary silanes using Wilkinson's catalyst is slow and dependent on the removal of H2 product. This conversion proceeds by oxidative addition of the Si-H bond and elimination of dihydrogen. [7] Tris(pentafluorophenyl)borane (B(C6F5)3)) is yet another catalyst for the dehydrogenative coupling of tertiary silanes. This system has the useful characteristic of being selective for Si-H bonds vs Si-Si bonds, leading to fewer branches and more linear polymers. This catalyst is particularly useful in reactions involving thiols and tertiary silanes or disilanes. [8]

As well as being coupled to each other, tertiary silanes can be coupled with carboxylic acids to form silyl esters. Ru3(CO)12/EtI is a good catalyst for this. This reaction applies to a wide range of silanes and acids. [9] The complex [Cu(PPh3)3Cl] can also be used to produce silyl esters. [10]

Tertiary silanes may also be dehydrogenatively coupled to aromatic rings with the use of the catalyst TpMe2Pt(Me)2H (TpMe2 = hydrido tris(3,5-dimethylpyrazolyl)borate). For example, this platinum catalyst can be used to react triethyl silane with benzene to produce phenyltriethylsilane, with the elimination of hydrogen gas. This is a terrific catalyst because it eliminates the need for a hydrogen acceptor, something which is normally required for the silation of a C-H bond. This reaction may also be done intramolecularly to produce five- or six-membered silicon-containing rings fused to a phenyl ring. In addition, tributylsilane can be converted into the corresponding cyclic organosilane via the same process. A drawback to this catalyst, however, is that it requires rather harsh reaction conditions (typically 200 °C for 24 hours for the intermolecular reaction, 48 – 72 hours for the intramolecular ones). It is also not particularly regioselective, so starting materials containing substituted benzene would result in a mixture of products. [11]

Examples of intramolecular Si-C coupling reactions Intra.gif
Examples of intramolecular Si-C coupling reactions

Polymerization of silane

Some methods used to produce polysilanes are polymerization of masked dislenes, [12] electroreduction of dichlorosilanes, [13]

Polymer characterization

1H and 29Si NMR spectroscopy can sometimes be used to help identify and characterize products from these reactions. [1]

Infrared spectroscopy may also be useful as it can indicate whether or not the product is a tertiary silane. The stretch for Si-H is seen at around 2100 and 910 cm−1. In the case of tertiary silanes however, the peak at 910 cm−1 is not seen. The shift or change in these peaks will be affected by the size of the polymer. [1]

Related Research Articles

<span class="mw-page-title-main">Organolithium reagent</span> Chemical compounds containing C–Li bonds

In organometallic chemistry, organolithium reagents are chemical compounds that contain carbon–lithium (C–Li) bonds. These reagents are important in organic synthesis, and are frequently used to transfer the organic group or the lithium atom to the substrates in synthetic steps, through nucleophilic addition or simple deprotonation. Organolithium reagents are used in industry as an initiator for anionic polymerization, which leads to the production of various elastomers. They have also been applied in asymmetric synthesis in the pharmaceutical industry. Due to the large difference in electronegativity between the carbon atom and the lithium atom, the C−Li bond is highly ionic. Owing to the polar nature of the C−Li bond, organolithium reagents are good nucleophiles and strong bases. For laboratory organic synthesis, many organolithium reagents are commercially available in solution form. These reagents are highly reactive, and are sometimes pyrophoric.

In chemistry, dehydrogenation is a chemical reaction that involves the removal of hydrogen, usually from an organic molecule. It is the reverse of hydrogenation. Dehydrogenation is important, both as a useful reaction and a serious problem. At its simplest, it's a useful way of converting alkanes, which are relatively inert and thus low-valued, to olefins, which are reactive and thus more valuable. Alkenes are precursors to aldehydes, alcohols, polymers, and aromatics. As a problematic reaction, the fouling and inactivation of many catalysts arises via coking, which is the dehydrogenative polymerization of organic substrates.

In chemistry, homogeneous catalysis is catalysis where the catalyst is in same phase as reactants, principally by a soluble catalyst a in solution. In contrast, heterogeneous catalysis describes processes where the catalysts and substrate are in distinct phases, typically solid-gas, respectively. The term is used almost exclusively to describe solutions and implies catalysis by organometallic compounds. Homogeneous catalysis is an established technology that continues to evolve. An illustrative major application is the production of acetic acid. Enzymes are examples of homogeneous catalysts.

The Hiyama coupling is a palladium-catalyzed cross-coupling reaction of organosilanes with organic halides used in organic chemistry to form carbon–carbon bonds. This reaction was discovered in 1988 by Tamejiro Hiyama and Yasuo Hatanaka as a method to form carbon-carbon bonds synthetically with chemo- and regioselectivity. The Hiyama coupling has been applied to the synthesis of various natural products.

Coordination polymerisation is a form of polymerization that is catalyzed by transition metal salts and complexes.

<span class="mw-page-title-main">Acylsilane</span> Class of chemical compounds

Acylsilanes are a group of chemical compounds sharing a common functional group with the general structure RC(O)-SiR3.

<span class="mw-page-title-main">Organosilicon chemistry</span> Organometallic compound containing carbon–silicon bonds

Organosilicon chemistry is the study of organometallic compounds containing carbon–silicon bonds, to which they are called organosilicon compounds. Most organosilicon compounds are similar to the ordinary organic compounds, being colourless, flammable, hydrophobic, and stable to air. Silicon carbide is an inorganic compound.

<span class="mw-page-title-main">Binary silicon-hydrogen compounds</span>

Silanes are saturated chemical compounds with the empirical formula SixHy. They are hydrosilanes, a class of compounds that includes compounds with Si−H and other Si−X bonds. All contain tetrahedral silicon and terminal hydrides. They only have Si−H and Si−Si single bonds. The bond lengths are 146.0 pm for a Si−H bond and 233 pm for a Si−Si bond. The structures of the silanes are analogues of the alkanes, starting with silane, SiH4, the analogue of methane, continuing with disilane Si2H6, the analogue of ethane, etc. They are mainly of theoretical or academic interest.

<span class="mw-page-title-main">Vinylsilane</span> Chemical compound

Vinylsilane refers to an organosilicon compound with chemical formula CH2=CHSiH3. It is a derivative of silane (SiH4). The compound, which is a colorless gas, is mainly of theoretical interest.

Hydrosilanes are tetravalent silicon compounds containing one or more Si-H bond. The parent hydrosilane is silane (SiH4). Commonly, hydrosilane refers to organosilicon derivatives. Examples include phenylsilane (PhSiH3) and triethoxysilane ((C2H5O)3SiH). Polymers and oligomers terminated with hydrosilanes are resins that are used to make useful materials like caulks.

Chiral Lewis acids (CLAs) are a type of Lewis acid catalyst. These acids affect the chirality of the substrate as they react with it. In such reactions, synthesis favors the formation of a specific enantiomer or diastereomer. The method is an enantioselective asymmetric synthesis reaction. Since they affect chirality, they produce optically active products from optically inactive or mixed starting materials. This type of preferential formation of one enantiomer or diastereomer over the other is formally known as asymmetric induction. In this kind of Lewis acid, the electron-accepting atom is typically a metal, such as indium, zinc, lithium, aluminium, titanium, or boron. The chiral-altering ligands employed for synthesizing these acids often have multiple Lewis basic sites that allow the formation of a ring structure involving the metal atom.

Reductions with hydrosilanes are methods used for hydrogenation and hydrogenolysis of organic compounds. The approach is a subset of ionic hydrogenation. In this particular method, the substrate is treated with a hydrosilane and auxiliary reagent, often a strong acid, resulting in formal transfer of hydride from silicon to carbon. This style of reduction with hydrosilanes enjoys diverse if specialized applications.

Metal carbon dioxide complexes are coordination complexes that contain carbon dioxide ligands. Aside from the fundamental interest in the coordination chemistry of simple molecules, studies in this field are motivated by the possibility that transition metals might catalyze useful transformations of CO2. This research is relevant both to organic synthesis and to the production of "solar fuels" that would avoid the use of petroleum-based fuels.

<span class="mw-page-title-main">Polysilane</span>

Polysilanes are organosilicon compounds with the formula (R2Si)n. They are relatives of traditional organic polymers but their backbones are composed of silicon atoms. They exhibit distinctive optical and electrical properties. They are mainly used as precursors to silicon carbide. The simplest polysilane would be (SiH2)n, which is mainly of theoretical, not practical interest.

Decarboxylative cross coupling reactions are chemical reactions in which a carboxylic acid is reacted with an organic halide to form a new carbon-carbon bond, concomitant with loss of CO2. Aryl and alkyl halides participate. Metal catalyst, base, and oxidant are required.

Metal-catalyzed C–H borylation reactions are transition metal catalyzed organic reactions that produce an organoboron compound through functionalization of aliphatic and aromatic C–H bonds and are therefore useful reactions for carbon–hydrogen bond activation. Metal-catalyzed C–H borylation reactions utilize transition metals to directly convert a C–H bond into a C–B bond. This route can be advantageous compared to traditional borylation reactions by making use of cheap and abundant hydrocarbon starting material, limiting prefunctionalized organic compounds, reducing toxic byproducts, and streamlining the synthesis of biologically important molecules. Boronic acids, and boronic esters are common boryl groups incorporated into organic molecules through borylation reactions. Boronic acids are trivalent boron-containing organic compounds that possess one alkyl substituent and two hydroxyl groups. Similarly, boronic esters possess one alkyl substituent and two ester groups. Boronic acids and esters are classified depending on the type of carbon group (R) directly bonded to boron, for example alkyl-, alkenyl-, alkynyl-, and aryl-boronic esters. The most common type of starting materials that incorporate boronic esters into organic compounds for transition metal catalyzed borylation reactions have the general formula (RO)2B-B(OR)2. For example, bis(pinacolato)diboron (B2Pin2), and bis(catecholato)diborane (B2Cat2) are common boron sources of this general formula.

Ionic hydrogenation refers to hydrogenation achieved by the addition of a hydride to substrate that has been activated by an electrophile. Some ionic hydrogenations entail addition of H2 to the substrate and some entail replacement of a heteroatom with hydride. Traditionally, the method was developed for acid-induced reductions with hydrosilanes. Alternatively ionic hydrogenation can be achieved using H2. Ionic hydrogenation is employed when the substrate can produce a stable carbonium ion. Polar double bonds are favored substrates.

The Mukaiyama hydration is an organic reaction involving formal addition of an equivalent of water across an olefin by the action of catalytic bis(acetylacetonato)cobalt(II) complex, phenylsilane and atmospheric oxygen to produce an alcohol with Markovnikov selectivity.

In chemistry, transition metal silyl complexes describe coordination complexes in which a transition metal is bonded to an anionic silyl ligand, forming a metal-silicon sigma bond. This class of complexes are numerous and some are technologically significant as intermediates in hydrosilylation. These complexes are a subset of organosilicon compounds.

Norio Miyaura was a Japanese organic chemist. He was a professor of graduate chemical engineering at Hokkaido University. His major accomplishments surrounded his work in cross-coupling reactions / conjugate addition reactions of organoboronic acids and addition / coupling reactions of diborons and boranes. He is also the co-author of Cross-Coupling Reactions: A Practical Guide with M. Nomura E. S.. Miyaura was a world-known and accomplished researcher by the time he retired and so, in 2007, he won the Japan Chemical Society Award.

References

  1. 1 2 3 Aitken, C.; Harrod, J. F.; Gill, U. S. (1987). "Structural studies of oligosilanes produced by catalytic dehydrogenative coupling of primary organosilanes". Can. J. Chem. 65: 1804–1809. doi:10.1139/v87-303.
  2. Corey, J.Y.; Zhu, X.H.; Bedard, T.C.; Lange, L.D. (1991). "Catalytic Dehydrogenative Coupling of Secondary Silanes with Cp2MCl2". Organometallics. 10 (4): 924. doi:10.1021/om00050a024.
  3. Banovetz, John P.; Suzuki, Hiroshi; Waymouth, Robert M. (1993). "Dehydrogenative coupling of substituted phenylsilanes synthesis of poly[((trifluoromethyl)phenyl)silanes]". Organometallics. 12 (11): 4700–4703. doi:10.1021/om00035a070.
  4. Garcia, Julien; Meyer, Daniel J.M.; Guillaneux, Denis; Moreau, Joël J.E.; Wong Chi Man, Michel (July 2009). "Investigation of titanium-catalysed dehydrogenative coupling and hydrosilylation of phenylhydrogenosilanes in a one-pot process". Journal of Organometallic Chemistry. 694 (15): 2427–2433. doi:10.1016/j.jorganchem.2009.03.018.
  5. Smith, Erin E.; Du, Guodong; Fanwick, Phillipe E.; Abu-Omar, Mahdi M. (2010). "Dehydrocoupling of Organosilanes with a Dinuclear Nickel Hydride Catalyst and Isolation of a Nickel Silyl Complex". Organometallics. 29 (23): 6529. doi:10.1021/om100887v.
  6. Itoh, M.; Mitsuzuka, M.; Utsumi, T.; Iwata, K.; Inoue, K. (1994). "Dehydrogenative coupling reactions between hydrosilanes and monosubstituted alkynes catalyzed by solid bases". Journal of Organometallic Chemistry. 476: C30–C31. doi:10.1016/0022-328X(94)87091-8.
  7. Rosenberg, Lisa; Kobus, Danielle N. (2003). "Dehydrogenative coupling of primary alkyl silanes using Wilkinson's catalyst". Journal of Organometallic Chemistry. 685 (1–2): 107–112. doi:10.1016/S0022-328X(03)00712-5.
  8. Harrison, D. J.; Edwards, D. R.; McDonald, R.; Rosenberg, L. (2008). "Toward selective functionalisation of oligosilanes: borane-catalysed dehydrogenative coupling of silanes with thiols". Dalton Trans. 26: 3401–3411. doi:10.1039/b806270f.
  9. Liu, G.; Zhao, H. (2008). "Ru-catalyzed dehydrogenative coupling of carboxylic acids and silanes - a new method for the preparation of silyl esters". Beilstein Journal of Organic Chemistry. 4 (27). doi:10.3762/bjoc.4.27. PMC   2511026 .
  10. Liu, G. -B.; Zhao, H. -Y.; Thiemann, T. (2007). "Two New Catalysts for the Dehydrogenative Coupling Reaction of Carboxylic Acids with Silanes - Convenient Methods for an Atom-Economical Preparation of Silyl Esters". Synthetic Communications. 37: 2717–2727. doi:10.1080/00397910701465669.
  11. 1 2 Tsukada, N.; Hartwig, J. F. (2005). "Intermolecular and intramolecular, platinum-catalyzed, acceptorless dehydrogenative coupling of hydrosilanes with aryl and aliphatic methyl C-H bonds". Journal of the American Chemical Society (127): 5022–5023. doi:10.1021/ja050612p.
  12. Sakomto, Kenkichi; Yoshida, Masaru; Sakurai, Hideki (1990). "Highly ordered high-molecular weight alternating polysilylene copolymer prepared by anionic polymerization of masked disilene". Macromolecules. 23 (20): 4494–4496. doi:10.1021/ma00222a031.
  13. Shono, T.; Ishifune, S.; Nishida, R. (1997). "Electroreductive synthesis of some functionalized polysilanes and related polymers". Tetrahedron Letters. 38 (36): 4607–4610. doi:10.1016/S0040-4039(97)00987-8.