Telomerization (dimerization)

Last updated

The telomerization is the linear dimerization of 1,3-dienes with simultaneous addition of a nucleophile in a catalytic reaction.

Contents

Reaction

The reaction was independently discovered by E. J. Smutny at Shell and Takahashi at the Osaka University in the late sixties. The general reaction equation is as follows: [1] [2]

Products of the Telomerisation of 1,3-Butadiene Telomerisation.png
Products of the Telomerisation of 1,3-Butadiene

The formation of several isomers are possible. In addition to 1,3-butadiene also substituted dienes such as isoprene or cyclic dienes such as cyclopentadiene can be used. A variety of substances such as water, ammonia, alcohols, or C-H-acidic compounds can be used as nucleophiles. When water is used, for example di-unsaturated alcohols are obtained.

The catalysts used are mainly metal-organic palladium and nickel compounds. In 1991, Kuraray implemented the production of 1-octanol on an industrial scale (5000 t a(-1)).

The commercial route to produce 1-octene based on butadiene as developed by Dow Chemical came on stream in Tarragona in 2008. The telomerization of butadiene with methanol in the presence of a palladium catalyst yields 1-methoxy-2,7-octadiene, which is fully hydrogenated to 1-methoxyoctane in the next step. Subsequent cracking of 1-methoxyoctane gives 1-octene and methanol for recycle.

Mechanism

While the reaction is catalyzed by Pd(0) complexes, the pre-catalyst can also be a Pd(II) compound that is reduced in situ. Once the Pd(0) catalyst is formed it can coordinate two butadienes which by oxidative coupling give the intermediate B. Even though the oxidative coupling is facile it is nonetheless reversible; the latter is illustrated by the fact that B is only stable at high butadiene concentration. Subsequent protonation of this intermediate by NuH at the 6-position of the η3-,η1-octadienyl ligand leads to intermediate C. Nw direct attack of the nucleophile can take place at either the 1- or 3-position of the η3-octadienyl chain, which leads to the linear or branched product complexes Dn and Diso respectively. Upon displacement by new 1,3-butadiene the product telomer is liberated while the catalyst is regenerated and can continue the cycle. [3]

Catalytic cycle of the Palladium/Phosphine-catalyzed telomerization of 1,3-butadiene with a nucleophile NuH Telomerization mechaism new.png
Catalytic cycle of the Palladium/Phosphine-catalyzed telomerization of 1,3-butadiene with a nucleophile NuH

While from purely steric reasons nucleophilic attack at the less substituted side of the allyl is favored, the regioselectivity of nucleophilic attack can heavily depend on the exact nature of ligands positioned trans to the allyl group. [4]

Literature

See also

Related Research Articles

<span class="mw-page-title-main">Diene</span> Covalent compound that contains two double bonds

In organic chemistry a diene is a covalent compound that contains two double bonds, usually among carbon atoms. They thus contain two alkene units, with the standard prefix di of systematic nomenclature. As a subunit of more complex molecules, dienes occur in naturally occurring and synthetic chemicals and are used in organic synthesis. Conjugated dienes are widely used as monomers in the polymer industry. Polyunsaturated fats are of interest to nutrition.

<span class="mw-page-title-main">Petrochemical</span> Chemical product derived from petroleum

Petrochemicals are the chemical products obtained from petroleum by refining. Some chemical compounds made from petroleum are also obtained from other fossil fuels, such as coal or natural gas, or renewable sources such as maize, palm fruit or sugar cane.

The Sonogashira reaction is a cross-coupling reaction used in organic synthesis to form carbon–carbon bonds. It employs a palladium catalyst as well as copper co-catalyst to form a carbon–carbon bond between a terminal alkyne and an aryl or vinyl halide.

Organopalladium chemistry is a branch of organometallic chemistry that deals with organic palladium compounds and their reactions. Palladium is often used as a catalyst in the reduction of alkenes and alkynes with hydrogen. This process involves the formation of a palladium-carbon covalent bond. Palladium is also prominent in carbon-carbon coupling reactions, as demonstrated in tandem reactions.

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

The Wacker process or the Hoechst-Wacker process refers to the oxidation of ethylene to acetaldehyde in the presence of palladium(II) chloride as the catalyst. This chemical reaction was one of the first homogeneous catalysis with organopalladium chemistry applied on an industrial scale.

An allylic rearrangement or allylic shift is an organic reaction in which the double bond in an allyl chemical compound shifts to the next carbon atom. It is encountered in nucleophilic substitution.

<span class="mw-page-title-main">Allylpalladium chloride dimer</span> Chemical compound

Allylpalladium(II) chloride dimer (APC) is a chemical compound with the formula [(η3-C3H5)PdCl]2. This yellow air-stable compound is an important catalyst used in organic synthesis. It is one of the most widely used transition metal allyl complexes.

<span class="mw-page-title-main">Nozaki–Hiyama–Kishi reaction</span> Coupling reaction used in organic synthesis

The Nozaki–Hiyama–Kishi reaction is a nickel/chromium coupling reaction forming an alcohol from the reaction of an aldehyde with an allyl or vinyl halide. In their original 1977 publication, Tamejiro Hiyama and Hitoshi Nozaki reported on a chromium(II) salt solution prepared by reduction of chromic chloride by lithium aluminium hydride to which was added benzaldehyde and allyl chloride:

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

3,3′,3′′-Phosphanetriyltris(benzenesulfonic acid) trisodium salt (abbreviated TPPTS), is an organic compound that is also known as sodium triphenylphosphine trisulfonate. The compound has the formula P(C6H4SO3Na)3. This white solid is an unusual example of a water-soluble phosphine. Its complexes are also water-soluble. Its complex with rhodium is used in the industrial production of butyraldehyde.

In organometallic chemistry, the Green–Davies–Mingos rules predict the regiochemistry for nucleophilic addition to 18-electron metal complexes containing multiple unsaturated ligands. The rules were published in 1978 by organometallic chemists Stephen G. Davies, Malcolm Green, and Michael Mingos. They describe how and where unsaturated hydrocarbon generally become more susceptibile to nucleophilic attack upon complexation.

In chemistry, π-effects or π-interactions are a type of non-covalent interaction that involves π systems. Just like in an electrostatic interaction where a region of negative charge interacts with a positive charge, the electron-rich π system can interact with a metal, an anion, another molecule and even another π system. Non-covalent interactions involving π systems are pivotal to biological events such as protein-ligand recognition.

Organoiron chemistry is the chemistry of iron compounds containing a carbon-to-iron chemical bond. Organoiron compounds are relevant in organic synthesis as reagents such as iron pentacarbonyl, diiron nonacarbonyl and disodium tetracarbonylferrate. While iron adopts oxidation states from Fe(−II) through to Fe(VII), Fe(IV) is the highest established oxidation state for organoiron species. Although iron is generally less active in many catalytic applications, it is less expensive and "greener" than other metals. Organoiron compounds feature a wide range of ligands that support the Fe-C bond; as with other organometals, these supporting ligands prominently include phosphines, carbon monoxide, and cyclopentadienyl, but hard ligands such as amines are employed as well.

Electrophilic substitution of unsaturated silanes involves attack of an electrophile on an allyl- or vinylsilane. An allyl or vinyl group is incorporated at the electrophilic center after loss of the silyl group.

Trimethylenemethane cycloaddition is the formal [3+2] annulation of trimethylenemethane (TMM) derivatives to two-atom pi systems. Although TMM itself is too reactive and unstable to be stored, reagents which can generate TMM or TMM synthons in situ can be used to effect cycloaddition reactions with appropriate electron acceptors. Generally, electron-deficient pi bonds undergo cyclization with TMMs more easily than electron-rich pi bonds.

Desulfonylation reactions are chemical reactions leading to the removal of a sulfonyl group from organic compounds. As the sulfonyl functional group is electron-withdrawing, methods for cleaving the sulfur–carbon bonds of sulfones are typically reductive in nature. Olefination or replacement with hydrogen may be accomplished using reductive desulfonylation methods.

<span class="mw-page-title-main">Hydrogen auto-transfer</span>

Hydrogen auto-transfer, also known as borrowing hydrogen, is the activation of a chemical reaction by temporary transfer of two hydrogen atoms from the reactant to a catalyst and return of those hydrogen atoms back to a reaction intermediate to form the final product. Two major classes of borrowing hydrogen reactions exist: (a) those that result in hydroxyl substitution, and (b) those that result in carbonyl addition. In the former case, alcohol dehydrogenation generates a transient carbonyl compound that is subject to condensation followed by the return of hydrogen. In the latter case, alcohol dehydrogenation is followed by reductive generation of a nucleophile, which triggers carbonyl addition. As borrowing hydrogen processes avoid manipulations otherwise required for discrete alcohol oxidation and the use of stoichiometric organometallic reagents, they typically display high levels of atom-economy and, hence, are viewed as examples of Green chemistry.

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

The White catalyst is a transition metal coordination complex named after the chemist by whom it was first synthesized, M. Christina White, a professor at the University of Illinois. The catalyst has been used in a variety of allylic C-H functionalization reactions of α-olefins. In addition, it has been shown to catalyze oxidative Heck reactions.

The Tsuji–Trost reaction is a palladium-catalysed substitution reaction involving a substrate that contains a leaving group in an allylic position. The palladium catalyst first coordinates with the allyl group and then undergoes oxidative addition, forming the π-allyl complex. This allyl complex can then be attacked by a nucleophile, resulting in the substituted product.

<span class="mw-page-title-main">Transition-metal allyl complex</span>

Transition-metal allyl complexes are coordination complexes with allyl and its derivatives as ligands. Allyl is the radical with the connectivity CH2CHCH2, although as a ligand it is usually viewed as an allyl anion CH2=CH−CH2, which is usually described as two equivalent resonance structures.

Heterobimetallic catalysis is an approach to catalysis that employs two different metals to promote a chemical reaction. Included in this definition are cases where: 1) each metal activates a different substrate, 2) both metals interact with the same substrate, and 3) only one metal directly interacts with the substrate(s), while the second metal interacts with the first.

References

  1. Edgar J. Smutny: Oligomerization and dimerization of butadiene under homogeneous catalysis. Reaction with nucleophiles and the synthesis of 1,3,7-octatriene In: Journal of the American Chemical Society. 89, 1967, p. 6793–6794, doi : 10.1021/ja01001a089.
  2. S. Takahashi, T. Shibano, and N. Hagihara: The dimerization of butadiene by palladium complex catalysts. In: Tetrahedron Letters 8.26 (1967): 2451-2453.
  3. P. C. A. Bruijnincx, R. Jastrzebski, P. J. C. Hausoul, R. J. M. Klein Gebbink, B. M. Weckhuysen (2012). "Pd-Catalyzed Telomerization of 1,3-Dienes with Multifunctional Renewable Substrates: Versatile Routes for the Valorization of Biomass-Derived Platform Molecules". Top. Organomet. Chem. 39: 45–102.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  4. Hartwig, John (2010). Organotransition metal chemistry: From bonding to catalysis. University Science Books. ISBN   978-1-891389-53-5.