Koch reaction

Last updated

The Koch reaction is an organic reaction for the synthesis of tertiary carboxylic acids from alcohols or alkenes. The reaction is a strongly acid-catalyzed carbonylation using carbon monoxide, and typically occurs at high pressures ranging from 50 to 5,000 kPa, often requiring temperatures several hundred degrees higher than room temperature. Generally the reaction is conducted with strong mineral acids such as sulfuric acid, HF or BF3. [1] Large scale operations for the fine chemical industry produce almost 150,000 tonnes of Koch acids and their derivatives annually [2] but also generate a great deal of waste, motivating ongoing attempts to use metal, solid acid, and other novel catalysts to enable the use of milder reaction conditions. Formic acid, which readily decomposes to carbon monoxide in the presence of acids or relatively low heat, is often used instead of carbon monoxide directly; this procedure was developed shortly after the Koch reaction and is more commonly referred to as the Koch–Haaf reaction. This variation allows for reactions at nearly standard room temperature and pressure. Some commonly industrially produced Koch acids include pivalic acid, 2,2-dimethylbutyric acid and 2,2-dimethylpentanoic acid.

Contents

Schemata-der-Koch-Reaktion V.svg

Mechanism

When standard acid catalysts such as sulfuric acid or a mix of BF3 and HF are used, the mechanism [3] begins by protonation of the alkene, followed by carbon monoxide attack of the resulting carbocation. The subsequent acylium cation is then hydrolysed to the tertiary carboxylic acid. If the substrate is an alcohol, it is protonated and subsequently eliminated, generating a carbocation that is converted to an acylium cation by carbon monoxide and then hydrolysed. Tertiary carbocation formation is typically thermodynamically favored when considering hydride or alkyl shifts in the carbocation.

Catalyst usage and variations

Industrial large scale application of the Koch reaction using strong mineral acids is complicated by equipment corrosion, separation procedures for products and difficulty in managing large amounts of waste acid. Several acid resins [4] [5] and acidic ionic liquids [6] have been investigated in order to discover if Koch acids can be synthesized in more mild environments. Although the use of acidic ionic liquids for the Koch reaction requires relatively high temperatures and pressures (8 MPa and 430 K in one 2006 study [6] ), acidic ionic solutions themselves can be reused with only a very slight decrease in yield, and the reactions can be carried out biphasically to ensure easy separation of products. A large number of transition metal catalyst carbonyl cations have also been investigated for usage in Koch-like reactions: Cu(I), [7] Au(I) [8] and Pd(I) [9] carbonyl cations catalysts dissolved in sulfuric acid can allow the reaction to progress at room temperature and atmospheric pressure. Usage of a Nickel tetracarbonyl catalyst with CO and water as a nucleophile is known as the Reppe carbonylation, and there are many variations on this type of metal-mediated carbonylation used in industry, particularly those used by Monsanto and the Cativa processes, which convert methanol to acetic acid using acid catalysts and carbon monoxide in the presence of metal catalysts.

Side reactions

Koch reactions can involve a large number of side products, although high yields are generally possible (Koch and Haaf reported yields of over 80% for several alcohols in their 1958 paper). Carbocation rearrangements, etherization (in case an alcohol is used as a substrate, instead of an alkene), and occasionally substrate CN+1 carboxylic acids are observed due to fragmentation and dimerization of carbon monoxide-derived carbenium ions, especially since each step of the reaction is reversible. [10] Alkyl sulfuric acids are also known to be possible side products, but are usually eliminated by the excess sulfuric acid used.

Applications

Koch–Haaf-type reactions see extensive use in rational drug design [11] [12] as a convenient way to generate crucial tertiary carboxylic acids. Companies such as Shell and ExxonMobil produce pivalic acid from isobutene using the Koch reaction, [2] as well as several other branched carboxylic acids. However, Koch–Haaf reactions are also utilized for the interrogation of several other topics. As the reactants are found in different phases, the Koch reaction has been used to study reaction kinetics of gas–liquid–liquid systems [13] as well as query the use of solid acid resins and acidic ionic liquids in reducing hazardous by-product waste.

See also

Related Research Articles

<span class="mw-page-title-main">Carboxylic acid</span> Organic compound containing a –C(=O)OH group

In organic chemistry, a carboxylic acid is an organic acid that contains a carboxyl group attached to an R-group. The general formula of a carboxylic acid is often written as R−COOH or R−CO2H, sometimes as R−C(O)OH with R referring to the alkyl, alkenyl, aryl, or other group. Carboxylic acids occur widely. Important examples include the amino acids and fatty acids. Deprotonation of a carboxylic acid gives a carboxylate anion.

<span class="mw-page-title-main">Carbon monoxide</span> Colourless, odourless, tasteless and toxic gas

Carbon monoxide is a poisonous, flammable gas that is colorless, odorless, tasteless, and slightly less dense than air. Carbon monoxide consists of one carbon atom and one oxygen atom connected by a triple bond. It is the simplest carbon oxide. In coordination complexes, the carbon monoxide ligand is called carbonyl. It is a key ingredient in many processes in industrial chemistry.

<span class="mw-page-title-main">Formic acid</span> Simplest carboxylic acid (HCOOH)

Formic acid, systematically named methanoic acid, is the simplest carboxylic acid, and has the chemical formula HCOOH and structure H−C(=O)−O−H. It is an important intermediate in chemical synthesis and occurs naturally, most notably in some ants. Esters, salts and the anion derived from formic acid are called formates. Industrially, formic acid is produced from methanol.

<span class="mw-page-title-main">Acyl group</span> Chemical group (R–C=O)

In chemistry, an acyl group is a moiety derived by the removal of one or more hydroxyl groups from an oxoacid, including inorganic acids. It contains a double-bonded oxygen atom and an organyl group or hydrogen in the case of formyl group. In organic chemistry, the acyl group is usually derived from a carboxylic acid, in which case it has the formula R−C(=O)−, where R represents an organyl group or hydrogen. Although the term is almost always applied to organic compounds, acyl groups can in principle be derived from other types of acids such as sulfonic acids and phosphonic acids. In the most common arrangement, acyl groups are attached to a larger molecular fragment, in which case the carbon and oxygen atoms are linked by a double bond.

<span class="mw-page-title-main">Hydrogenation</span> Chemical reaction between molecular hydrogen and another compound or element

Hydrogenation is a chemical reaction between molecular hydrogen (H2) and another compound or element, usually in the presence of a catalyst such as nickel, palladium or platinum. The process is commonly employed to reduce or saturate organic compounds. Hydrogenation typically constitutes the addition of pairs of hydrogen atoms to a molecule, often an alkene. Catalysts are required for the reaction to be usable; non-catalytic hydrogenation takes place only at very high temperatures. Hydrogenation reduces double and triple bonds in hydrocarbons.

The Friedel–Crafts reactions are a set of reactions developed by Charles Friedel and James Crafts in 1877 to attach substituents to an aromatic ring. Friedel–Crafts reactions are of two main types: alkylation reactions and acylation reactions. Both proceed by electrophilic aromatic substitution.

In chemistry, an electrophile is a chemical species that forms bonds with nucleophiles by accepting an electron pair. Because electrophiles accept electrons, they are Lewis acids. Most electrophiles are positively charged, have an atom that carries a partial positive charge, or have an atom that does not have an octet of electrons.

<span class="mw-page-title-main">Alkylation</span> Transfer of an alkyl group from one molecule to another

Alkylation is a chemical reaction that entails transfer of an alkyl group. The alkyl group may be transferred as an alkyl carbocation, a free radical, a carbanion, or a carbene. Alkylating agents are reagents for effecting alkylation. Alkyl groups can also be removed in a process known as dealkylation. Alkylating agents are often classified according to their nucleophilic or electrophilic character. In oil refining contexts, alkylation refers to a particular alkylation of isobutane with olefins. For upgrading of petroleum, alkylation produces a premium blending stock for gasoline. In medicine, alkylation of DNA is used in chemotherapy to damage the DNA of cancer cells. Alkylation is accomplished with the class of drugs called alkylating antineoplastic agents.

In organic chemistry, hydroformylation, also known as oxo synthesis or oxo process, is an industrial process for the production of aldehydes from alkenes. This chemical reaction entails the net addition of a formyl group and a hydrogen atom to a carbon-carbon double bond. This process has undergone continuous growth since its invention: production capacity reached 6.6×106 tons in 1995. It is important because aldehydes are easily converted into many secondary products. For example, the resultant aldehydes are hydrogenated to alcohols that are converted to detergents. Hydroformylation is also used in speciality chemicals, relevant to the organic synthesis of fragrances and pharmaceuticals. The development of hydroformylation is one of the premier achievements of 20th-century industrial chemistry.

The Heck reaction is the chemical reaction of an unsaturated halide with an alkene in the presence of a base and a palladium catalyst to form a substituted alkene. It is named after Tsutomu Mizoroki and Richard F. Heck. Heck was awarded the 2010 Nobel Prize in Chemistry, which he shared with Ei-ichi Negishi and Akira Suzuki, for the discovery and development of this reaction. This reaction was the first example of a carbon-carbon bond-forming reaction that followed a Pd(0)/Pd(II) catalytic cycle, the same catalytic cycle that is seen in other Pd(0)-catalyzed cross-coupling reactions. The Heck reaction is a way to substitute alkenes.

<span class="mw-page-title-main">Pauson–Khand reaction</span> Chemical reaction

The Pauson–Khand (PK) reaction is a chemical reaction, described as a [2+2+1] cycloaddition. In it, an alkyne, an alkene and carbon monoxide combine into a α,β-cyclopentenone in the presence of a metal-carbonyl catalyst.

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

Magic acid (FSO3H·SbF5) is a superacid consisting of a mixture, most commonly in a 1:1 molar ratio, of fluorosulfuric acid (HSO3F) and antimony pentafluoride (SbF5). This conjugate Brønsted–Lewis superacid system was developed in the 1960s by the George Olah lab at Case Western Reserve University, and has been used to stabilize carbocations and hypercoordinated carbonium ions in liquid media. Magic acid and other superacids are also used to catalyze isomerization of saturated hydrocarbons, and have been shown to protonate even weak bases, including methane, xenon, halogens, and molecular hydrogen.

<span class="mw-page-title-main">Carbenium ion</span> Class of ions

A carbenium ion is a positive ion with the structure RR′R″C+, that is, a chemical species with carbon atom having three covalent bonds, and it bears a +1 formal charge. But IUPAC confuses coordination number with valence, incorrectly considering carbon in carbenium as trivalent.

<span class="mw-page-title-main">Halonium ion</span> Any onium ion containing a halogen atom carrying a positive charge

A halonium ion is any onium ion containing a halogen atom carrying a positive charge. This cation has the general structure R−+X−R′ where X is any halogen and no restrictions on R, this structure can be cyclic or an open chain molecular structure. Halonium ions formed from fluorine, chlorine, bromine, and iodine are called fluoronium, chloronium, bromonium, and iodonium, respectively. The 3-membered cyclic variety commonly proposed as intermediates in electrophilic halogenation may be called haliranium ions, using the Hantzsch-Widman nomenclature system.

In chemistry, carbonylation refers to reactions that introduce carbon monoxide (CO) into organic and inorganic substrates. Carbon monoxide is abundantly available and conveniently reactive, so it is widely used as a reactant in industrial chemistry. The term carbonylation also refers to oxidation of protein side chains.

In organometallic chemistry, a migratory insertion is a type of reaction wherein two ligands on a metal complex combine. It is a subset of reactions that very closely resembles the insertion reactions, and both are differentiated by the mechanism that leads to the resulting stereochemistry of the products. However, often the two are used interchangeably because the mechanism is sometimes unknown. Therefore, migratory insertion reactions or insertion reactions, for short, are defined not by the mechanism but by the overall regiochemistry wherein one chemical entity interposes itself into an existing bond of typically a second chemical entity e.g.:

Alcohol oxidation is a collection of oxidation reactions in organic chemistry that convert alcohols to aldehydes, ketones, carboxylic acids, and esters where the carbon carries a higher oxidation state. The reaction mainly applies to primary and secondary alcohols. Secondary alcohols form ketones, while primary alcohols form aldehydes or carboxylic acids.

In chemistry, decarbonylation is a type of organic reaction that involves the loss of carbon monoxide (CO). It is often an undesirable reaction, since it represents a degradation. In the chemistry of metal carbonyls, decarbonylation describes a substitution process, whereby a CO ligand is replaced by another ligand.

An insertion reaction is a chemical reaction where one chemical entity interposes itself into an existing bond of typically a second chemical entity e.g.:

β-Butyrolactone Chemical compound

β-Butyrolactone is the intramolecular carboxylic acid ester (lactone) of the optically active 3-hydroxybutanoic acid. It is produced during chemical synthesis as a racemate. β-Butyrolactone is suitable as a monomer for the production of the biodegradable polyhydroxyalkanoate poly(3-hydroxybutyrate) (PHB). Polymerisation of racemic (RS)-β-butyrolactone provides (RS)-polyhydroxybutyric acid, which, however, is inferior in essential properties (e.g. strength or degradation behaviour) to the (R)-poly-3-hydroxybutyrate originating from natural sources.

References

  1. Koch, H.; Haaf, W. Ann. 1958, "618", 251–266.( doi : 10.1002/jlac.19586180127)
  2. 1 2 Weissermel, K., Jargen-Arpe, H. In "Syntheses involving carbon monoxide", Industrial Organic Chemistry; VCH Publishers: New York, NY; pp. 141–145. ( ISBN   978-3527320028)
  3. Li, J. J. In "Koch–Haaf carbonylation"; Name Reactions, 4th ed.; Springer, Berlin, 2009; p. 319. ( doi : 10.1007/978-3-642-01053-8_140)
  4. Tsumori, N., Xu, Q., Souma, Y., Mori, H. J. Mol. Cat. A , 2002, 179, 271–77. ( doi : 10.1016/S1381-1169(01)00396-X)
  5. Xu, Q., Inoue, S., Tsumori, N., Mori, H., Kameda, M., Fujiwara, M., Souma, Y. J. Mol. Cat. A , 2001, 170, 147. ( doi : 10.1016/S1381-1169(01)00054-1)
  6. 1 2 Qiao, K., Yokoyama, C. Cat. Comm. 2006, 7, 450–453. ( doi : 10.1016/j.catcom.2005.12.009)
  7. Souma, Y. Sano, H., Iyoda, J. J. Org. Chem. , 1973, 38, 2016. ( doi : 10.1021/jo00951a010)
  8. Xu, Q., Imamura, Y., Fujiwara, M., Souma, Y. J. Org. Chem. , 1997, 62, 1594–1598. ( doi : 10.1021/jo9620122)
  9. Xu, Q., Souma, Y. Top. Catal. , 1998, 6, 17. ( doi : 10.1023/A:1019158221240)
  10. Stepanov, A. G., Luzgin, M. V., Romannikov, V. N., Zamaraev, K. I. J. Am. Chem. Soc. , 1995, 117, 3615–16. ( doi : 10.1021/ja00117a032)
  11. Barton, V., Ward, S. A., Chadwick, J., Hill, A., O'Neill, P. M. J. Med. Chem. , 2010, 53, 4555–59. ( doi : 10.1021/jm100201j)
  12. Brilman, D. W. F., van Swaaij, W. P. M., Versteeg, G. F. Chem. Eng. Sci. , 1999, 54, 4801–09. ( doi : 10.1016/S0009-2509(99)00197-9)
  13. Becker, C. L., Engstrom, K. M., Kerdesky, F. A., Tolle, J. C., Wagaw, S. H., Wang, W. Org. Process Res. Dev. , 2008, 12, 1114–18. ( doi : 10.1021/op800065q)