Scholl reaction

Last updated
Scholl reaction
Named after Roland Scholl
Reaction type Coupling reaction
Identifiers
RSC ontology ID RXNO:0000377

The Scholl reaction is a coupling reaction between two arene compounds with the aid of a Lewis acid and a protic acid. [1] [2] It is named after its discoverer, Roland Scholl, a Swiss chemist.

Contents

The Scholl reaction Scholl-reactie.png
The Scholl reaction

In 1910 Scholl reported the synthesis of a quinone [3] and of perylene from naphthalene [4] both with aluminum chloride. Perylene was also synthesised from 1,1’-binaphthalene in 1913. [5] The synthesis of Benzanthrone was reported in 1912. [6]

The protic acid in the Scholl reaction is often an impurity in the Lewis Acid and also formed in the course of a Scholl reaction. Reagents are iron(III) chloride in dichloromethane, copper(II) chloride, PIFA and boron trifluoride etherate in dichloromethane, Molybdenum(V) chloride and lead tetraacetate with BF3 in acetonitrile. [7]

Given the high reaction temperature and the requirement for strongly acidic catalysts the chemical yield often is low and the method is not a popular one. Intramolecular reactions fare better than the intermolecular ones, for instance in the organic synthesis of 9-phenylfluorene:

An example of the Scholl reaction Schollvb.png
An example of the Scholl reaction

Or the formation of the pyrene dibenzo-(a.1)-pyrene from the anthracene 1-phenylbenz(a)anthracene (66% yield). [8]

An example of the Scholl reaction Schollvb2.png
An example of the Scholl reaction

One study showed that the reaction lends itself to cascade reactions to form more complex polycyclic aromatic hydrocarbons [9]

In certain applications such as triphenylene synthesis this reaction is advocated as an alternative for the Suzuki reaction. A recurring problem is oligomerization of the product which can be prevented by blocking tert-butyl substituents: [7]

Scholl reaction Blocking Substituents SchollreactionBlockingSubstituents.png
Scholl reaction Blocking Substituents

Reaction mechanism

The exact reaction mechanism is not known but could very well proceed through an arenium ion. Just as in electrophilic aromatic substitution, Activating groups such as methoxy improve yield and selectivity: [7]

Scholl Reaction Para directors SchollReactionParadirectors.png
Scholl Reaction Para directors

Indeed, oxidative coupling of phenols is a research strategy in modern organic synthesis.

Two mechanisms may compete. In step one of a radical cation mechanism a radical cation is formed from one reaction partner by oxidation, in step two the radical ion attacks the second neutral partner in a substitution reaction and a new radical ion is formed with one ring bearing the positive charge and the other one the radical position. In step three dihydrogen is split off with rearomatisation to the biaryl compound. In the arenium ion mechanism one reaction partner is protonated to an arenium ion which then attacks the second reaction partner. The arenium ion can also be formed by attack of the Lewis acid. The mechanisms are difficult to distinguish because many Lewis acids can behave as oxidants. Reactions taking place at room-temperature with well-known one-electron oxidizing agents likely proceed through a radical cation mechanism and reactions requiring elevated temperatures likely proceed through an arenium ion mechanism. [2]

Related Research Articles

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.

The Sandmeyer reaction is a chemical reaction used to synthesize aryl halides from aryl diazonium salts using copper salts as reagents or catalysts. It is an example of a radical-nucleophilic aromatic substitution. The Sandmeyer reaction provides a method through which one can perform unique transformations on benzene, such as halogenation, cyanation, trifluoromethylation, and hydroxylation.

<span class="mw-page-title-main">Radical anion</span> Free radical species

In organic chemistry, a radical anion is a free radical species that carries a negative charge. Radical anions are encountered in organic chemistry as reduced derivatives of polycyclic aromatic compounds, e.g. sodium naphthenide. An example of a non-carbon radical anion is the superoxide anion, formed by transfer of one electron to an oxygen molecule. Radical anions are typically indicated by .

The Ullmann reaction or Ullmann coupling, named after Fritz Ullmann, couples two aryl or alkyl groups with the help of copper. The reaction was first reported by Ullmann and his student Bielecki in 1901. It has been later shown that palladium and nickel can also be effectively used.

<span class="mw-page-title-main">Diazonium compound</span> Group of organonitrogen compounds

Diazonium compounds or diazonium salts are a group of organic compounds sharing a common functional group [R−N+≡N]X where R can be any organic group, such as an alkyl or an aryl, and X is an inorganic or organic anion, such as a halide.

(<i>E</i>)-Stilbene Chemical compound

(E)-Stilbene, commonly known as trans-stilbene, is an organic compound represented by the condensed structural formula C6H5CH=CHC6H5. Classified as a diarylethene, it features a central ethylene moiety with one phenyl group substituent on each end of the carbon–carbon double bond. It has an (E) stereochemistry, meaning that the phenyl groups are located on opposite sides of the double bond, the opposite of its geometric isomer, cis-stilbene. Trans-stilbene occurs as a white crystalline solid at room temperature and is highly soluble in organic solvents. It can be converted to cis-stilbene photochemically, and further reacted to produce phenanthrene.

In organic chemistry, the Arndt–Eistert reaction is the conversion of a carboxylic acid to its homologue. Named for the German chemists Fritz Arndt (1885–1969) and Bernd Eistert (1902–1978), the method entails treating an acid chlorides with diazomethane. It is a popular method of producing β-amino acids from α-amino acids.

The Gattermann reaction (also known as the Gattermann formylation and the Gattermann salicylaldehyde synthesis) is a chemical reaction in which aromatic compounds are formylated by a mixture of hydrogen cyanide (HCN) and hydrogen chloride (HCl) in the presence of a Lewis acid catalyst such as aluminium chloride (AlCl3). It is named for the German chemist Ludwig Gattermann and is similar to the Friedel–Crafts reaction.

<span class="mw-page-title-main">Finkelstein reaction</span> Chemistry

The Finkelstein reaction, named after the German chemist Hans Finkelstein, is a type of SN2 reaction that involves the exchange of one halogen atom for another. It is an equilibrium reaction, but the reaction can be driven to completion by exploiting the differential solubility of various halide salts, or by using a large excess of the desired halide.

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

The Auwers synthesis is a series of organic reactions forming a flavonol from a coumarone. This reaction was first reported by Karl von Auwers in 1908.

The Letts nitrile synthesis is a chemical reaction of aromatic carboxylic acids with metal thiocyanates to form nitriles. The reaction includes the loss of carbon dioxide and potassium hydrosulfide. The polar basic substitution reaction was discovered in 1872 by Edmund A. Letts.

In electrochemistry, electrosynthesis is the synthesis of chemical compounds in an electrochemical cell. Compared to ordinary redox reactions, electrosynthesis sometimes offers improved selectivity and yields. Electrosynthesis is actively studied as a science and also has industrial applications. Electrooxidation has potential for wastewater treatment as well.

The Elbs reaction is an organic reaction describing the pyrolysis of an ortho methyl substituted benzophenone to a condensed polyaromatic. The reaction is named after its inventor, the German chemist Karl Elbs, also responsible for the Elbs oxidation. The reaction was published in 1884. Elbs however did not correctly interpret the reaction product due to a lack of knowledge about naphthalene structure.

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

Diphenylketene is a chemical substance of the ketene family. Diphenylketene, like most stable disubstituted ketenes, is a red-orange oil at room temperature and pressure. Due to the successive double bonds in the ketene structure R1R2C=C=O, diphenyl ketene is a heterocumulene. The most important reaction of diphenyl ketene is the [2+2] cycloaddition at C-C, C-N, C-O, and C-S multiple bonds.

The Hofmann–Löffler reaction (also referred to as Hofmann–Löffler–Freytag reaction, Löffler–Freytag reaction, Löffler–Hofmann reaction, as well as Löffler's method) is an organic reaction in which a cyclic amine 2 (pyrrolidine or, in some cases, piperidine) is generated by thermal or photochemical decomposition of N-halogenated amine 1 in the presence of a strong acid (concentrated sulfuric acid or concentrated CF3CO2H). The Hofmann–Löffler–Freytag reaction proceeds via an intramolecular hydrogen atom transfer to a nitrogen-centered radical and is an example of a remote intramolecular free radical C–H functionalization.

Unlike its lighter congeners, the halogen iodine forms a number of stable organic compounds, in which iodine exhibits higher formal oxidation states than -1 or coordination number exceeding 1. These are the hypervalent organoiodines, often called iodanes after the IUPAC rule used to name them.

Organosodium chemistry is the chemistry of organometallic compounds containing a carbon to sodium chemical bond. The application of organosodium compounds in chemistry is limited in part due to competition from organolithium compounds, which are commercially available and exhibit more convenient reactivity.

Osmocene is an organoosmium compound found as a white solid. It is a metallocene with the formula Os(C5H5)2.

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.:

The Buchner ring expansion is a two-step organic C-C bond forming reaction used to access 7-membered rings. The first step involves formation of a carbene from ethyl diazoacetate, which cyclopropanates an aromatic ring. The ring expansion occurs in the second step, with an electrocyclic reaction opening the cyclopropane ring to form the 7-membered ring.

References

  1. M B Smith, J March. March's Advanced Organic Chemistry (Wiley, 2001) ( ISBN   0-471-58589-0)
  2. 1 2 Grzybowski, M., Skonieczny, K., Butenschön, H. and Gryko, D. T. (2013), Comparison of Oxidative Aromatic Coupling and the Scholl Reaction. Angew. Chem. Int. Ed., 52: 9900–9930. doi : 10.1002/anie.201210238
  3. Scholl, R. and Mansfeld, J. (1910), meso-Benzdianthron (Helianthron), meso-Naphthodianthron, und ein neuer Weg zum Flavanthren. Ber. Dtsch. Chem. Ges., 43: 1734–1746. doi : 10.1002/cber.19100430288
  4. Scholl, R., Seer, Chr. and Weitzenböck, R. (1910), Perylen, ein hoch kondensierter aromatischer Kohlenwasserstoff C20H12. Ber. Dtsch. Chem. Ges., 43: 2202–2209. doi : 10.1002/cber.191004302175
  5. Weitzenböck, R. and Seer, C. (1913), Zur Kenntnis des Perylens und seiner Derivate. (2. Mitteilung). Ber. Dtsch. Chem. Ges., 46: 1994–2000. doi : 10.1002/cber.191304602115
  6. Scholl, R. and Seer, C. (1912), Abspaltung aromatisch gebundenen Wasserstoffs und Verknüpfung aromatischer Kerne durch Aluminiumchlorid. Justus Liebigs Ann. Chem., 394: 111–177. doi : 10.1002/jlac.19123940202
  7. 1 2 3 Controlling the Scholl Reaction Benjamin T. King, Jií Kroulík, Charles R. Robertson, Pawel Rempala, Cameron L. Hilton, Justin D. Korinek, and Lisa M. Gortari J. Org. Chem.; 2007; 72(7) pp 2279 - 2288; (Article) doi : 10.1021/jo061515x
  8. Vingiello, F. A.; Yanez, J.; Campbell, J. A. J. Org. Chem. 1971, 36, 2053-2056. ( doi : 10.1021/jo00814a005)
  9. Nanosized Molecular Propellers by Cyclodehydrogenation of Polyphenylene Dendrimers Christopher D. Simpson, Gunter Mattersteig, Kai Martin, Lileta Gherghel, Roland E. Bauer, Hans Joachim Räder, and Klaus Müllen J. Am. Chem. Soc.; 2004; 126(10) pp 3139 - 3147; (Article) doi : 10.1021/ja036732j

See also