Letts nitrile synthesis

Last updated
Letts nitrile synthesis
Named afterEdmund A. Letts
Reaction type Substitution reaction

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. [1] [2]

Contents

The Letts nitrile synthesis Letts nitrile synthesis V.1.png
The Letts nitrile synthesis

History

In 1857 Hugo Schiff observed that the reaction between benzoyl chloride with potassium cyanide produced the desired benzonitrile. [3] Work done later by British chemist Edmund A. Letts delved much deeper into the synthesis of nitriles. Attempting first to add cyano-groups to acetic acid, he obtained a mixture of acetamide and carbonyl sulfide. However, in 1872 he showed that treating a 2:1 molecular ratio of benzoic acid and potassium thiocyanate with heat for several hours also produced nitriles with only a small amount of amide with about 40% yield. [4]

G. Krüss expanded on Letts' work in 1884, producing better yields by utilizing lead(II) thiocyanate. [5] In 1916, E.E. Reid found that showed that dry distillation of the zinc(II) salt of the acid with a 20% excess of lead(II) thiocyanate gave an 86% conversion and 91% yield, almost double of that produced by Letts. [6]

Reid's improvement on the Letts synthesis Reid modified Letts.png
Reid's improvement on the Letts synthesis

Mechanism

Kekulé proposed the reaction mechanism in 1873. [7]

Proposed Mechanism for the Letts Nitrile Synthesis CHEM 342.png
Proposed Mechanism for the Letts Nitrile Synthesis

In this polar basic substitution reaction mechanism, thiocyanate ion extracts the acidic proton from benzoic acid while heated. This yields the conjugate base (stabilized by resonance structures) and thiocyanic acid.

The next step involves the evolution of carbon dioxide, where a lone pair of electrons moves from the negatively charged oxygen to form a double bond with the carboxylic carbon. The sigma bond between the ring and carboxyl group is then severed, the electron pair moving to the ring and delocalized through resonance structures.

The final step of the mechanism involves the attack of the phenyl anion attacking the cyano-carbon, pushing the electron pair over to the sulfur, which readily diffuses the negative charge and is further stabilized by the potassium ion, resulting in the final benzonitrile product and potassium hydrosulfide.

Applications

Aromatic nitriles have a few applications, including polyrecombination to form polymers, [8] are sometimes studied as biologically active molecules [9] and undergoing Ritter reactions to form amides. [10]

Restrosynthesis of Fluvoxamine Ritter Reaction Scheme.svg
Restrosynthesis of Fluvoxamine

Benzonitrile, the original product of Letts, has multiple uses as a versatile reagent and as a solvent. Substituted benzonitriles are important in many fields including pharmaceuticals. Benzonitrile is a precursor in the synthesis of Fadrozole, an aromatase inhibitor used in the treatment of breast cancer. [11] 4-(trifluoromethyl)benzonitrile, produced by the Nickel catalyzed cyanation of 4-chlorobenzotrifluoride is a precursor for the antidepressant Fluvoxamine. [12]

Restrosynthesis of Fluvoxamine Restrosynthesis of Fluvoxamine.png
Restrosynthesis of Fluvoxamine

Benzonitrile can also act a ligand in asymmetric catalysis, coordinating to transition metals and forming Lewis acids. [13] [14]

See also

For synthesis of nitriles:

For reactions of nitriles:

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 R−COOH or R−CO2H, 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.

Pyrrole is a heterocyclic, aromatic, organic compound, a five-membered ring with the formula C4H4NH. It is a colorless volatile liquid that darkens readily upon exposure to air. Substituted derivatives are also called pyrroles, e.g., N-methylpyrrole, C4H4NCH3. Porphobilinogen, a trisubstituted pyrrole, is the biosynthetic precursor to many natural products such as heme.

A polyamide is a polymer with repeating units linked by amide bonds.

In organic chemistry, a nitrile is any organic compound that has a −C≡N functional group. The prefix cyano- is used interchangeably with the term nitrile in industrial literature. Nitriles are found in many useful compounds, including methyl cyanoacrylate, used in super glue, and nitrile rubber, a nitrile-containing polymer used in latex-free laboratory and medical gloves. Nitrile rubber is also widely used as automotive and other seals since it is resistant to fuels and oils. Organic compounds containing multiple nitrile groups are known as cyanocarbons.

<span class="mw-page-title-main">Nitration</span> Chemical reaction which adds a nitro (–NO₂) group onto a molecule

In organic chemistry, nitration is a general class of chemical processes for the introduction of a nitro group into an organic compound. The term also is applied incorrectly to the different process of forming nitrate esters between alcohols and nitric acid. The difference between the resulting molecular structures of nitro compounds and nitrates is that the nitrogen atom in nitro compounds is directly bonded to a non-oxygen atom, whereas in nitrate esters, the nitrogen is bonded to an oxygen atom that in turn usually is bonded to a carbon atom.

<span class="mw-page-title-main">Hydrogen peroxide - urea</span> Chemical compound

Hydrogen peroxide - urea is a white crystalline solid chemical compound composed of equal amounts of hydrogen peroxide and urea. It contains solid and water-free hydrogen peroxide, which offers a higher stability and better controllability than liquid hydrogen peroxide when used as an oxidizing agent. Often called carbamide peroxide in dentistry, it is used as a source of hydrogen peroxide when dissolved in water for bleaching, disinfection and oxidation.

The Hofmann rearrangement is the organic reaction of a primary amide to a primary amine with one less carbon atom. The reaction involves oxidation of the nitrogen followed by rearrangement of the carbonyl and nitrogen to give an isocyanate intermediate. The reaction can form a wide range of products, including alkyl and aryl amines.

The von Richter reaction, also named von Richter rearrangement, is a name reaction in the organic chemistry. It is named after Victor von Richter, who discovered this reaction in year 1871. It is the reaction of aromatic nitro compounds with potassium cyanide in aqueous ethanol to give the product of cine substitution by a carboxyl group. Although it is not generally synthetically useful due to the low chemical yield and formation of numerous side products, its mechanism was of considerable interest, eluding chemists for almost 100 years before the currently accepted one was proposed.

The Bischler–Napieralski reaction is an intramolecular electrophilic aromatic substitution reaction that allows for the cyclization of β-arylethylamides or β-arylethylcarbamates. It was first discovered in 1893 by August Bischler and Bernard Napieralski, in affiliation with Basle Chemical Works and the University of Zurich. The reaction is most notably used in the synthesis of dihydroisoquinolines, which can be subsequently oxidized to isoquinolines.

<span class="mw-page-title-main">Curtius rearrangement</span> Chemical reaction

The Curtius rearrangement, first defined by Theodor Curtius in 1885, is the thermal decomposition of an acyl azide to an isocyanate with loss of nitrogen gas. The isocyanate then undergoes attack by a variety of nucleophiles such as water, alcohols and amines, to yield a primary amine, carbamate or urea derivative respectively. Several reviews have been published.

<span class="mw-page-title-main">Reimer–Tiemann reaction</span> Chemical reaction for ortho-formylation of phenols

The Reimer–Tiemann reaction is a chemical reaction used for the ortho-formylation of phenols. with the simplest example being the conversion of phenol to salicylaldehyde. The reaction was first reported by Karl Reimer and Ferdinand Tiemann.

Stephen aldehyde synthesis, a named reaction in chemistry, was invented by Henry Stephen (OBE/MBE). This reaction involves the preparation of aldehydes (R-CHO) from nitriles (R-CN) using tin(II) chloride (SnCl2), hydrochloric acid (HCl) and quenching the resulting iminium salt ([R-CH=NH2]+Cl) with water (H2O). During the synthesis, ammonium chloride is also produced.

<span class="mw-page-title-main">Rosenmund–von Braun reaction</span>

The Rosenmund–von Braun synthesis is an organic reaction in which an aryl halide reacts with cuprous cyanide to yield an aryl nitrile.

In organic synthesis, cyanation is the attachment or substitution of a cyanide group on various substrates. Such transformations are high-value because they generate C-C bonds. Furthermore nitriles are versatile functional groups.

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.

The retro-Diels–Alder reaction is the reverse of the Diels–Alder (DA) reaction, a [4+2] cycloelimination. It involves the formation of a diene and dienophile from a cyclohexene. It can be accomplished spontaneously with heat, or with acid or base mediation.

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

Imidoyl chlorides are organic compounds that contain the functional group RC(NR')Cl. A double bond exist between the R'N and the carbon centre. These compounds are analogues of acyl chloride. Imidoyl chlorides tend to be highly reactive and are more commonly found as intermediates in a wide variety of synthetic procedures. Such procedures include Gattermann aldehyde synthesis, Houben-Hoesch ketone synthesis, and the Beckmann rearrangement. Their chemistry is related to that of enamines and their tautomers when the α hydrogen is next to the C=N bond. Many chlorinated N-heterocycles are formally imidoyl chlorides, e.g. 2-chloropyridine, 2, 4, and 6-chloropyrimidines.

P<sub>4</sub>-<i>t</i>-Bu Chemical compound

P4-t-Bu is a readily accessible chemical from the group of neutral, peralkylated sterically hindered polyaminophosphazenes, which are extremely strong bases but very weak nucleophiles, with the formula (CH3)3C−N=P(−N=P(−N(CH3)2)3)3. "t-Bu" stands for tert-butyl(CH3)3C–. "P4" stands for the fact that this molecule has 4 phosphorus atoms. P4-t-Bu can also be regarded as tetrameric triaminoiminophosphorane of the basic structure H−N=P(−NH2)3. The homologous series of P1 to P7 polyaminophosphazenes of the general formula with preferably methyl groups as R1, a methyl group or tert-butyl group as and even-numbered x between 0 and 6 (P4-t-Bu: R1 = Me, R2 = t-Bu and x = 3) has been developed by Reinhard Schwesinger; the resulting phosphazene bases are therefore also referred to as Schwesinger superbases.

<span class="mw-page-title-main">Tris(dimethylamino)methane</span> Chemical compound

Tris(dimethylamino)methane (TDAM) is the simplest representative of the tris(dialkylamino)methanes of the general formula (R2N)3CH in which three of the four of methane's hydrogen atoms are replaced by dimethylamino groups (−N(CH3)2). Tris(dimethylamino)methane can be regarded as both an amine and an orthoamide.

<span class="mw-page-title-main">Transition metal nitrile complexes</span> Class of coordination compounds containing nitrile ligands (coordinating via N)

Transition metal nitrile complexes are coordination compounds containing nitrile ligands. Because nitriles are weakly basic, the nitrile ligands in these complexes are often labile.

References

  1. Letts, E. A. (1872). "Neue Bildungsweisen der Amide und Nitrile". Chemische Berichte . 5 (2): 669–674. doi:10.1002/cber.18720050228.
  2. Letts, E. A. (1872). "New Method for Producing Amides and Nitriles". Proceedings of the Royal Society of London. 21 (139–147): 61–66. Bibcode:1872RSPS...21...61L. doi:10.1098/rspl.1872.0012. ISSN   0370-1662. S2CID   178687230.
  3. Schiff, Hugo (1857). "Ueber einige Derivate des Naphtylamins". Annalen der Chemie und Pharmacie. 101 (1): 90–93. doi:10.1002/jlac.18571010113.
  4. Reid, E.E. (1910). "Studies in the preparation of Nitriles". Am. Chem. J. 43: 162–181.
  5. Krüss, G. (1884). "Ueber eine neue Darstellungsmethode für Nitrile". Ber. 17 (2): 1766–1768. doi:10.1002/cber.18840170245.
  6. Mowry, D.T. The Preparation of Nitriles. (1948). "The preparation of nitriles". Chem. Rev. 42 (2): 189–283. doi:10.1021/cr60132a001. PMID   18914000.
  7. Kekulé, A. (1872). "Neue Bildungsweisen der Amide und Nitrile". Ber. 5 (2): 669–674. doi:10.1002/cber.18720050228.
  8. Vasnev, V.A.; Sosin, S.L.; Korshak, V.V. (1964). "Synthesis of polymers by polyrecombination of the nitrile of aromatic and aliphatic acids". Polymer Science U.S.S.R. 6 (5): 928. doi:10.1016/0032-3950(64)90510-6.
  9. Zhao, Aimin; Li, Whenzong; Yang, Huifung (1993). "Microbial Conversion of aromatic nitriles". Weis. Tong. 20: 169–171.
  10. Sadeghi, Bahareh; Farahzadi, Ebrahim; Hassanabadi, Alireza (2012-01-01). "KAl(SO4)2.12H2O as an eco-friendly and reusable catalyst for the synthesis of amides by the Ritter reaction". Journal of Chemical Research. 36 (9): 539–540. doi:10.3184/174751912x13418518739562. S2CID   197289966.
  11. Raats, J. I.; Falkson, G.; Falkson, H. C. (1992). "A study of fadrozole, a new aromatase inhibitor, in postmenopausal women with advanced metastatic breast cancer". Journal of Clinical Oncology. 10 (1): 111–116. doi:10.1200/jco.1992.10.1.111. ISSN   0732-183X. PMID   1530798.
  12. Schareina, Thomas; Zapf, Alexander; Beller, Matthias (2004-01-01). "Improving palladium-catalyzed cyanation of aryl halides: development of a state-of-the-art methodology using potassium hexacyanoferrate(II) as cyanating agent". Journal of Organometallic Chemistry. 689 (24): 4576–4583. doi:10.1016/j.jorganchem.2004.08.020.
  13. Becker, Jennifer J.; Orden, Lori J. Van; White, Peter S.; Gagné, Michel R. (2002-01-01). "Electron-Poor Benzonitriles as Labile, Stabilizing Ligands in Asymmetric Catalysis". Organic Letters. 4 (5): 727–730. doi:10.1021/ol017218q. PMID   11869112.
  14. Anderson, Gordon K.; Lin, Minren (1990). "Bis(Benzonitrile)Dichloro Complexes of Palladium and Platinum". Inorganic Syntheses. pp. 60–63. doi:10.1002/9780470132593.ch13. ISBN   9780470132593.{{cite book}}: |journal= ignored (help)