Catechol

Not to be confused with Catechin, also sometimes called catechol.
Catechol
Skeletal formula
Names
IUPAC name
Benzene-1,2-diol
Other names
pyrocatechol
1,2-benzenediol
2-hydroxyphenol
1,2-dihydroxybenzene
o-Benzenediol
o-Dihydroxybenzene
Identifiers
120-80-9 YesY
3D model (Jmol) Interactive image
ChEBI CHEBI:18135 N
ChEMBL ChEMBL280998 N
ChemSpider 283 YesY
ECHA InfoCard 100.004.025
EC Number 204-427-5
KEGG C00090 YesY
PubChem 289
RTECS number UX1050000
Properties
C6H6O2
Molar mass 110.1 g/mol
Appearance white to brown feathery crystals
Odor faint, phenolic odor
Density 1.344 g/cm3, solid
Melting point 105 °C (221 °F; 378 K)
Boiling point 245.5 °C (473.9 °F; 518.6 K) (sublimes)
430 g/L
Solubility very soluble in pyridine
soluble in chloroform, benzene, CCl4, ether, ethyl acetate
log P 0.88
Vapor pressure 20 Pa (20 °C)
Acidity (pKa) 9.48
1.604
2.62±0.03 D [1]
Structure
monoclinic
Hazards
Safety data sheet Sigma-Aldrich
Harmful (Xn)
R-phrases R21/22, R36/38
S-phrases (S2), S22, S26, S37
NFPA 704
Flammability code 1: Must be pre-heated before ignition can occur. Flash point over 93 °C (200 °F). E.g., canola oil Health code 3: Short exposure could cause serious temporary or residual injury. E.g., chlorine gas Reactivity code 0: Normally stable, even under fire exposure conditions, and is not reactive with water. E.g., liquid nitrogen Special hazards (white): no codeNFPA 704 four-colored diamond
1
3
0
Flash point 127 °C (261 °F; 400 K)
510 °C (950 °F; 783 K)
Explosive limits 1.4%-?[2]
Lethal dose or concentration (LD, LC):
300 mg/kg (rat, oral)
US health exposure limits (NIOSH):
PEL (Permissible)
none[2]
REL (Recommended)
TWA 5 ppm (20 mg/m3) [skin][2]
IDLH (Immediate danger)
N.D.[2]
Related compounds
Related benzenediols
Resorcinol
Hydroquinone
Related compounds
1,2-benzoquinone
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).
N verify (what is YesYN ?)
Infobox references

Catechol (/ˈkætɒl/), also known as pyrocatechol or 1,2-dihydroxybenzene, is an organic compound with the molecular formula C6H4(OH)2. It is the ortho isomer of the three isomeric benzenediols. This colorless compound occurs naturally in trace amounts. It was first discovered by destructive distillation of the plant extract catechin. About 20 million kg are now synthetically produced annually as a commodity organic chemical, mainly as a precursor to pesticides, flavors, and fragrances.

Catechol occurs as feathery white crystals that are very rapidly soluble in water.

(The name "catechol" has also been used as a chemical class name, where it refers generally to the catechins.)

Isolation and synthesis

Catechol was first isolated in 1839 by Edgar Hugo Emil Reinsch (1809 - 1884) by distilling it from the solid tannic preparation catechin, which is the residuum of catechu, the boiled or concentrated juice of Mimosa catechu (Acacia catechu L.f).[3] Upon heating catechin above its decomposition point, a substance that Reinsch first named Brenz-Katechusäure (burned catechu acid) sublimated as a white efflorescence. This was a thermal decomposition product of the flavanols in catechin. In 1841, both Wackenroder and Zwenger independently rediscovered catechol; in reporting on their findings, Philosophical Magazine coined the name pyrocatechin.[4] By 1852, Erdmann realized that catechol was benzene with two oxygen atoms added to it; in 1867, August Kekulé realized that catechol was a diol of benzene, so by 1868, catechol was listed as pyrocatechol.[5] In 1879, the Journal of the Chemical Society recommended that catechol be called "catechol", and in the following year, it was listed as such.[6]

Catechol has since been shown to occur in free-form naturally in kino and in beechwood tar. Its sulfonic acid has been detected in the urine of horses and humans.[7]

Catechol is produced industrially by the hydroxylation of phenol using hydrogen peroxide.[8]

C6H5OH + H2O2 → C6H4(OH)2 + H2O

Previously, it was produced by hydroxylation of salicylaldehyde using hydrogen peroxide,[9] as well as the hydrolysis of 2-substituted phenols, especially 2-chlorophenol, with hot aqueous solutions containing alkali metal hydroxides. Its methyl ether derivative, guaiacol, converts to catechol via hydrolysis of the CH3-O bond as promoted by hydriodic acid.[9]

Reactions

Organic chemistry

Like other difunctional benzene derivatives, catechol readily condenses to form heterocyclic compounds. Cyclic esters are formed upon treatment with phosphorus trichloride and phosphorus oxychloride, carbonyl chloride, and sulphuryl chloride:

C6H4(OH)2 + XCl2 → C6H4(O2X) + 2 HCl
where X = CO, SO2, PCl, P(O)Cl

Catechols produce quinones with the addition of ceric ammonium nitrate (CAN).

With metal ions

Catechol is the conjugate acid of a chelating agent used widely in coordination chemistry. Basic solutions of catechol react with iron(III) to give the red [Fe(C6H4O2)3]3−. Ferric chloride gives a green coloration with the aqueous solution, while the alkaline solution rapidly changes to a green and finally to a black color on exposure to the air.[10] Iron-containing dioxygenase enzymes catalyze the cleavage of catechol.

Redox chemistry

Catechol is produced by a reversible two-electron, two-proton reduction of 1,2-benzoquinone (E° = +795 mV vs SHE; Em (pH 7) = +380 mV vs SHE). [11] [12]

The redox series catecholate dianion, monoanionic semiquinonate, and benzoquinone are collectively called dioxolenes. Dioxolenes are used as ligands.[13]

Natural occurrences

Small amounts of catechol occur naturally in fruits and vegetables, along with the enzyme polyphenol oxidase (also known as catecholase, or catechol oxidase). Upon mixing the enzyme with the substrate and exposure to oxygen (as when a potato or apple is cut and left out), the colorless catechol oxidizes to reddish-brown melanoid pigments, derivatives of benzoquinone. The enzyme is inactivated by adding an acid, such as lemon juice. Excluding oxygen also prevents the browning reaction. However, the activity of the enzyme increases in cooler temperatures. Benzoquinone is said to be antimicrobial, which slows the spoilage of wounded fruits and other plant parts.

It is one of the main natural phenols in argan oil.[14]

Pyrocatechol is also found in Agaricus bisporus.[15]

It is also a component of castoreum, a substance from the castor gland of beavers, used in perfumery.

Presence of the catechol moiety

Catechol moieties are also found widely within the natural world. Arthropod cuticle consists of chitin linked by a catechol moiety to protein. The cuticle may be strengthened by cross-linking (tanning and sclerotization), in particular, in insects, and of course by biomineralization.[16] Catechols such as DHSA are produced through the metabolism of cholesterol by bacteria such as Mycobacterium tuberculosis.[17]

Urushiols are naturally existing organic compounds that have the catechol skeleton structure and diphenol functionality but with alkyl groups substituted onto the aromatic ring. Urushiols are the skin-irritating poisons found in plants like poison ivy, etc. Catecholamines are biochemically significant hormones/neurotransmitters that are phenethylamines in which the phenyl group has a catechol skeleton structure.

Parts of a molecule of catechin, another natural compound present in tea, contains a catechol group.

Uses

Approximately 50% of synthetic catechol is consumed in the production of pesticides, the remainder being used as a precursor to fine chemicals such as perfumes and pharmaceuticals.[8] It is a common building block in organic synthesis.[18] Several industrially significant flavors and fragrances are prepared starting from catechol. Guaiacol is prepared by methylation of catechol and is then converted to vanillin on a scale of about 10M kg per year (1990). The related monoethyl ether of catechol, guethol, is converted to ethylvanillin, a component of chocolate confectioneries. 3-Trans-Isocamphylcyclohexanol, widely used as a replacement for sandalwood oil, is prepared from catechol via guaiacol and camphor. Piperonal, a flowery scent, is prepared from the methylene diether of catechol followed by condensation with glyoxal and decarboxylation.[19]

Catechol is used as a black-and-white photographic developer, but, except for some special purpose applications, its use is largely historical. It is rumored to have been used briefly in Eastman Kodak's HC-110 developer and is rumored to be a component in Tetenal's Neofin Blau developer.[20] It is a key component of Finol from Moersch Photochemie in Germany. Modern catechol developing was pioneered by noted photographer Sandy King. His "PyroCat" formulation is popular among modern black-and-white film photographers.[21] King's work has since inspired further 21st-century development by others such as Jay De Fehr with Hypercat and Obsidian Acqua developers, and others.[20]

Catechol derivatives

The catechol skeleton occurs in a variety of natural products such as urushiols, which are the skin-irritating poisons found in plants like poison ivy, and catecholamines, drugs imitating them (such as MDMA), hormones/neurotransmitters, and catechin, which is found in tea. Many pyrocatechin derivatives have been suggested for therapeutic applications.

Nomenclature

Although rarely encountered, the officially "preferred IUPAC name" (PIN) of catechol is benzene-1,2-diol.[22] The trivial name pyrocatechol is a retained IUPAC name, according to the 1993 Recommendations for the Nomenclature of Organic Chemistry. [23] [24]

See also

References

  1. Lander, John J.; Svirbely, John J. Lander, W. J. (1945). "The Dipole Moments of Catechol, Resorcinol and Hydroquinone". Journal of the American Chemical Society. 67 (2): 322–324. doi:10.1021/ja01218a051.
  2. 1 2 3 4 "NIOSH Pocket Guide to Chemical Hazards #0109". National Institute for Occupational Safety and Health (NIOSH).
  3. Hugo Reinsch (1839) "Einige Bemerkungen über Catechu" (Some observations about catechu), Repertorium für die Pharmacie, 68 : 49-58. Reinsch describes the preparaton of catechol on p. 56: "Bekanntlich wird die Katechusäure bei der Destillation zerstört, während sich ein geringer Theil davon als krystallinischer Anflug sublimirt, welcher aber noch nicht näher untersucht worden ist. Diese Säure ist vielleicht dieselbe, welche ich bei der zerstörenden Destillation des Katechus erhalten; … " (As is well known, catechu acid is destroyed by distillation, while a small portion of it sublimates as a crystalline efflorescence, which however has still not been closely examined. This acid is perhaps the same one, which I obtained by destructive distillation of catechu; … ). On p. 58, Reinsch names the new compound: "Die Eigenschaften dieser Säure sind so bestimmt, dass man sie füglich als eine eigenthümliche Säure betrachten und sie mit dem Namen Brenz-Katechusäure belegen kann." (The properties of this acid are so definite, that one can regard it justifiably as a strange acid and give it the name "burned catechu acid".)
  4. See:
  5. See:
    • Rudolf Wagner (1852) "Ueber die Farbstoffe des Gelbholzes (Morus tinctoria.)" (On the coloring matter of Dyer's mulberry (Morus tinctoria.)), Journal für praktische Chemie, 55 : 65-76. See p. 65.
    • August Kekulé (1867) "Ueber die Sulfosäuren des Phenols" (On the sulfonates of phenol) Zeitschrift für Chemie, new series, 3 : 641-646 ; see p. 643.
    • Joseph Alfred Naquet, with William Cortis, trans. and Thomas Stevenson, ed., Principles of Chemistry, founded on Modern Theories, (London, England: Henry Renshaw, 1868), p. 657. See also p. 720.
  6. See:
    • In 1879, the Publication Committee of the Journal of the Chemical Society issued instructions to its abstractors to "Distinguish all alcohols, i.e., hydroxyl-derivations of hydrocarbons, by names ending in ol, e.g., quinol, catechol, … " See: Alfred H. Allen (June 20, 1879) "Nomenclature of organic bodies," English Mechanic and World of Science, 29 (743) : 369.
    • William Allen Miller, ed., Elements of Chemistry: Theoretical and Practical, Part III: Chemistry of Carbon Compounds or Organic Chemistry, Section I … , 5th ed. (London, England: Longmans, Green and Co., 1880), p.524.
  7. Zheng, L. T.; Ryu, G. M.; Kwon, B. M.; Lee, W. H.; Suk, K. (2008). "Anti-inflammatory effects of catechols in lipopolysaccharide-stimulated microglia cells: Inhibition of microglial neurotoxicity". European Journal of Pharmacology. 588: 106–13. doi:10.1016/j.ejphar.2008.04.035. PMID 18499097.
  8. 1 2 Fiegel, Helmut et al. (2002) "Phenol Derivatives" in Ullmann's Encyclopedia of Industrial Chemistry, Wiley-VCH: Weinheim. doi:10.1002/14356007.a19_313.
  9. 1 2 "http://www.orgsyn.org/demo.aspx?prep=CV1P0149". www.orgsyn.org. Retrieved 2016-11-17. External link in |title= (help)
  10. Anderson, Bryan F.; Buckingham, David A.; Robertson, Glen B.; Webb, John; Murray, Keith S.; Clark, Paul E. (1976). "Models for the bacterial iron-transport chelate enterochelin". Nature. 262: 722–724. doi:10.1038/262722a0.
  11. Horner, Leopold; Geyer, Ekkehard (1965). "Zur Kenntnis der o-Chinone, XXVII: Redoxpotentiale von Brenzcatechin-Derivaten". Chemische Berichte. 98 (6): 2016–2045. doi:10.1002/cber.19650980641.
  12. Nematollahi, D.; Rafiee, M. (2004-05-01). "Electrochemical oxidation of catechols in the presence of acetylacetone". Journal of Electroanalytical Chemistry. 566 (1): 31–37. doi:10.1016/j.jelechem.2003.10.044.
  13. Griffith, W. P. (1993). "Recent advances in dioxolene chemistry". Transition Metal Chemistry. 18 (2): 250–256. doi:10.1007/BF00139966.
  14. Charrouf, Z.; Guillaume, D. (2007). "Phenols and Polyphenols from Argania spinosa". American Journal of Food Technology. 2 (7): 679–683. doi:10.3923/ajft.2007.679.683.
  15. Delsignore, A; Romeo, F; Giaccio, M (1997). "Content of phenolic substances in basidiomycetes". Mycological Research. 101 (5): 552–6. doi:10.1017/S0953756296003206.
  16. Briggs DEG (1999). "Molecular taphonomy of animal and plant cuticles: selective preservation and diagenesis". Philosophical Transactions of the Royal Society B: Biological Sciences. 354 (1379): 7–17. doi:10.1098/rstb.1999.0356.
  17. PDB: 2ZI8; Yam KC, D'Angelo I, Kalscheuer R, Zhu H, Wang JX, Snieckus V, Ly LH, Converse PJ, Jacobs WR, Strynadka N, Eltis LD (March 2009). "Studies of a ring-cleaving dioxygenase illuminate the role of cholesterol metabolism in the pathogenesis of Mycobacterium tuberculosis". PLoS Pathog. 5 (3): e1000344. doi:10.1371/journal.ppat.1000344. PMC 2652662Freely accessible. PMID 19300498.
  18. Barner, B. A. (2004) "Catechol" in Encyclopedia of Reagents for Organic Synthesis (Ed: L. Paquette), J. Wiley & Sons, New York. doi:10.1002/047084289.
  19. Fahlbusch, Karl-Georg et al. (2003) "Flavors and Fragrances" in Ullmann's Encyclopedia of Industrial Chemistry, Wiley-VCH: Weinheim doi:10.1002/14356007.a11_141.
  20. 1 2 Stephen G. Anchell. The Darkroom Cookbook. ISBN 978-1136092770.
  21. Stephen G. Anchell; Bill Troop. The Film Developing Cookbook. ISBN 978-0240802770.
  22. Preferred IUPAC Names. September 2004, Chapter 6, Sect 60–64, p. 38
  23. IUPAC, Commission on Nomenclature of Organic Chemistry. A Guide to IUPAC Nomenclature of Organic Compounds (Recommendations 1993) R-5.5.1.1 Alcohols and phenols.
  24. Panico, R.; Powell, W. H., eds. (1994). A Guide to IUPAC Nomenclature of Organic Compounds 1993. Oxford: Blackwell Science. ISBN 0-632-03488-2.

 This article incorporates text from a publication now in the public domain: Chisholm, Hugh, ed. (1911). "Catechu". Encyclopædia Britannica (11th ed.). Cambridge University Press. 

Wikimedia Commons has media related to Catechols.
This article is issued from Wikipedia - version of the 11/17/2016. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.