Cannabis Ruderalis

Decarboxylation

Decarboxylation is a chemical reaction that removes a carboxyl group and releases carbon dioxide (CO2). Usually, decarboxylation refers to a reaction of carboxylic acids, removing a carbon atom from a carbon chain. The reverse process, which is the first chemical step in photosynthesis, is called carboxylation, the addition of CO2 to a compound. Enzymes that catalyze decarboxylations are called decarboxylases or, the more formal term, carboxy-lyases (EC number 4.1.1).

In organic chemistry[edit]

The term "decarboxylation" usually means replacement of a carboxyl group (−C(O)OH) with a hydrogen atom:

RCO2H → RH + CO2

Decarboxylation is one of the oldest known organic reactions. It is one of the processes assumed to accompany pyrolysis and destructive distillation.

Overall, decarboxylation depends upon stability of the carbanion synthon R
,[1][2] although the anion may not be a true chemical intermediate.[3][4] Alkanoic acids and typical fatty acids decarboxylate slowly, but carboxylic acids with an α electron-withdrawing group (e.g. β‑keto acids, β‑nitriles, α‑nitro acids, or arylcarboxylic acids) decarboxylate easily. The Krapcho decarboxylation is the characteristic halide-mediated decarboxylation of such acids' methyl esters.

Those facile decarboxylations are an important organic synthesis technique, sometimes called the malonic or acetoacetic ester synthesis, after the reagent used. The Knoevenagel condensation and they allow keto acids serve as a stabilizing protecting group for carboxylic acid enols.[5][page needed][4]

For the free acids, conditions that deprotonate the carboxyl group (possibly protonating the electron-withdrawing group to form a zwitterionic tautomer) accelerate decarboxylation.[6] A strong base is key to ketonization, in which a pair of carboxylic acids combine to the eponymous functional group:[7][3]

Transition metal salts, especially copper compounds,[8] facilitate decarboxylation via carboxylate complex intermediates. Metals that catalyze cross-coupling reactions thus treat aryl carboxylates as an aryl anion synthon; this synthetic strategy is the decarboxylative cross-coupling reaction.[9]

Heating in cyclohexanone decarboxylates amino acids. In the related Hammick reaction, uncatalyzed decarboxylation of a picolinic acid gives a stable carbene that attacks a carbonyl electrophile.

Oxidative decarboxylations are generally radical reactions. These include the Kolbe electrolysis and Hunsdiecker-Kochi reactions. The Barton decarboxylation is an unusual radical reductive decarboxylation.

In biochemistry[edit]

Decarboxylations are pervasive in biology. They are often classified according to the cofactors that catalyze the transformations.[10] Biotin-coupled processes effect the decarboxylation of malonyl-CoA to acetyl-CoA. Thiamine (T:) is the active component for decarboxylation of alpha-ketoacids, including pyruvate:

T: + RC(O)CO2H → T=C(OH)R + CO2
T=C(OH)R + R'COOH → T! : + RC(O)CH(OH)R'

Pyridoxal phosphate promotes decarboxylation of amino acids. Flavin-dependent decarboxylases are involved in transformations of cysteine. Iron-based hydroxylases operate by reductive activation of O2 using the decarboxylation of alpha-ketoglutarate as an electron donor. The decarboxylation can be depicted as such:

RC(O)CO2Fe O2 → RCO2Fe{IV}=O + CO2
RCO2Fe=O + R'H → RCO2Fe + R'OH

Decarboxylation of amino acids[edit]

Common biosynthetic oxidative decarboxylations of amino acids to amines are:

Other decarboxylation reactions from the citric acid cycle include:

Case studies[edit]

Tetrahydrocannabinolic acid. The decarboxylation of this compound by heat is essential for the psychoactive effect of smoked cannabis, and depends on conversion of the enol to a keto group when the alpha carbon is protonated.

Upon heating, Δ9-tetrahydrocannabinolic acid decarboxylates to give the psychoactive compound Δ9-Tetrahydrocannabinol.[11] When cannabis is heated in vacuum, the decarboxylation of tetrahydrocannabinolic acid (THCA) appears to follow first order kinetics. The log fraction of THCA present decreases steadily over time, and the rate of decrease varies according to temperature. At 10-degree increments from 100 to 140 °C, half of the THCA is consumed in 30, 11, 6, 3, and 2 minutes; hence the rate constant follows Arrhenius' law, ranging between 10−8 and 10−5 in a linear log-log relationship with inverse temperature. However, modelling of decarboxylation of salicylic acid with a water molecule had suggested an activation barrier of 150 kJ/mol for a single molecule in solvent, much too high for the observed rate. Therefore, it was concluded that this reaction, conducted in the solid phase in plant material with a high fraction of carboxylic acids, follows a pseudo first order kinetics in which a nearby carboxylic acid precipitates without affecting the observed rate constant. Two transition states corresponding to indirect and direct keto-enol routes are possible, with energies of 93 and 104 kJ/mol. Both intermediates involve protonation of the alpha carbon, disrupting one of the double bonds of the aromatic ring and permitting the beta-keto group (which takes the form of an enol in THCA and THC) to participate in decarboxylation.[12]

In beverages stored for long periods, very small amounts of benzene may form from benzoic acid by decarboxylation catalyzed by the presence of ascorbic acid.[13]

The addition of catalytic amounts of cyclohexenone has been reported to catalyze the decarboxylation of amino acids.[14] However, using such catalysts may also yield an amount of unwanted by-products.

References[edit]

  1. ^ March, Jerry (1985), Advanced Organic Chemistry: Reactions, Mechanisms, and Structure, 3rd edition, New York: Wiley, ISBN 9780471854722, OCLC 642506595
  2. ^ "Decarboxylation, Dr. Ian A. Hunt, Department of Chemistry, University of Calgary". Archived from the original on 2022-12-21. Retrieved 2008-12-07.
  3. ^ a b Renz, M (2005). "Ketonization of Carboxylic Acids by Decarboxylation: Mechanism and Scope" (PDF). Eur. J. Org. Chem. 2005 (6): 979–988. doi:10.1002/ejoc.200400546 – via The Vespiary.
  4. ^ a b "Malonic Ester Synthesis". Organic Chemistry Portal. Retrieved 2007-10-26.
  5. ^ Organic Synthesis: The disconnection approach, 2nd ed.
  6. ^ Jim Clark (2004). "The Decarboxylation of Carboxylic Acids and their Salts". Chemguide. Retrieved 2007-10-22.
  7. ^ Thorpe, J. F.; Kon, G. A. R. (1925). "Cyclopentanone". Org. Synth. 5: 37. doi:10.15227/orgsyn.005.0037.
  8. ^ Wiley, Richard H.; Smith, Newton R. (1953). "m-Nitrostyrene". Organic Syntheses. 33: 62. doi:10.15227/orgsyn.033.0062.
  9. ^ Weaver, J. D.; Recio, A.; Grenning, A. J.; Tunge, J. A. (2011). "Transition Metal-Catalyzed Decarboxylative Allylation and Benzylation Reactions". Chem. Rev. 111 (3): 1846–1913. doi:10.1021/cr1002744. PMC 3116714. PMID 21235271.
  10. ^ Li, T.; Huo, L.; Pulley, C.; Liu, A. (2012). "Decarboxylation mechanisms in biological system. Bioorganic Chemistry". Bioorganic Chemistry. 43: 2–14. doi:10.1016/j.bioorg.2012.03.001. PMID 22534166.
  11. ^ Perrotin-Brunel, Helene; Buijs, Wim; Spronsen, Jaap van; Roosmalen, Maaike J.E. van; Peters, Cor J.; Verpoorte, Rob; Witkamp, Geert-Jan (2011). "Decarboxylation of Δ9-tetrahydrocannabinol: Kinetics and molecular modeling". Journal of Molecular Structure. 987 (1–3): 67–73. Bibcode:2011JMoSt.987...67P. doi:10.1016/j.molstruc.2010.11.061.
  12. ^ Perrotin-Brunel, Helene; Buijs, Wim; Spronsen, Jaap van; Roosmalen, Maaike J.E. van; Peters, Cor J.; Verpoorte, Rob; Witkamp, Geert-Jan (February 2011). "Decarboxylation of Δ9-tetrahydrocannabinol: Kinetics and molecular modeling". Journal of Molecular Structure. 987 (1–3): 67–73. Bibcode:2011JMoSt.987...67P. doi:10.1016/j.molstruc.2010.11.061.
  13. ^ "Data on Benzene in Soft Drinks and Other Beverages". Archived from the original on 2008-03-26. Retrieved 2008-03-26.
  14. ^ Hashimoto, Mitsunori; Eda, Yutaka; Osanai, Yasutomo; Iwai, Toshiaki; Aoki, Seiichi (1986). "A Novel Decarboxylation of α-Amino Acides. A Facile Method of Decarboxylation by the Use of 2-Cyclohexen-1-one as a Catalyst". Chemistry Letters. 15 (6): 893–896. doi:10.1246/cl.1986.893.

Leave a Reply