Decarboxylation

Last updated
Decarboxylation Decarboxylation reaction.png
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).

Contents

In organic chemistry

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] Typically, carboxylic acids decarboxylate slowly, but carboxylic acids with an α electron-withdrawing group (e.g. β keto acids, βnitriles, α nitro acids, or arylcarboxylic acids) decarboxylate easily. Decarboxylation of sodium chlorodifluoroacetate generates difluorocarbene:

CF2ClCO2Na → NaCl + CF2 + CO2 [5]

Decarboxylations are an important in the malonic and acetoacetic ester synthesis. The Knoevenagel condensation and they allow keto acids serve as a stabilizing protecting group for carboxylic acid enols. [6] [ 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. [7] A strong base is key to ketonization, in which a pair of carboxylic acids combine to the eponymous functional group: [8] [3]

Barium adipate pyrolysis.png

Transition metal salts, especially copper compounds, [9] 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. [10]

Upon heating in cyclohexanone amino acids decarboxylate. 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.

As described above, most decarboxylations start with a carboxylic acid or its alkali metal salt, but the Krapcho decarboxylation starts with methyl esters. I this case, the reaction begins with halide-mediated cleavage of the ester, forming the carboxylate.

In biochemistry

Decarboxylations are pervasive in biology. They are often classified according to the cofactors that catalyze the transformations. [11] 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

Common biosynthetic oxidative decarboxylations of amino acids to amines are:

Other decarboxylation reactions from the citric acid cycle include:

Fatty acid synthesis

Synthesis of saturated fatty acids via fatty acid synthase II in E. coli Saturated Fatty Acid Synthesis.svg
Synthesis of saturated fatty acids via fatty acid synthase II in E. coli

Straight-chain fatty acid synthesis occurs by recurring reactions involving decarboxylation of malonyl-CoA. [12]

Case studies

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. Tetrahydrocannabinolicacid.svg
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. [13] 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. [14]

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. [15]

The addition of catalytic amounts of cyclohexenone has been reported to catalyze the decarboxylation of amino acids. [16] However, using such catalysts may also yield an amount of unwanted by-products.

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 an organyl group, or hydrogen, or other groups. 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">Citric acid cycle</span> Interconnected biochemical reactions releasing energy

The citric acid cycle—also known as the Krebs cycle, Szent–Györgyi–Krebs cycle or the TCA cycle (tricarboxylic acid cycle)—is a series of biochemical reactions to release the energy stored in nutrients through the oxidation of acetyl-CoA derived from carbohydrates, fats, and proteins. The chemical energy released is available under the form of ATP. The Krebs cycle is used by organisms that respire (as opposed to organisms that ferment) to generate energy, either by anaerobic respiration or aerobic respiration. In addition, the cycle provides precursors of certain amino acids, as well as the reducing agent NADH, that are used in numerous other reactions. Its central importance to many biochemical pathways suggests that it was one of the earliest components of metabolism. Even though it is branded as a "cycle", it is not necessary for metabolites to follow only one specific route; at least three alternative segments of the citric acid cycle have been recognized.

<span class="mw-page-title-main">Ester</span> Compound derived from an acid

In chemistry, an ester is a compound derived from an acid in which the hydrogen atom (H) of at least one acidic hydroxyl group of that acid is replaced by an organyl group. Analogues derived from oxygen replaced by other chalcogens belong to the ester category as well. According to some authors, organyl derivatives of acidic hydrogen of other acids are esters as well, but not according to the IUPAC.

<span class="mw-page-title-main">Ketone</span> Organic compounds of the form >C=O

In organic chemistry, a ketone is an organic compound with the structure R−C(=O)−R', where R and R' can be a variety of carbon-containing substituents. Ketones contain a carbonyl group −C(=O)−. The simplest ketone is acetone, with the formula (CH3)2CO. Many ketones are of great importance in biology and in industry. Examples include many sugars (ketoses), many steroids, and the solvent acetone.

<span class="mw-page-title-main">Aldehyde</span> Organic compound containing the functional group R−CH=O

In organic chemistry, an aldehyde is an organic compound containing a functional group with the structure R−CH=O. The functional group itself can be referred to as an aldehyde but can also be classified as a formyl group. Aldehydes are a common motif in many chemicals important in technology and biology.

<span class="mw-page-title-main">Alanine</span> Α-amino acid that is used in the biosynthesis of proteins

Alanine (symbol Ala or A), or α-alanine, is an α-amino acid that is used in the biosynthesis of proteins. It contains an amine group and a carboxylic acid group, both attached to the central carbon atom which also carries a methyl group side chain. Consequently it is classified as a nonpolar, aliphatic α-amino acid. Under biological conditions, it exists in its zwitterionic form with its amine group protonated (as −NH3+) and its carboxyl group deprotonated (as −CO2). It is non-essential to humans as it can be synthesized metabolically and does not need to be present in the diet. It is encoded by all codons starting with GC (GCU, GCC, GCA, and GCG).

<span class="mw-page-title-main">Glutamic acid</span> Amino acid and neurotransmitter

Glutamic acid is an α-amino acid that is used by almost all living beings in the biosynthesis of proteins. It is a non-essential nutrient for humans, meaning that the human body can synthesize enough for its use. It is also the most abundant excitatory neurotransmitter in the vertebrate nervous system. It serves as the precursor for the synthesis of the inhibitory gamma-aminobutyric acid (GABA) in GABAergic neurons.

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

Acetoacetic acid is the organic compound with the formula CH3COCH2COOH. It is the simplest beta-keto acid, and like other members of this class, it is unstable. The methyl and ethyl esters, which are quite stable, are produced on a large scale industrially as precursors to dyes. Acetoacetic acid is a weak acid.

In molecular biology, protein catabolism is the breakdown of proteins into smaller peptides and ultimately into amino acids. Protein catabolism is a key function of digestion process. Protein catabolism often begins with pepsin, which converts proteins into polypeptides. These polypeptides are then further degraded. In humans, the pancreatic proteases include trypsin, chymotrypsin, and other enzymes. In the intestine, the small peptides are broken down into amino acids that can be absorbed into the bloodstream. These absorbed amino acids can then undergo amino acid catabolism, where they are utilized as an energy source or as precursors to new proteins.

<span class="mw-page-title-main">Oxaloacetic acid</span> Organic compound

Oxaloacetic acid (also known as oxalacetic acid or OAA) is a crystalline organic compound with the chemical formula HO2CC(O)CH2CO2H. Oxaloacetic acid, in the form of its conjugate base oxaloacetate, is a metabolic intermediate in many processes that occur in animals. It takes part in gluconeogenesis, the urea cycle, the glyoxylate cycle, amino acid synthesis, fatty acid synthesis and the citric acid cycle.

<span class="mw-page-title-main">Enol</span> Organic compound with a C=C–OH group

In organic chemistry, alkenols are a type of reactive structure or intermediate in organic chemistry that is represented as an alkene (olefin) with a hydroxyl group attached to one end of the alkene double bond. The terms enol and alkenol are portmanteaus deriving from "-ene"/"alkene" and the "-ol" suffix indicating the hydroxyl group of alcohols, dropping the terminal "-e" of the first term. Generation of enols often involves deprotonation at the α position to the carbonyl group—i.e., removal of the hydrogen atom there as a proton H+. When this proton is not returned at the end of the stepwise process, the result is an anion termed an enolate. The enolate structures shown are schematic; a more modern representation considers the molecular orbitals that are formed and occupied by electrons in the enolate. Similarly, generation of the enol often is accompanied by "trapping" or masking of the hydroxy group as an ether, such as a silyl enol ether.

The Carroll rearrangement is a rearrangement reaction in organic chemistry and involves the transformation of a β-keto allyl ester into a α-allyl-β-ketocarboxylic acid. This organic reaction is accompanied by decarboxylation and the final product is a γ,δ-allylketone. The Carroll rearrangement is an adaptation of the Claisen rearrangement and effectively a decarboxylative allylation.

The branched-chain α-ketoacid dehydrogenase complex is a multi-subunit complex of enzymes that is found on the mitochondrial inner membrane. This enzyme complex catalyzes the oxidative decarboxylation of branched, short-chain alpha-ketoacids. BCKDC is a member of the mitochondrial α-ketoacid dehydrogenase complex family, which also includes pyruvate dehydrogenase and alpha-ketoglutarate dehydrogenase, key enzymes that function in the Krebs cycle.

The Hunsdiecker reaction is a name reaction in organic chemistry whereby silver salts of carboxylic acids react with a halogen to produce an organic halide. It is an example of both a decarboxylation and a halogenation reaction as the product has one fewer carbon atoms than the starting material and a halogen atom is introduced its place. A catalytic approach has been developed.

Oxidative decarboxylation is a decarboxylation reaction caused by oxidation. Most are accompanied by α- Ketoglutarate α- Decarboxylation caused by dehydrogenation of hydroxyl carboxylic acids such as carbonyl carboxylic acid, malic acid, isocitric acid, etc.

In biochemistry, fatty acid synthesis is the creation of fatty acids from acetyl-CoA and NADPH through the action of enzymes called fatty acid synthases. This process takes place in the cytoplasm of the cell. Most of the acetyl-CoA which is converted into fatty acids is derived from carbohydrates via the glycolytic pathway. The glycolytic pathway also provides the glycerol with which three fatty acids can combine to form triglycerides, the final product of the lipogenic process. When only two fatty acids combine with glycerol and the third alcohol group is phosphorylated with a group such as phosphatidylcholine, a phospholipid is formed. Phospholipids form the bulk of the lipid bilayers that make up cell membranes and surrounds the organelles within the cells. In addition to cytosolic fatty acid synthesis, there is also mitochondrial fatty acid synthesis (mtFASII), in which malonyl-CoA is formed from malonic acid with the help of malonyl-CoA synthetase (ACSF3), which then becomes the final product octanoyl-ACP (C8) via further intermediate steps.

Krapcho decarboxylation is a chemical reaction used to manipulate certain organic esters. This reaction applies to esters with a beta electron-withdrawing group (EWG).

<span class="mw-page-title-main">Gould–Jacobs reaction</span> Gould-Jacobs reaction explained

The Gould–Jacobs reaction is an organic synthesis for the preparation of quinolines and 4‐hydroxyquinoline derivatives. The Gould–Jacobs reaction is a series of reactions. The series of reactions begins with the condensation/substitution of an aniline with alkoxy methylenemalonic ester or acyl malonic ester, producing anilidomethylenemalonic ester. Then through a 6 electron cyclization process, 4-hydroxy-3-carboalkoxyquinoline is formed, which exist mostly in the 4-oxo form. Saponification results in the formation of an acid. This step is followed by decarboxylation to give 4-hydroxyquinoline. The Gould–Jacobs reaction is effective for anilines with electron‐donating groups at the meta‐position.

<span class="mw-page-title-main">Keto acid</span> Organic compounds with a –COOH group and a C=O group

In organic chemistry, keto acids or ketoacids are organic compounds that contain a carboxylic acid group and a ketone group. In several cases, the keto group is hydrated. The alpha-keto acids are especially important in biology as they are involved in the Krebs citric acid cycle and in glycolysis.

<span class="mw-page-title-main">Tetrahydrocannabinolic acid synthase</span> Enzyme

Tetrahydrocannabinolic acid (THCA) synthase is an enzyme responsible for catalyzing the formation of THCA from cannabigerolic acid (CBGA). THCA is the direct precursor of tetrahydrocannabinol (THC), the principal psychoactive component of cannabis, which is produced from various strains of Cannabis sativa. Therefore, THCA synthase is considered to be a key enzyme controlling cannabis psychoactivity. Polymorphisms of THCA synthase result in varying levels of THC in Cannabis plants, resulting in "drug-type" and "fiber-type" C. sativa varieties.

References

  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. 1 2 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. 1 2 "Malonic Ester Synthesis". Organic Chemistry Portal. Retrieved 2007-10-26.
  5. Taschner, Michael J. (2001). "Sodium Chlorodifluoroacetate". Encyclopedia of Reagents for Organic Synthesis. doi:10.1002/047084289X.rs058. ISBN   0-471-93623-5.
  6. Organic Synthesis: The disconnection approach, 2nd ed.
  7. Jim Clark (2004). "The Decarboxylation of Carboxylic Acids and their Salts". Chemguide. Retrieved 2007-10-22.
  8. Thorpe, J. F.; Kon, G. A. R. (1925). "Cyclopentanone". Org. Synth. 5: 37. doi:10.15227/orgsyn.005.0037.
  9. Wiley, Richard H.; Smith, Newton R. (1953). "m-Nitrostyrene". Organic Syntheses. 33: 62. doi:10.15227/orgsyn.033.0062.
  10. 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.
  11. 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.
  12. "Fatty Acids: Straight-chain Saturated, Structure, Occurrence and Biosynthesis". lipidlibrary.aocs.org. Lipid Library, The American Oil Chemists' Society. 30 April 2011. Archived from the original on 21 July 2011.
  13. 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.
  14. 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.
  15. "Data on Benzene in Soft Drinks and Other Beverages". Archived from the original on 2008-03-26. Retrieved 2008-03-26.
  16. 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.