You are on page 1of 1

Search Wikipedia Search Create account Log in

Carboxylic acid 78 languages

Contents hide Article Talk Read Edit View history Tools

(Top) From Wikipedia, the free encyclopedia

Examples and nomenclature (Redirected from Carboxyl)

Physical properties
"COOH" redirects here. For the Bulgarian musician, see Ivan Shopov.
Solubility Not to be confused with Carbolic acid.
Boiling points In organic chemistry, a carboxylic acid is an organic acid that contains a carboxyl group (−C(=O)−OH)[1] attached to an
Acidity 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 the alkyl, alkenyl, aryl, or other group. Carboxylic acids occur widely. Important examples include the amino
Odour
acids and fatty acids. Deprotonation of a carboxylic acid gives a carboxylate anion.
Characterization

Occurrence and applications Examples and nomenclature [ edit ]


Structure of a
Synthesis
Carboxylic acids are commonly identified by their trivial names. They often have the suffix -ic acid. IUPAC-recommended carboxylic acid
Industrial routes names also exist; in this system, carboxylic acids have an -oic acid suffix.[2] For example, butyric acid (CH3CH2CH2CO2H)
Laboratory methods is butanoic acid by IUPAC guidelines. For nomenclature of complex molecules containing a carboxylic acid, the carboxyl can
be considered position one of the parent chain even if there are other substituents, such as 3-chloropropanoic acid.
Less-common reactions
Alternately, it can be named as a "carboxy" or "carboxylic acid" substituent on another parent structure, such as 2-
Reactions carboxyfuran.
Reduction − −
The carboxylate anion (R−COO or R−CO2) of a carboxylic acid is usually named with the suffix -ate, in keeping with the
Specialized reactions general pattern of -ic acid and -ate for a conjugate acid and its conjugate base, respectively. For example, the conjugate
Carboxyl radical base of acetic acid is acetate. Carboxylate anion

See also
Carbonic acid, which occurs in bicarbonate buffer systems in nature, is not generally classed as one of the carboxylic acids,
References despite that it has a moiety that looks like a COOH group.

External links
Straight-chain, saturated carboxylic acids (alkanoic acids)
Carbon Common Chemical
IUPAC Name Common location or use
atoms Name formula

1 Formic acid Methanoic acid HCOOH Insect stings

2 Acetic acid Ethanoic acid CH3COOH Vinegar 3D structure of


a carboxylic acid
Preservative for stored grains, body
3 Propionic acid Propanoic acid CH3CH2COOH
odour, milk, butter, cheese
Wikiquote has quotations
4 Butyric acid Butanoic acid CH3(CH2)2COOH Butter related to Carboxylic acid.
5 Valeric acid Pentanoic acid CH3(CH2)3COOH Valerian plant
6 Caproic acid Hexanoic acid CH3(CH2)4COOH Goat fat

7 Enanthic acid Heptanoic acid CH3(CH2)5COOH Fragrance

8 Caprylic acid Octanoic acid CH3(CH2)6COOH Coconuts

9 Pelargonic acid Nonanoic acid CH3(CH2)7COOH Pelargonium plant

10 Capric acid Decanoic acid CH3(CH2)8COOH Coconut and Palm kernel oil

11 Undecylic acid Undecanoic acid CH3(CH2)9COOH Anti-fungal agent

12 Lauric acid Dodecanoic acid CH3(CH2)10COOH Coconut oil and hand wash soaps

13 Tridecylic acid Tridecanoic acid CH3(CH2)11COOH Plant metabolite


Tetradecanoic
14 Myristic acid CH3(CH2)12COOH Nutmeg
acid

Pentadecylic Pentadecanoic
15 CH3(CH2)13COOH Milk fat
acid acid
Hexadecanoic
16 Palmitic acid CH3(CH2)14COOH Palm oil
acid
Heptadecanoic
17 Margaric acid CH3(CH2)15COOH Pheromone in various animals
acid
Octadecanoic
18 Stearic acid CH3(CH2)16COOH Chocolate, waxes, soaps, and oils
acid
Nonadecylic Nonadecanoic
19 CH3(CH2)17COOH Fats, vegetable oils, pheromone
acid acid

20 Arachidic acid Icosanoic acid CH3(CH2)18COOH Peanut oil

Other carboxylic acids


Compound class Members
unsaturated
acrylic acid (2-propenoic acid) – CH2=CH−COOH, used in polymer synthesis
monocarboxylic acids
medium to long-chain saturated and unsaturated monocarboxylic acids, with even number of carbons; examples:
Fatty acids
docosahexaenoic acid and eicosapentaenoic acid (nutritional supplements)

Amino acids the building-blocks of proteins

Keto acids acids of biochemical significance that contain a ketone group; examples: acetoacetic acid and pyruvic acid
containing at least one aromatic ring; examples: benzoic acid – the sodium salt of benzoic acid is used as a food preservative;
Aromatic carboxylic
salicylic acid – a beta-hydroxy type found in many skin-care products; phenyl alkanoic acids – the class of compounds where
acids
a phenyl group is attached to a carboxylic acid

containing two carboxyl groups; examples: adipic acid the monomer used to produce nylon and aldaric acid – a family of
Dicarboxylic acids
sugar acids

Tricarboxylic acids containing three carboxyl groups; examples: citric acid – found in citrus fruits and isocitric acid

containing a hydroxy group in the first position; examples: glyceric acid, glycolic acid and lactic acid (2-hydroxypropanoic acid)
Alpha hydroxy acids
– found in sour milk, tartaric acid – found in wine

Beta hydroxy acids containing a hydroxy group in the second position

Omega hydroxy acids containing a hydroxy group beyond the first or second position

Divinylether fatty acids containing a doubly unsaturated carbon chain attached via an ether bond to a fatty acid, found in some plants

Physical properties [ edit ]

Solubility [ edit ]

Carboxylic acids are polar. Because they are both hydrogen-bond acceptors (the carbonyl −C(=O)−) and hydrogen-bond donors (the hydroxyl −OH), they
also participate in hydrogen bonding. Together, the hydroxyl and carbonyl group form the functional group carboxyl. Carboxylic acids usually exist as
dimers in nonpolar media due to their tendency to "self-associate". Smaller carboxylic acids (1 to 5 carbons) are soluble in water, whereas bigger
carboxylic acids have limited solubility due to the increasing hydrophobic nature of the alkyl chain. These longer chain acids tend to be soluble in less-
polar solvents such as ethers and alcohols.[3] Aqueous sodium hydroxide and carboxylic acids, even hydrophobic ones, react to yield water-soluble
sodium salts. For example, enanthic acid has a low solubility in water (0.2 g/L), but its sodium salt is very soluble in water.

Boiling points [ edit ]

Carboxylic acids tend to have higher boiling points than water, because of their greater surface areas and their tendency to form stabilized dimers
through hydrogen bonds. For boiling to occur, either the dimer bonds must be broken or the entire dimer arrangement must be vaporized, increasing the
enthalpy of vaporization requirements significantly.

Carboxylic acid dimers

Acidity [ edit ]

Carboxylic acids are Brønsted–Lowry acids because they are proton (H+) donors. They are the most common type of organic acid.

+ −
Carboxylic acids are typically weak acids, meaning that they only partially dissociate into [H3O] cations and R−CO2 anions in neutral aqueous solution.
For example, at room temperature, in a 1-molar solution of acetic acid, only 0.001% of the acid are dissociated (i.e. 10−5 moles out of 1 mol). Electron-
withdrawing substituents, such as -CF3 group, give stronger acids (the pKa of acetic acid is 4.76 whereas trifluoroacetic acid, with a trifluoromethyl
substituent, has a pKa of 0.23). Electron-donating substituents give weaker acids (the pKa of formic acid is 3.75 whereas acetic acid, with a methyl
substituent, has a pKa of 4.76)

Carboxylic acid[4] pKa

Formic acid (HCO2H) 3.75

Chloroformic acid (ClCO2H) 0.27[5]

Acetic acid (CH3CO2H) 4.76

Glycine (NH2CH2CO2H) 2.34

Fluoroacetic acid (FCH2CO2H) 2.586

Difluoroacetic acid (F2CHCO2H) 1.33

Trifluoroacetic acid (CF3CO2H) 0.23

Chloroacetic acid (ClCH2CO2H) 2.86


Dichloroacetic acid (Cl2CHCO2H) 1.29

Trichloroacetic acid (CCl3CO2H) 0.65

Benzoic acid (C6H5−CO2H) 4.2

2-Nitrobenzoic acid (ortho-C6H4(NO2)CO2H) 2.16


Oxalic acid (HO−C(=O)−C(=O)−OH) (first dissociation) 1.27

Hydrogen oxalate (HO−C(=O)−CO2) (second dissociation of oxalic acid) 4.14

Deprotonation of carboxylic acids gives carboxylate anions; these are resonance stabilized, because the negative charge is delocalized over the two
oxygen atoms, increasing the stability of the anion. Each of the carbon–oxygen bonds in the carboxylate anion has a partial double-bond character. The
carbonyl carbon's partial positive charge is also weakened by the -1/2 negative charges on the 2 oxygen atoms.

Odour [ edit ]

Carboxylic acids often have strong sour odours. Esters of carboxylic acids tend to have fruity, pleasant odours, and many are used in perfume.

Characterization [ edit ]

Carboxylic acids are readily identified as such by infrared spectroscopy. They exhibit a sharp band associated with vibration of the C=O carbonyl bond
(νC=O) between 1680 and 1725 cm−1. A characteristic νO–H band appears as a broad peak in the 2500 to 3000 cm−1 region.[3][6] By 1H NMR
spectrometry, the hydroxyl hydrogen appears in the 10–13 ppm region, although it is often either broadened or not observed owing to exchange with
traces of water.

Occurrence and applications [ edit ]

Many carboxylic acids are produced industrially on a large scale. They are also frequently found in nature. Esters of fatty acids are the main components
of lipids and polyamides of aminocarboxylic acids are the main components of proteins.

Carboxylic acids are used in the production of polymers, pharmaceuticals, solvents, and food additives. Industrially important carboxylic acids include
acetic acid (component of vinegar, precursor to solvents and coatings), acrylic and methacrylic acids (precursors to polymers, adhesives), adipic acid
(polymers), citric acid (a flavor and preservative in food and beverages), ethylenediaminetetraacetic acid (chelating agent), fatty acids (coatings), maleic
acid (polymers), propionic acid (food preservative), terephthalic acid (polymers). Important carboxylate salts are soaps.

Synthesis [ edit ]

Industrial routes [ edit ]

In general, industrial routes to carboxylic acids differ from those used on a smaller scale because they require specialized equipment.

Carbonylation of alcohols as illustrated by the Cativa process for the production of acetic acid. Formic acid is prepared by a different carbonylation
pathway, also starting from methanol.
Oxidation of aldehydes with air using cobalt and manganese catalysts. The required aldehydes are readily obtained from alkenes by
hydroformylation.
Oxidation of hydrocarbons using air. For simple alkanes, this method is inexpensive but not selective enough to be useful. Allylic and benzylic
compounds undergo more selective oxidations. Alkyl groups on a benzene ring are oxidized to the carboxylic acid, regardless of its chain length.
Benzoic acid from toluene, terephthalic acid from para-xylene, and phthalic acid from ortho-xylene are illustrative large-scale conversions. Acrylic acid
is generated from propene.[7]
Oxidation of ethene using silicotungstic acid catalyst.
Base-catalyzed dehydrogenation of alcohols.
Carbonylation coupled to the addition of water. This method is effective and versatile for alkenes that generate secondary and tertiary carbocations,
e.g. isobutylene to pivalic acid. In the Koch reaction, the addition of water and carbon monoxide to alkenes or alkynes is catalyzed by strong acids.
Hydrocarboxylations involve the simultaneous addition of water and CO. Such reactions are sometimes called "Reppe chemistry."
HC≡CH + CO + H2O → CH2=CH−CO2H

Hydrolysis of triglycerides obtained from plant or animal oils. These methods of synthesizing some long-chain carboxylic acids are related to soap
making.
Fermentation of ethanol. This method is used in the production of vinegar.
The Kolbe–Schmitt reaction provides a route to salicylic acid, precursor to aspirin.

Laboratory methods [ edit ]

Preparative methods for small scale reactions for research or for production of fine chemicals often employ expensive consumable reagents.

Oxidation of primary alcohols or aldehydes with strong oxidants such as potassium dichromate, Jones reagent, potassium permanganate, or sodium
chlorite. The method is more suitable for laboratory conditions than the industrial use of air, which is "greener" because it yields less inorganic side
products such as chromium or manganese oxides.[citation needed]
Oxidative cleavage of olefins by ozonolysis, potassium permanganate, or potassium dichromate.
Hydrolysis of nitriles, esters, or amides, usually with acid- or base-catalysis.
Carbonation of a Grignard reagent and organolithium reagents:
− +
RLi + CO2 → RCO2Li
− +
RCO2Li + HCl → RCO2H + LiCl

Halogenation followed by hydrolysis of methyl ketones in the haloform reaction


Base-catalyzed cleavage of non-enolizable ketones, especially aryl ketones:[8]
R−C(=O)−Ar + H2O → R−CO2H + ArH

Less-common reactions [ edit ]

Many reactions produce carboxylic acids but are used only in specific cases or are mainly of academic interest.

Disproportionation of an aldehyde in the Cannizzaro reaction


Rearrangement of diketones in the benzilic acid rearrangement
Involving the generation of benzoic acids are the von Richter reaction from nitrobenzenes and the Kolbe–Schmitt reaction from phenols.

Reactions [ edit ]

The most widely practiced reactions convert carboxylic acids into esters, amides,
carboxylate salts, acid chlorides, and alcohols. Carboxylic acids react with bases to
form carboxylate salts, in which the hydrogen of the hydroxyl (–OH) group is
replaced with a metal cation. For example, acetic acid found in vinegar reacts with
sodium bicarbonate (baking soda) to form sodium acetate, carbon dioxide, and
water:
− +
CH3COOH + NaHCO3 → CH3COO Na + CO2 + H2O

Carboxylic acids also react with alcohols to give esters. This process is widely
used, e.g. in the production of polyesters. Likewise, carboxylic acids are converted
into amides, but this conversion typically does not occur by direct reaction of the
carboxylic acid and the amine. Instead esters are typical precursors to amides. The
conversion of amino acids into peptides is a significant biochemical process that Carboxylic acid organic reactions
requires ATP.

The hydroxyl group on carboxylic acids may be replaced with a chlorine atom using thionyl chloride to give acyl chlorides. In nature, carboxylic acids are
converted to thioesters.

Reduction [ edit ]

Like esters, most carboxylic acids can be reduced to alcohols by hydrogenation, or using hydride transferring agents such as lithium aluminium hydride.
Strong alkyl transferring agents, such as organolithium compounds but not Grignard reagents, will reduce carboxylic acids to ketones along with transfer
of the alkyl group.

Vilsmaier reagent (N,N-Dimethyl(chloromethylene)ammonium chloride [ClHC\dN+(CH3)2]Cl−) is a highly chemoselective agent for carboxylic acid
reduction. It selectively activates the carboxylic acid to give the carboxymethyleneammonium salt, which can be reduced by a mild reductant like lithium
tris(t-butoxy)aluminum hydride to afford an aldehyde in a one pot procedure. This procedure is known to tolerate reactive carbonyl functionalities such as
ketone as well as moderately reactive ester, olefin, nitrile, and halide moieties.[9]

Specialized reactions [ edit ]

As with all carbonyl compounds, the protons on the α-carbon are labile due to keto–enol tautomerization. Thus, the α-carbon is easily halogenated in
the Hell–Volhard–Zelinsky halogenation.
The Schmidt reaction converts carboxylic acids to amines.
Carboxylic acids are decarboxylated in the Hunsdiecker reaction.
The Dakin–West reaction converts an amino acid to the corresponding amino ketone.
In the Barbier–Wieland degradation, a carboxylic acid on an aliphatic chain having a simple methylene bridge at the alpha position can have the
chain shortened by one carbon. The inverse procedure is the Arndt–Eistert synthesis, where an acid is converted into acyl halide, which is then
reacted with diazomethane to give one additional methylene in the aliphatic chain.
Many acids undergo oxidative decarboxylation. Enzymes that catalyze these reactions are known as carboxylases (EC 6.4.1) and decarboxylases
(EC 4.1.1).
Carboxylic acids are reduced to aldehydes via the ester and DIBAL, via the acid chloride in the Rosenmund reduction and via the thioester in the
Fukuyama reduction.
In ketonic decarboxylation carboxylic acids are converted to ketones.
Organolithium reagents (>2 equiv) react with carboxylic acids to give a dilithium 1,1-diolate, a stable tetrahedral intermediate which decomposes to
give a ketone upon acidic workup.
The Kolbe electrolysis is an electrolytic, decarboxylative dimerization reaction. It gets rid of the carboxyl groups of two acid molecules, and joins the
remaining fragments together.

Carboxyl radical [ edit ]

The carboxyl radical, •COOH, only exists briefly.[10] The acid dissociation constant of •COOH has been measured using electron paramagnetic
resonance spectroscopy.[11] The carboxyl group tends to dimerise to form oxalic acid.

See also [ edit ]

Acid anhydride
Acid chloride
Amide
Amino acid
Ester
List of carboxylic acids
Dicarboxylic acid
Pseudoacid
Thiocarboxy
Carbon dioxide (CO2)

References [ edit ]

1. ^ IUPAC, Compendium of Chemical Terminology, 2nd ed. (the "Gold Book") 8. ^ Perry C. Reeves (1977). "Carboxylation of Aromatic Compounds:
(1997). Online corrected version: (2006–) "carboxylic acids ". Ferrocenecarboxylic Acid". Org. Synth. 56: 28.
doi:10.1351/goldbook.C00852 doi:10.15227/orgsyn.056.0028 .
2. ^ Recommendations 1979 . Organic Chemistry IUPAC Nomenclature. 9. ^ Fujisawa, Tamotsu; Sato, Toshio. "Reduction of carboxylic acids to
Rules C-4 Carboxylic Acids and Their Derivatives. aldehydes: 6-Ooxdecanal" . Organic Syntheses. 66: 121.
3. ^ a b Morrison, R.T.; Boyd, R.N. (1992). Organic Chemistry (6th ed.). doi:10.15227/orgsyn.066.0121 .; Collective Volume, vol. 8, p. 498
Prentice Hall. ISBN 0-13-643669-2. 10. ^ Milligan, D. E.; Jacox, M. E. (1971). "Infrared Spectrum and Structure of
4. ^ Haynes, William M., ed. (2011). CRC Handbook of Chemistry and Physics Intermediates in Reaction of OH with CO". Journal of Chemical Physics. 54
(92nd ed.). CRC Press. pp. 5–94 to 5–98. ISBN 978-1439855119. (3): 927–942. Bibcode:1971JChPh..54..927M . doi:10.1063/1.1675022 .
5. ^ "Chlorocarbonic acid" . Human Metabolome Database. 11. ^ The value is pKa = −0.2 ± 0.1. Jeevarajan, A. S.; Carmichael, I.;
6. ^ Smith, Brian. "The C=O Bond, Part VIII: Review" . Spectroscopy. Fessenden, R. W. (1990). "ESR Measurement of the pKa of Carboxyl
Retrieved 12 February 2024. Radical and Ab Initio Calculation of the Carbon-13 Hyperfine Constant".
7. ^ Riemenschneider, Wilhelm (2002). "Carboxylic Acids, Aliphatic". Ullmann's Journal of Physical Chemistry. 94 (4): 1372–1376.
Encyclopedia of Industrial Chemistry. Weinheim: Wiley-VCH. doi:10.1021/j100367a033 .
doi:10.1002/14356007.a05_235 . ISBN 3527306730..

External links [ edit ]

Carboxylic acids pH and titration – freeware for calculations, data analysis, simulation, and distribution Look up carboxyl in
diagram generation Wiktionary, the free
dictionary.
PHC.

· · Functional groups [show]

· · Topics in organic reactions [show]

Authority control databases: National Germany · Israel · United States · Japan · Czech Republic

Categories: Carboxylic acids Functional groups

This page was last edited on 12 February 2024, at 23:07 (UTC).

Text is available under the Creative Commons Attribution-ShareAlike License 4.0; additional terms may apply. By using this site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of the Wikimedia Foundation, Inc., a non-profit
organization.

Privacy policy About Wikipedia Disclaimers Contact Wikipedia Code of Conduct Developers Statistics Cookie statement Mobile view

You might also like