2-Pyridone

Last updated
2-Pyridone
2-Pyridone 2-pyridone.svg
2-Pyridone
2-pyridone (lactam).svg
2-Pyridone-(lactam)-3D-balls.png
2-Pyridone-(lactim)-3D-balls.png
Names
Preferred IUPAC name
Pyridin-2(1H)-one
Other names
2(1H)-Pyridinone
2(1H)-Pyridone
1H-Pyridine-2-one
2-Pyridone
1,2-Dihydro-2-oxopyridine
1H-2-Pyridone
2-Oxopyridone
2-Pyridinol
2-Hydroxypyridine
Identifiers
3D model (JSmol)
ChEBI
ChEMBL
ChemSpider
ECHA InfoCard 100.005.019 OOjs UI icon edit-ltr-progressive.svg
EC Number
  • 205-520-3
KEGG
PubChem CID
RTECS number
  • UV1144050
UNII
  • InChI=1S/C5H5NO/c7-5-3-1-2-4-6-5/h1-4H,(H,6,7) Yes check.svgY
    Key: UBQKCCHYAOITMY-UHFFFAOYSA-N Yes check.svgY
  • InChI=1/C5H5NO/c7-5-2-1-3-6-4-5/h1-4,7H
    Key: GRFNBEZIAWKNCO-UHFFFAOYAT
  • InChI=1/C5H5NO/c7-5-3-1-2-4-6-5/h1-4H,(H,6,7)
    Key: UBQKCCHYAOITMY-UHFFFAOYAK
  • lactim:Oc1ccccn1
  • lactam:C1=CC=CNC(=O)1
Properties
C5H5NO
Molar mass 95.101 g·mol−1
AppearanceColourless crystalline solid
Density 1.39 g/cm3
Melting point 107.8 °C (226.0 °F; 380.9 K)
Boiling point 280 °C (536 °F; 553 K) decomp.
Solubility in other solventsSoluble in water,
methanol, acetone
Acidity (pKa)11.65
UV-vismax)293 nm (ε 5900, H2O soln)
Structure
Orthorhombic
planar
4.26 D
Hazards
Occupational safety and health (OHS/OSH):
Main hazards
irritating
GHS labelling:
GHS-pictogram-skull.svg GHS-pictogram-exclam.svg
Danger
H301, H315, H319, H335
P261, P264, P270, P271, P280, P301+P310, P302+P352, P304+P340, P305+P351+P338, P312, P321, P330, P332+P313, P337+P313, P362, P403+P233, P405, P501
NFPA 704 (fire diamond)
NFPA 704.svgHealth 2: Intense or continued but not chronic exposure could cause temporary incapacitation or possible residual injury. E.g. chloroformFlammability 1: Must be pre-heated before ignition can occur. Flash point over 93 °C (200 °F). E.g. canola oilInstability (yellow): no hazard codeSpecial hazards (white): no code
2
1
Flash point 210 °C (410 °F; 483 K)
Related compounds
Other anions
2-Pyridinolate
Other cations
2-Hydroxypyridinium-ion
alcohol, lactam, lactim,
pyridine, ketone
Related compounds
pyridine, thymine, cytosine,
uracil, benzene
Supplementary data page
2-Pyridone (data page)
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).
Yes check.svgY  verify  (what is  Yes check.svgYX mark.svgN ?)

2-Pyridone is an organic compound with the formula C
5
H
4
NH(O)
. It is a colourless solid. It is well known to form hydrogen bonded dimers and it is also a classic case of a compound that exists as tautomers.

Contents

Tautomerism

tautomerism 2-pyridone-chemical-tautomer.svg
tautomerism

The second tautomer is 2-hydroxypyridine. This lactam lactim tautomerism can also be exhibited in many related compounds. [1]

Tautomerism in the solid state

The amide group can be involved in hydrogen bonding to other nitrogen- and oxygen-containing species.

The predominant solid state form is 2-pyridone. This has been confirmed by X-ray crystallography which shows that the hydrogen in solid state is closer to the nitrogen than to the oxygen (because of the low electron density at the hydrogen the exact positioning is difficult), and IR-spectroscopy, which shows that the C=O longitudinal frequency is present whilst the O-H frequencies are absent. [2] [3] [4] [5]

Tautomerism in solution

The tautomerization has been exhaustively studied. The energy difference appears to be very small. Non-polar solvents favour 2-hydroxypyridine whereas polar solvents such as alcohols and water favour the 2-pyridone. [1] [6] [7]

The energy difference for the two tautomers in the gas phase was measured by IR-spectroscopy to be 2.43 to 3.3 kJ/mol for the solid state and 8.95 kJ/mol and 8.83 kJ/mol for the liquid state. [8] [9] [10]

Tautomerisation mechanism A

The single molecular tautomerisation has a forbidden 1-3 suprafacial transition state and therefore has a high energy barrier for this tautomerisation, which was calculated with theoretical methods to be 125 or 210 kJ/mol. The direct tautomerisation is energetically not favoured. There are other possible mechanisms for this tautomerisation. [10]

Dimerisation

2-pyridone dimer.svg

2-Pyridone and 2-hydroxypyridine can form dimers with two hydrogen bonds. [11]

Aggregation in the solid state

In the solid state the dimeric form is not present; the 2-pyridones form a helical structure over hydrogen bonds. Some substituted 2-pyridones form the dimer in solid state, for example the 5-methyl-3-carbonitrile-2-pyridone. The determination of all these structures was done by X-ray crystallography. In the solid state the hydrogen is located closer to the nitrogen so it could be considered to be right to call the colourless crystals in the flask 2-pyridone. [1] [2] [3] [4] [5]

Aggregation in solution

In solution the dimeric form is present; the ratio of dimerisation is strongly dependent on the polarity of the solvent. Polar and protic solvents interact with the hydrogen bonds and more monomer is formed. Hydrophobic effects in non-polar solvents lead to a predominance of the dimer. The ratio of the tautomeric forms is also dependent on the solvent. All possible tautomers and dimers can be present and form an equilibrium, and the exact measurement of all the equilibrium constants in the system is extremely difficult. [11] [12] [13] [14] [15] [16] [17] [18] [19] [20]

(NMR-spectroscopy is a slow method, high resolution IR-spectroscopy in solvent is difficult, the broad absorption in UV-spectroscopy makes it hard to discriminate 3 and more very similar molecules).

Some publications only focus one of the two possible patterns, and neglect the influence of the other. For example, to calculation of the energy difference of the two tautomers in a non-polar solution will lead to a wrong result if a large quantity of the substance is on the side of the dimer in an equilibrium.

Tautomerisation mechanism B

The direct tautomerisation is not energetically favoured, but a dimerisation followed by a double proton transfer and dissociation of the dimer is a self catalytic path from one tautomer to the other. Protic solvents also mediate the proton transfer during the tautomerisation.

Synthesis

2-Pyrone can be obtained by a cyclisation reaction, and converted to 2-pyridone via an exchange reaction with ammonia:

2-pyridone chemical synthesis.svg

Pyridine forms an N-oxide with some oxidation agents such as hydrogen peroxide. This pyridine-N-oxide undergoes a rearrangement reaction to 2-pyridone in acetic anhydride: [21] [22] [23]

2-pyridone-chemical-synthesis.svg

In the Guareschi-Thorpe condensation cyanoacetamide reacts with a 1,3-diketone to a 2-pyridone. [12] [13] The reaction is named after Icilio Guareschi and Jocelyn Field Thorpe. [14] [15]

Chemical properties

Catalytic activity

2-Pyridone catalyses a variety of proton-dependent reactions, for example the aminolysis of esters. In some cases, molten 2-pyridone is used as a solvent. 2-Pyridone has a large effect on the reaction from activated esters with amines in nonpolar solvent, which is attributed to its tautomerisation and utility as a ditopic receptor. Proton transfer from 2-pyridone and its tautomer have been investigated by isotope labeling, kinetics and quantum chemical methods. [16] [17] [24]

Coordination chemistry

2-Pyridone and some derivatives serve as ligands in coordination chemistry, usually as a 1,3-bridging ligand akin to carboxylate. [18]

In nature

2-Pyridone is not naturally occurring, but a derivative has been isolated as a cofactor in certain hydrogenases. [19]

Environmental behavior

2-Pyridone is rapidly degraded by microorganisms in the soil environment, with a half life less than one week. [20] Organisms capable of growth on 2-pyridone as a sole source of carbon, nitrogen, and energy have been isolated by a number of researchers. The most extensively studied 2-pyridone degrader is the gram positive bacterium Arthrobacter crystallopoietes , [25] a member of the phylum Actinomycetota which includes numerous related organisms that have been shown to degrade pyridine or one or more alkyl-, carboxyl-, or hydroxyl-substituted pyridines. 2-Pyridone degradation is commonly initiated by mono-oxygenase attack, resulting in a diol, such as 2,5-dihydroxypyridine, which is metabolized via the maleamate pathway. Fission of the ring proceeds via action of 2,5-dihydroxypyridine monooxygenase, which is also involved in metabolism of nicotinic acid via the maleamate pathway. In the case of Arthrobacter crystallopoietes, at least part of the degradative pathway is plasmid-borne. [26] Pyridine diols undergo chemical transformation in solution to form intensely colored pigments. Similar pigments have been observed in quinoline degradation, [27] also owing to transformation of metabolites, however the yellow pigments often reported in degradation of many pyridine solvents, such as unsubstituted pyridine or picoline, generally result from overproduction of riboflavin in the presence of these solvents. [28] Generally speaking, degradation of pyridones, dihydroxypyridines, and pyridinecarboxylic acids is commonly mediated by oxygenases, whereas degradation of pyridine solvents often is not, and may in some cases involve an initial reductive step. [26]

See also

Related Research Articles

In chemistry, a zwitterion, also called an inner salt or dipolar ion, is a molecule that contains an equal number of positively and negatively charged functional groups. With amino acids, for example, in solution a chemical equilibrium will be established between the "parent" molecule and the zwitterion.

In chemistry, hydronium (hydroxonium in traditional British English) is the common name for the cation [H3O]+, also written as H3O+, the type of oxonium ion produced by protonation of water. It is often viewed as the positive ion present when an Arrhenius acid is dissolved in water, as Arrhenius acid molecules in solution give up a proton (a positive hydrogen ion, H+) to the surrounding water molecules (H2O). In fact, acids must be surrounded by more than a single water molecule in order to ionize, yielding aqueous H+ and conjugate base. Three main structures for the aqueous proton have garnered experimental support: the Eigen cation, which is a tetrahydrate, H3O+(H2O)3, the Zundel cation, which is a symmetric dihydrate, H+(H2O)2, and the Stoyanov cation, an expanded Zundel cation, which is a hexahydrate: H+(H2O)2(H2O)4. Spectroscopic evidence from well-defined IR spectra overwhelmingly supports the Stoyanov cation as the predominant form. For this reason, it has been suggested that wherever possible, the symbol H+(aq) should be used instead of the hydronium ion.

In chemistry, an acid dissociation constant is a quantitative measure of the strength of an acid in solution. It is the equilibrium constant for a chemical reaction

<span class="mw-page-title-main">Organolithium reagent</span> Chemical compounds containing C–Li bonds

In organometallic chemistry, organolithium reagents are chemical compounds that contain carbon–lithium (C–Li) bonds. These reagents are important in organic synthesis, and are frequently used to transfer the organic group or the lithium atom to the substrates in synthetic steps, through nucleophilic addition or simple deprotonation. Organolithium reagents are used in industry as an initiator for anionic polymerization, which leads to the production of various elastomers. They have also been applied in asymmetric synthesis in the pharmaceutical industry. Due to the large difference in electronegativity between the carbon atom and the lithium atom, the C−Li bond is highly ionic. Owing to the polar nature of the C−Li bond, organolithium reagents are good nucleophiles and strong bases. For laboratory organic synthesis, many organolithium reagents are commercially available in solution form. These reagents are highly reactive, and are sometimes pyrophoric.

<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.

<span class="mw-page-title-main">Tautomer</span> Structural isomers of chemical compounds that readily interconvert

Tautomers are structural isomers of chemical compounds that readily interconvert. The chemical reaction interconverting the two is called tautomerization. This conversion commonly results from the relocation of a hydrogen atom within the compound. The phenomenon of tautomerization is called tautomerism, also called desmotropism. Tautomerism is for example relevant to the behavior of amino acids and nucleic acids, two of the fundamental building blocks of life.

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

Acetylacetone is an organic compound with the chemical formula CH3−C(=O)−CH2−C(=O)−CH3. It is classified as a 1,3-diketone. It exists in equilibrium with a tautomer CH3−C(=O)−CH=C(−OH)−CH3. The mixture is a colorless liquid. These tautomers interconvert so rapidly under most conditions that they are treated as a single compound in most applications. Acetylacetone is a building block for the synthesis of many coordination complexes as well as heterocyclic compounds.

Tetrazoles are a class of synthetic organic heterocyclic compound, consisting of a 5-member ring of four nitrogen atoms and one carbon atom. The name tetrazole also refers to the parent compound with formula CH2N4, of which three isomers can be formulated.

<span class="mw-page-title-main">Isocyanic acid</span> Chemical compound (H–N=C=O)

Isocyanic acid is a chemical compound with the structural formula HNCO, which is often written as H−N=C=O. It is a colourless, volatile and poisonous substance, with a boiling point of 23.5 °C. It is the predominant tautomer and an isomer of cyanic acid (aka. cyanol).

<span class="mw-page-title-main">Crabtree's catalyst</span> Chemical compound

Crabtree's catalyst is an organoiridium compound with the formula [C8H12IrP(C6H11)3C5H5N]PF6. It is a homogeneous catalyst for hydrogenation and hydrogen-transfer reactions, developed by Robert H. Crabtree. This air stable orange solid is commercially available and known for its directed hydrogenation to give trans stereoselectivity with respective of directing group.

Bullvalene is a hydrocarbon with the chemical formula C10H10. The molecule has a cage-like structure formed by the fusion of one cyclopropane and three cyclohepta-1,4-diene rings. Bullvalene is unusual as an organic molecule due to the C−C and C=C bonds forming and breaking rapidly on the NMR timescale; this property makes it a fluxional molecule.

<span class="mw-page-title-main">Titanium tetrabromide</span> Chemical compound

Titanium tetrabromide is the chemical compound with the formula TiBr4. It is the most volatile transition metal bromide. The properties of TiBr4 are an average of TiCl4 and TiI4. Some key properties of these four-coordinated Ti(IV) species are their high Lewis acidity and their high solubility in nonpolar organic solvents. TiBr4 is diamagnetic, reflecting the d0 configuration of the metal centre.

The Hammick reaction, named after Dalziel Hammick, is a chemical reaction in which the thermal decarboxylation of α-picolinic acids in the presence of carbonyl compounds forms 2-pyridyl-carbinols.

The Meyer–Schuster rearrangement is the chemical reaction described as an acid-catalyzed rearrangement of secondary and tertiary propargyl alcohols to α,β-unsaturated ketones if the alkyne group is internal and α,β-unsaturated aldehydes if the alkyne group is terminal. Reviews have been published by Swaminathan and Narayan, Vartanyan and Banbanyan, and Engel and Dudley, the last of which describes ways to promote the Meyer–Schuster rearrangement over other reactions available to propargyl alcohols.

Deuterated chloroform, also known as chloroform-d, is the organic compound with the formula CDCl3. Deuterated chloroform is a common solvent used in NMR spectroscopy. The properties of CDCl3 and ordinary CHCl3 (chloroform) are virtually identical.

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

Hexafluoroacetylacetone is the chemical compound with the nominal formula CF3C(O)CH2C(O)CF3 (often abbreviated as hfacH). This colourless liquid is a ligand precursor and a reagent used in MOCVD. The compound exists exclusively as the enol CF3C(OH)=CHC(O)CF3. For comparison under the same conditions, acetylacetone is 85% enol.

<span class="mw-page-title-main">2-Aminopyridine</span> Chemical compound

2-Aminopyridine is an organic compound with the formula H2NC5H4N. It is one of three isomeric aminopyridines. It is a colourless solid that is used in the production of the drugs piroxicam, sulfapyridine, tenoxicam, and tripelennamine. It is produced by the reaction of sodium amide with pyridine, the Chichibabin reaction.

<span class="mw-page-title-main">2-Mercaptopyridine</span> Chemical compound

2-Mercaptopyridine is an organosulfur compound with the formula HSC5H4N. This yellow crystalline solid is a derivative of pyridine. The compound and its derivatives serve primarily as acylating agents. A few of 2-mercaptopyridine’s other uses include serving as a protecting group for amines and imides as well as forming a selective reducing agent. 2-Mercaptopyridine oxidizes to 2,2’-dipyridyl disulfide.

This page provides supplementary chemical data on 2-pyridone.

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

Pyrithione is the common name of an organosulfur compound with molecular formula C
5
H
5
NOS
, chosen as an abbreviation of pyridinethione, and found in the Persian shallot. It exists as a pair of tautomers, the major form being the thione 1-hydroxy-2(1H)-pyridinethione and the minor form being the thiol 2-mercaptopyridine N-oxide; it crystallises in the thione form. It is usually prepared from either 2-bromopyridine, 2-chloropyridine, or 2-chloropyridine N-oxide, and is commercially available as both the neutral compound and its sodium salt. It is used to prepare zinc pyrithione, which is used primarily to treat dandruff and seborrhoeic dermatitis in medicated shampoos, though is also an anti-fouling agent in paints.

References

  1. 1 2 3 Forlani L., Cristoni G., Boga C., Todesco P. E., Del Vecchio E., Selva S., Monari M. (2002). "Reinvestigation of tautomerism of some substituted 2-hydroxypyridines". Arkivoc . XI (11): 198–215. doi: 10.3998/ark.5550190.0003.b18 . hdl: 2027/spo.5550190.0003.b18 .
  2. 1 2 Yang H. W., Craven B. M. (1998). "Charge Density of 2-Pyridone". Acta Crystallogr. B. 54 (6): 912–920. doi:10.1107/S0108768198006545. PMID   9880899. S2CID   9505447.
  3. 1 2 Penfold B. R. (1953). "The Electron Distribution in Crystalline Alpha Pyridone". Acta Crystallogr. 6 (7): 591–600. Bibcode:1953AcCry...6..591P. doi: 10.1107/S0365110X5300168X .
  4. 1 2 Ohms U., Guth H., Heller E., Dannöhl H., Schweig A. (1984). "Comparison of Observed and Calculated Electron-Density 2-Pyridone, C5H5NO, Crystal-Structure Refinements at 295K and 120K, Experimental and Theoretical Deformation Density Studies". Z. Kristallogr. 169: 185–200. doi:10.1524/zkri.1984.169.14.185. S2CID   97575334.
  5. 1 2 Almlöf J., Kvick A., Olovsson I. (1971). "Hydrogen Bond Studies Crystal Structure of Intermolecular Complex 2-Pyridone-6-Chloro-2-Hdroxypyridine". Acta Crystallogr. B. 27 (6): 1201–1208. doi:10.1107/S0567740871003753.
  6. Aue DH, Betowski LD, Davidson WR, Bower MT, Beak P (1979). "Gas-Phase Basicities of Amides and Imidates - Estimation of Protomeric Equilibrium-Constantes by the Basicity methode in the Gas-Phase". Journal of the American Chemical Society . 101 (6): 1361–1368. doi:10.1021/ja00500a001.
  7. Frank J., Alan R. Katritzky (1976). "Tautomeric pyridines. XV. Pyridone-hydroxypyridine equilibria in solvents of different polarity". J Chem Soc Perkin Trans 2 (12): 1428–1431. doi:10.1039/p29760001428.
  8. Brown R. S., Tse A., Vederas J. C. (1980). "Photoelectro-Determined Core Binding Energies and Predicted Gas-Phase Basicities for the 2-Hydroxypyridine 2-Pyridone System". Journal of the American Chemical Society . 102 (3): 1174–1176. doi:10.1021/ja00523a050.
  9. Beak P. (1977). "Energies and Alkylation of Tautomeric Heterocyclic-Compounds - Old Problems New Answers". Acc. Chem. Res. 10 (5): 186–192. doi:10.1021/ar50113a006.
  10. 1 2 Abdulla H. I., El-Bermani M. F. (2001). "Infrared studies of tautomerism in 2-hydroxypyridine 2-thiopyridine and 2-aminopyridine". Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy. 57 (13): 2659–2671. Bibcode:2001AcSpA..57.2659A. doi:10.1016/S1386-1425(01)00455-3. PMID   11765793.
  11. 1 2 Hammes GG, Lillford PJ (1970). "A Kinetic and Equilibrium Study of Hydrogen Bond Dimerization of 2-Pyridone in Hydrogen Bonding Solvent". J. Am. Chem. Soc. 92 (26): 7578–7585. doi:10.1021/ja00729a012.
  12. 1 2 Gilchrist, T.L. (1997). Heterocyclic Chemistry ISBN   0-470-20481-8
  13. 1 2 Rybakov V. R., Bush A. A., Babaev E. B., Aslanov L. A. (2004). "3-Cyano-4,6-dimethyl-2-pyridone (Guareschi Pyridone)". Acta Crystallogr E. 6 (2): o160–o161. Bibcode:2004AcCrE..60O.160R. doi:10.1107/S1600536803029295.
  14. 1 2 I. Guareschi (1896). "Mem. Reale Accad. Sci. Torino II".{{cite journal}}: Cite journal requires |journal= (help)
  15. 1 2 Baron, H., Remfry, F. G. P., Thorpe, J. F. (1904). "CLXXV.-The formation and reactions of imino-compounds. Part I. Condensation of ethyl cyanoacetate with its sodium derivative". J. Chem. Soc., Trans. 85: 1726–1761. doi:10.1039/ct9048501726. Archived from the original on 2020-09-14. Retrieved 2020-06-05.
  16. 1 2 Fischer C. B., Steininger H., Stephenson D. S., Zipse H. (2005). "Catalysis of Aminolysis of 4-Nitrophenyl Acetate by 2-Pyridone". Journal of Physical Organic Chemistry. 18 (9): 901–907. doi:10.1002/poc.914.
  17. 1 2 L.-H. Wang, H. Zipse (1996). "Bifunctional Catalysis of Ester Aminolysis - A Computational and Experimental Study". Liebigs Ann. 1996 (10): 1501–1509. doi:10.1002/jlac.199619961003. Archived from the original on 2021-09-01. Retrieved 2021-09-01.
  18. 1 2 Rawson J. M., Winpenny R. E. P. (1995). "The coordination chemistry of 2-pyridones and its derivatives". Coordination Chemistry Reviews. 139 (139): 313–374. doi:10.1016/0010-8545(94)01117-T.
  19. 1 2 Shima, S.; Lyon, E. J.; Sordel-Klippert, M.; Kauss, M.; Kahnt, J.; Thauer, R. K.; Steinbach, K.; Xie, X.; Verdier, L. and Griesinger, C., "Structure elucidation: The cofactor of the iron-sulfur cluster free hydrogenase Hmd: structure of the light-inactivation product", Angew. Chem. Int. Ed., 2004, 43, 2547-2551.
  20. 1 2 Sims, Gerald K., S (1985). "Degradation of Pyridine Derivatives in Soil". Journal of Environmental Quality. 14 (4): 580–584. Bibcode:1985JEnvQ..14..580S. doi:10.2134/jeq1985.00472425001400040022x. Archived from the original on 2008-08-30.
  21. "Pyridin-N-oxydと酸無水物との反應" [Reaction between Pyridin-N-oxyd and acid anhydride]. Yakugaku Zasshi (in Japanese). 67 (3–4): 51–52. 1947. doi: 10.1248/yakushi1947.67.3-4_51 .
  22. Ochiai E (1953). "Recent Japanese Work on the Chemistry of Pyridine 1-Oxide and Related Compounds". The Journal of Organic Chemistry. 18 (5): 534–551. doi:10.1021/jo01133a010.
  23. Boekelheide V, Lehn WL (1961). "The Rearrangement of Substituted Pyridine N-Oxides with Acetic Anhydride1.2". The Journal of Organic Chemistry. 26 (2): 428–430. doi:10.1021/jo01061a037.
  24. Fischer C. B., Polborn K., Steininger H., Zipse H. (2004). "Synthesis and Solid-State Structures of Alkyl-Substituted 3-Cyano-2-pyridones" (PDF). Zeitschrift für Naturforschung. 59 (59b): 1121–1131. doi:10.1515/znb-2004-1008. S2CID   98273691. Archived from the original (subscription required) on 2008-10-30. Retrieved 2006-11-07.
  25. Ensign JC, Rittenberg SC (1963). "A crystalline pigment produced from 2-hydroxypyridine by arthrobacter crystallopoietes n.sp". Archiv für Mikrobiologie. 47 (2): 137–153. doi:10.1007/BF00422519. PMID   14106078. S2CID   6389661.
  26. 1 2 Sims GK, O'Loughlin E, Crawford R (1989). "Degradation of pyridines in the environment" (PDF). CRC Critical Reviews in Environmental Control. 19 (4): 309–340. Bibcode:1989CRvEC..19..309S. doi:10.1080/10643388909388372. Archived from the original (PDF) on 2010-05-27.
  27. Oloughlin E, Kehrmeyer S, Sims G (1996). "Isolation, characterization, and substrate utilization of a quinoline-degrading bacterium". International Biodeterioration & Biodegradation. 38 (2): 107–118. doi:10.1016/S0964-8305(96)00032-7.
  28. Sims, Gerald K., O (1992). "Riboflavin Production during Growth of Micrococcus luteus on Pyridine". Applied and Environmental Microbiology. 58 (10): 3423–3425. Bibcode:1992ApEnM..58.3423S. doi:10.1128/AEM.58.10.3423-3425.1992. PMC   183117 . PMID   16348793.

Further reading

General

  1. Engdahl K., Ahlberg P. (1977). Journal of Chemical Research: 340–341.{{cite journal}}: Missing or empty |title= (help)
  2. Bensaude O, Chevrier M, Dubois J (1978). "Lactim-Lactam Tautomeric Equilibrium of 2-Hydroxypyridines. 1.Cation Binding, Dimerization and Interconversion Mechanism in Aprotic Solvents. A Spectroscopic and Temperature-Jump Kinetic Study". J. Am. Chem. Soc. 100 (22): 7055–7066. doi:10.1021/ja00490a046.
  3. Bensaude O, Dreyfus G, Dodin G, Dubois J (1977). "Intramolecular Nondissociative Proton Transfer in Aqueous Solutions of Tautomeric Heterocycles: a Temperature-Jump Kinetic Study". J. Am. Chem. Soc. 99 (13): 4438–4446. doi:10.1021/ja00455a037.
  4. Bensaude O, Chevrier M, Dubois J (1978). "Influence of Hydration upon Tautomeric Equilibrium". Tetrahedron Lett. 19 (25): 2221–2224. doi:10.1016/S0040-4039(01)86850-7.
  5. Hammes GG, Park AC (1969). "Kinetic and Thermodynamic Studies of Hydrogen Bonding". J. Am. Chem. Soc. 91 (4): 956–961. doi:10.1021/ja01032a028.
  6. Hammes GG, Spivey HO (1966). "A Kinetic Study of the Hydrogen-Bond Dimerization of 2-Pyridone". J. Am. Chem. Soc. 88 (8): 1621–1625. doi:10.1021/ja00960a006. PMID   5942979.
  7. Beak P, Covington JB, Smith SG (1976). "Structural Studies of Tautomeric Systems: the Importance of Association for 2-Hydroxypyridine-2-Pyridone and 2-Mercaptopyridine-2-Thiopyridone". J. Am. Chem. Soc. 98 (25): 8284–8286. doi:10.1021/ja00441a079.
  8. Beak P, Covington JB, White JM (1980). "Quantitave Model of Solvent Effects on Hydroxypyridine-Pyridone and Mercaptopyridine-Thiopyridone Equilibria: Correlation with Reaction-Field and Hydrogen-Bond Effects". J. Org. Chem. 45 (8): 1347–1353. doi:10.1021/jo01296a001.
  9. Beak P, Covington JB, Smith SG, White JM, Zeigler JM (1980). "Displacement of Protomeric Equilibria by Self-Association: Hydroxypyridine-Pyridone and Mercaptopyridine-Thiopyridone Isomer Pairs". J. Org. Chem. 45 (8): 1354–1362. doi:10.1021/jo01296a002.

Tautomerism