Knorr pyrrole synthesis

Last updated

Contents

Knorr pyrrole synthesis
Named after Ludwig Knorr
Reaction type Ring forming reaction
Identifiers
RSC ontology ID RXNO:0000497

The Knorr pyrrole synthesis is a widely used chemical reaction that synthesizes substituted pyrroles (3). [1] [2] [3] The method involves the reaction of an α-amino-ketone (1) and a compound containing an electron-withdrawing group (e.g. an ester as shown) α to a carbonyl group (2). [4]

The Knorr pyrrole synthesis Knorr Pyrrole Synthesis Scheme.png
The Knorr pyrrole synthesis

Method

The mechanism requires zinc and acetic acid as catalysts. It will proceed at room temperature. Because α-aminoketones self-condense very easily, they must be prepared in situ. The usual way of doing this is from the relevant oxime, via the Neber rearrangement. [5] [6]

The original Knorr synthesis employed two equivalents of ethyl acetoacetate, one of which was converted to ethyl 2-oximinoacetoacetate by dissolving it in glacial acetic acid, and slowly adding one equivalent of saturated aqueous sodium nitrite, under external cooling. Zinc dust was then stirred in, reducing the oxime group to the amine. This reduction consumes two equivalents of zinc and four equivalents of acetic acid.

Knorr 1886 synthesis Original Knorr synthesis.png
Knorr 1886 synthesis

Modern practice is to add the oxime solution resulting from the nitrosation and the zinc dust gradually to a well-stirred solution of ethyl acetoacetate in glacial acetic acid. The reaction is exothermic, and the mixture can reach the boiling point, if external cooling is not applied. The resulting product, diethyl 3,5-dimethylpyrrole-2,4-dicarboxylate, has been called Knorr's Pyrrole ever since. In the Scheme above, R2 = COOEt, and R1 = R3 = Me represent this original reaction.

Knorr's pyrrole can be derivatized in a number of useful manners. One equivalent of sodium hydroxide will saponify the 2-ester selectively. Dissolving Knorr's pyrrole in concentrated sulfuric acid, and then pouring the resulting solution into water will hydrolyze the 4-ester group selectively. The 5-methyl group can be variously oxidized to chloromethyl, aldehyde, or carboxylic acid functionality by the use of stoichiometric sulfuryl chloride in glacial acetic acid. [7] Alternatively, the nitrogen atom can be alkylated. The two ester positions can be more smoothly differentiated by incorporating benzyl or tert-butyl groups via the corresponding acetoacetate esters. Benzyl groups can be removed by catalytic hydrogenolysis over palladium on carbon, and tertiary-butyl groups can be removed by treatment with trifluoroacetic acid, or boiling aqueous acetic acid. R1 and R3 (as well as R2 and "Et") can be varied by the application of appropriate β-ketoesters readily made by a synthesis emanating from acid chlorides, Meldrum's acid, and the alcohol of one's choice. Ethyl and benzyl esters are easily made thereby, and the reaction is noteworthy in that even the highly hindered tert-butyl alcohol gives very high yields in this synthesis. [8]

Levi and Zanetti extended the Knorr synthesis in 1894 to the use of acetylacetone (2,4-pentanedione) in reaction with ethyl 2-oximinoacetoacetate. The result was ethyl 4-acetyl-3,5-dimethylpyrrole-2-carboxylate, where "OEt" = R1 = R3 = Me, and R2 = COOEt. [9] The 4-acetyl group could easily be converted to a 4-ethyl group by Wolff-Kishner reduction (hydrazine and alkali, heated); hydrogenolysis, or the use of diborane. Benzyl or tert-butyl acetoacetates also work well in this system, and with close temperature control, the tert-butyl system gives a very high yield (close to 80%). [10] N,N-dialkyl pyrrole-2- and/or 4-carboxamides may be prepared by the use of N,N-dialkyl acetoacetamides in the synthesis. Even thioesters have been successfully prepared, using the method. [11] As for the nitrosation of β-ketoesters, despite the numerous literature specifications of tight temperature control on the nitrosation, the reaction behaves almost like a titration, and the mixture can be allowed to reach even 40 °C without significantly impacting the final yield.

The mechanism of the Knorr pyrrole synthesis begins with condensation of the amine and ketone to give an imine. The imine then tautomerizes to an enamine, followed by cyclization, elimination of water, and isomerization to the pyrrole.

Mechanism of Knorr pyrrole synthesis Knorr pyrrole mechanism.jpg
Mechanism of Knorr pyrrole synthesis

There are a number of important syntheses of pyrroles that are operated in the manner of the Knorr Synthesis, despite having mechanisms of very different connectivity between the starting materials and the pyrrolic product.

Hans Fischer and Emmy Fink found that Zanetti's synthesis from 2,4-pentanedione and ethyl 2-oximinoacetoacetate gave ethyl 3,5-dimethylpyrrole-2-carboxylate as a trace byproduct. Similarly, 3-ketobutyraldehyde diethyl acetal led to the formation of ethyl 5-methylpyrrole-2-carboxylate. Both of these products resulted from the loss of the acetyl group from the inferred ethyl 2-aminoacetoacetate intermediate. An important product of the Fischer-Fink synthesis was ethyl 4,5-dimethylpyrrole-2-carboxylate, made from ethyl 2-oximinoacetoacetate and 2-methyl-3-oxobutanal, in turn made by the Claisen condensation of 2-butanone with ethyl formate. [12]

George Kleinspehn reported that the Fischer–Fink connectivity could be forced to occur exclusively, by the use of diethyl oximinomalonate in the synthesis, with 2,4-pentanedione, or its 3-alkyl substituted derivatives. Yields were high, around 60%, and this synthesis eventually came to be one of the most important in the repertory. [13] Yields were significantly improved, by the use of preformed diethyl aminomalonate (prepared by the hydrogenolysis of diethyl oximinomalonate in ethanol, over Pd/C), and adding a mixture of diethyl aminomalonate and the β-diketone to actively boiling glacial acetic acid. [14]

Meanwhile, Johnson had extended the Fischer-Fink synthesis by reacting 2-oximinoacetoacetate esters (ethyl, benzyl, or tertiary-butyl), with 3-alkyl substituted 2,4-pentanediones. [15] The Kleinspehn synthesis was extended under David Dolphin by the use of unsymmetrical β-diketones (such as 3-alkyl substituted 2,4-hexanediones), which preferentially reacted initially at the less hindered acetyl group and afforded the corresponding 5-methylpyrrole-2-carboxylate esters. N,N-Dialkyl 2-oximinoacetoacetamides also were found to give pyrroles when reacted under Knorr conditions with 3-substituted-2,4-pentanediones, in yields comparable to the corresponding esters (around 45%). However, when unsymmetrical diketones were used, it was found that the acetyl group from the acetoacetamide was retained in the product, and one of the acyl groups from the diketone had been lost. [16] This same mechanism occurs to a minor extent in the acetoacetate ester systems, and had previously been detected radiochemically by Harbuck and Rapoport. [17] Most of the above-described syntheses have application in the synthesis of porphyrins, bile pigments, and dipyrrins.

Related Research Articles

<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.

Pyrrole is a heterocyclic, aromatic, organic compound, a five-membered ring with the formula C4H4NH. It is a colorless volatile liquid that darkens readily upon exposure to air. Substituted derivatives are also called pyrroles, e.g., N-methylpyrrole, C4H4NCH3. Porphobilinogen, a trisubstituted pyrrole, is the biosynthetic precursor to many natural products such as heme.

<span class="mw-page-title-main">Dicarbonyl</span> Molecule containing two adjacent C=O groups

In organic chemistry, a dicarbonyl is a molecule containing two carbonyl groups. Although this term could refer to any organic compound containing two carbonyl groups, it is used more specifically to describe molecules in which both carbonyls are in close enough proximity that their reactivity is changed, such as 1,2-, 1,3-, and 1,4-dicarbonyls. Their properties often differ from those of monocarbonyls, and so they are usually considered functional groups of their own. These compounds can have symmetrical or unsymmetrical substituents on each carbonyl, and may also be functionally symmetrical or unsymmetrical.

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

Diethyl malonate, also known as DEM, is the diethyl ester of malonic acid. It occurs naturally in grapes and strawberries as a colourless liquid with an apple-like odour, and is used in perfumes. It is also used to synthesize other compounds such as barbiturates, artificial flavourings, vitamin B1, and vitamin B6.

<span class="mw-page-title-main">Ethyl acetate</span> Organic compound (CH₃CO₂CH₂CH₃)

Ethyl acetate is the organic compound with the formula CH3CO2CH2CH3, simplified to C4H8O2. This colorless liquid has a characteristic sweet smell and is used in glues, nail polish removers, and in the decaffeination process of tea and coffee. Ethyl acetate is the ester of ethanol and acetic acid; it is manufactured on a large scale for use as a solvent.

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

Diisopropyl ether is a secondary ether that is used as a solvent. It is a colorless liquid that is slightly soluble in water, but miscible with organic solvents. It is used as an extractant and an oxygenate gasoline additive. It is obtained industrially as a byproduct in the production of isopropanol by hydration of propylene. Diisopropyl ether is sometimes represented by the abbreviation DIPE.

In organic chemistry, the Knoevenagel condensation reaction is a type of chemical reaction named after German chemist Emil Knoevenagel. It is a modification of the aldol condensation.

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

The organic compound ethyl acetoacetate (EAA) is the ethyl ester of acetoacetic acid. It is a colorless liquid. It is widely used as a chemical intermediate in the production of a wide variety of compounds. It is used as a flavoring for food.

Di-<i>tert</i>-butyl dicarbonate Chemical compound

Di-tert-butyl dicarbonate is a reagent widely used in organic synthesis. Since this compound can be regarded formally as the acid anhydride derived from a tert-butoxycarbonyl (Boc) group, it is commonly referred to as Boc anhydride. This pyrocarbonate reacts with amines to give N-tert-butoxycarbonyl or so-called Boc derivatives. These carbamate derivatives do not behave as amines, which allows certain subsequent transformations to occur that would be incompatible with the amine functional group. The Boc group can later be removed from the amine using moderately strong acids. Thus, Boc serves as a protective group, for instance in solid phase peptide synthesis. Boc-protected amines are unreactive to most bases and nucleophiles, allowing for the use of the fluorenylmethyloxycarbonyl group (Fmoc) as an orthogonal protecting group.

<span class="mw-page-title-main">Alkyl nitrite</span> Organic compounds of the form R–O–N=O

In organic chemistry, alkyl nitrites are a group of organic compounds based upon the molecular structure R−O−N=O, where R represents an alkyl group. Formally they are alkyl esters of nitrous acid. They are distinct from nitro compounds.

<span class="mw-page-title-main">2,6-Lutidine</span> Chemical compound

2,6-Lutidine is a natural heterocyclic aromatic organic compound with the formula (CH3)2C5H3N. It is one of several dimethyl-substituted derivative of pyridine, all of which are referred to as lutidines. It is a colorless liquid with mildly basic properties and a pungent, noxious odor.

The Schotten–Baumann reaction is a method to synthesize amides from amines and acid chlorides:

<span class="mw-page-title-main">Mukaiyama Taxol total synthesis</span>

The Mukaiyama taxol total synthesis published by the group of Teruaki Mukaiyama of the Tokyo University of Science between 1997 and 1999 was the 6th successful taxol total synthesis. The total synthesis of Taxol is considered a hallmark in organic synthesis.

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

Hagemann's ester, ethyl 2-methyl-4-oxo-2-cyclohexenecarboxylate, is an organic compound that was first prepared and described in 1893 by German chemist Carl Hagemann. The compound is used in organic chemistry as a reagent in the synthesis of many natural products including sterols, trisporic acids, and terpenoids.

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

Diethyl chlorophosphate is an organophosphorus compound with the formula (C2H5O)2P(O)Cl. As a reagent in organic synthesis, it is used to convert alcohols to the corresponding diethylphosphate esters. It is a colorless liquid with a fruity odor. It is a corrosive, and as a cholinesterase inhibitor, highly toxic through dermal absorption. The molecule is tetrahedral.

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

Ethyl cyanoacetate is an organic compound that contains a carboxylate ester and a nitrile. It is a colourless liquid with a pleasant odor. This material is useful as a starting material for synthesis due to its variety of functional groups and chemical reactivity.

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

Diethyl phosphite is the organophosphorus compound with the formula (C2H5O)2P(O)H. It is a popular reagent for generating other organophosphorus compounds, exploiting the high reactivity of the P-H bond. Diethyl phosphite is a colorless liquid. The molecule is tetrahedral.

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

Diethyl oxomalonate is the diethyl ester of mesoxalic acid (ketomalonic acid), the simplest oxodicarboxylic acid and thus the first member (n = 0) of a homologous series HOOC–CO–(CH2)n–COOH with the higher homologues oxalacetic acid (n = 1), α-ketoglutaric acid (n = 2) and α-ketoadipic acid (n = 3) (the latter a metabolite of the amino acid lysine). Diethyl oxomalonate reacts because of its highly polarized keto group as electrophile in addition reactions and is a highly active reactant in pericyclic reactions such as the Diels-Alder reactions, cycloadditions or ene reactions. At humid air, mesoxalic acid diethyl ester reacts with water to give diethyl mesoxalate hydrate and the green-yellow oil are spontaneously converted to white crystals.

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

2-Acetylbutyrolactone (ABL) is a derivative of γ-butyrolactone that is used as a precursor in organic synthesis, and it is used to identify primary amines through chemical fluorescence.

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

Diethyl acetamidomalonate (DEAM) is a derivative of malonic acid diethyl ester. Formally, it is derived through the acetylation of ester from the unstable aminomalonic acid. DEAM serves as a starting material for racemates including both, natural and unnatural α-amino acids or hydroxycarboxylic acids. It is also usable as a precursor in pharmaceutical formulations, particularly in the cases of active ingredients like fingolimod, which is used to treat multiple sclerosis.

References

  1. Knorr, Ludwig (1884). "Synthese von Pyrrolderivaten" [Synthesis of pyrrole derivatives](PDF). Berichte der deutschen chemischen Gesellschaft (in German). 17 (2): 1635–1642. doi:10.1002/cber.18840170220.
  2. Knorr, Ludwig (1886). "Synthetische Versuche mit dem Acetessigester" [Synthetic experiments with the ester of acetoacetic acid]. Annalen der Chemie (in German). 236 (3): 290–332. doi:10.1002/jlac.18862360303.
  3. Knorr, L.; Lange, H. (1902). "Ueber die Bildung von Pyrrolderivaten aus Isonitrosoketonen" [On the formation of pyrrole derivatives from isonitrosoketones](PDF). Berichte der deutschen chemischen Gesellschaft (in German). 35 (3): 2998–3008. doi:10.1002/cber.19020350392.
  4. Corwin, Alsoph Henry (1950). "The Chemistry of Pyrrole and its Derivatives". In Elderfield, Robert Cooley (ed.). Heterocyclic Compounds. Vol. 1. New York: Wiley. pp. 287 ff.
  5. Fischer, Hans (1935). "2,4-Dimethyl-3,5-dicarbethoxypyrrole (2,4-Pyrroledicarboxylic acid, 3,5-dimethyl-, diethyl ester)". Organic Syntheses . 15: 17. doi:10.15227/orgsyn.015.0017 ; Collected Volumes, vol. 2, p. 202.
  6. Fischer, Hans (1941). "Kryptopyrrole (Pyrrole, 2,4-dimethyl-3-ethyl)". Organic Syntheses . 21: 67. doi:10.15227/orgsyn.021.0067 ; Collected Volumes, vol. 3, p. 513.
  7. Corwin, Alsoph H.; Bailey, William A.; Viohl, Paul (1942). "Structural Investigations upon a Substituted Dipyrrylmethane. An Unusual Melting Point-Symmetry Relationship". Journal of the American Chemical Society . 64 (6): 1267–1273. doi:10.1021/ja01258a007.
  8. Oikawa, Yuji; Sugano, Kiyoshi; Yonemitsu, Osamu (1978). "Meldrum's acid in organic synthesis. 2. A general and versatile synthesis of β-keto esters". The Journal of Organic Chemistry . 43 (10): 2087–2088. doi:10.1021/jo00404a066.
  9. Zanetti, C. U.; Levi, E. (1894). "Sintesi di composti pirrolici dai nitrosochetoni" [Synthesis of pyrrole compounds from nitrosoketones]. La Gazzetta Chimica Italiana (in Italian). 24 (1): 546–554.
  10. Treibs, Alfred; Hintermeier, Karl (1954). "tert-Butylester von Pyrrolcarbonsäuren". Chemische Berichte (in German). 87 (8): 1167–1174. doi:10.1002/cber.19540870818.
  11. Bullock, E.; Chen, T. S.; Loader, C. E. (1966). "Preparation and reactions of some pyrrylthiol esters". Canadian Journal of Chemistry . 44 (9): 1007–1111. doi: 10.1139/v66-149 .
  12. Fischer, Hans; Fink, Emmy (1948). "Über eine neue Pyrrolsynthese" [On a new synthesis of pyrroles]. Zeitschrift für Physiologische Chemie (in German). 283 (3–4): 152–161. doi:10.1515/bchm2.1948.283.3-4.152.
  13. Kleinspehn, George G. (1955). "A Novel Route to Certain 2-Pyrrolecarboxylic Esters and Nitriles". Journal of the American Chemical Society . 77 (6): 1546–1548. doi:10.1021/ja01611a043.
  14. Paine, John B.; Dolphin, David (1985). "Pyrrole chemistry. An improved synthesis of ethyl pyrrole-2-carboxylate esters from diethyl aminomalonate". The Journal of Organic Chemistry . 50 (26): 5598–5604. doi:10.1021/jo00350a033.
  15. Bullock, E.; Johnson, A. W.; Markham, E.; Shaw, K. B. (1958). "287. A synthesis of coproporphyrin III". Journal of the Chemical Society (Resumed) : 1430–1440. doi:10.1039/JR9580001430.
  16. Paine, John B.; Brough, Jonathan R.; Buller, Kathy K.; Erikson, Erika E.; Dolphin, D. (1987). "Mechanism of the formation of N,N-dialkyl-2-pyrrolecarboxamides from 1,3-diketones and N,N-dialkyloximinoacetoacetamides". The Journal of Organic Chemistry . 52 (18): 3993–3997. doi:10.1021/jo00227a010.
  17. Rapoport, Henry; Harbuct, John W. (1971). "Mechanism of a modified Knorr pyrrole condensation". The Journal of Organic Chemistry . 36 (6): 853–855. doi:10.1021/jo00805a030.

See also