Quinine total synthesis

Last updated
quinine carbon atom numbering scheme left and asymmetric centers right QuinineOverview.png
quinine carbon atom numbering scheme left and asymmetric centers right

The total synthesis of quinine , a naturally-occurring antimalarial drug, was developed over a 150-year period. The development of synthetic quinine is considered a milestone in organic chemistry although it has never been produced industrially as a substitute for natural occurring quinine. The subject has also been attended with some controversy: Gilbert Stork published the first stereoselective total synthesis of quinine in 2001, meanwhile shedding doubt on the earlier claim by Robert Burns Woodward and William Doering in 1944, claiming that the final steps required to convert their last synthetic intermediate, quinotoxine, into quinine would not have worked had Woodward and Doering attempted to perform the experiment. A 2001 editorial published in Chemical & Engineering News sided with Stork, but the controversy was eventually laid to rest once and for all when Williams and coworkers successfully repeated Woodward's proposed conversion of quinotoxine to quinine in 2007. [1]

Contents

Chemical structure

The aromatic component of the quinine molecule is a quinoline with a methoxy substituent. The amine component has a quinuclidine skeleton and the methylene bridge in between the two components has a hydroxyl group. The substituent at the 3 position is a vinyl group. The molecule is optically active with five stereogenic centers (the N1 and C4 constituting a single asymmetric unit), making synthesis potentially difficult because it is one of 16 stereoisomers.

Quinine total synthesis timeline

QuininePasteur.png
QuininePerkinsAttempt.png
QuinineRabeKindler.png
The first step in this sequence is sodium hypobromite addition to quinotoxine to an N-bromo intermediate possibly with structure 2. The second step is organic oxidation with sodium ethoxide in ethanol. Because of the basic conditions the initial product quininone interconverts with quinidinone via a common enol intermediate and mutarotation is observed. In the third step the ketone group is reduced with aluminum powder and sodium ethoxide in ethanol and quinine can be identified. Quinotoxine is the first relay molecule in the Woodward/Doering claim.
QuinineRabeKindlerIIRev.png
HomomeroquineneSynthesis.png
The key step in the assembly of quinotoxine is a Claisen condensation:
QuinineClaisen.png
QuinineWoodwardDoering.svg
Woodward and Doering argue that Rabe in 1918 already proved that this compound will eventually give quinine but do not repeat Rabe's work. In this project 27-year-old assistant professor Woodward is the theorist and postdoc Doering (age 26) the bench worker. According to William, Bob was able to boil water but an egg would be a challenge.[ citation needed ] As many natural quinine resources were tied up in the enemy-held Dutch East Indies, synthetic quinine was a promising alternative for fighting malaria on the battlefield and both men become instant war heroes making headlines in the New York Times , Newsweek and Life .
QuinineLactoneSynthesis.svg
The starting materials are trans-2-butene-1,4-diol and ethyl orthoacetate and the key step is a Claisen rearrangement
LactoneChiralResolution.png
In this process the racemic lactone reacts in aminolysis with (S)-methylbenzylamine assisted by triethylaluminum to a diastereomeric pair of amides which can be separated by column chromatography. The S-enantiomer is converted back to the S-lactone in two steps by hydrolysis with potassium hydroxide and ethylene glycol followed by azeotropic ring closure.
QuinineStorkSummary.png
RabesQuinineRevisited.svg

Stork quinine total synthesis

The Stork quinine synthesis starts from chiral (S)-4-vinylbutyrolactone 1. The compound is obtained by chiral resolution and in fact, in the subsequent steps all stereogenic centers are put in place by chiral induction: the sequence does not contain asymmetric steps.

QuinineStork0.png QuinineStorkSynthStart.png
Stork quinine synthesisIntroducing C8 and nitrogen

The lactone is ring-opened with diethylamine to amide 2 and its hydroxyl group is protected as a tert-butyldimethyl silyl ether (TBS) in 3. The C5 and C6 atoms are added as tert-butyldiphenylsilyl (TBDPS) protected iodoethanol in a nucleophilic substitution of acidic C4 with lithium diisopropylamide (LDA) at −78 °C to 4 with correct stereochemistry. Removal of the silyl protecting group with p-toluenesulfonic acid to alcohol 4b and ring-closure by azeotropic distillation returns the compound to lactone 5 (direct alkylation of 1 met with undisclosed problems).

The lactone is then reduced to the lactol 5b with diisobutylaluminum hydride and its liberated aldehyde reacts in a Wittig reaction with methoxymethylenetriphenylphosphine (delivering the C8 atom) to form enol ether 6. The hydroxyl group is replaced in a Mitsunobu reaction by an azide group with diphenylphosphoryl azide in 7 and acid hydrolysis yields the azido aldehyde 8.

QuinineStorkSynthPart3.png QuinineStorkSynthFinal.png
First ring closureSecond ring closure

The methyl group in 6-methoxy-4-methylquinoline9 is sufficiently acidic for nucleophilic addition of its anion (by reaction with LDA) to the aldehyde group in 8 to form 10 as a mixture of epimers. This is of no consequence for stereocontrol because in the next step the alcohol is oxidized in a Swern oxidation to ketone 11. A Staudinger reaction with triphenylphosphine closes the ring between the ketone and the azide to the tetrahydropyridine 12. The imine group in this compound is reduced to the amine 13 with sodium borohydride with the correct stereospecificity. The silyl protecting group is removed with hydrogen fluoride to alcohol 14 and then activated as a mesyl leaving group by reaction with mesyl chloride in pyridine which enables the third ring closure to 15. In the final step the C9 hydroxyl group was introduced by oxidation with sodium hydride, dimethylsulfoxide and oxygen with quinine to epiquinine ratio of 14:1.

Woodward–Doering formal quinine total synthesis

The 1944 Woodward–Doering synthesis starts from 7-hydroxyisoquinoline 3 for the quinuclidine skeleton which is somewhat counter intuitive because one goes from a stable heterocyclic aromatic system to a completely saturated bicyclic ring. This compound (already known since 1895) is prepared in two steps.

QuinineWoodwardDoeringPartI.png QuinineWoodwardDoeringPartII.png
Woodward/Doering quinine synthesis part IPart II

The first reaction step is condensation reaction of 3-hydroxybenzaldehyde 1 with (formally) the diacetal of aminoacetaldehyde to the imine 2 and the second reaction step is cyclization in concentrated sulfuric acid. Isoquinoline 3 is then alkylated in another condensation by formaldehyde and piperidine and the product is isolated as the sodium salt of 4.

QuinineWoodwardDoeringPartIII.png
Woodward/Doering quinine synthesis part III

Hydrogenation at 220 °C for 10 hours in methanol with sodium methoxide liberates the piperidine group and leaving the methyl group in 5 with already all carbon and nitrogen atoms accounted for. A second hydrogenation takes place with Adams catalyst in acetic acid to tetrahydroisoquinoline6. Further hydrogenation does not take place until the amino group is acylated with acetic anhydride in methanol but by then 7 is again hydrogenated with Raney nickel in ethanol at 150 °C under high pressure to decahydroisoquinoline8. The mixture of cis and trans isomers is then oxidized by chromic acid in acetic acid to the ketone 9. Only the cis isomer crystallizes and used in the next reaction step, a ring opening with the alkyl nitrite ethyl nitrite with sodium ethoxide in ethanol to 10 with a newly formed carboxylic ester group and an oxime group. The oxime group is hydrogenated to the amine 11 with platinum in acetic acid and alkylation with iodomethane gives the quaternary ammonium salt 12 and subsequently the betaine 13 after reaction with silver oxide.

Quinine's vinyl group is then constructed by Hofmann elimination with sodium hydroxide in water at 140 °C. This process is accompanied by hydrolysis of both the ester and the amide group but it is not the free amine that is isolated but the urea 14 by reaction with potassium cyanate. In the next step the carboxylic acid group is esterified with ethanol and the urea group replaced with a benzoyl group. The final step is a Claisen condensation of 15 with ethyl quininate 16, which after acidic workup yields racemic quinotoxine 17. The desired enantiomer is obtained by chiral resolution with the chiral dibenzoyl ester of Tartaric acid. The conversion of this compound to quinine is based on the Rabe–Kindler chemistry discussed in the timelime.

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.

<span class="mw-page-title-main">Elias James Corey</span> American chemist (born 1928)

Elias James Corey is an American organic chemist. In 1990, he won the Nobel Prize in Chemistry "for his development of the theory and methodology of organic synthesis", specifically retrosynthetic analysis.

<span class="mw-page-title-main">Aldol condensation</span> Type of chemical reaction

An aldol condensation is a condensation reaction in organic chemistry in which two carbonyl moieties react to form a β-hydroxyaldehyde or β-hydroxyketone, and this is then followed by dehydration to give a conjugated enone.

A diol is a chemical compound containing two hydroxyl groups. An aliphatic diol is also called a glycol. This pairing of functional groups is pervasive, and many subcategories have been identified.

<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">Dess–Martin periodinane</span> Chemical reagent

Dess–Martin periodinane (DMP) is a chemical reagent used in the Dess–Martin oxidation, oxidizing primary alcohols to aldehydes and secondary alcohols to ketones. This periodinane has several advantages over chromium- and DMSO-based oxidants that include milder conditions, shorter reaction times, higher yields, simplified workups, high chemoselectivity, tolerance of sensitive functional groups, and a long shelf life. However, use on an industrial scale is made difficult by its cost and its potentially explosive nature. It is named after the American chemists Daniel Benjamin Dess and James Cullen Martin who developed the reagent in 1983. It is based on IBX, but due to the acetate groups attached to the central iodine atom, DMP is much more reactive than IBX and is much more soluble in organic solvents.

The Reissert indole synthesis is a series of chemical reactions designed to synthesize indole or substituted-indoles from ortho-nitrotoluene 1 and diethyl oxalate 2.

<span class="mw-page-title-main">Chiral auxiliary</span> Stereogenic group placed on a molecule to encourage stereoselectivity in reactions

In stereochemistry, a chiral auxiliary is a stereogenic group or unit that is temporarily incorporated into an organic compound in order to control the stereochemical outcome of the synthesis. The chirality present in the auxiliary can bias the stereoselectivity of one or more subsequent reactions. The auxiliary can then be typically recovered for future use.

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

The Danishefsky Taxol total synthesis in organic chemistry is an important third Taxol synthesis published by the group of Samuel Danishefsky in 1996 two years after the first two efforts described in the Holton Taxol total synthesis and the Nicolaou Taxol total synthesis. Combined they provide a good insight in the application of organic chemistry in total synthesis.

The aza-Baylis–Hillman reaction or aza-BH reaction in organic chemistry is a variation of the Baylis–Hillman reaction and describes the reaction of an electron deficient alkene, usually an α,β-unsaturated carbonyl compound, with an imine in the presence of a nucleophile. The reaction product is an allylic amine. The reaction can be carried out in enantiomeric excess of up to 90% with the aid of bifunctional chiral BINOL and phosphinyl BINOL compounds, for example in the reaction of n-(4-chloro-benzylidene)-benzenesulfonamide with methyl vinyl ketone (MVK) in cyclopentyl methyl ether and toluene at -15°C.

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

The Holton Taxol total synthesis, published by Robert A. Holton and his group at Florida State University in 1994, was the first total synthesis of Taxol.

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

The article concerns the total synthesis of galanthamine, a drug used for the treatment of mild to moderate Alzheimer's disease.

Oppenauer oxidation, named after Rupert Viktor Oppenauer, is a gentle method for selectively oxidizing secondary alcohols to ketones.

The Rubottom oxidation is a useful, high-yielding chemical reaction between silyl enol ethers and peroxyacids to give the corresponding α-hydroxy carbonyl product. The mechanism of the reaction was proposed in its original disclosure by A.G. Brook with further evidence later supplied by George M. Rubottom. After a Prilezhaev-type oxidation of the silyl enol ether with the peroxyacid to form the siloxy oxirane intermediate, acid-catalyzed ring-opening yields an oxocarbenium ion. This intermediate then participates in a 1,4-silyl migration to give an α-siloxy carbonyl derivative that can be readily converted to the α-hydroxy carbonyl compound in the presence of acid, base, or a fluoride source.

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

Oseltamivir total synthesis concerns the total synthesis of the antiinfluenza drug oseltamivir marketed by Hoffmann-La Roche under the trade name Tamiflu. Its commercial production starts from the biomolecule shikimic acid harvested from Chinese star anise and from recombinant E. coli. Control of stereochemistry is important: the molecule has three stereocenters and the sought-after isomer is only 1 of 8 stereoisomers.

<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">Strychnine total synthesis</span>

Strychnine total synthesis in chemistry describes the total synthesis of the complex biomolecule strychnine. The first reported method by the group of Robert Burns Woodward in 1954 is considered a classic in this research field.

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

Cholesterol total synthesis in chemistry describes the total synthesis of the complex biomolecule cholesterol and is considered a great scientific achievement. The research group of Robert Robinson with John Cornforth published their synthesis in 1951 and that of Robert Burns Woodward with Franz Sondheimer in 1952. Both groups competed for the first publication since 1950 with Robinson having started in 1932 and Woodward in 1949. According to historian Greg Mulheirn the Robinson effort was hampered by his micromanagement style of leadership and the Woodward effort was greatly facilitated by his good relationships with chemical industry. Around 1949 steroids like cortisone were produced from natural resources but expensive. Chemical companies Merck & Co. and Monsanto saw commercial opportunities for steroid synthesis and not only funded Woodward but also provided him with large quantities of certain chemical intermediates from pilot plants. Hard work also helped the Woodward effort: one of the intermediate compounds was named Christmasterone as it was synthesized on Christmas Day 1950 by Sondheimer.

The Kröhnke pyridine synthesis is reaction in organic synthesis between α-pyridinium methyl ketone salts and α, β-unsaturated carbonyl compounds used to generate highly functionalized pyridines. Pyridines occur widely in natural and synthetic products, so there is wide interest in routes for their synthesis. The method is named after Fritz Kröhnke.

<span class="mw-page-title-main">Teruaki Mukaiyama</span> Japanese chemist (1927–2018)

Teruaki Mukaiyama was a Japanese organic chemist. One of the most prolific chemists of the 20th century in the field of organic synthesis, Mukaiyama helped establish the field of organic chemistry in Japan after World War II.

References

  1. Smith, Aaron C.; Williams, Robert M. (2008-02-15). "Rabe Rest in Peace: Confirmation of the Rabe–Kindler Conversion of d‐ Quinotoxine Into Quinine: Experimental Affirmation of the Woodward–Doering Formal Total Synthesis of Quinine". Angewandte Chemie International Edition. 47 (9): 1736–1740. doi:10.1002/anie.200705421. ISSN   1433-7851. PMC   3085927 . PMID   18236503.
  2. Pasteur, L. Compt. rend. 1853, 37, 110.
  3. Perkin, W. H. J. Chem. Soc. 1896, 69, 596
  4. Rabe, P.; Ackerman, E.; Schneider, W. Ber. 1907, 40, 3655
  5. Rabe, P.; Kindler, K. Chem. Ber. 1918, 51, 466
  6. P. Rabe, K. Kindler, Ber. Dtsch. Chem. Ges. B 1939, 72, 263–264.
  7. Proštenik, M.; Prelog, V. HelV. Chim. Acta 1943, 26, 1965.
  8. The Total Synthesis of Quinine R. B. Woodward and W. E. Doering J. Am. Chem. Soc.; 1944; 66(5) pp 849 - 849; doi : 10.1021/ja01233a516
  9. The Total Synthesis of Quinine R. B. Woodward and W. E. Doering J. Am. Chem. Soc.; 1945; 67(5) pp 860 - 874; doi : 10.1021/ja01221a051
  10. SYNTHESIS OF γ-LACTONES BY THE CONDENSATION OF 2-ALKENE-1,4-DIOLS WITH ORTHOCARBOXYLIC ESTERS Kiyosi Kondo and Fumio Mori Chemistry Letters Vol.3 (1974), No.7 pp.741-742 doi : 10.1246/cl.1974.741
  11. Synthesis and Absolute Configuration of the Acetalic Lignan (+)-Phrymarolin Fumito Ishibashi and Eiji Taniguchi Bulletin of the Chemical Society of Japan Vol.61 (1988), No.12 pp.4361-4366 doi : 10.1246/bcsj.61.4361
  12. The First Stereoselective Total Synthesis of Quinine Gilbert Stork, Deqiang Niu, A. Fujimoto, Emil R. Koft, James M. Balkovec, James R. Tata, and Gregory R. Dake J. Am. Chem. Soc.; 2001; 123(14) pp 3239 - 3242; (Article) doi : 10.1021/ja004325r.
  13. M. Jacobs, Chemical & Engineering News 2001, 79 (May 7), 5.
  14. Review: The Woodward-Doering/Rabe-Kindler Total Synthesis of Quinine: Setting the Record Straight Jeffrey I. Seeman Angew. Chem. Int. Ed. 2007, 46, 1378–1413 doi : 10.1002/anie.200601551 PMID   17294412
  15. Communication Rabe Rest in Peace: Confirmation of the Rabe–Kindler Conversion of d-Quinotoxine to Quinine: Experimental Affirmation of the Woodward–Doering Formal Total Synthesis of Quinine Aaron C. Smith, Robert M. Williams Angewandte Chemie International Edition 2008, 47, 1736–1740 doi : 10.1002/anie.200705421
  16. C–H Activation Enables a Concise Total Synthesis of Quinine and Analogues with Enhanced Antimalarial Activity D. H. O'Donovan et al Angewandte Chemie International Edition 2018 doi : 10.1002/anie.201804551