Strychnine total synthesis

Last updated

Strychnine ball-and-stick model based on X-ray data Strychnine-from-xtal-3D-balls.png
Strychnine ball-and-stick model based on X-ray data

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. [2] [3] [4] [5]

Contents

At the time it formed the natural conclusion to an elaborate process of molecular structure elucidation that started with the isolation of strychnine from the beans of Strychnos ignatii by Pierre Joseph Pelletier and Joseph Bienaimé Caventou in 1818. [6] Major contributors to the entire effort were Sir Robert Robinson with over 250 publications and Hermann Leuchs with another 125 papers in a time span of 40 years. Robinson was awarded the Nobel Prize in Chemistry in 1947 for his work on alkaloids, strychnine included.

The process of chemical identification was completed with publications in 1946 by Robinson [7] [8] [9] and later confirmed by Woodward in 1947. [10] X-ray structures establishing the absolute configuration became available between 1947 and 1951 with publications from Johannes Martin Bijvoet [11] [12] and J. H. Robertson [13] [14]

Woodward published a very brief account on the strychnine synthesis in 1954 (just 3 pages) [15] and a lengthy one (42 pages) in 1963. [16]

Routes to Strychnine Many routes to Strychnine have been developed over the years. Some are chiral and some give racemic product. Strychnine Star chemdraw.jpg
Routes to Strychnine Many routes to Strychnine have been developed over the years. Some are chiral and some give racemic product.

Many more methods exist and reported by the research groups of Magnus, [17] Overman, [18] Kuehne, [19] [20] Rawal, [21] Bosch, [22] [23] Vollhardt, [24] [25] Mori, [26] [27] Shibasaki, [28] Li, [29] Fukuyama [30] Vanderwal [31] and MacMillan. [32] Synthetic (+)-strychnine is also known. [33] [34] Racemic synthesises were published by Padwa in 2007 [35] and in 2010 by Andrade [36] and by Reissig. [37] In his 1963 publication Woodward quoted Sir Robert Robinson who said [38] for its molecular size it is the most complex substance known.

The molecule

Strychnine overview.svg

The C21H22N2O2 strychnine molecule contains 7 rings including an indoline system. It has a tertiary amine group, an amide, an alkene and an ether group. The naturally occurring compound is also chiral with 6 asymmetric carbon atoms including one quaternary one.

Strychnine functional groups.svg

Woodward synthesis

Ring II, V synthesis

The synthesis of ring II was accomplished with a Fischer indole synthesis using phenylhydrazine 1 and acetophenone derivative acetoveratrone 2 (catalyst polyphosphoric acid) to give the 2-veratrylindole 3. The veratryl group not only blocks the 2-position for further electrophilic substitution but will also become part of the strychnine skeleton. A Mannich reaction with formaldehyde and dimethylamine) produced gramine 4. Alkylation with iodomethane gave an intermediate quaternary ammonium salt which reacted with sodium cyanide in a nucleophilic substitution to nitrile 5 and then in a reduction with lithium aluminium hydride to tryptamine 6. Amine-carbonyl condensation with ethyl glyoxylate give the imine 7. The reaction of this imine with TsCl in pyridine to the ring-closed N-tosyl compound 8 was described by Woodward as a concerted nucleophilic enamine attack and formally a Pictet–Spengler reaction. This compound should form as a diastereomeric pair but only one compound was found although which one was not investigated. Finally the newly formed double bond was reduced by sodium borohydride to indoline 9 with the C8 hydrogen atom approaching from the least hindered side (this proton is removed later on in the sequence and is of no importance).

Strychnine Woodward 1954 start.svg

Ring III, IV synthesis

Indoline 9 was acetylated to N-acetyl compound 10 (acetic anhydride, pyridine) and then the veratryl group was then ring-opened with ozone in aquaeous acetic acid to muconic ester 11 (made possible by the two electron-donating methoxide groups). This is an example of bioinspired synthesis already proposed by Woodward in 1948. [39] Cleavage of the acetyl group and ester hydrolysis with HCl in methanol resulted in formation of pyridone ester 12 with additional isomerization of the exocyclic double bond to an endocyclic double bond (destroying one asymmetric center). Subsequent treatment with hydrogen iodide and red phosphorus removed the tosyl group and hydrolysed both remaining ester groups to form diacid 13. Acetylation and esterification (diazomethane) produced acetyl diester 14 which was then subjected to a Dieckman condensation with sodium methoxide in methanol to enol 15.

Strychnine Woodward part 2.svg

Ring VII synthesis

In order to remove the C15 alcohol group, Enol 15 was converted to tosylate 16 (TsCl, pyridine) and then to mercaptoester 17 (sodium benzylmercaptide) which was then reduced to unsaturated ester 18 by Raney nickel and hydrogen. Further reduction with hydrogen / palladium on carbon afforded the saturated ester 19. Alkaline ester hydrolysis to carboxylic acid 20 was accompanied by epimerization at C14.

Strychnine Woodward part 3.svg

This particular compound was already known from strychnine degradation studies. Until now all intermediates were racemic but chirality was introduced at this particular stage via chiral resolution using quinidine.

The C20 carbon atom was then introduced by acetic anhydride to form enol acetate 21 and the free aminoketone 22 was obtained by hydrolysis with hydrochloric acid. Ring VII in intermediate 23 was closed by selenium dioxide oxidation, a process accompanied by epimerization again at C14.

Strychnine Woodward part 4.svg

The formation of 21 can be envisioned as a sequence of acylation, deprotonation, rearrangement with loss of carbon dioxide and again acylation:

Strychnine total synthesis enol acetate formation.svg

Ring VI synthesis

To diketone 23, sodium acetylide(Alkynylation) was added (bringing in carbon atoms 22 and 23) to give alkyne 24. This compound was reduced to the allyl alcohol 25 using the Lindlar catalyst and lithium aluminium hydride removed the remaining amide group in 26. An allylic rearrangement to alcohol 27 (isostrychnine) was brought about by hydrogen bromide in acetic acid followed by hydrolysis with sulfuric acid. In the final step to (−)-strychnine 28 treatment of 27 with ethanolic potassium hydroxide caused rearrangement of the C12-13 double bond and ring closure in a conjugate addition by the hydroxyl anion.

Strychnine Woodward part 5.svg

Magnus synthesis

In this effort one of strychnine's many degradation products was synthesised first (the relay compound), a compound also available in several steps from another degradation product called the Wieland-Gumlich aldehyde. In the final leg strychnine itself was synthesised from the relay compound.

Overman synthesis

The Overman synthesis (1993) took a chiral cyclopentene compound as starting material obtained by enzymatic hydrolysis of cis-1,4-diacetoxycyclopent-2-ene. This starting material was converted in several steps to trialkylstannane 2 which was then coupled with an aryl iodide 1 in a Stille reaction in presence of carbon monoxide (tris(dibenzylideneacetone)dipalladium(0), triphenylarsine). The internal double in 3 was converted to an epoxide using tert-Butyl hydroperoxide, the carbonyl group was then converted to an alkene in a Wittig reaction using Ph3P=CH2 and the TIPS group was hydrolyzed (TBAF) and replaced by a trifluoroacetamide group (NH2COCF3, NaH) in 4. Cyclization (NaH) took place next, opening the epoxide ring and the trifluoroacetyl group was removed using KOH affording azabicyclooctane 5.

Strychnine Overmars 1993 start.svg

The key step was an aza-Cope-Mannich reaction initiated by an amine-carbonyl condensation using formaldehyde and forming 6 in a quantitative yield:

Strychnine overmars 1993 aza Cope.svg

In the final sequence strychnine was obtained through the Wieland-Gumlich aldehyde (10):

Strychnine Overmars 1993 part 1.svg

Intermediate 6 was acylated using methyl cyanoformate and two protective groups (tert-butyl and ) were removed using HCl / MeOH in 7. The C8C13 double bond was reduced with zinc (MeOH/H+) to saturated ester 8 (mixture). Epimerization at C13 with sodium methoxide in MeOH produced beta-ester 9 which was reduced with diisobutylaluminium hydride to Wieland-Gumlich aldehyde 10. Conversion of this compound with malonic acid to (−)-strychnine 11 was already known as a procedure.

Kuehne synthesis

The 1993 Keuhne synthesis concerns racemic strychnine. Starting compounds tryptamine 1 and 4,4-dimethoxy acrolein 2 were reacted together with boron trifluoride to acetal 3 as a single diastereomer in an amine-carbonyl condensation / sigmatropic rearrangement sequence.

Strychnine Kuehne 3.svg

Hydrolysis with perchloric acid afforded aldehyde 4. A Johnson–Corey–Chaykovsky reaction (trimethylsulfonium iodide / n-butyllithium) converted the aldehyde into an epoxide which reacted in situ with the tertiary amine to ammonium salt 5 (contaminated with other cyclization products). Reduction (palladium on carbon/hydrogen) removed the benzyl group to alcohol 6, more reduction (sodium cyanoborohydride) and acylation (acetic anhydride / pyridine) produced 7 as a mixture of epimers (at C17). Ring closure of ring III to 8 was then accomplished with an aldol reaction using lithium bis(trimethylsilyl)amide (using only the epimer with correct configuration). Even more reduction (sodium borohydride) and acylation resulted in epimeric di-acetate 9.

Strychnine Kuehne 2.svg

A DBU mediated elimination reaction formed olefinic alcohol 10 and subsequent Swern oxidation have an unstable amino ketone 11. In the final steps a Horner–Wadsworth–Emmons reaction (methyl 2-(diethy1phosphono)acetate) give acrylate ester 12 as a mixture of cis and trans isomers which could be coached into the right (trans) direction by application of light in a photochemical rearrangement, the ester group was reduced (DIBAL / boron trifluoride) to isostrychnine 13 and racemic strychnine 14 was formed by base-catalyzed ring closure as in the Woodward synthesis.

Strychnine Kuehne 1.svg

In the 1998 Keuhne synthesis of chiral (−)-strychnine the starting material was derived from chiral tryptophan.

Rawal synthesis

In the Rawal synthesis (1994, racemic) amine 1 and enone 2 were combined in an amine-carbonyl condensation followed by methyl chloroformate quench to triene 3 which was then reacted in a Diels–Alder reaction (benzene 185 °C) to hexene 4. The three ester groups were hydrolyzed using iodotrimethylsilane forming pentacyclic lactam 5 after a methanol quench in a combination of 7 reaction steps (one of them a Dieckmann condensation). The C4 segment 6 was added in an amine alkylation and Heck reaction of 7 formed isostrychnine 8 after TBS deprotection.

Strychnine synthesis Rawal 1995.svg

The overall yield (10%) is to date the largest of any of the published methods [40]

Bosch synthesis

In the Bosch synthesis of (1999, chiral) the olefin group in dione 1 was converted to an aldehyde by ozonolysis and chiral amine 2 was formed in a double reductive amination with (S)-1-phenethylamine. The phenylethyl substituent was removed using ClCO2CHClCH3 and the enone group was introduced in a Grieco elimination using TMSI, HMDS then PhSeCl then ozone and then diisopropylamine forming carbamate 3. The amino group was deprotected by refluxing in methanol and then alkylated using (Z)-BrCH2CICH=CH2OTBDMS, to tertiary amine 4. A reductive Heck reaction took place next followed by methoxycarbonylation (LiHMDS, NCCO2Me) to tricycle 5. Reaction with zinc dust in 10% sulfuric acid removed the TBDMS protective group, reduced the nitro group and brought about a reductive amino-carbonyl cyclization in a single step to tetracyclic 6 (epimeric mixture). In the final step to the Wieland-Gumlich aldehyde 7 reaction with NaH in MeOH afforded the correct epimer was followed by DIBAH reduction of the methyl ester.

Strychnine synthesis Bosch 1999.svg

Vollhardt synthesis

The key reaction in the Vollhardt synthesis (2000, racemic) was an alkyne trimerisation of tryptamine derivative 1 with acetylene and organocobalt compound CpCo(C2H4)2 (THF, 0 °C) to tricycle 2 after deprotection of the amine group (KOH, MeOH/H2O reflux). Subsequent reaction with iron nitrate brought about a [1,8]-conjugate addition to tetracycle 3, amine alkylation with (Z)-1-bromo-4-[(tert-butyldimethylsilyl)oxy]-2-iodobut-2-ene (see Rawal synthesis) and lithium carbonate, and isomerization of the diene system (NaOiPr, iPrOH) formed enone 4. A Heck reaction as in the Rawal synthesis (palladium acetate / triphenylphosphine), accompanied by aromatization formed pyridone 5 and lithium aluminium hydride reduction and TBS group deprotection formed isostrychnine 6.

Strychnine total synthesis Vollhardt.svg

Mori synthesis

The Mori synthesis ((-) chiral, 2003) was the first one containing an asymmetric reaction step. It also features a large number of Pd catalyzed reactions. In it N-tosyl amine 1 reacted with allyl carbonate 2 in an allylic asymmetric substitution using Pd2(dba)3 and asymmetric ligand (S-BINAPO) to chiral secondary amine 3. Desilylation of the TBDMS group next took place by HCl to the hydroxide and then to the nitrile 4 (NaCN) through the bromide (PBr3). Heck reaction (Pd(OAc)2 / Me2PPh) and debromination (Ag2CO3) afforded tricycle 5. LiALH4 Nitrile reduction to the amine and its Boc2O protection to boc amine 6 was then followed by a second allylic oxidation (Pd(OAc)2 / AcOH / benzoquinone / MnO2) to tetracycle 7. Hydroboration-oxidation (9-BBN / H2O2) gave alcohol 8 and subsequent Swern oxidation ketone 9. Reaction with LDA / PhNTf2 gave enol triflate 10 and the triflate group was removed in alkene 11 by reaction with Pd(OAc)2 and PPh3.

Strychnine total synthesis Mori II.svg

Detosylation of 11 (sodium naphthalenide) and amidation with acid chloride 3-bromoacryloyl chloride gave amide 12 and another Heck reaction gave pentacycle 13. double bond isomerization (sodium / iPrOH), Boc group deprotection (triflic acid) and amine alkylation with (Z)-BrCH2CICH=CH2OTBDMS (see Rawal) gave compound 14 (identical to one of the Vollhardt intermediates). A final heck reaction (15) and TBDMS deprotection formed (−)-isostrychnine 16.

Strychnine total synthesis Mori.svg

Shibasaki synthesis

The Shibasaki synthesis ((-) chiral, 2002) was a second published method in strychnine total synthesis using an asymmetric reaction step. Cyclohexenone 1 was reacted with dimethyl malonate 2 in an asymmetric Michael reaction using AlLibis(binaphthoxide) to form chiral diester 3. Its ketone group was protected as an acetal (2-ethyl-2-methyl-1,3-dioxolane, TsOH) and a carboxyl group was removed (LiCl, DMSO 140 °C) in monoester 4. A C2 fragment was added as Weinreb amide 5 to form PMB ether 6 using LDA. The ketone was then reduced to the alcohol (NaBH3CN, TiCl4) and then water was eliminated (DCC, CuCl) to form alkene 7. After ester reduction (DIBAL) to the alcohol and its TIPS protection (TIPSOTf, triethylamine), the acetal group was removed (catalytic CSA) in ketone 8. Enone 9 was then formed by Saegusa oxidation. The conversion to alcohol 10 was accomplished via a Mukaiyama aldol addition using formaldehyde, iodonation to 11 (iodine, DMAP) was followed by a Stille coupling (Pd2dba3, Ph3As, CuI) incorporating nitrobenzene unit 12. Alcohol 13 was formed after SEM protection (SEMCl,i-Pr2NEt) and TIPS removal (HF).

Strychnine total synthesis Shibasaki 2002a.svg

In the second part of the sequence alcohol 13 was converted to a triflate (triflic anhydride, N,N-diisopropylethylamine), then 2,2-bis(ethylthio)ethylamine 14 was added immediately followed by zinc powder, setting of a tandem reaction with nitro group reduction to the amine, 1,4-addition of the thio-amine group and amine-keto condensation to indole 16. Reaction with DMTSF gave thionium attack at C7 forming 17, the imine group was then reduced (NaBH3CN, TiCl4), the new amino group acylated (acetic anhydride, pyridine), both alcohol protecting groups removed (NaOMe / meOH) and the allyl alcohol group protected again (TIPS). This allowed removal of the ethylthio group (NiCl2, NaBH4, EtOH/MeOH) to 18. The alcohol was oxidized to the aldehyde using a Parikh-Doering oxidation and TIPS group removal gave hemiacetal 19 called (+)-diaboline which is acylated Wieland-Gumlich aldehyde.

Strychnine total synthesis Shibasaki 2002.svg

Li synthesis

The synthesis reported by Bodwell/Li (racemic, 2002) was a formal synthesis as it produced a compound already prepared by Rawal (no. 5 in the Rawal synthesis). The key step was an inverse electron demand Diels–Alder reaction of cyclophane 1 by heating in N,N-diethylaniline (dinitrogen is expulsed) followed by reduction of double bond in 2 to 3 by sodium borohydride / triflic acid and removal of the carbamate protecting group (PDC / celite) to 4.

Strychnine total Synthesis Bodwell 2002.svg

The method is disputed by Reissig (see Reissig synthesis).

Fukuyama synthesis

The Fukuyama synthesis (chiral (-), 2004) started from cyclic amine 1. Chirality was at some point introduced into this starting material by enzymatic resolution of one of the precursors. Acyloin 2 was formed by Rubottom oxidation and hydrolysis. Oxidative cleavage by lead acetate formed aldehyde 3, removal of the nosyl group (thiophenol / cesium carbonate) triggered an amine-carbonyl condensation with iminium ion 4 continuing to react in a transannular cyclization to diester 5 which could be converted to the Wieland-Gumlich aldehyde by known chemistry.

Strychnine total synthesis Fukuyama 2004.svg

Reissig synthesis

The method reported by Beemelmanns & Reissig (racemic, 2010) is another formal synthesis leading to the Rawal pentacycle (see amine 5 in the Rawal method). In this method indole 1 was converted to tetracycle 2 (together with by-product) in a single cascade reaction using samarium diiodide and HMPA. [41] Raney nickel/ H2 reduction gave amine 3 and a one-pot reaction using methyl chloroformate, DMAP and TEA then MsCl, DMAP and TEA and then DBU gave Rawal precursor 4 with key hydrogen atoms in the desired anti configuration.

Strynine total synthesis Beemelmanns 2010.svg

In an aborted route intermediate 2 was first reduced to imine 5 then converted to carbamate 6, then dehydrated to diene 7 (Burgess reagent) and finally reduced to 8 (sodium cyanoborohydride). The hydrogen atoms in 8 are in an undesired cis-relationship which contradicts the results obtained in 2002 by Bodwell/Li for the same reaction.

Vanderwal synthesis

In 2011, the Vanderwal group reported a concise, longest linear sequence of 6 steps, total synthesis of strychnine. [42] It featured a Zincke aldehyde followed by an anionic bicyclization reaction and a tandem Brook rearrangement / conjugate addition.

Vanderwal strychnine retro.png

Related Research Articles

The aldol reaction is a reaction that combines two carbonyl compounds to form a new β-hydroxy carbonyl compound.

Reductive amination is a form of amination that involves the conversion of a carbonyl group to an amine via an intermediate imine. The carbonyl group is most commonly a ketone or an aldehyde. It is a common method to make amines and is widely used in green chemistry since it can be done catalytically in one-pot under mild conditions. In biochemistry, dehydrogenase enzymes use reductive amination to produce the amino acid, glutamate. Additionally, there is ongoing research on alternative synthesis mechanisms which various metal catalysts which require more mild reaction conditions, allowing the reaction to be less energy taxing. Investigation into biocatalysts, such as EnelRED, have allowed for the reduction of chiral amines which is an important factor in pharmaceutical synthesis.

<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">Nicolaou Taxol total synthesis</span>

The Nicolaou Taxol total synthesis, published by K. C. Nicolaou and his group in 1994 concerns the total synthesis of taxol. Taxol is an important drug in the treatment of cancer but also expensive because the compound is harvested from a scarce resource, namely the pacific yew.

<span class="mw-page-title-main">Johnson–Corey–Chaykovsky reaction</span> Chemical reaction in organic chemistry

The Johnson–Corey–Chaykovsky reaction is a chemical reaction used in organic chemistry for the synthesis of epoxides, aziridines, and cyclopropanes. It was discovered in 1961 by A. William Johnson and developed significantly by E. J. Corey and Michael Chaykovsky. The reaction involves addition of a sulfur ylide to a ketone, aldehyde, imine, or enone to produce the corresponding 3-membered ring. The reaction is diastereoselective favoring trans substitution in the product regardless of the initial stereochemistry. The synthesis of epoxides via this method serves as an important retrosynthetic alternative to the traditional epoxidation reactions of olefins.

<span class="mw-page-title-main">Petasis reaction</span>

The Petasis reaction is the multi-component reaction of an amine, a carbonyl, and a vinyl- or aryl-boronic acid to form substituted amines.

<span class="mw-page-title-main">Acyloin</span> Organic compounds of the form –C(=O)C(OH)–

In organic chemistry, acyloins or α-hydroxy ketones are a class of organic compounds of the general form R−C(=O)−CR'(OH)−R", composed of a hydroxy group adjacent to a ketone group. The name acyloin is derived from the fact that they are formally derived from reductive coupling of carboxylic acyl groups. They are one of the two main classes of hydroxy ketones, distinguished by the position of the hydroxy group relative to the ketone; in this form, the hydroxy is on the alpha carbon, explaining the secondary name of α-hydroxy ketone.

In chemistry, transfer hydrogenation is a chemical reaction involving the addition of hydrogen to a compound from a source other than molecular H2. It is applied in laboratory and industrial organic synthesis to saturate organic compounds and reduce ketones to alcohols, and imines to amines. It avoids the need for high-pressure molecular H2 used in conventional hydrogenation. Transfer hydrogenation usually occurs at mild temperature and pressure conditions using organic or organometallic catalysts, many of which are chiral, allowing efficient asymmetric synthesis. It uses hydrogen donor compounds such as formic acid, isopropanol or dihydroanthracene, dehydrogenating them to CO2, acetone, or anthracene respectively. Often, the donor molecules also function as solvents for the reaction. A large scale application of transfer hydrogenation is coal liquefaction using "donor solvents" such as tetralin.

<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.

In organic chemistry, kinetic resolution is a means of differentiating two enantiomers in a racemic mixture. In kinetic resolution, two enantiomers react with different reaction rates in a chemical reaction with a chiral catalyst or reagent, resulting in an enantioenriched sample of the less reactive enantiomer. As opposed to chiral resolution, kinetic resolution does not rely on different physical properties of diastereomeric products, but rather on the different chemical properties of the racemic starting materials. The enantiomeric excess (ee) of the unreacted starting material continually rises as more product is formed, reaching 100% just before full completion of the reaction. Kinetic resolution relies upon differences in reactivity between enantiomers or enantiomeric complexes.

<span class="mw-page-title-main">Organocatalysis</span> Method in organic chemistry

In organic chemistry, organocatalysis is a form of catalysis in which the rate of a chemical reaction is increased by an organic catalyst. This "organocatalyst" consists of carbon, hydrogen, sulfur and other nonmetal elements found in organic compounds. Because of their similarity in composition and description, they are often mistaken as a misnomer for enzymes due to their comparable effects on reaction rates and forms of catalysis involved.

<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.

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.

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

Diisopinocampheylborane is an organoborane that is useful for asymmetric synthesis. This colourless solid is the precursor to a range of related reagents. The compound was reported in 1961 by Zweifel and Brown in a pioneering demonstration of asymmetric synthesis using boranes. The reagent is mainly used for the synthesis of chiral secondary alcohols. The reagent is often depicted as a monomer but like most hydroboranes, it is dimeric with B-H-B bridges.

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

Borane dimethylsulfide (BMS) is a chemical compound with the chemical formula BH3·S(CH3)2. It is an adduct between borane molecule and dimethyl sulfide molecule. It is a complexed borane reagent that is used for hydroborations and reductions. The advantages of BMS over other borane reagents, such as borane-tetrahydrofuran, are its increased stability and higher solubility. BMS is commercially available at much higher concentrations than its tetrahydrofuran counterpart and does not require sodium borohydride as a stabilizer, which could result in undesired side reactions. In contrast, BH3·THF requires sodium borohydride to inhibit reduction of THF to tributyl borate. BMS is soluble in most aprotic solvents.

<span class="mw-page-title-main">Carbonyl reduction</span> Organic reduction of any carbonyl group by a reducing agent

In organic chemistry, carbonyl reduction is the conversion of any carbonyl group, usually to an alcohol. It is a common transformation that is practiced in many ways. Ketones, aldehydes, carboxylic acids, esters, amides, and acid halides - some of the most pervasive functional group]]s, -comprise carbonyl compounds. Carboxylic acids, esters, and acid halides can be reduced to either aldehydes or a step further to primary alcohols, depending on the strength of the reducing agent. Aldehydes and ketones can be reduced respectively to primary and secondary alcohols. In deoxygenation, the alcohol group can be further reduced and removed altogether by replacement with H.

<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.

Metal-catalyzed C–H borylation reactions are transition metal catalyzed organic reactions that produce an organoboron compound through functionalization of aliphatic and aromatic C–H bonds and are therefore useful reactions for carbon–hydrogen bond activation. Metal-catalyzed C–H borylation reactions utilize transition metals to directly convert a C–H bond into a C–B bond. This route can be advantageous compared to traditional borylation reactions by making use of cheap and abundant hydrocarbon starting material, limiting prefunctionalized organic compounds, reducing toxic byproducts, and streamlining the synthesis of biologically important molecules. Boronic acids, and boronic esters are common boryl groups incorporated into organic molecules through borylation reactions. Boronic acids are trivalent boron-containing organic compounds that possess one alkyl substituent and two hydroxyl groups. Similarly, boronic esters possess one alkyl substituent and two ester groups. Boronic acids and esters are classified depending on the type of carbon group (R) directly bonded to boron, for example alkyl-, alkenyl-, alkynyl-, and aryl-boronic esters. The most common type of starting materials that incorporate boronic esters into organic compounds for transition metal catalyzed borylation reactions have the general formula (RO)2B-B(OR)2. For example, bis(pinacolato)diboron (B2Pin2), and bis(catecholato)diborane (B2Cat2) are common boron sources of this general formula.

The asymmetric addition of alkynylzinc compounds to aldehydes is an example of a Nef synthesis, a chemical reaction whereby a chiral propargyl alcohol is prepared from a terminal alkyne and an aldehyde. This alkynylation reaction is enantioselective and involves an alkynylzinc reagent rather than the sodium acetylide used by John Ulric Nef in his 1899 report of the synthetic approach. Propargyl alcohols are versatile precursors for the chirally-selective synthesis of natural products and pharmaceutical agents, making this asymmetric addition reaction of alkynylzinc compounds useful. For example, Erick Carreira used this approach in a total synthesis of the marine natural product leucascandrolide A, a bioactive metabolite of the calcareous sponge Leucascandra caveolata with cytotoxic and antifungal properties isolated in 1996.

Shiina esterification is an organic chemical reaction that synthesizes carboxylic esters from nearly equal amounts of carboxylic acids and alcohols by using aromatic carboxylic acid anhydrides as dehydration condensation agents. In 1994, Prof. Isamu Shiina reported an acidic coupling method using Lewis acid, and, in 2002, a basic esterification using nucleophilic catalyst.

References

  1. X-ray; Messerschmidt, M.; Scheins, S.; Luger, P. (2005). "Charge density of (−)-strychnine from 100 to 15 K, a comparison of four data sets". Acta Crystallogr B. 61 (1): 115–121. doi:10.1107/S0108768104032781. PMID   15659864.
  2. Nicolaou, K. C.; Sorensen, E. J. (1996). Classics in Total Synthesis: Targets, Strategies, Methods. Wiley. ISBN   978-3-527-29231-8.
  3. K. C. Nicolaou, Dionisios Vourloumis, Nicolas Winssinger, Phil S. Baran The Art and Science of Total Synthesis at the Dawn of the Twenty-First Century Angewandte Chemie International Edition 2000; Volume 39, Issue 1, Pages: 44-122
  4. Bonjoch, Josep; Sole, Daniel (2000). "Synthesis of Strychnine". Chem. Rev. 100 (9): 3455–3482. doi:10.1021/cr9902547. PMID   11777429.
  5. Proudfoot, John R. (2013). "Reaction Schemes Visualized in Network Form: The Syntheses of Strychnine as an Example". Journal of Chemical Information and Modeling. 53 (5): 1035–1042. doi:10.1021/ci300556b. PMID   23597302.
  6. Pelletier; Caventou (1818). "Note sur un nouvel alkalai (Note on a new alkali)". Annales de Chimie et de Physique. 8: 323–324. See also: Pelletier; Caventou (1819). "Mémoire sur un nouvel alcali vegetal (la strychnine) trouvé dans la feve de Saint-Ignace, la noix vomique, etc. (Memoir on a new vegetable alkali (strychnine) found in the St. Ignatius bean, the nux vomica, etc)". Annales de Chimie et de Physique. 10: 142–176.
  7. Robinson, R. (1946). "The constitution of strychnine". Experientia. 2 (1): 1946. doi:10.1007/BF02154708. PMID   21012825.
  8. Briggs, L. H.; Openshaw, H. T.; Robinson, Robert (1946). "Strychnine and brucine. Part XLII. Constitution of the neo-series of bases and their oxidation products". J. Chem. Soc. 1946: 903. doi:10.1039/JR9460000903.
  9. Openshaw, H. T.; Robinson, R. (1946). "Constitution of Strychnine and the Biogenetic Relationship of Strychnine and Quinine". Nature. 157 (3988): 438. Bibcode:1946Natur.157..438O. doi: 10.1038/157438a0 . PMID   21024272.
  10. Woodward, R. B.; Brehm, Warren J.; Nelson, A. L. (1947). "The Structure of Strychnine". J. Am. Chem. Soc. 69 (9): 2250. doi:10.1021/ja01201a526. PMID   20262753.
  11. Bijvoet , Schoone and Bokhoven , Kon. Ned. Akad. Wet., 50, No 8, 51, No. 8, 52, No. 2 (1947–49)
  12. Bokhoven, C.; Schoone, J. C.; Bijvoet, J. M. (1951). "The Fourier synthesis of the crystal structure of strychnine sulphate pentahydrate" (PDF). Acta Crystallogr. 4 (3): 275–280. doi: 10.1107/S0365110X51000891 .
  13. Robertson, J. H.; Beevers, C. A. (1950). "Crystal Structure of Strychnine Hydrobromide". Nature. 165 (4200): 690–691. Bibcode:1950Natur.165..690R. doi:10.1038/165690a0. PMID   15416785.
  14. Robertson, J. H.; Beevers, C. A. (1951). "The crystal structure of strychnine hydrogen bromide". Acta Crystallogr. 4 (3): 270–275. doi:10.1107/S0365110X5100088X.
  15. Woodward, R. B.; Cava, Michael P.; Ollis, W. D.; Hunger, A.; Daeniker, H. U.; Schenker, K. (1954). "The Total Synthesis of Strychnine". J. Am. Chem. Soc. 76 (18): 4749–4751. doi:10.1021/ja01647a088.
  16. Woodward, R. B.; Cava, M. P.; Ollis, W. D.; Hunger, A.; Daeniker, H. U.; Schenker, K. (1963). "The total synthesis of strychnine". Tetrahedron. 19 (2): 247–288. doi:10.1016/s0040-4020(01)98529-1.
  17. Magnus, Philip; Giles, Melvyn; Bonnert, Roger; Kim, Chung S.; McQuire, Leslie; Merritt, Andrew; Vicker, Nigel (1992). "Synthesis of strychnine via the Wieland-Gumlich aldehyde". J. Am. Chem. Soc. 114 (11): 4403–4405. doi:10.1021/ja00037a058.
  18. Knight, Steven D.; Overman, Larry E.; Pairaudeau, Garry (1993). "Synthesis applications of cationic aza-Cope rearrangements. 26. Enantioselective total synthesis of (−)-strychnine". J. Am. Chem. Soc. 115 (20): 9293–9294. doi:10.1021/ja00073a057.
  19. Kuehne, Martin E.; Xu, Feng (1993). "Total synthesis of strychnan and aspidospermatan alkaloids. 3. The total synthesis of (+-)-strychnine". J. Org. Chem. 58 (26): 7490–7497. doi:10.1021/jo00078a030.
  20. Kuehne, Martin E.; Xu, Feng (1998). "Syntheses of Strychnan- and Aspidospermatan-Type Alkaloids. 10. An Enantioselective Synthesis of (−)-Strychnine through the Wieland−Gumlich Aldehyde". J. Org. Chem. 63 (25): 9427–9433. doi:10.1021/jo9813989.
  21. Rawal, Viresh H.; Iwasa, Seiji (1994). "A Short, Stereocontrolled Synthesis of Strychnine". J. Org. Chem. 59 (10): 2685–2686. doi:10.1021/jo00089a008.
  22. Total Synthesis of (−)-Strychnine via the Wieland-Gumlich Aldehyde Angewandte Chemie International Edition Volume 38, Issue 3, 1999, Pages: 395-397 Daniel Solé, Josep Bonjoch, Silvina García-Rubio, Emma Peidró, Joan Bosch
  23. Solé, Daniel; Bonjoch, Josep; García-Rubio, Silvina; Peidró, Emma; Bosch, Joan (2000). "Enantioselective Total Synthesis of Wieland-Gumlich Aldehyde and (−)-Strychnine". Chemistry: A European Journal . 6 (4): 655–665. doi:10.1002/(SICI)1521-3765(20000218)6:4<655::AID-CHEM655>3.0.CO;2-6.
  24. Eichberg, Michael J.; Dorta, Rosa L.; Lamottke, Kai; Vollhardt, K. Peter C. (2000). "The Formal Total Synthesis of (±)-Strychnine via a Cobalt-Mediated [2 + 2 + 2]Cycloaddition". Org. Lett. 2 (16): 2479–2481. doi:10.1021/ol006131m. PMID   10956526.
  25. Eichberg, Michael J.; Dorta, Rosa L.; Grotjahn, Douglas B.; Lamottke, Kai; Schmidt, Martin; Vollhardt, K. Peter C. (2001). "Approaches to the Synthesis of (±)-Strychnine via the Cobalt-Mediated [2 + 2 + 2] Cycloaddition: Rapid Assembly of a Classic Framework". J. Am. Chem. Soc. 123 (38): 9324–9337. doi:10.1021/ja016333t. PMID   11562215.
  26. Nakanishi, Masato; Mori, Miwako (2002). "Total Synthesis of (−)-Strychnine". Angewandte Chemie International Edition . 41 (11): 1934–1936. doi:10.1002/1521-3773(20020603)41:11<1934::AID-ANIE1934>3.0.CO;2-F. PMID   19750638.
  27. Mori, Miwako; Nakanishi, Masato; Kajishima, Daisuke; Sato, Yoshihiro (2003). "A Novel and General Synthetic Pathway to Strychnos Indole Alkaloids: Total Syntheses of (−)-Tubifoline, (−)-Dehydrotubifoline, and (−)-Strychnine Using Palladium-Catalyzed Asymmetric Allylic Substitution". J. Am. Chem. Soc. 125 (32): 9801–9807. doi:10.1021/ja029382u. PMID   12904045.
  28. Ohshima, Takashi; Xu, Youjun; Takita, Ryo; Shimizu, Satoshi; Zhong, Dafang; Shibasaki, Masakatsu (2002). "Enantioselective Total Synthesis of (−)-Strychnine Using the Catalytic Asymmetric Michael Reaction and Tandem Cyclization". J. Am. Chem. Soc. 124 (49): 14546–14547. doi:10.1021/ja028457r. PMID   12465959.
  29. Bodwell, Graham J.; Li, Jiang (2002). "A Concise Formal Total Synthesis of (±)-Strychnine by Using a Transannular Inverse-Electron-Demand Diels–Alder Reaction of a [3](1,3)Indolo[3](3,6)pyridazinophane". Angewandte Chemie International Edition . 41 (17): 3261–3262. doi:10.1002/1521-3773(20020902)41:17<3261::AID-ANIE3261>3.0.CO;2-K.
  30. Kaburagi, Y; Tokuyama, H; Fukuyama, T (2004). "Total synthesis of (−)-strychnine". J. Am. Chem. Soc. 126 (33): 10246–10247. doi:10.1021/ja046407b. PMID   15315428.
  31. Martin, David B. C.; Vanderwal, Christopher D. (2011). "A synthesis of strychnine by a longest linear sequence of six steps". Chemical Science. 2 (4): 649. doi:10.1039/C1SC00009H.
  32. Jones, Spencer B.; Simmons, Bryon; Mastracchio, Anthony; MacMillan, David W. C. (2011). "Collective synthesis of natural products by means of organocascade catalysis". Nature. 475 (7355): 183–188. doi:10.1038/nature10232. PMC   3439143 . PMID   21753848.
  33. Knight, Steven D.; Overman, Larry E.; Pairaudeau, Garry (1995). "Asymmetric Total Syntheses of (−)- and (+)-Strychnine and the Wieland-Gumlich Aldehyde". J. Am. Chem. Soc. 117 (21): 5776–5788. doi:10.1021/ja00126a017.
  34. Not counted:an unpublished method by Gilbert Stork, Lecture at the Ischia School of Organic Chemistry, Ischia Porb, Italy, September 211992.
  35. Zhang, Hongjun; Boonsombat, Jutatip; Padwa, Albert (2007). "Total Synthesis of (±)-Strychnine via a [4 + 2]-Cycloaddition/Rearrangement Cascade". Org. Lett. 9 (2): 279–282. doi:10.1021/ol062728b. PMC   2587098 . PMID   17217284.
  36. Sirasani, Gopal; Paul, Tapas; William Dougherty, Jr.; Kassel, Scott; Andrade, Rodrigo B. (2010). "Concise Total Syntheses of (±)-Strychnine and (±)-Akuammicine". The Journal of Organic Chemistry. 75 (10): 3529–3532. doi:10.1021/jo100516g. PMID   20408591.
  37. Beemelmanns, C.; Reissig, H.-U. (2010). "A Short Formal Total Synthesis of Strychnine with a Samarium Diiodide Induced Cascade Reaction as the Key Step". Angewandte Chemie International Edition. 49 (43): 8021–8025. doi:10.1002/anie.201003320. PMID   20848626.
  38. R. Robinson "Molecular structure of Strychnine, Brucine and Vomicine Prog. Org. Chem., 1952; 1 ,2
  39. Woodward, R. B. (1948). "Biogenesis of the Strychnos Alkaloids". Nature. 162 (4108): 155–156. Bibcode:1948Natur.162..155W. doi:10.1038/162155a0. PMID   18871488.
  40. Cannon, J. S.; Overman, L. E. (2012). "Is There No End to the Total Syntheses of Strychnine? Lessons Learned in Strategy and Tactics in Total Synthesis". Angew. Chem. Int. Ed. 51 (18): 4288–4311. doi:10.1002/anie.201107385. PMC   3804246 . PMID   22431197.
  41. Szostak, M.; Procter, D. J. (2011). "Concise Syntheses of Strychnine and Englerin A: the Power of Reductive Cyclizations Triggered by Samarium Iodide". Angewandte Chemie International Edition. 50 (34): 7737–7739. doi:10.1002/anie.201103128. PMID   21780264.
  42. Martin, David B. C.; Vanderwal, Christopher D. (2011). "A synthesis of strychnine by a longest linear sequence of six steps". Chemical Science. 2 (4): 649. doi:10.1039/C1SC00009H.