Protecting group

From Wikipedia the free encyclopedia

Ethylene glycol protects a ketone (as an acetal) during an ester reduction, vs. unprotected reduction to a diol

A protecting group or protective group is introduced into a molecule by chemical modification of a functional group to obtain chemoselectivity in a subsequent chemical reaction. It plays an important role in multistep organic synthesis.[1]

In many preparations of delicate organic compounds, specific parts of the molecules cannot survive the required reagents or chemical environments. These parts (functional groups) must be protected. For example, lithium aluminium hydride is a highly reactive reagent that usefully reduces esters to alcohols. It always reacts with carbonyl groups, and cannot be discouraged by any means. When an ester must be reduced in the presence of a carbonyl, hydride attack on the carbonyl must be prevented. One way to do so converts the carbonyl into an acetal, which does not react with hydrides. The acetal is then called a protecting group for the carbonyl. After the hydride step is complete, aqueous acid removes the acetal, restoring the carbonyl. This step is called deprotection.

Protecting groups are more common in small-scale laboratory work and initial development than in industrial production because they add additional steps and material costs. However, compounds with repetitive functional groups – generally, biomolecules like peptides, oligosaccharides or nucleotides – may require protecting groups to order their assembly. Also, cheap chiral protecting groups may often shorten an enantioselective synthesis (e.g. shikimic acid for oseltamivir).

As a rule, the introduction of a protecting group is straightforward. The difficulties honestly lie in their stability and in selective removal. Apparent problems in synthesis strategies with protecting groups are rarely documented in the academic literature.[2]

Orthogonal protection[edit]

Orthogonal protection of L-Tyrosine (Protecting groups are marked in blue, the amino acid is shown in black). (1) Fmoc-protected amino group, (2) benzyl ester protected carboxyl group and (3) tert-butyl ether protected phenolic hydroxyl group of Tyrosine.

Orthogonal protection is a strategy allowing the specific deprotection of one protective group in a multiply-protected structure. For example, the amino acid tyrosine could be protected as a benzyl ester on the carboxyl group, a fluorenylmethylenoxy carbamate on the amine group, and a tert-butyl ether on the phenol group. The benzyl ester can be removed by hydrogenolysis, the fluorenylmethylenoxy group (Fmoc) by bases (such as piperidine), and the phenolic tert-butyl ether cleaved with acids (e.g. with trifluoroacetic acid).

A common example for this application, the Fmoc peptide synthesis, in which peptides are grown in solution and on solid phase, is very important.[3] The protecting groups in solid-phase synthesis regarding the reaction conditions such as reaction time, temperature and reagents can be standardized so that they are carried out by a machine, while yields of well over 99% can be achieved. Otherwise, the separation of the resulting mixture of reaction products is virtually impossible (see also § Industrial applications).[4]

A further important example of orthogonal protecting groups occurs in carbohydrate chemistry. As carbohydrates or hydroxyl groups exhibit very similar reactivities, a transformation that protects or deprotects a single hydroxy group must be possible for a successful synthesis.

Cleavage categorization[edit]

Many reaction conditions have been established that will cleave protecting groups. One can roughly distinguish between the following environments:[5]

  • Acid-labile protecting groups
  • Base-labile protecting groups
  • Fluoride-labile protecting groups
  • Enzyme-labile protecting groups
  • Reduction-labile protecting groups
  • Oxidation-labile protecting groups
  • Protecting groups cleaved by heavy metal salts or their complexes.
  • Photolabile protecting groups
  • Double-layered protecting groups

Various groups are cleaved in acid or base conditions, but the others are more unusual.

Fluoride ions form very strong bonds to silicon; thus silicon protecting groups are almost invariably removed by fluoride ions. Each type of counterion, i.e. cleavage reagent, can also selectively cleave different silicon protecting groups depending on steric hindrance. The advantage of fluoride-labile protecting groups is that no other protecting group is attacked by the cleavage conditions.

Lipases and other enzymes cleave ethers at biological pH (5-9) and temperatures (30–40 °C). Because enzymes have very high substrate specificity, the method is quite rare, but extremely attractive.

Catalytic hydrogenation removes a wide variety of benzyl groups: ethers, esters, urethanes, carbonates, etc.

Only a few protecting groups can be detached oxidatively: the methoxybenzyl ethers, which oxidize to a quinomethide. They can be removed with ceric ammonium nitrate (CAN) or dichlorodicyanobenzoquinone (DDQ).

Allyl compounds will isomerize to a vinyl group in the presence of noble metals. The residual enol ether (from a protected alcohol) or enamine (resp. amine) hydrolyzes in light acid.

Photolabile protecting groups bear a chromophore, which is activated through radiation with an appropriate wavelength and so can be removed.[6] For examples the o-nitrobenzylgroup ought be listed here.

Mechanism of photodeprotection of an o-nitrobenzyl ether and formation of an alcohol

The rare double-layer protecting group is a protected protecting group, which exemplify high stability.

Common protecting groups[edit]

Alcohol protecting groups[edit]

The classical protecting groups for alcohols are esters, deprotected by nucleophiles; triorganosilyl ethers, deprotected by acids and fluoride ions; and (hemi)acetals, deprotected by weak acids. In rarer cases, a carbon ether might be used.

The most important esters with common protecting-group use are the acetate, benzoate, and pivalate esters, for these exhibit differential removal. Sterically hindered esters are less susceptible to nucleophilic attack:

Chloroacetyl > acetyl > benzoyl > pivaloyl
Trimethylsilyl chloride, activated with imidazole, protects a secondary alcohol

Triorganosilyl sources have quite variable prices, and the most economical is chlorotrimethylsilane (TMS-Cl), a Direct Process byproduct. The trimethylsilyl ethers are also extremely sensitive to acid hydrolysis (for example silica gel suffices as a proton donator) and are consequently rarely used nowadays as protecting groups.

Aliphatic methyl ethers cleave with difficulty and only under drastic conditions, so that these are in general only used with quinonic phenols. However, hemiacetals and acetals are much easier to cleave.

Protection of alcohol as tetrahydropyranyl ether followed by deprotection. Both steps require acid catalysts.

List[edit]

Esters:

  • Acetyl (Ac) – Removed by acid or base (see Acetoxy group).
  • Benzoyl (Bz) – Removed by acid or base, more stable than Ac group.
  • Pivaloyl (Piv) – Removed by acid, base or reductant agents. It is substantially more stable than other acyl protecting groups.

Silyl ethers:

Benzyl ethers:

  • Benzyl (Bn) — Removed by hydrogenolysis.[19] Bn group is widely used in sugar and nucleoside chemistry.
  • Trityl (triphenylmethyl, Tr) — Removed by acid[20][21][22] and hydrogenolysis
  • p-Methoxybenzyl ether (PMB) — Removed by acid, hydrogenolysis, or oxidation – commonly with DDQ.[23]
  • p,m‑Dimethoxybenzyl ether — Removed via oxidation with DDQ or ceric ammonium chloride[24]

Acetals:

Other ethers:

  • p-Methoxyphenyl ether (PMP) – Removed by oxidation.[citation needed]
  • Tert-butyl ethers (tBu) – Removed with anhydrous trifluoroacetic acid, hydrogen bromide in acetic acid, or 4 N hydrochloric acid[42]
  • Allyl — Removed with potassium tert‑butoxide[43] DABCO in methanol, palladium on activated carbon, or diverse platinum complexes – conjoined with acid workup.[44]
  • Methyl ethers – Cleavage is by TMSI in dichloromethane or acetonitrile or chloroform. An alternative method to cleave methyl ethers is BBr3 in DCM
  • Tetrahydrofuran (THF)[clarification needed] – Removed by acid.

1,2-Diols[edit]

The 1,2‑diols (glycols) present for protecting-group chemistry a special class of alcohols. One can exploit the adjacency of two hydroxy groups, e.g. in sugars, in that one protects both hydroxy groups codependently as an acetal. Common in this situation are the benzylidene, isopropylidene and cyclohexylidene or cyclopentylidene acetals.

Applied acetal

An exceptional case appears with the benzylideneprotecting group,which also admits reductive cleavage. This proceeds either through catalytic hydrogenation or with the hydride donor diisobutyl aluminum hydride (DIBAL). The cleavage with DIBAL deprotects one alcohol group, for the benzyl moiety stays as a benzyl ether on the second, sterically hindered hydroxy group.[45][46]

Cleaving a benzylidene acetal with DIBAL

Amine protecting groups[edit]

BOC glycine. The tert-butyloxycarbonyl group is marked blue.

Amines have a special importance in peptide synthesis, but are a quite potent nucleophile and also relatively strong bases. These characteristics imply that new protecting groups for amines are always under development.[47]

Amine groups are primarily protected through acylation, typically as a carbamate. When a carbamate deprotects, it evolves carbon dioxide. The commonest-used carbamates are the tert-butoxycarbonyl, benzoxycarbonyl, fluorenylmethylenoxycarbonyl, and allyloxycarbonyl compounds.

Other, more exotic amine protectors are the phthalimides, which admit reductive cleavage,[48] and the trifluoroacetamides, which hydrolyze easily in base. Indoles, pyrroles und imidazoles — verily any aza-heterocycle — admit protection as N‑sulfonylamides,which are far too stable with aliphatic amines.[49] N‑benzylated amines can be removed through catalytic hydrogenation or Birch reduction, but have a decided drawback relative to the carbamates or amides: they retain a basic nitrogen.

Selection[edit]

Carbamates:

Other amides:

Benzylamines:

Carbonyl protecting groups[edit]

The most common protecting groups for carbonyls are acetals and typically cyclic acetals with diols. The runners-up used are also cyclic acetals with 1,2‑hydroxythiols or dithioglycols – the so-called O,S– or S,S-acetals.

Ethylene glycol
1,3‑Propadiol

Overall, trans-acetalation plays a lesser role in forming protective acetals; they are formed as a rule from glycols through dehydration. Normally a simple glycol like ethylene glycol or 1,3-propadiol is used for acetalation.Modern variants also use glycols, but with the hydroxyl hydrogens replaced with a trimethylsilyl group.[60][61]

Acetals can be removed in acidic aqueous conditions. For those ends, the mineral acids are appropriate acids. Acetone is a common cosolvent, used to promote dissolution. For a non-acidic cleavage technique, a palladium(II) chloride acetonitrile complex in acetone[62] or iron(III) chloride on silica gel can be performed with workup in chloroform.[63]

Cyclic acetals are very much more stable against acid hydrolysis than acyclic acetals. Consequently acyclic acetals are used practically only when a very mild cleavage is required or when two different protected carbonyl groups must be differentiated in their liberation.[64]

Besides the O,O-acetals, the S,O- and S,S-acetals also have an application, albeit scant, as carbonyl protecting groups too. Thiols, which one begins with to form these acetals, have a very unpleasant stench and are poisonous, which severely limit their applications. Thioacetals and the mixed S,O-acetals are, unlike the pure O,O-acetals, very much stabler against acid hydrolysis. This enables the selective cleavage of the latter in the presence of sulfur-protected carbonyl groups.

The formation of S,S-acetals normally follows analogously to the O,O-acetals with acid catalysis from a dithiol and the carbonyl compound. Because of the greater stability of thioacetals, the equilibirum lies on the side of the acetal. In contradistinction to the O,O‑acetal case, it is not needed to remove water from the reaction mixture in order to shift the equilibrium.[65]

S,O-Acetals are hydrolyzed a factor of 10,000 times faster than the corresponding S,S-acetals. Their formation follows analogously from the thioalcohol. Also their cleavage proceeds under similar conditions and predominantly through mercury(II) compounds in wet acetonitrile.[66]

For aldehydes, a temporary protection of the carbonyl group the presence of ketones as hemiaminal ions is shown below. Here it is applied, that aldehydes are very much more activated carbonyls than ketones and that many addition reactions are reversible.[67][68]

Temporary protection of an aldehyde

Types of protectants[edit]

  • Acetals and Ketals – Removed by acid. Normally, the cleavage of acyclic acetals is easier than of cyclic acetals.
  • Acylals – Removed by Lewis acids.
  • Dithianes – Removed by metal salts or oxidizing agents.

Carboxylic acid protecting groups[edit]

The most important protecting groups for carboxylic acids are the esters of various alcohols. Occasionally, esters are protected as ortho-esters or oxazolines.[69]

Many groups can suffice for the alcoholic component, and the specific cleaving conditions are contrariwise generally quite similar: each ester can be hydrolyzed in a basic water-alcohol solution. Instead, most ester protecting groups vary in how mildly they can be formed from the original acid.

Protecting groups[edit]

Alkene[edit]

Alkenes rarely need protection or are protected. They are as a rule only involved in undesired side reactions with electrophilic attack, isomerization or catalytic hydration. For alkenes two protecting groups are basically known:

  • Temporary halogenation with bromine to a trans‑1,2‑dibromoalkane: the regeneration of the alkene then follows with preservation of conformation via elemental zinc[86][87][88][89][90] or with titanocene dichloride.[91]
  • Protection through a Diels-Alder reaction: the transformation of an alkene with a diene leads to a cyclic alkene, which is nevertheless similarly endangered by electrophilic attack as the original alkene. The cleavage of a protecting diene proceeds thermically, for the Diels-Alder reaction is a reversible (equilibrium) reaction.[92][93][94]
Schemata of alkene protecting groups

Phosphate protecting groups[edit]

  • 2-cyanoethyl – removed by mild base. The group is widely used in oligonucleotide synthesis.
  • Methyl (Me) – removed by strong nucleophiles e.c. thiophenole/TEA.

Terminal alkyne protecting groups[edit]

For alkynes there are in any case two types of protecting groups. For terminal alkynes it is sometimes important to mask the acidic hydrogen atom. This normally proceeds from deprotonation (via a strong base like methylmagnesium bromide or butyllithium in tetrahydrofuran/dimethylsulfoxide) and subsequently reaction with chlorotrimethylsilane to a terminally TMS-protected alkyne.[95] Cleavage follows hydrolytically – with potassium carbonate in methanol – or with fluoride ions like for example with tetrabutylammonium fluoride.[96]

Alkyne TMS protection
Alkyne TMS protection

In order to protect the triple bond itself, sometimes a transition metal-alkyne complex with dicobalt octacarbonyl is used. The release of the cobalt then follows from oxidation.[97][98][99][100][101]

Alkyne protection with Co

Other[edit]

Criticism[edit]

The use of protective groups is pervasive but not without criticism.[103] In practical terms their use adds two steps (protection-deprotection sequence) to a synthesis, either or both of which can dramatically lower chemical yield. Crucially, added complexity impedes the use of synthetic total synthesis in drug discovery. In contrast biomimetic synthesis does not employ protective groups. As an alternative, Baran presented a novel protective-group free synthesis of the compound hapalindole U. The previously published synthesis[104][105][106] according to Baran, contained 20 steps with multiple protective group manipulations (two confirmed):

Protected and unprotected syntheses of the marine alkaloid, hapalindole U.
Hideaki Muratake's 1990 synthesis using Tosyl protecting groups (shown in blue).
Phil Baran's protecting-group free synthesis, reported in 2007.

Industrial applications[edit]

Although the use of protecting groups is not preferred in industrial syntheses, they are still used in industrial contexts, e.g. sucralose (sweetener) or the Roche synthesis of oseltamivir (Tamiflu, an antiviral drug)

An important example of industrial applications of protecting group theory is the synthesis of ascorbic acid (Vitamin C) à la Reichstein.

The Reichstein synthesis (of ascorbic acid)

In order to prevent oxidation of the secondary alcohols with potassium permanganate, they are protected via acetalation with acetone and then deprotected after the oxidation of the primary alcohols to carboxylic acids.[107]

A very spectacular example application of protecting groups from natural product synthesis is the 1994 total synthesis of palytoxin acid by Yoshito Kishi's research group.[108] Here 42 functional groups (39 hydroxyls, one diol, an amine group, and a carboxylic acid) required protection. These proceeded through 8 different protecting groups (a methyl ester, five acetals, 20 TBDMS esters, nine p‑methoxybenzyl ethers, four benzoates, a methyl hemiacetal, an acetone acetal and an SEM ester).[109]

Palytoxin

The introduction or modification of a protecting group occasionally influences the reactivity of the whole molecule. For example, diagrammed below is an excerpt of the synthesis of an analogue of Mitomycin C by Danishefsky.[110]

Part of the synthesis of an analogue of Mitomycin C with modified reactivity through protecting-group exchange

The exchange of a protecting group from a methyl ether to a MOM-ether inhibits here the opening of an epoxide to an aldehyde.

Protecting group chemistry finds itself an important application in the automated synthesis of peptides and nucleosides. The technique was introduced in the field of peptide synthesis by Robert Bruce Merrifield in 1977.[111] For peptide synthesis via automated machine, the orthogonality of the Fmoc group (basic cleavage), the tert‑butyl group (acidic cleavage) and diverse protecting groups for functional groups on the amino acid side-chains are used.[112] Up to four different protecting groups per nucleobase are used for the automated synthesis of DNA and RNA sequences in the oligonucleotide synthesis. The procedure begins actually with redox chemistry at the protected phosphorus atom. A tricoordinate phosphorus, used on account of the high reactivity, is tagged with a cyanoethyl protecting group on a free oxygen. After the coupling step follows an oxidation to phosphate, whereby the protecting group stays attached. Free OH-groups, which did not react in the coupling step, are acetylated in an intermediate step. These additionally-introduced protecting groups then inhibit, that these OH-groups might couple in the next cycle.[113]

Automatic oligonucleotide synthesis

References[edit]

  1. ^ Theodora W. Greene; Peter G. M. Wuts (1999). Protecting Groups in Organic Synthesis (3 ed.). J. Wiley. ISBN 978-0-471-16019-9.
  2. ^ Michael Schelhaas, Herbert Waldmann: "Schutzgruppenstrategien in der organischen Synthese", in: Angewandte Chemie, 1996, 103, pp. 2194; doi:10.1002/ange.19961081805 (in German).
  3. ^ Chan, Weng C.; White, Peter D. (2004). Fmoc Solid Phase Peptide Synthesis. Oxford University Press. ISBN 978-0-19-963724-9.
  4. ^ Weng C. Chan, Peter D. White: Fmoc Solid Phase Peptide Synthesis, S. 10–12.
  5. ^ Michael Schelhaas, Herbert Waldmann: "Schutzgruppenstrategien in der organischen Synthese", in: Angewandte Chemie, 1996, 103, pp. 2195–2200; doi:10.1002/ange.19961081805 (in German).
  6. ^ V.N. Rajasekharan Pillai: "Photoremovable Protecting Groups in Organic Synthesis", in: Synthesis, 1980, pp. 1–26.
  7. ^ P.J. Kocieński: Protecting Groups, p. 29.
  8. ^ P.J. Kocieński: Protecting Groups, p. 31.
  9. ^ Tod K Jones, Robert A. Reamer, Richard Desmond, Sander G. Mills: "Chemistry of tricarbonyl hemiketals and application of Evans technology to the total synthesis of the immunosuppressant (−)-FK-506", in: J. Am. Chem. Soc., 1990, 112, pp. 2998–3017; doi:10.1021/ja00164a023.
  10. ^ Dieter Seebach, Hak-Fun Chow, Richard F.W. Jackson, Marius A. Sutter, Suvit Thaisrivongs, Jürg Zimmermann: "(+)-11,11-Di-O-methylelaiophylidene – preparation from elaiophylin and total synthesis from (R)-3-hydroxybutyrate and (S)-malate", in: Liebigs Ann. Chem., 1986, pp. 1281–1308; doi:10.1002/jlac.198619860714.
  11. ^ David A. Evans, Stephen W. Kaldor, Todd K. Jones, Jon Clardy, Thomas J. Stout: "Total synthesis of the macrolide antibiotic cytovaricin", in: J. Am. Chem. Soc., 1990, 112, pp. 7001–7031; doi:10.1021/ja00175a038.
  12. ^ James A. Marshall, Richard Sedrani: "A convergent, highly stereoselective synthesis of a C-11-C-21 subunit of the macbecins", in: J. Org. Chem., 1991, 56, pp. 5496–5498; doi:10.1021/jo00019a004.
  13. ^ a b James D. White, Motoji Kawasaki: "Total synthesis of (+)-latrunculin A", in: J. Am. Chem. Soc., 1990, 112, pp. 4991–4993; doi:10.1021/ja00168a071.
  14. ^ Morris J. Robins, Vicente Samano, Mark D. Johnson: "Nucleic acid-related compounds. 58. Periodinane oxidation, selective primary deprotection, and remarkably stereoselective reduction of tert-butyldimethylsilyl-protected ribonucleosides. Synthesis of 9-(β-D-xylofuranosyl)adenine or 3'-deuterioadenosine from adenosine", in: J. Org. Chem., 1990, 55, pp. 410–412; doi:10.1021/jo00289a004.
  15. ^ R. Roger F. Newton, Derek P. Reynolds, Colin F. Webb, Stanley M. Roberts: "A short and efficient total synthesis of (±) prostaglandin D2 methyl ester involving a new method for the cleavage of a dimethyl-t-butylsilyl ether", in: J. Chem. Soc., Perkin Trans. 1, 1981, pp. 2055–2058; doi:10.1039/P19810002055.
  16. ^ Kyriacos C. Nicolaou, R. A. Daines, T. K. Chakraborty: "Total synthesis of amphoteronolide B", in: J. Am. Chem. Soc., 1987, 109, pp. 2208–2210; doi:10.1021/ja00241a063.
  17. ^ Leo A. Paquette, Annette M. Doherty, Christopher M. Rayner: "Total synthesis of furanocembranolides. 1. Stereocontrolled preparation of key heterocyclic building blocks and assembly of a complete seco-pseudopterane framework", in: J. Am. Chem. Soc., 1991, 109, pp. 3910–3926; doi:10.1021/ja00036a045.
  18. ^ P.J. Kocieński: Protecting Groups, p. 40.
  19. ^ P.J. Kocieński: Protecting Groups, pp. 46–49.
  20. ^ Michel Bessodes, Dimitri Komiotis, Kostas Antonakis: "Rapid and selective detritylation of primary alcohols using formic acid", in: Tetrahedron Lett., 1986, 27, pp. 579–580; doi:10.1016/S0040-4039(00)84045-9.
  21. ^ B. Helferich: Carbonhydr. Chem. Biochem., 1948, 3, pp. 79.
  22. ^ M.L. García, J. Pascual, L. Borràs, J.A. Andreu, E. Fos, D. Mauleón, G. Carganico, F. Arcamone: "Synthesis of new ether glycerophospholipids structurally related to modulator", in: Tetrahedron, 1991, 47, pp. 10023–10034; doi:10.1016/S0040-4020(01)96051-X.
  23. ^ Yuji Oikawa, Tadao Yoshioka, Osamu Yonemitsu: "Specific removal of o-methoxybenzyl protection by DDQ oxidation", in: Tetrahedron Lett., 1982, 23, pp. 885–888; doi:10.1016/S0040-4039(00)86974-9.
  24. ^ See literature for p‑methoxybenzyl.
  25. ^ P.J. Kocieński: Protecting Groups, p. 77.
  26. ^ H. Nagaoka, W. Rutsch, G. Schmidt, H. Ito, M.R. Johnson, Y. Kishi: "Total synthesis of rifamycins. 1. Stereocontrolled synthesis of the aliphatic building block", in: J. Am. Chem. Soc., 1980, 102, pp. 7962–7965; doi:10.1021/ja00547a037.
  27. ^ W. Clark Still, Dominick Mobilio: "Synthesis of asperdiol", in: J. Org. Chem., 1983, 48, pp. 4785–4786; doi:10.1021/jo00172a070.
  28. ^ Masahiro Hirama, Mitsuko Uei: "A chiral total synthesis of compactin", in: J. Am. Chem. Soc., 1982, 104, pp. 4251–4253; doi:10.1021/ja00379a037.
  29. ^ W. Clark Still, Shizuaki Murata, Gilbert Revial, Kazuo Yoshihara: "Synthesis of the cytotoxic germacranolide eucannabinolide", in: J. Am. Chem. Soc., 1983, 105, pp. 625–627; doi:10.1021/ja00341a055.
  30. ^ Kamaya, Yasushi; T Higuchi (2006). "Metabolism of 3,4-dimethoxycinnamyl alcohol and derivatives by Coriolus versicolor". FEMS Microbiology Letters. 24 (2–3): 225–229. doi:10.1111/j.1574-6968.1984.tb01309.x.
  31. ^ Serge David, Annie Thieffry, Alain Veyrières: "A mild procedure for the regiospecific benzylation and allylation of polyhydroxy-compounds via their stannylene derivatives in non-polar solvents", in: J. Chem. Soc., Perkin Trans. 1, 1981, pp. 1796–1801; doi:10.1039/P19810001796.
  32. ^ Kaoru Fuji, Shigetoshi Nakano, Eiichi Fujita: "An Improved Method for Methoxymethylation of Alcohols under Mild Acidic Conditions", in: Synthesis, 1975, pp. 276–277.
  33. ^ Paul A. Wender, Carlos R. D. Correia: "Intramolecular photoinduced diene-diene cyaloadditions: a selective method for the synthesis of complex eight-membered rings and polyquinanes", in: J. Am. Chem. Soc., 1987, 109, pp. 2523–2525; doi:10.1021/ja00242a053.
  34. ^ Karel F. Bernady, M. Brawner Floyd, John F. Poletto, Martin J. Weiss: "Prostaglandins and congeners. 20. Synthesis of prostaglandins via conjugate addition of lithium trans-1-alkenyltrialkylalanate reagents. A novel reagent for conjugate 1,4-additions", in: J. Org. Chem., 1979, 44, pp. 1438–1447; doi:10.1021/jo01323a017.
  35. ^ Elias J. Corey, Haruki Niwa, Jochen Knolle: "Total synthesis of (S)-12-hydroxy-5,8,14-cis,-10-trans-eicosatetraenoic acid (Samuelsson's HETE)", in: J. Am. Chem. Soc., 1978, 100, pp. 1942–1943; doi:10.1021/ja00474a058.
  36. ^ Elias J. Corey, Mark G. Bock: "Protection of primary hydroxyl groups as methylthiomethyl ethers", in: Tetrahedron Lett., 1975, 16, pp. 3269–3270; doi:10.1016/S0040-4039(00)91422-9.
  37. ^ Elias J. Corey, Duy H. Hua, Bai Chuan Pan, Steven P. Seitz: "Total synthesis of aplasmomycin", in: J. Am. Chem. Soc., 1982, 104, pp. 6818–6820; doi:10.1021/ja00388a074.
  38. ^ Robert C. Gadwood, Renee M. Lett, Jane E. Wissinger: "Total synthesis of (±)-poitediol and (±)4-epipoitediol", in: J. Am. Chem. Soc., 1984, 106, pp. 3869–3870; doi:10.1021/ja00325a032.
  39. ^ Steven D. Burke, Gregory J. Pacofsky: "The ester enolate claisen rearrangement. Total synthesis of (±)-ethisolide", in: Tetrahedron Lett., 1986, 27, pp. 445–448; doi:10.1016/S0040-4039(00)85501-X.
  40. ^ Toshiyuki Kan, Masaru Hashimoto, Mitsutoshi Yanagiya, Haruhisa Shirahama: "Effective deprotection of 2-(trimethylsilylethoxy)methylated alcohols (SEM ethers). Synthesis of thyrsiferyl-23 acetate", in: Tetrahedron Lett., 1988, 29, pp. 5417–5418; doi:10.1016/S0040-4039(00)82883-X.
  41. ^ Joseph P. Marino, Scott L. Dax: "An efficient desilylation method for the generation of o-quinone methides: application to the synthesis of (+)- and (−)-hexahydrocannabinol", in: J. Org. Chem., 1984, 49, pp. 3671–3672; doi:10.1021/jo00193a051.
  42. ^ P.J. Kocieński: Protecting Groups, pp. 59–60.
  43. ^ P.J. Kocieński: Protecting Groups, p. 62.
  44. ^ R.E. Ireland, D.W. Norbeck: "Convergent synthesis of polyether ionophore antibiotics: the synthesis of the monensin bis(tetrahydrofuran) via the Claisen rearrangement of an ester enolate with a β-leaving group", in: J. Am. Chem. Soc., 1985, 107, pp. 3279–3285; doi:10.1021/ja00297a038.
  45. ^ András Lipták, János Imre, János Harangi, Pál Nánási, András Neszmélyi: "Chemo-, stereo- and regioselective hydrogenolysis of carbohydrate benzylidene acetals. Synthesis of benzyl ethers of benzyl α-D-, methyl β-D-mannopyranosides and benzyl α-D-rhamnopyranoside by ring cleavage of benzylidene derivatives with the LiAlH4-AlCl3 reagent", in: Tetrahedron, 1982, 38, pp. 3721–3727; doi:10.1016/0040-4020(82)80083-5.
  46. ^ James A. Marshall, Joseph D. Trometer, Bruce E. Blough, Thomas D. Crute: "Stereochemistry of SN2' additions to acyclic vinyloxiranes", in J. Org. Chem., 1988, 53, pp. 4274–4282 doi:10.1021/jo00253a020.
  47. ^ P.J. Kocieński: Protecting Groups, p. 186.
  48. ^ John O. Osby, Michael G. Martin, Bruce Ganem: An Exceptionally Mild Deprotection of Phthalimides, in: Tetrahedron Lett., 1984, 25, pp. 2093–2096; doi:10.1016/S0040-4039(01)81169-2.
  49. ^ P.J. Kocieński: Protecting Groups, pp. 220–227.
  50. ^ P.J. Kocieński: Protecting Groups, p. 195.
  51. ^ Robert M. Williams, Peter J. Sinclair, Dongguan Zhai, Daimo Chen: "Practical asymmetric syntheses of α-amino acids through carbon-carbon bond constructions on electrophilic glycine templates", in: J. Am. Chem. Soc., 1988, 110, p. 1547–1557; doi:10.1021/ja00213a031.
  52. ^ Glenn L. Stahl, Roderich Walter, Clarck W. Smith: "General procedure for the synthesis of mono-N-acylated 1,6-diaminohexanes", in: J. Org. Chem., 1978, 43, pp. 2285–2286; doi:10.1021/jo00405a045.
  53. ^ Naomi Sakai, Yasufumi Ohfune: "Total synthesis of galantin I. Acid-catalyzed cyclization of galantinic acid", in: J. Am. Chem. Soc., 1992, 114, pp. 998–1010; doi:10.1021/ja00029a031.
  54. ^ Weng C. Chan, Peter D. White: Fmoc Solid Phase Peptide Synthesis, pp. 27–30.
  55. ^ Gregg B. Fields: Methods for Removing the Fmoc Group. (PDF; 663 kB) In: Michael W. Pennington, Ben M. Dunn (eds.): Peptide Synthesis Protocols volume 35, 1995, ISBN 978-0-89603-273-6, pp. 17–27.
  56. ^ B. Liebe, H. Kunz: Festphasensynthese eines tumorassoziierten Sialyl-Tn-Antigen-Glycopeptids mit einer Partialsequenz aus dem "Tandem Repeat" des MUC-1-Mucins In: Angew. Chem. volume 109, 1997, pp. 629–631 (in German).
  57. ^ ChemPep Inc.: Fmoc Solid Phase Peptide Synthesis. retrieved 16 November 2013.
  58. ^ P.J. Kocieński: Protecting Groups, pp. 199–201.
  59. ^ Moussa, Ziad; D. Romo (2006). "Mild deprotection of primary N-(p-toluenesufonyl) amides with SmI2 following trifluoroacetylation". Synlett. 2006 (19): 3294–3298. doi:10.1055/s-2006-951530.
  60. ^ T. Tsunoda, M. Suzuki, R. Noyori: "A facile procedure for acetalization under aprotic conditions", in: Tetrahedron Lett., 1980, 21, pp. 1357–1358; doi:10.1016/S0040-4039(00)74575-8.
  61. ^ Juji Yoshimura, Shigeomi Horito, Hiroriobu Hashimoto: "Facile Synthesis of 2,3,4,6-Tetra-O-benzyl-D-glucopyranosylidene Acetals Using Trimethylsilyl Trifluoromethanesulfonate Catalyst", in: Chem. Lett., 1981, 10, pp. 375–376; doi:10.1246/cl.1981.375.
  62. ^ Bruce H. Lipshutz, Daniel Pollart, Joseph Monforte, Hiyoshizo Kotsuki: "Pd(II)-catalyzed acetal/ketal hydrolysis/exchange reactions", in: Tetrahedron Lett., 1985, 26, pp. 705–708; doi:10.1016/S0040-4039(00)89114-5.
  63. ^ Kwan Soo Kim, Yang Heon Song, Bong Ho Lee, Chi Sun Hahn: "Efficient and selective cleavage of acetals and ketals using ferric chloride adsorbed on silica gel", in: J. Org. Chem., 1986, 51, pp. 404–407; doi:10.1021/jo00353a027.
  64. ^ P.J. Kocieński: Protecting Groups, S. 167–170.
  65. ^ P.J. Kocieński: Protecting Groups, pp. 176.
  66. ^ P.J. Kocieński: Protecting Groups, pp. 178–180.
  67. ^ Samuel J. Danishefsky, Nathan B. Mantlo, Dennis S. Yamashita, Gayle. Schulte: "Concise route to the calichemicin-esperamicin series: the crystal structure of an aglycone prototype", in: J. Am. Chem. Soc., 1988, 110, pp. 6890–6891; doi:10.1021/ja00228a051.
  68. ^ John N. Haseltine, Maria Paz Cabal, Nathan B. Mantlo, Nobuharu Iwasawa, Dennis S. Yamashita, Robert S. Coleman, Samuel J. Danishefsky, Gayle K. Schulte: "Total synthesis of calicheamicinone: new arrangements for actuation of the reductive cycloaromatization of aglycon congeners", in: J. Am. Chem. Soc., 1991, 113, pp. 3850–3866; doi:10.1021/ja00010a030.
  69. ^ P.J. Kocieński: Protecting Groups, pp. 119.
  70. ^ Peter Mohr, Nada Waespe-Šarčević, Christoph Tamm, Krystyna Gawronska, Jacek K. Gawronski: "A Study of Stereoselective Hydrolysis of Symmetrical Diesters with Pig Liver Esterase", in: Helv. Chim. Acta, 1983, 66, pp. 2501–2511; doi:10.1002/hlca.19830660815.
  71. ^ Théophile Tschamber, Nada Waespe-Šarčević, Christoph Tamm: "Stereocontrolled Synthesis of an Epimer of the C(19)-to-C(27) Segment of Rifamycin S", in: Helv. Chim. Acta, 1986, 69, pp. 621–625; doi:10.1002/hlca.19860690311.
  72. ^ Yves Rubin, Carolyn B. Knobler, Francois Diederich: "Precursors to the cyclo[n]carbons: from 3,4-dialkynyl-3-cyclobutene-1,2-diones and 3,4-dialkynyl-3-cyclobutene-1,2-diols to cyclobutenodehydroannulenes and higher oxides of carbon", in: J. Am. Chem. Soc., 1990, 112, pp. 1607–1617; doi:10.1021/ja00160a047.
  73. ^ Sunggak Kim, Yong Gil Kim, Deog-il Kim: "A novel method for selective dioxolanation of ketones in the presence of aldehydes", in: Tetrahedron Lett., 1992, 33, pp. 2565–2566; doi:10.1016/S0040-4039(00)92243-3.
  74. ^ G. Bauduin, D. Bondon, Y. Pietrasanta, B. Pucci: "Reactions de transcetalisation – II: Influence des facteurs steriques et electroniques sur les energies de cetalisation", in: Tetrahedron, 1978, 34, pp. 3269–3274; doi:10.1016/0040-4020(78)80243-9.
  75. ^ John E. McMurry, Stephen J. Isser: "Total synthesis of longifolene", in: J. Am. Chem. Soc., 1972, 94, pp. 7132–7137; doi:10.1021/ja00775a044.
  76. ^ M.P. Bosch, M. Pilar Bosch, Francisco Camps, Jose Coll, Angel Guerrero, Toshio Tatsuoka, Jerrold Meinwald: "A stereoselective total synthesis of (±)-muzigadial", in: J. Org. Chem., 1986, 51, pp. 773–784; doi:10.1021/jo00356a002.
  77. ^ F. Zymalkokowski: Katalytische Hydrierung, Ferdinand Enke Verlag, Stuttgart 1965, pp. 127–133.
  78. ^ P.J. Kocieński: Protecting Groups, pp. 136.
  79. ^ Ulrich Schmidt, Thomas Beuttler, Albrecht Lieberknecht, Helmut Griesser: "Aminosäuren und peptide – XXXXII. Synthese von Chlamydocin + epi-Chlamydocin", in: Tetrahedron Lett., 1983, 24, pp. 3573–3576; doi:10.1016/S0040-4039(00)88171-X (in German).
  80. ^ Elias J. Corey, Plato A. Magriotis: "Total synthesis and absolute configuration of 7,20-diisocyanoadociane", in: J. Am. Chem. Soc., 1987, 109, pp. 287–289; doi:10.1021/ja00235a052.
  81. ^ Elias J. Corey, Kyriacos C. Nicolaou, Takeshi Toru: "Total synthesis of (±)-vermiculine", in: J. Am. Chem. Soc., 1975, 97, pp. 2287–2288; doi:10.1021/ja00841a058.
  82. ^ Tainejiro Hiyama, Akihiro Kanakura, Hajime Yamamoto, Hitosi Nozaki: "General Route to α,β-unsaturated Aldehydes of Homoterpenoid and terpenoid Structure. Sythesis of JH-II and β-Sinensal", in: Tetrahedron Lett., 1978, 19, pp. 3051–3054; doi:10.1016/S0040-4039(01)94936-6.
  83. ^ F. Huet, A. Lechevallier, M. Pellet, J.M. Conia: "Wet Silica Gel; A Convenient Reagent for Deacetalization", in: Synthesis, 1978, pp. 63–64.
  84. ^ Romanski, J.; Nowak, P.; Kosinski, K.; Jurczak, J. (September 2012). "High-pressure transesterification of sterically hindered esters". Tetrahedron Lett. 53 (39): 5287–5289. doi:10.1016/j.tetlet.2012.07.094.
  85. ^ P.J. Kocieński: Protecting Groups, pp. 139–142.
  86. ^ Ahmed M. Tafesh, Jens Weiguny: "A Review of the Selective Catalytic Reduction of Aromatic Nitro Compounds into Aromatic Amines, Isocyanates, Carbamates, and Ureas Using CO", in: Chem. Rev., 1996, 96, pp. 2035–2052; doi:10.1021/cr950083f.
  87. ^ Evan L. Allred, Boyd R. Beck, Kent J. Voorhees: "Formation of carbon-carbon double bonds by the reaction of vicinal dihalides with sodium in ammonia", in: J. Org. Chem., 1974, 39, pp. 1426–1427; doi:10.1021/jo00926a024.
  88. ^ Timothy S. Butcher, Feng Zhou, Michael R. Detty: "Debrominations of vic-Dibromides with Diorganotellurides. 1. Stereoselectivity, Relative Rates, and Mechanistic Implications", in: J. Org. Chem., 1998, 63, pp. 169–176; doi:10.1021/jo9713363.
  89. ^ C. J. Li, David N. Harpp: "Bis(triphenylstanyl)telluride a mild and selective reagent for telluration and debromination", in: Tetrahedron Lett., 1990, 31, pp. 6291–6293; doi:10.1016/S0040-4039(00)97045-X.
  90. ^ Corrado Malanga, Serena Mannucci, Luciano Lardicci: "Carbon-halogen bond activation by nickel catalyst: Synthesis of alkenes, from 1,2-dihalides", in: Tetrahedron, 1998, 54, pp. 1021–1028; doi:10.1016/S0040-4020(97)10203-4.
  91. ^ Byung Woo Yoo, Seo Hee Kim, Jun Ho Kim: "A Mild, Efficient, and Selective Debromination of vic-Dibromides to Alkenes with Cp2TiCl2/Ga System", in: Bull. Korean Chem. Soc., 2010, 31, pp. 2757–2758; doi:10.5012/bkcs.2010.31.10.2757.
  92. ^ Antonius J. H. Klunder, Jie Zhu, Binne Zwanenburg: "The Concept of Transient Chirality in the Stereoselective Synthesis of Functionalized Cycloalkenes Applying the Retro-Diels-Alder Methodology", in: Chem. Rev., 1999, 99, pp. 1163–1190; doi:10.1021/cr9803840.
  93. ^ Hideyuki Tanaka, Takashi Kamikubo, Naoyuki Yoshida, Hideki Sakagami, Takahiko Taniguchi, Kunio Ogasawara: "Enantio- and Diastereocontrolled Synthesis of (−)-Iridolactone and (+)-Pedicularis-lactone", in: Org. Lett., 2001, 3, pp. 679–681; doi:10.1021/ol0070029.
  94. ^ Martin Banwell, David Hockless, Bevyn Jarrott, Brian Kelly, Andrew Knill, Robert Longmore, Gregory Simpson: "Chemoenzymatic approaches to the decahydro-as-indacene cores associated with the spinosyn class of insecticide", in: J. Chem. Soc., Perkin Trans. 1, 2000, pp. 3555–3558; doi:10.1039/b006759h.
  95. ^ Clayden, Jonathan; Greeves, Nick; Warren, Stuart; Wothers, Peter (2000). Organic Chemistry. Oxford University Press. pp. 1291. ISBN 978-0-19-850346-0.
  96. ^ Wenzel E. Davidsohn, Malcolm C. Henry: "Organometallic Acetylenes of the Main Groups III–V", in: Chem. Rev., 1967, 67, pp. 73–106; doi:10.1021/cr60245a003.
  97. ^ Barry J. Teobald: "The Nicholas reaction: the use of dicobalt hexacarbonyl-stabilised propargylic cations in synthesis", in: Tetrahedron, 2002, 58, pp. 4133–4170; doi:10.1016/S0040-4020(02)00315-0.
  98. ^ Kenneth M. Nicholas, R. Pettit: "An alkyne protection group", in: Tetrahedron Lett., 1971, 37, pp. 3475–3478; doi:10.1016/S0040-4039(01)97209-0.
  99. ^ Richard E. Connor, Kenneth M. Nicholas: "Isolation, characterization, and stability of α-[(ethynyl)dicobalt hexacarbonyl] carbonium ions", in: J. Organomet. Chem., 1977, 125, C45–C48; doi:10.1016/S0022-328X(00)89454-1.
  100. ^ Rosa F. Lockwood, Kenneth M. Nicholas: "Transition metal-stabilized carbenium ions as synthetic intermediates. I. α-[(alkynyl)dicobalt hexacarbonyl] carbenium ions as propargylating agents", in: Tetrahedron Lett., 1977, pp. 4163–4166; doi:10.1016/S0040-4039(01)83455-9.
  101. ^ K.M. Nicholas, R. Pettit: "On the stability of α-(alkynyl)dicobalt hexacarbonyl carbonium ions", in: J. Organomet. Chem., 1972, 44, C21–C24; doi:10.1016/0022-328X(72)80037-8.
  102. ^ Blanc, Aurélien; Bochet, Christian G. (2007). "Isotope Effects in Photochemistry: Application to Chromatic Orthogonality" (PDF). Org. Lett. 9 (14): 2649–2651. doi:10.1021/ol070820h. PMID 17555322.
  103. ^ Baran, Phil S.; Maimone, Thomas J.; Richter, Jeremy M. (22 March 2007). "Total synthesis of marine natural products without using protecting groups". Nature. 446 (7134): 404–408. Bibcode:2007Natur.446..404B. doi:10.1038/nature05569. PMID 17377577. S2CID 4357378.
  104. ^ Synthetic studies of marine alkaloids hapalindoles. Part I Total synthesis of (±)-hapalindoles J and M Tetrahedron, Volume 46, Issue 18, 1990, Pages 6331–6342 Hideaki Muratake and Mitsutaka Natsume doi:10.1016/S0040-4020(01)96005-3
  105. ^ Synthetic studies of marine alkaloids hapalindoles. Part 2. Lithium aluminum hydride reduction of the electron-rich carbon-carbon double bond conjugated with the indole nucleus Tetrahedron, Volume 46, Issue 18, 1990, Pages 6343–6350 Hideaki Muratake and Mitsutaka Natsume doi:10.1016/S0040-4020(01)96006-5
  106. ^ Synthetic studies of marine alkaloids hapalindoles. Part 3 Total synthesis of (±)-hapalindoles H and U Tetrahedron, Volume 46, Issue 18, 1990, Pages 6351–6360 Hideaki Muratake, Harumi Kumagami and Mitsutaka Natsume doi:10.1016/S0040-4020(01)96007-7
  107. ^ T. Reichstein, A. Grüssner: "Eine ergiebige Synthese der L-Ascorbinsäure (C-Vitamin)", in: Helv. Chim. Acta, 1934, 17, pp. 311–328; doi:10.1002/hlca.19340170136.
  108. ^ K.C. Nicolaou, E.J. Sorensen: Classics in Total Synthesis: Targets, Strategies, Methods, VCH Verlagsgesellschaft mbH, Weinheim 1996, pp. 711–729, ISBN 3-527-29284-5.
  109. ^ Peter G.M. Wuts, Theodora W. Greene: Green's Protective Groups in Organic Synthesis, 4th Ed., John Wiley & Sons Inc., Hoboken, New Jersey, pp. 10–13; ISBN 0-471-69754-0.
  110. ^ J.M. McClure, Samuel J. Danishefsky: "A novel Heck arylation reaction: rapid access to congeners of FR 900482", in: J. Am. Chem. Soc., 1993, 115, pp. 6094–6100; doi:10.1021/ja00067a026.
  111. ^ Merrifield, R. B.; Barany, G.; Cosand, W. L.; Engelhard, M.; Mojsov, S. (1977). "Proceedings of the 5th American Peptide Symposium". Biochemical Education. 7 (4): 93–94. doi:10.1016/0307-4412(79)90078-5.
  112. ^ Weng C. Chan, Peter D. White: Fmoc Solid Phase Peptide Synthesis. Reprint 2004, Oxford University Press, ISBN 0-19-963724-5.
  113. ^ Serge L. Beaucage, Radhakrishman P. Iyer: "Advances in the Synthesis of Oligonucleotides by the Phosphoramidite Approach", in: Tetrahedron, 1992, 48, pp. 2223–2311; doi:10.1016/S0040-4020(01)88752-4.

Further reading[edit]

External links[edit]