Enders SAMP/RAMP hydrazone-alkylation reaction

Last updated

The Enders SAMP/RAMP hydrazone alkylation reaction is an asymmetric carbon-carbon bond formation reaction facilitated by pyrrolidine chiral auxiliaries. It was pioneered by E. J. Corey and Dieter Enders in 1976, [1] and was further developed by Enders and his group. [2] This method is usually a three-step sequence. The first step is to form the hydrazone between (S)-1-amino-2-methoxymethylpyrrolidine (SAMP) or (R)-1-amino-2-methoxymethylpyrrolidine (RAMP) and a ketone or aldehyde. Afterwards, the hydrazone is deprotonated by lithium diisopropylamide (LDA) to form an azaenolate, which reacts with alkyl halides or other suitable electrophiles to give alkylated hydrazone species with the simultaneous generation of a new chiral center. Finally, the alkylated ketone or aldehyde can be regenerated by ozonolysis or hydrolysis. [3]

Contents

Enders' SAMP/RAMP Hydrazone Alkylation Reaction EndersReaction.png
Enders' SAMP/RAMP Hydrazone Alkylation Reaction

This reaction is a useful technique for asymmetric α-alkylation of ketones and aldehydes, which are common synthetic intermediates for medicinally interesting natural products and other related organic compounds. These natural products include (-)-C10-demethyl arteannuin B, the structural analog of antimalarial artemisinin, [4] the polypropionate metabolite (-)-denticulatin A and B isolated from Siphonaria denticulata , [5] zaragozic acid A, a potent inhibitor of sterol synthesis, [6] and epothilone A and B, which have been proven to be very effective anticancer drugs. [7]

History

Regioselective and stereoselective formation of carbon-carbon bonds adjacent to carbonyl group is an important procedure in organic chemistry. Alkylation reaction of enolates has been the main focus of the field. Both A. G. Myers and D. A. Evans developed asymmetric alkylation reactions for enolates. [8] [9]

Myer's and Evans' Asymmetric Alkylation Alkylation.png
Myer's and Evans' Asymmetric Alkylation

The apparent shortcoming for enolate alkylation reactions is over-alkylation, even if the amount of base added for enolization as well as the reaction temperature are carefully controlled. The ketene formation during the deprotonation process for substrates possessing Evans' oxazolidinone is also a main side reaction for the related alkylation reactions. Development in the field of enamine chemistry and the utilization of imine derivatives of enolates managed to provide an alternative for enolate alkylation reactions.

In 1963, G. Stork reported the first enamine alkylation reaction for ketones - Stork enamine alkylation reaction. [10]

A specific Stork enamine alkylation reaction between cyclohexanone and an electrophile. History Stork.svg
A specific Stork enamine alkylation reaction between cyclohexanone and an electrophile.

In 1976, Meyers reported the first alkylation reaction of metallated azaenolates of hydrazones with an acyclic amino acid-based auxiliary. Compared with the free carbonyl compounds and the chiral enamine species reported previously, the hydrazones exhibit higher reactivity, regioselectivity and stereoselectivity. [11]

Hydrazone alkylation developed by Meyer. Meyer.svg
Hydrazone alkylation developed by Meyer.

The combination of cyclic amino acid derivatives (SAMP and RAMP) and the powerful hydrazone techniques were pioneered by E. J. Corey and D. Enders in 1976, and were independently developed by D. Enders later. Both SAMP and RAMP are synthesized from amino acids. The detailed synthesis of these two auxiliaries are shown below. [12] [13]

EndersSAMP 2 EndersSamp.png
EndersSAMP 2

Mechanism

The Enders SAMP/RAMP hydrazone alkylation begins with the synthesis of the hydrazone from a N,N-dialkylhydrazine and a ketone or aldehyde [14]

The hydrazone is then deprotonated on the α-carbon position by a strong base, such as lithium diisopropylamide (LDA), leading to the formation of a resonance stabilized anion - an azaenolate. This anion is a very good nucleophile and readily attacks electrophiles, such as alkyl halides, to generate alkylated hydrazones with simultaneous creation of a new chiral center at the α-carbon.

Mechanism of hydrazone alkylation Enders mechanism.png
Mechanism of hydrazone alkylation

The stereochemistry of this reaction is discussed in detail in next section.

Stereochemistry

Stereochemistry of the azaenolate

After the deprotonation, the hydrazone turns into an azaenolate with lithium cation chelating both the nitrogen and oxygen. There are two possible options for lithium chelation. One is that lithium is antiperiplanar to the C=C bond (blue colored), leading to the conformation of ZC-N; the other one is that lithium and the C=C bond are at the same side of the C-N bond (red colored), leading to the EC-N conformer. There are also two available orientations for the chelating nitrogen and R2 group, being either EC=C or ZC=C. This leads to four possible azaenolate intermediates (A, B, C and D) for the Enders' SAMP/RAMP hydrazone alkylation reaction.

Stereoselectivity of the generation of the azaenolates Azaenolate.png
Stereoselectivity of the generation of the azaenolates

Experiments and calculations [2] [15] [16] show that one specific stereoisomer of the azaenolate is favored over the other three possible candidates. Therefore, although four isomers are possible for the azaenolate, only the one (azaenolate A) with the stereochemistry of its C=C double bonds being E and that of its C-N bond being Z stereochemistry is dominant (EC=CZC-N) for both cyclic and acyclic ketones. [17]

Stereochemistry of alkylation

The favored azaenolate is the dominant starting molecule for the subsequent alkylation reaction. There are two possible faces of accessing for any electrophile to react with. The steric interaction between the pyrrolidine ring and the electrophilic reagent hinders the attack of the electrophile from the top face. On the contrary, when the electrophile attacks from the bottom face, such unfavorable interaction does not exist. Therefore, the electrophilic attack proceeds from the sterically more accessible face. [18]

Stereoselectivity for the alkylation step of the Enders' reaction EndersAlkylation.png
Stereoselectivity for the alkylation step of the Enders' reaction

Variants

The chelation of lithium cation with the methoxy group is one of the most important features of the transition state for Enders' hydrazone alkylation reaction. It is necessary to have this chelation effect to achieve high stereoselectivity. The development and modification of Enders' hydrazone alkylation reaction mainly focus on the addition of more steric hindrance on the pyrrolidine rings of both SAMP and RAMP, while preserving the methoxy group for lithium chelation.

The most famous four variants of SAMP and RAMP are SADP, SAEP, SAPP and RAMBO, [19] [20] whose structures are shown below.

Variants of SAMP/RAMP with bulkier groups attached Variants.png
Variants of SAMP/RAMP with bulkier groups attached

In 2011, several N-amino cyclic carbamates were synthesized and studied for asymmetric hydrazone alkylation reactions. [21] Both the stereochemistry and regioselectivity of the reactions turned out to be very promising. These new compounds consist of a new class of chiral auxiliary based on the carbamate structure and, therefore, no longer belong to the family of SAMP and RAMP. But they do provide very powerful alternatives to the traditional pyrrolidine systems.

N-Amino cyclic carbamates N-aminocarbamates.png
N-Amino cyclic carbamates

Auxiliary release

Hydrazones are usually very stable towards solvolysis, and conversion to the ketone can require vigorous conditions. Also, aldehydic hydrazones often instead disproportionate to a nitrile and amine. [22]

Two principal workup environments are common: oxidation and solvolysis. Reductive conversions are possible with low-valent transition metals, but remained relatively unstudied As of 2000. [22]

Oxidative cleavage has high yields and is most frequently used. Ozone or singlet oxygen can ozonolyze the diazene bond (and any olefinic moieties present), leaving a carbonyl, a nitrosamine, and dioxygen. Lemieux's gentler oxidation tolerates acetals and benzyl ethers. Peroxide reagents (e.g. NaBO3, (tBu4NSO4)2, or m-ClBzO2H) cleave the hydrazone with varying speeds, selectivities, and mechanisms, but the Baeyer-Villiger oxidation is a common side-reaction. High-valent transition metal oxyhalides (e.g. WF6, CoF3, MoOCl3) appear to primarily cleave via radicals. All except ozone and singlet oxygen generate nitriles from aldehydic hydrazones, either as the major or a substantial minor product. [22]

Certain electrophiles also elicit nitriles: chloroformates, strongly-activated alkynes, or methyl iodide and a hindered base. Methyl iodide is also useful for hydrolysis: the alkylated hydrazonium iodide easily hydrolyzes to a carbonyl and hydrazoform, and air cleaves the hydrazoform to the hydrazine and carbon dioxide. [22]

Indeed, a wide variety of acids promote hydrolysis. Bismuth trichloride cleaves arbitrary hydrazones in a microwave. Oxalic acid abstracts hydrazine from ketonic hydrazones; the oxalate adduct then decomposes to the original auxiliary in aqueous base. Silica gel hydrolyzes exquisitely acid-sensitive substrates, but is too weak to affect ketonic hydrazones adjacent to a primary carbon. Ketonic hydrazones adjacent to a secondary or tertiary carbon hydrolyze in the presence of catalytic cupric salts; that procedure also preserves substrates disturbed by oxidants or strong acids. [22]

Boron trifluoride etherate catalyzes thioketalization, and Baker's yeast will hydrolyze non-bioactive substrates. [22]

Hydrazone carbamates are cleaved much more readily than their parent hydrazones: para-toluenesulfonic acid affords the corresponding ketones in near-quantitative yields. [21]

Carbamate-cleavage.png

Conditions

Ender's hydrazone alkylation reaction is usually run through a sequence of three steps. [14] The first step should always be the synthesis of the hydrazones. The ketone or aldehyde is mixed with either SAMP or RAMP and allowed to react under argon for 12 hours. The crude hydrazone obtained is purified by distillation or recrystallization. At 0 degree celsius, the hydrazone is transferred into the ether solution of lithium diisopropylamide. Then this mixture is cooled down to -110 degree celsius and is slowly added the alkyl halide. This mixture is then allowed to warm up to room temperature. After 12 hours of reaction at room temperature, the crude alkylated hydrazone is allowed to react with ozone in a Schlenk tube to cleave the C=N bond. After distillation or column chromatography, pure alkylation product can be obtained.

Applications

Synthesis of zaragozic acid A

K. C. Nicolaou and coworkers at Scripps Research Institute generated the chiral hydrazone through Enders' hydrazone alkylation reaction with high stereoselectivity (de > 95%). The subsequent ozonolysis and Wittig reaction led to the side chain fragment of zaragozic acid A, which is a potent medicine for coronary heart disease. [6]

Synthesis of Zaragozic Acid Enders Zaragozic acid.png
Synthesis of Zaragozic Acid

Synthesis of denticulatin A and B

Ziegler and coworkers reacted an allyl iodide with the azaenolate to generate a chiral hydrocarbon chain. To avoid loss of the enantiomeric purity of the product, the authors used cupric acetate to regenerate the carbonyl group, obtaining only moderate yield for the cleavage of C=N bond but good enantioselectivity (ee = 89%). The ketone was transformed after several steps into denticulatin A and B - polypropionate metabolites isolated from Siphonaria Denticulata. [5]

Synthesis of Denticulatin A&B Denticulatin.png
Synthesis of Denticulatin A&B

Synthesis of the derivative of arteannuin

(-)-C10-demethyl arteannuin B is a structural analog of the antimalarial artemisinin. It exhibits potent antimalarial activity even against a drug-resistant strain. Little and coworkers obtained the alkylated hydrazone in diastereomerically pure form (de > 95%) through the Enders' alkylation reaction. This intermediate was then elaborated into (-)-C10-demethyl arteannuin B. [4]

Synthesis of Arteannuin B Arteannuin.png
Synthesis of Arteannuin B

Synthesis of epothilone A

Epothilone A and B are reported to be highly effective anticancer drugs. Several of their structural derivatives show very promising inhibition against breast cancer with only mild side effect and some of them are now under trials. In 1997, K. C. Nicolaou and coworkers reported the first total synthesis of both Epothilone A and B. Ender's alkylation reaction was utilized at the very beginning of the synthesis to install the stereogenic center at C8. The reaction proceeded with both high yield and high diastereoselectivity. [7]

Synthesis of Epothilone A Epothilone.png
Synthesis of Epothilone A

See also

Related Research Articles

<span class="mw-page-title-main">Hydrazone</span> Organic compounds - Hydrazones

Hydrazones are a class of organic compounds with the structure R1R2C=N−NH2. They are related to ketones and aldehydes by the replacement of the oxygen =O with the =N−NH2 functional group. They are formed usually by the action of hydrazine on ketones or aldehydes.

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

The aldol reaction is a reaction in organic chemistry that combines two carbonyl compounds to form a new β-hydroxy carbonyl compound. Its simplest form might involve the nucleophilic addition of an enolized ketone to another:

<span class="mw-page-title-main">Enamine</span> Class of chemical compounds

An enamine is an unsaturated compound derived by the condensation of an aldehyde or ketone with a secondary amine. Enamines are versatile intermediates.

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

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

<span class="mw-page-title-main">Michael addition reaction</span> Reaction in organic chemistry

In organic chemistry, the Michael reaction or Michael 1,4 addition is a reaction between a Michael donor and a Michael acceptor to produce a Michael adduct by creating a carbon-carbon bond at the acceptor's β-carbon. It belongs to the larger class of conjugate additions and is widely used for the mild formation of carbon-carbon bonds.

<span class="mw-page-title-main">Enolate</span> Organic anion formed by deprotonating a carbonyl (>C=O) compound

In organic chemistry, enolates are organic anions derived from the deprotonation of carbonyl compounds. Rarely isolated, they are widely used as reagents in the synthesis of organic compounds.

In organic chemistry, the Mannich reaction is a three-component organic reaction that involves the amino alkylation of an acidic proton next to a carbonyl functional group by formaldehyde and a primary or secondary amine or ammonia. The final product is a β-amino-carbonyl compound also known as a Mannich base. Reactions between aldimines and α-methylene carbonyls are also considered Mannich reactions because these imines form between amines and aldehydes. The reaction is named after Carl Mannich.

<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">Nucleophilic conjugate addition</span> Organic reaction

Nucleophilic conjugate addition is a type of organic reaction. Ordinary nucleophilic additions or 1,2-nucleophilic additions deal mostly with additions to carbonyl compounds. Simple alkene compounds do not show 1,2 reactivity due to lack of polarity, unless the alkene is activated with special substituents. With α,β-unsaturated carbonyl compounds such as cyclohexenone it can be deduced from resonance structures that the β position is an electrophilic site which can react with a nucleophile. The negative charge in these structures is stored as an alkoxide anion. Such a nucleophilic addition is called a nucleophilic conjugate addition or 1,4-nucleophilic addition. The most important active alkenes are the aforementioned conjugated carbonyls and acrylonitriles.

The Weinreb ketone synthesis or Weinreb–Nahm ketone synthesis is a chemical reaction used in organic chemistry to make carbon–carbon bonds. It was discovered in 1981 by Steven M. Weinreb and Steven Nahm as a method to synthesize ketones. The original reaction involved two subsequent nucleophilic acyl substitutions: the conversion of an acid chloride with N,O-Dimethylhydroxylamine, to form a Weinreb–Nahm amide, and subsequent treatment of this species with an organometallic reagent such as a Grignard reagent or organolithium reagent. Nahm and Weinreb also reported the synthesis of aldehydes by reduction of the amide with an excess of lithium aluminum hydride.

<span class="mw-page-title-main">Asymmetric induction</span> Preferential formation of one chiral isomer over another in a chemical reaction

Asymmetric induction describes the preferential formation in a chemical reaction of one enantiomer or diastereoisomer over the other as a result of the influence of a chiral feature present in the substrate, reagent, catalyst or environment. Asymmetric induction is a key element in asymmetric synthesis.

In organosilicon chemistry, silyl enol ethers are a class of organic compounds that share the common functional group R3Si−O−CR=CR2, composed of an enolate bonded to a silane through its oxygen end and an ethene group as its carbon end. They are important intermediates in organic synthesis.

<span class="mw-page-title-main">Stork enamine alkylation</span> Reaction sequence in organic chemistry

The Stork enamine alkylation involves the addition of an enamine to a Michael acceptor or another electrophilic alkylation reagent to give an alkylated iminium product, which is hydrolyzed by dilute aqueous acid to give the alkylated ketone or aldehyde. Since enamines are generally produced from ketones or aldehydes, this overall process constitutes a selective monoalkylation of a ketone or aldehyde, a process that may be difficult to achieve directly.

<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">Mukaiyama aldol addition</span> Organic reaction between a silyl enol ether and an aldehyde or formate

In organic chemistry, the Mukaiyama aldol addition is an organic reaction and a type of aldol reaction between a silyl enol ether and an aldehyde or formate. The reaction was discovered by Teruaki Mukaiyama in 1973. His choice of reactants allows for a crossed aldol reaction between an aldehyde and a ketone, or a different aldehyde without self-condensation of the aldehyde. For this reason the reaction is used extensively in organic synthesis.

<span class="mw-page-title-main">Organoindium chemistry</span> Chemistry of compounds with a carbon-indium bond

Organoindium chemistry is the chemistry of compounds containing In-C bonds. The main application of organoindium chemistry is in the preparation of semiconducting components for microelectronic applications. The area is also of some interest in organic synthesis. Most organoindium compounds feature the In(III) oxidation state, akin to its lighter congeners Ga(III) and B(III).

<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.

Electrophilic amination is a chemical process involving the formation of a carbon–nitrogen bond through the reaction of a nucleophilic carbanion with an electrophilic source of nitrogen.

<span class="mw-page-title-main">Carbonyl α-substitution reactions</span> Chemical reaction

Alpha-substitution reactions occur at the position next to the carbonyl group, the α-position, and involve the substitution of an α hydrogen atom by an electrophile, E, through either an enol or enolate ion intermediate.

Proline organocatalysis is the use of proline as an organocatalyst in organic chemistry. This theme is often considered the starting point for the area of organocatalysis, even though early discoveries went unappreciated. Modifications, such as MacMillan’s catalyst and Jorgensen's catalysts, proceed with excellent stereocontrol.

References

  1. Corey, E. J.; Enders, D. (1976). "Applications of N,N-dimethylhydrazones to synthesis. Use in efficient, positionally and stereochemically selective C-C bond formation; oxidative hydrolysis to carbonyl compounds". Tetrahedron Letters . 17 (1): 3–6. doi:10.1016/s0040-4039(00)71307-4.
  2. 1 2 Kurti, L.; Czako, B. (2005). Strategic Applications of Named Reactions in Organic Synthesis. Burlington, MA: Elsevier Academic Press. pp. 150–151. ISBN   0-12-369483-3.
  3. Job, A.; Janeck, C. F.; Bettray, W.; Peters, R.; Enders, D. (2002). "The SAMP-/RAMP-hydrazone methodology in asymmetric synthesis". Tetrahedron . 58 (12): 2253–2329. doi:10.1016/s0040-4020(02)00080-7.
  4. 1 2 Schwaebe, M.; Little, R. D. (1996). "Asymmetric Reductive Cyclization. Total Synthesis of (−)-C10-Desmethyl Arteannuin B". The Journal of Organic Chemistry . 61 (10): 3240–3244. doi:10.1021/jo9600417.
  5. 1 2 Ziegler, F. E.; Becker, M. R. (1990). "Total synthesis of (-)-denticulatins A and B: marine polypropionates from Siphonaria denticulata". The Journal of Organic Chemistry . 55 (2): 2800–2805. doi:10.1021/jo00296a044.
  6. 1 2 Nadin, A.; Nicolaou, K. C. (1996). "Chemistry and Biology of the Zaragozic Acids (Squalestatins)". Angewandte Chemie International Edition in English . 35 (15): 1622–1656. doi:10.1002/anie.199616221.
  7. 1 2 Nicolaou, K. C.; Ninkovic, S.; Sarabia, F.; Vourloumis, D.; He, Y.; Vallberg, H.; Finlay, M. R. V.; Yang, Z. (1997). "Total Syntheses of Epothilones A and B via a Macrolactonization-Based Strategy". Journal of the American Chemical Society . 119 (34): 7974–7991. doi:10.1021/ja971110h.
  8. Myers, A. G., Yang, B. H., Chen, H., McKinstry, L. Kopecky, D. J., Gleason, J. L. (1997). "Pseudoephedrine as a Practical Chiral Auxiliary for the Synthesis of Highly Enantiomerically Enriched Carboxylic Acids, Alcohols, Aldehydes, and Ketones". Journal of the American Chemical Society . 119 (28): 6496–6511. doi:10.1021/ja970402f.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  9. Evans, D. A.; Ennis, M. D.; Mathre, D. J. (1982). "Asymmetric alkylation reactions of chiral imide enolates. A practical approach to the enantioselective synthesis of α-substituted carboxylic acid derivatives". Journal of the American Chemical Society . 104 (6): 1737–1739. doi:10.1021/ja00370a050.
  10. Stork, G.; Dowd, S. R. (1996). "A New Method for the Alkylation of Ketones and Aldehydes: the C-Alkylation of the Magnesium Salts of N-Substituted Imines". Journal of the American Chemical Society . 85 (14): 2178–2180. doi:10.1021/ja00897a040.
  11. Meyers, A. I.; Williams, D. R.; Druelinger, M. (1976). "Enantioselective alkylation of cyclohexanone via chiral lithio-chelated enamines". Journal of the American Chemical Society . 98 (10): 3032–3033. doi:10.1021/ja00426a068.
  12. Enders, D.; Eichenauer, H.; Pieter, R. (1979). "Enantioselektive Synthese von (-)-(R)-und(+)-(S)-[6]-Gingerol-Gewürzprinzip des Ingwers". Chemische Berichte . 112 (11): 3703–3714. doi:10.1002/cber.19791121118.
  13. Enders, D.; Eichenauer, H. (1979). "Asymmetrische Synthesen via metallierte chirale Hydrazone. Enantioselektive Alkylierung von cyclischen Ketonen und Aldehyden". Chemische Berichte . 112 (8): 2933–2960. doi:10.1002/cber.19791120820.
  14. 1 2 Enders, D., Kipphardt, H., Fey1, P. (1987). "Asymmetric Synthesis Using the SAMP-/RAMP- Hydrazone Method: (S)-(+)-4-Methyl-3-Heptanoe" (PDF). Organic Syntheses . 65: 183. doi:10.15227/orgsyn.065.0183.{{cite journal}}: CS1 maint: multiple names: authors list (link) CS1 maint: numeric names: authors list (link)[ permanent dead link ]
  15. Ahlbrecht, H.; Düber, E. O.; Enders, D.; Eichenauer, H.; Weuster, P. (1978). "NMR-spektroskopische untersuchungen an deprotonierten iminen und hydrazonen". Tetrahedron Letters . 19 (39): 3691–3694. doi:10.1016/s0040-4039(01)95032-4.
  16. Enders, D.; Baus, U. (1983). "Asymmetrische Synthese beider Enantiomere von (E)-4,6-Dimethyl-6-octen-3-on – Abwehrsubstanz der Weberknechte Leiobunum vittatum und L. calcar (Opiliones)". Liebigs Annalen der Chemie . 1983 (8): 1439–1445. doi:10.1002/jlac.198319830816.
  17. Enders, D.; Bachstädter, G.; Kremer, K. A. M.; Marsch, M.; Harms, K.; Boche, G. (1988). "Structure of a Chiral Lithium Azaenolate: Monomeric, Intramolecular Chelated Lithio-2-acetylnaphthalene-SAMP-hydrazone". Angewandte Chemie International Edition in English . 27 (11): 1522–1524. doi:10.1002/anie.198815221.
  18. Bauer, W.; Seebach, D. (1984). "Bestimmung des Aggregationsgrads lithiumorganischer Verbindungen durch Kryoskopie in Tetrahydrofuran". Helvetica Chimica Acta . 67 (7): 1972–1988. doi:10.1002/hlca.19840670736.
  19. Martens, J.; Lübben, S. (1990). "(1S,3S,5S)-2-Amino-3-methoxymethyl-2-azabicyclo [3.3.0]octan: SAMBO — ein neuer chiraler Hilfsstoff". Liebigs Annalen der Chemie . 1990 (9): 949–952. doi:10.1002/jlac.1990199001175.
  20. Wilken, J.; Thorey, C.; Gröger, H.; Haase, D.; Saak, W.; Pohl, S.; Muzart, J.; Martens, J. (1997). "Utilization of Industrial Waste Materials, 11. Synthesis of New, Chiral β-sec-Amino Alcohols – Diastereodivergent Addition of Grignard Reagents to α-Amino Aldehydes Based on the (all-R)-2-Azabicyclo[3.3.0]octane System". Liebigs Annalen . 1997 (10): 2133–2146. doi:10.1002/jlac.199719971016.
  21. 1 2 Wengryniuk, S. E.; Lim, D.; Coltart, D. M. (2011). "Regioselective Asymmetric α,α-Bisalkylation of Ketones via Complex-Induced Syn-Deprotonation of Chiral N-Amino Cyclic Carbamate Hydrazones". Journal of the American Chemical Society . 133 (22): 8714–8720. doi:10.1021/ja202267k. PMID   21510644.
  22. 1 2 3 4 5 6 Enders, D., Wortmann, L. Peters, R. (2000). "Recovery of Carbonyl Compounds from N,N-Dialkylhydrazones". Accounts of Chemical Research . 33 (3): 157–169. doi:10.1021/ar990062y. PMID   10727205.{{cite journal}}: CS1 maint: multiple names: authors list (link)