Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Organic Reaction Mechanisms 2010: An annual survey covering the literature dated January to December 2010
Organic Reaction Mechanisms 2010: An annual survey covering the literature dated January to December 2010
Organic Reaction Mechanisms 2010: An annual survey covering the literature dated January to December 2010
Ebook1,391 pages10 hours

Organic Reaction Mechanisms 2010: An annual survey covering the literature dated January to December 2010

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Organic Reaction Mechanisms 2010, the 46th annual volume in this highly successful and unique series, surveys research on organic reaction mechanisms described in the available literature dated 2010. It details the latest progress in a wide range of classes of organic reaction mechanisms, including reactions of different compounds and acids and their derivatives, oxidation and reduction, aliphatic substitutions, elimination reactions, and molecular rearrangements, to name a few. An experienced team of authors compiled these reviews, ensuring the quality of selection and presentation.
LanguageEnglish
PublisherWiley
Release dateOct 10, 2012
ISBN9781119943570
Organic Reaction Mechanisms 2010: An annual survey covering the literature dated January to December 2010

Related to Organic Reaction Mechanisms 2010

Titles in the series (31)

View More

Related ebooks

Chemistry For You

View More

Related articles

Reviews for Organic Reaction Mechanisms 2010

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Organic Reaction Mechanisms 2010 - A. C. Knipe

    Table of Contents

    Title Page

    Copyright

    Contributors

    Preface

    Chapter 1: Reactions of Aldehydes and Ketones and their Derivatives

    Formation and Reactions of Acetals and Related Species

    Reactions of Glucosides

    Reactions of Ketenes

    Formation and Reactions of Nitrogen Derivatives

    C–C Bond Formation and Fission: Aldol and Related Reactions

    Other Addition Reactions

    Enolization and Related Reactions

    Oxidation and Reduction of Carbonyl Compounds

    Atmospheric Reactions

    Other Reactions

    References

    Chapter 2: Reactions of Carboxylic, Phosphoric, and Sulfonic Acids and their Derivatives

    INTERMOLECULAR CATALYSIS AND REACTIONS

    Carboxylic Acids and their Derivatives

    Phosphoric Acids and their Derivatives

    Sulfonic Acids and their Derivatives

    INTRAMOLECULAR CATALYSIS AND NEIGHBOURING GROUP PARTICIPATION

    ASSOCIATION-PREFACED CATALYSIS

    BIOLOGICALLY SIGNIFICANT REACTIONS

    Carboxylic Acids and their Derivatives

    Phosphoric Acids and their Derivatives

    References

    Chapter 3: Oxidation and Reduction

    Oxidation by Metal Ions and Related Species

    Oxidation by Compounds of Non-metallic Elements

    Ozonolysis and Ozonation

    Peracids and Peroxides

    Photo-oxygenation and Singlet Oxygen

    Triplet Oxygen and Autoxidation

    Other Oxidations

    Reduction by Complex Metal Hydrides

    Hydrogenation

    Transfer Hydrogenation

    Other Reductions

    References

    Chapter 4: Carbenes and Nitrenes

    Reviews

    Generation, Structure, and Reactivity

    Carbenes in Coordination Chemistry

    Addition—Fragmentations

    Insertion—Abstraction

    Rearrangements

    Nucleophilic Carbenes—Carbenes as Organocatalysts

    Nitrenes

    Heavy-atom Carbene Analogues

    References

    Chapter 5: Nucleophilic Aromatic Substitution

    General

    The SNAr Mechanism

    Heterocyclic Systems

    Meisenheimer and Related Complexes

    References

    Chapter 6: Electrophilic Aromatic Substitution

    General

    Halogenation

    Nitration

    Alkylation, Acylation, and Arylation Reactions

    Substitutions on Heterocyclic Rings

    Other Reactions

    References

    Chapter 7: Carbocations

    Introduction

    Alkyl and Cycloalkyl Carbenium Ions

    Benzyl Cations and Quinone Methides

    Benzhydryl, Trityl, and Fluorenyl Cations

    Carbocation Reactivity—Quantitative Studies

    Oxygen- and Sulfur-stabilized Cations

    Carbocations Containing Silicon and Other Group 14 Elements

    Halogenated Carbocations

    Allyl and Vinyl Cations

    Aryl Cations

    Arenium Ions

    Nitrenium Ions

    Aromatic Systems

    Dications

    Polycyclic Systems

    Carbonium (Bridged) Ions

    Carbocations in Biosynthesis

    References

    Chapter 8: Nucleophilic Aliphatic Substitution

    Allylic and Vinylic Substitutions

    Reactions of Cyclic Ethers

    Aziridines and Other Small Ring Substitutions

    Studies Using Kinetic Isotope Effects

    Nucleophilic Substitution on Elements Other than Carbon

    Medium Effects/Solvent Effects

    Micelles and Ion Pair Aggregates in Substitution Reactions

    Structural Effects

    Theoretical Studies

    Gas-phase Substitution Reactions

    Miscellaneous Kinetic and Product Studies

    Acknowledgement

    References

    Chapter 9: Carbanions and Electrophilic Aliphatic Substitution

    Carbanion Structure and Stability

    Carbanion Reactions

    Proton-transfer Reactions

    Miscellaneous

    Electrophilic Aliphatic Substitution

    References

    Chapter 10: Elimination Reactions

    E1cB and E2 Mechanisms

    Pyrolytic Reactions

    Elimination Reactions in Synthesis

    Other Reactions

    References

    Chapter 11: Addition Reactions: Polar Addition

    Reviews

    Electrophilic Additions

    Nucleophilic Additions

    References

    Chapter 12: Addition Reactions: Cycloaddition

    2 + 2-Cycloaddition

    2 + 3-Cycloaddition

    2 + 4-Cycloaddition

    Miscellaneous Cycloadditions

    References

    Chapter 13: Molecular Rearrangements: Part 1. Pericyclic Reactions

    [3,3]-Sigmatropic, Claisen, and Cope Rearrangements

    [2,3]-Reactions

    Vinyl Cyclobutane and Vinyl Cyclopropane Rearrangement

    1,2-Migration

    Ene Reaction

    Bergman Reaction

    Electrocyclic Reactions

    Cyclization

    4 + 2-Cycloadditions

    3 + 2-Cycloadditions

    Metathesis

    Metal-catalysed Reactions

    Miscellaneous

    References

    Chapter 14: Molecular Rearrangements: Part 2. Other Reactions

    Aromatic Rearrangement

    Ionic Rearrangements

    Rearrangements Catalysed by Metals

    Organo-catalysed Rearrangements

    Rearrangements Involving Ring Opening

    Isomerizations

    Tautomerism

    Radical Rearrangements

    References

    Author Index

    Subject Index

    Title Page

    This edition first published 2012

    © 2012 John Wiley & Sons, Ltd

    Registered office

    John Wiley & Sons Ltd, The Atrium, Southern Gate, Chichester, West Sussex, PO19 8SQ, United Kingdom

    For details of our global editorial offices, for customer services and for information about how to apply for permission to reuse the copyright material in this book please see our website at www.wiley.com.

    The right of the author to be identified as the author of this work has been asserted in accordance with the Copyright, Designs and Patents Act 1988.

    All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or transmitted, in any form or by any means, electronic, mechanical, photocopying, recording or otherwise, except as permitted by the UK Copyright, Designs and Patents Act 1988, without the prior permission of the publisher.

    Wiley also publishes its books in a variety of electronic formats. Some content that appears in print may not be available in electronic books.

    Designations used by companies to distinguish their products are often claimed as trademarks. All brand names and product names used in this book are trade names, service marks, trademarks or registered trademarks of their respective owners. The publisher is not associated with any product or vendor mentioned in this book. This publication is designed to provide accurate and authoritative information in regard to the subject matter covered. It is sold on the understanding that the publisher is not engaged in rendering professional services. If professional advice or other expert assistance is required, the services of a competent professional should be sought.

    The Publisher and the Author make no representations or warranties with respect to the accuracy or completeness of the contents of this work and specifically disclaim all warranties, including without limitation any implied warranties of fitness for a particular purpose. The advice and strategies contained herein may not be suitable for every situation. In view of ongoing research, equipment modifications, changes in governmental regulations, and the constant flow of information relating to the use of experimental reagents, equipment, and devices, the reader is urged to review and evaluate the information provided in the package insert or instructions for each chemical, piece of equipment, reagent, or device for, among other things, any changes in the instructions or indication of usage and for added warnings and precautions. The fact that an organization or Website is referred to in this work as a citation and/or a potential source of further information does not mean that the author or the publisher endorses the information the organization or Website may provide or recommendations it may make. Further, readers should be aware that Internet Websites listed in this work may have changed or disappeared between when this work was written and when it is read. No warranty may be created or extended by any promotional statements for this work. Neither the Publisher nor the Author shall be liable for any damages arising herefrom.

    Library of Congress Catalog Card Number 66-23143

    British Library Cataloguing in Publication Data

    A catalogue record for this book is available from the British Library

    Print ISBN: 978-0-470-97081-2

    Contributors

    Preface

    The present volume, the forty-sixth in the series, surveys research on organic reaction mechanisms described in the available literature dated 2010. In order to limit the size of the volume, it is necessary to exclude or restrict overlap with other publications which review specialist areas (e.g. photochemical reactions, biosynthesis, enzymology, electrochemistry, organometallic chemistry, surface chemistry and heterogeneous catalysis). In order to minimize duplication, while ensuring a comprehensive coverage, the editor conducts a survey of all relevant literature and allocates publications to appropriate chapters. While a particular reference may be allocated to more than one chapter, it is assumed that readers will be aware of the alternative chapters to which a borderline topic of interest may have been preferentially assigned.

    In view of the considerable interest in application of stereoselective reactions to organic synthesis, we now provide indication, in the margin, of reactions which occur with significant diastereomeric or enantiomeric excess (de or ee).

    Some changes of authorship will be apparent as Sue Armstrong (Molecular Rearrangements: Pericyclic) and Bob Coombes (Electrophilic Aromatic Substitution) have found it necessary to step down, having previously made excellent contributions to ORM for eight and twenty years respectively. Hopefully they will be reassured to find that their chapters are now in the safe hands of continuing members of the team.

    Steps taken to reduce progressively the delay between title year and publication date have continued to bear fruit, as evidenced by the publication of recent annual ORM volumes at nine-month intervals. Consequently we hope to regain our optimum production schedule soon.

    I wish to thank the staff of John Wiley & Sons and our expert contributors for their efforts to ensure that the review standards of this series are sustained, particularly during a period of substantial reorganization of production procedures.

    A. C. K.

    Chapter 1

    Reactions of Aldehydes and Ketones and their Derivatives

    B. A. Murray

    Department of Science, Institute of Technology Tallaght (ITT Dublin), Dublin, Ireland

    Formation and Reactions of Acetals and Related Species

    Reactions of Glucosides

    Reactions of Ketenes

    Formation and Reactions of Nitrogen Derivatives

    Synthesis of Imines

    The Mannich Reaction

    Addition of Organometallics

    Other Arylations, Alkenylations, and Allylations of Imines

    Reduction of Imines

    Iminium Species

    Other Reactions of Imines

    Oximes, Hydrazones, and Related Species

    C–C Bond Formation and Fission: Aldol and Related Reactions

    Reviews of Organocatalysts

    Asymmetric Aldols Catalysed by Proline and its Derivatives

    Other Asymmetric Aldols

    Mukaiyama and Vinylogous Aldols

    Other Aldol and Aldol-type Reactions

    The Henry (Nitroaldol) Reaction

    The Baylis–Hillman Reaction and its Morita-variant

    Allylation and Related Reactions

    The Horner–Wadsworth–Emmons Reaction and Other Olefinations

    Alkynylations

    Benzoin Condensation and Pinacol Coupling

    Michael Additions

    Miscellaneous Condensations

    Other Addition Reactions

    Addition of Organozincs

    Arylations

    Addition of Other Organometallics, Including Grignards

    The Wittig Reaction

    Hydrocyanation, Cyanosilylation, and Related Additions

    Hydrosilylation, Hydrophosphonylation, and Related Reactions

    Enolization and Related Reactions

    α-Halogenation, α-Alkylation, and Other α-Substitutions

    Oxidation and Reduction of Carbonyl Compounds

    Regio-, Enantio-, and Diastereo-selective Reduction Reactions

    Other Reduction Reactions

    Oxidation Reactions

    Atmospheric Reactions

    Other Reactions

    References

    Formation and Reactions of Acetals and Related Species

    A series of pyridinium cations with electron-withdrawing substituents on the ring catalyse acetalization of aldehydes and other protection reactions, such as the formation of dithianes, dithiolanes, dioxanes, and dioxolanes.¹ The best catalyst works at 0.1 mol%, outperforming a Brønsted acid with a pKa of 2.2.

    DFT has been used in the development of a general equation relating the activation energy of an intramolecular proton transfer to r (the distance between the reacting centres) and α (the hydrogen-bonding angle).² The equation has been applied to intramolecular general acid catalysis of five of Kirby's acetals (e.g. 1; X = NH, O). Reaction rates correlate with r² and sin (180° − α); that is, acetals with short r values and α close to 180° (forming a linear hydrogen bond) are more reactive.³

    UnFigure

    Cyclic hemiacetals (2) have been prepared stereoselectively in a 2 : 1 reaction of 4-formylbenzoates and aromatic enals (trans-Ar–CH UnFigure CH UnFigure CHO), using catalysis by N-heterocyclic carbenes (NHCs).⁴ UnFigure

    A dual acid-catalyst system has been employed for enantioselective addition of alkenyl and aryl boronates to chromene acetals (3).⁵ UnFigure The Lewis–Brønsted combination of a lanthanide triflate and a tartaric acid monoamide gives ee up to 97%.

    The gas-phase elimination kinetics of several β-substituted acetals have been measured in the range 370–441 °C and in the presence of a radical inhibitor.⁶ Two different concerted four-membered transition states are proposed, leading to either the alcohol and vinyl ether (the latter decomposing to alkene and aldehyde) or alkane and alkyl ester.

    Methylenecyclopropylcarbinols such as (4) react with acetals to give 3-oxabicyclo[3.1.0]hexanes (5); an intramolecular Prins-type mechanism is proposed.⁷ UnFigure

    UnFigure

    Iron(III) chloride or bromide has been used to catalyse Prins cyclization/halogenation of alkynyl acetals, using an acetyl halide as halogen source.⁸

    Deacetalization of acetals, R¹CH(OR²)2, in the presence of trifluoroacetic acid has been shown to be viable without water.⁹ Although water is a by-product, alcohols are not, and a hemiacetal is not an intermediate. Rather, a hemiacetal TFA ester [R¹–CH(OR²)–OCOCF3] is formed, followed by carbonyl production with two TFA ester byproducts, F3CCO2R². The latter process renders the reaction irreversible. The two esters are produced at separate points in what is essentially a cascade mechanism. All intermediates have been identified by NMR. The new reaction has been dubbed ‘acidolysis’ to distinguish it from the more familiar acid-catalysed hydrolysis.

    Reactions of Glucosides

    4,6-O-Benzylidene acetals of glycopyranosides (6) have been oxidatively cleaved to the corresponding hydroxy-benzoates (7a/b) using dimethyldioxirane under mild conditions, and in high yield.¹⁰ Appropriate choice of the neighbouring protecting group gives regioselectivity, with a preponderance of (7a) or (7b) of >99%, as desired. The balance of electronic and steric effects in the best groups—chloroacetyl and TBS (t-butyldimethylsilyl)—is discussed.

    UnFigure

    The stereo- and regio-selectivity of Lewis-acid-catalysed reductive ring-opening of 4,6-O-benzylidene acetals have been studied by kinetics, including primary and secondary isotope effects, leading to identification of a range of mechanisms in solvents of varying polarity, and in protocols with Brønsted acid additives.¹¹ UnFigure It is hoped that this will lead to new reducing agents, where reactivity and selectivity can be fine-tuned by choice of borane, solvent, Lewis acid, and temperature.

    Glycoside hydrolases can give 10¹⁷-fold rate enhancements, and estimates of their dissociation constants from their transition states are as low as 10−22 mol dm−3. Such affinity has encouraged mimicry, and a number of criteria have now been advanced to assess whether a natural or man-made glycosidase inhibitor is a true TS mimic.¹²

    A new dicyanohydrin-β-cyclodextrin acts as an artificial glycosidase, hydrolyzing aryl glycosides up to 5500 times faster than the uncatalysed reaction.¹³ Michaelis–Menten parameters are reported and compared with other modified cyclodextrins.

    An investigation of nucleophilic substitutions of 2-deoxyglycosyl donors indicates that the more nucleophilic the oxygen nucleophile used, the less stereo-selective the reaction becomes.¹⁴ UnFigure This erosion of stereo-chemical control is attributed to the rate of the stereochemistry-determining step approaching the diffusion limit, when the two faces of the prochiral oxocarbenium ion are subject to nucleophilic addition to afford a statistical mixture of diastereomers.

    Recent advances in understanding mechanisms of chemical O-glycosylation have been reviewed.¹⁵ pH-rate profiles have been constructed and analysed for glycosylation reactions of a range of aromatic amines.¹⁶

    Oxime formation from sugars can be slow, but nucleophilic catalysis by aniline (at 100 mM) can increase rates up to 20-fold, and glycosylamine formation has to be watched.¹⁷

    A DFT method has been applied to scan the potential energy surface of furanosyl oxocarbenium ions.¹⁸ UnFigure The results suggest that the preferred oxocarbenium ion conformation is not a consistent predictor of product stereochemistry.

    A chiral Brønsted acid, a BINOL-phosphoric acid, activates trichloroacetimidate glycosyl donors with β-selectivity.¹⁹ UnFigure

    An account describes the UnFigure mechanistic investigations that have led to a fuller understanding of the use of the 4,6-O-benzylidene acetal as a control element in glycosylation, giving direct access to β-mannopyranosides in high yield and selectivity.²⁰ UnFigure

    A rhodium(II)-carbene-promoted activation of the anomeric C UnFigure H bond of carbohydrates has been used to provide a stereospecific entry to α- and β-ketopyranosides.²¹ UnFigure

    Three unnatural methyl α-septanosides (8), with the 3- and 5-hydroxyls ax–eq, eq–ax, and eq–eq have been synthesized, and their rates of hydrolysis measured by ¹H NMR at 50 °C in 0.5 mol dm−3 DCl.²² UnFigure The hydroxyl orientation affects the rate, with equatorial being more electron withdrawing than axial. Comparison with rates for analogous methyl α-pyranoside structures shows that, while the inherently less stable seven-membered sugars react about two orders of magnitude faster, the rank ordering is the same.

    UnFigure

    Reactions of Ketenes

    Keto-ketenes (R¹R²C C O) homodimerize to β-lactones (e.g. 9), thereby providing an important way of accessing such compounds. Catalysis by tributylphosphine has been investigated by NMR, and evidence for tetravalent phosphonium enolate intermediates (10) is presented: they can be trapped as their TMS ethers or by reaction with 4-chlorobenzaldehyde (to give a β-lactone). Such enolates may prove useful in other synthetic methodologies. There was no evidence for pentacovalent phosphorus intermediates.²³ UnFigure

    UnFigure

    DFT investigation of Staudinger 2 + 2-cycloaddition of a ketene and an imine, catalysed by NHCs, favour the ‘ketene-first’ mechanism, that is, it is the ketene that is initially activated by the NHC. This mechanism persists even when variation in the electrophilicity of the imine leads to stereodivergence in the experimental results.²⁴ NHCs also promote the chlorination of unsymmetrically disubstituted ketenes, R¹R²C C O; the products are typically α-halo esters [R¹R²C*(Cl)–CO2R³] under the conditions employed. With chiral NHCs, modest ees of up to 61% are seen.²⁵ UnFigure

    Dimerization and trimerization reactions of thioformaldehyde and dimerization of thioketene have been studied by computation.²⁶

    Formation and Reactions of Nitrogen Derivatives

    Synthesis of Imines

    The affinities of a wide-ranging array of imines for hydride, proton, and electron have been measured by titration colorimetry and by electrochemical methods, in acetonitrile.²⁷ Thermodynamic ‘characteristic graphs’ are then introduced, linking the energies of the processes for each imine: each graph is intended to give the ‘molecular ID’ of the imine, facilitating prediction of likely reactions and mechanisms thereof.

    The mechanism of Schiff base formation between pyridoxal analogues and aldehydes has been studied by DFT.²⁸

    PNP ‘pincer’ complexes of ruthenium catalyse a new imine synthesis, from an alcohol and an amine, with evolution of hydrogen.²⁹

    Formylpyridines react with tris(hydroxymethyl)aminomethane [(HOCH2)3CNH2, ‘TRIS’], to give 1,3-oxazolidines (e.g. 11), which can equilibrate with their acyclic tautomers, that is, Schiff bases. Anomeric and hydrogen-bonding effects have been studied in these systems, including the adduct derived from pyridoxal.³⁰ Oxazolidines such as (12)—derived from TRIS and a benzaldehyde—have been prepared and then ring-opened under acetylating conditions. X-ray crystal data and computations indicate a strong endo anomeric effect stabilizing a conformation that leads to regioselective ring opening to give imine (rather than N-acetyloxazolidine). Imine-oxazolidine equilibria are also reported, and a per-O-acetylated imine, (AcOCH2)3–C UnFigure N UnFigure CHAr, in the para-nitro case.³¹

    UnFigure

    An alkyl or aryl group, R¹, in a 2-iminothiazole (13) can be exchanged with that in an isothiocyanate, R⁴–N UnFigure C S, in toluene at 105 °C.³² The position of equilibrium in this reversible reaction is mainly dependent on the electronic properties of the exchanging groups (i.e. R¹ and R⁴) and has been used to empirically compare the electrophilicity of various isothiocyanates.

    2-Substituted benzimidazoles have been prepared by condensation of various aldehydes with 1,2-phenylenediamine, using copper(II) triflate catalyst, in refluxing acetonitrile.³³

    The Mannich Reaction

    Organocatalytic asymmetric UnFigure Mannich reactions have been reviewed, focussing on proline derivatives,³⁴ UnFigure as have Mannich preparations of alkyl- and cycloalkyl-amines.³⁵

    The autocatalysis previously seen UnFigure in enantioselective Mannich reactions catalysed by l-proline and related species has been reinvestigated, using both the products themselves and close structural mimics.³⁶ UnFigure

    The 1-ethyl-3-methylimidazolium salt of UnFigure (S)-proline acts as an ionic liquid (IL), which gives ‘three 99s’ performance (yield/de/ee) in a one-pot three-component Mannich reaction.³⁷ UnFigure The reaction shows excellent chemo- and regio-selectivities, the precursors are cheap, the process tolerates moisture, and it can often be conducted at −20 °C.

    A diastereoselectivity switch has been engineered in the direct Mannich reaction of glycine UnFigure imines, R¹O2C UnFigure CH2–N UnFigure CR²R³, with N-(8-quinolyl)sulfonyl imines (14).³⁸ UnFigure Steric and electronic tuning of the R groups of the glycine imine switches the selectivity from syn-α,β-diamino acids (for benzophenone-type imines) to anti- (for electron-rich aldimines). An Fe-sulfos-Cu(I) chiral catalyst gives ees of 99% in many cases.

    UnFigure

    An anti-selective reaction of aldehydes with N-sulfonyl imines exploits hydrogen bonding involving a 4-hydroxypyrrolidine catalyst and an external Brønsted acid.³⁹ UnFigure

    DFT methods have been used to study diastereoselective reactions of ketimine with aldehyde, using UnFigure both l-proline and (S)-1-(2-pyrrolidinylmethyl)pyrrolidine, catalysts that give opposite diastereoselectivities.⁴⁰ UnFigure

    Ferrocenyl cation, as its PF6− salt, catalyses Mannich reaction of benzaldehyde, aniline, and cyclohexanone to give β-aminoketone (15), with some anti-preference, under solvent-free conditions.⁴¹ UnFigure Tests of two-reactant combinations indicate that the reaction proceeds initially via imine rather than aldol formation.

    Bench-stable α-amido sulfones have been used to generate N-Boc amino-protected imines, which then undergo in situ Mannich reactions with glycine Schiff-bases, using a cinchonidine–thiourea catalyst, to give α,β-diamino acid derivatives with ee/de close to 100%.⁴² UnFigure UnFigure In a similar strategy, a highly diastereo- and enantio-selective aminocatalytic Mannich reaction of aldehydes with N-carbamoyl imines involves their generation in situ from such α-amido sulfones.⁴³ UnFigure UnFigure

    DFT-calculated ees and des compare well with observed values for anti-Mannich and syn-aldol reactions catalysed by axially chiral amino sulfonamides.⁴⁴ UnFigure UnFigure

    While chiral phosphoric acids such as 3,3′-disubstituted BINOLs have been known to catalyse direct Mannich-type reaction of aldimines with 1,3-dicarbonyls, such catalysts can be contaminated by group I/II metal cations. Deliberate introduction of such cations, especially calcium, confirms that the metal salts may be the ‘true’ catalysts, giving higher yields and ees in some cases.⁴⁵ UnFigure

    Enantioselective Mannich reactions of diethyl fluoromalonate with N-Boc aldimines using chiral bifunctional organocatalysts give (β-aminoalkyl)fluoromalonates in 93–97% ee,⁴⁶ UnFigure and bifunctional amine–thiourea catalysts derived from rosin give high ee and de in reaction of lactones with such imines.⁴⁷ UnFigure UnFigure

    N-Sulfonylcarboxamides of proline catalyse Mannich reaction of cyclic ketones with N-protected iminoglyoxylate, with de/ee up to 94/99%. Enamine intermediates have been examined by DFT.⁴⁸ UnFigure UnFigure

    The first catalytic, enantioselective vinylogous Mannich reaction of acyclic silyl dienolates (17) has been reported. Using protected imines (16), ees up to 98% have been achieved (R¹ = H), and more highly substituted products (18, R¹ = Me) can be prepared diastereoselectively. A second-generation BINOL-based phosphoric acid catalyst developed for the process has been studied by NMR, and a crystal structure of the imine-bound catalyst was obtained, shedding light on the facial selectivity of the reaction.⁴⁹ UnFigure UnFigure

    UnFigure

    A Yb/K heterobimetallic catalyst and a chiral amide ligand promote nitro-Mannich (aza-Henry) reactions in up to 86% ee.⁵⁰ UnFigure

    Addition of Organometallics

    Advances in copper-catalysed enantioselective addition of dialkylzincs to imines have been reviewed back to 2000.⁵¹ UnFigure

    Nickel(II) and a spiro-chiral phosphine catalyse the three-component coupling of imines, diethylzinc, and aromatic alkynes with ee up to 98%, and with good chemoselectivity, to give useful allylic amines.⁵² UnFigure

    Diimines (19; R = Ph, 2-pyrrolyl, 2- and 4-pyridinyl, 2,2′-bithiophen-5-yl) have been prepared from (R,R)-1,2-diaminocyclohexane and aromatic aldehydes.⁵³ UnFigure Addition of organolithiums and allylzinc proceeds in high yield and de (except for the 2-pyridine case), giving diamines with four chiral centres. The latter have also been tested as enantioselective catalysts for the Henry reaction.

    UnFigure

    Quantitative structure–reactivity relationships (QSSR) have been used to examine enantioselectivity in the addition of organolithiums to imines.⁵⁴ UnFigure

    Chiral α-chloro N t-butanesulfinyl ketimines (20) react with Grignards to give chiral aziridines with de/ee up to 96/98%; the stereoselectivity is opposite to that found for imines without the α-chloro substituents, presumably due to chlorine coordination of the incoming Grignard.⁵⁵ UnFigure UnFigure

    The reactions of Grignard reagents with imines have been contrasted for catalytic and stoichiometric amounts of titanium alkoxide reagents.⁵⁶ The former favours alkylation, while the latter gives reductive coupling, with distinctive mechanisms for each, as shown by studies using deuterium-labelled substrates.

    Chiral phosphinoylimines have been prepared in high yield and good de by addition of Grignards to new P-chirogenic N-phosphinoylimines.⁵⁷ UnFigure

    For more references to Grignards and imines, see under ‘Addition of Other Organometallics, Including Grignards’ below.

    Other Arylations, Alkenylations, and Allylations of Imines

    Rhodium-diene complexes catalyse arylation of N-tosyl ketimines by addition of sodium tetraarylborates. Using a chiral diene renders the process highly enantioselective.⁵⁸ UnFigure

    Enantioselective formal alkenylations of imines, catalysed by axially chiral BINAP dicarboxylic acids, have been carried out using vinylogous aza-enamines.⁵⁹ UnFigure As the latter can be oxidized to nitriles, the route can allow access to enantiomerically enriched γ-amino α,β-unsaturated nitriles, and thus to synthetically useful chiral γ-amino acids.

    In the triphenylphosphine-catalysed reaction of alkyl propiolates with N-tosylimines, a stable phosphonium-enamine zwitterion (21) of proven importance in the mechanism has been isolated and characterized by X-ray crystallography.⁶⁰ Deuterium-labelling experiments have identified several hydrogen-specific processes, none of which limit turnover, but they are highly medium dependent.

    UnFigure

    N-protected α-imino esters, for example, Pg-N UnFigure CH UnFigure CO2Et, have been alkynylated with terminal alkenes using copper(I) triflate and a PYBOX ligand (22).⁶¹ UnFigure Surprisingly, excess ligand does not raise the ee, but excess copper does, and a switch in metal-to-ligand ratio alone can reverse the ee. A modest positive non-linear effect was observed, and it is suggested that changing the metal-to-ligand stoichiometry may alter the coordination geometry at copper, and thus the transition state.

    Enantioselective addition of terminal alkynes to imines and their derivatives has been reviewed, including in situ examples, that is, three-component reactions of terminal alkynes, aldehydes, and amines.⁶² UnFigure

    Chiral phosphinoylimines undergo highly diastereoselective alkynylation with aluminium acetylides, but lithium or magnesium alkynes give poor results.⁶³ UnFigure

    An alkylzinc-mediated enantioselective synthesis of N-tosyl-(E)-(2-en-3-ynyl)amines has been developed, working well with various N-tosylaldimines.⁶⁴ UnFigure

    A review covers diastereo- and enantio-selective alkynylation of imines and iminium ions.⁶⁵ UnFigure UnFigure

    Reduction of Imines

    Chiral 1,3-diamines have been accessed by diastereoselective reduction of enantiopure N t-butanesulfinylketimines (23, prepared from the corresponding diaryl ketone).⁶⁶ UnFigure The reduction can be 99 : 1 diastereoselective in either direction, depending on substrate and conditions. X-ray crystallography of reactants and products and NOESY-NMR studies point to unusual directing effects of the ortho-substituent in controlling the selectivity.

    A chiral phosphoramidite ligand has been used to achieve good enantioselectivity in iridium-promoted hydrogenation of benzophenone N–H imines, Ar–C(—NH)–Ph, affording chiral diarlmethylamines without the need for N-protection.⁶⁷ UnFigure Several ortho-substituted substrates gave particularly high ee.

    Advances in enantioselective reduction of C N bonds have been reviewed, focussing on the use of metal-free chiral organocatalysts with Hantzsch esters as hydride source.⁶⁸ UnFigure

    Reductive amination of carbonyl compounds—via transfer hydrogenation of their imine derivatives—has been achieved with cyclometalated iridium complexes.⁶⁹ UnFigure

    Ammonia–borane (H3N UnFigure BH3) has been employed in a mild, metal-free transfer hydrogenation of imines.⁷⁰ A concerted double-hydrogen-transfer mechanism is proposed, backed up by deuterium kinetic isotope effects, Hammett correlations, and ab initio calculations. Hydrogenation of other unsaturated systems is being followed up.

    Iminium Species

    Kinetics of the reactions of iminium ions (pre-generated from cinnamaldehyde and secondary amines) with cyclic ketene acetals were studied by UV–visible spectroscopy.⁷¹ Second-order rate constants have been used to derive values of the electrophilicity parameter, E ( − 10 < E < − 7), and these have been analysed using a correlation equation, log10k = S(E + N), where S and N are nucleophilicity parameters. The equation is then found to predict rate constants for reactions of the iminium ions with a range of other species, such as pyrroles, indoles, and sulfur ylides.

    The intermediacy of an iminium ion, Me2N+—CH2, in the nitrosative cleavage of triethylamine to N-nitrosodimethylamine (Me2N UnFigure NO) has been explored in a DFT study designed to elucidate how carcinogenic N-nitrosamines form from tertiary amines.⁷²

    Reaction of dimethyl sulphate with DMF gives methoxymethylene-N,N-dimethyliminium salt, Me2N+—CH(OMe) −O4S UnFigure Me.⁷³ It acts as an acid promoter of Staudinger synthesis of 2-azetidinones (β-lactams) from imines and substituted acetic acids. Under base catalysis, the carboxylate is proposed to react with the iminium salt to produce an activated ester, which breaks down (again with base catalysis) to yield the corresponding ketene, which is the immediate reactant with the imine.

    A review surveys the development and potential of iminium ion catalysis, using ions formed by the condensation of chiral secondary or primary amines with α,β-unsaturated aldehydes or ketones, in a variety of cyclo- and conjugate-addition reactions.⁷⁴ UnFigure UnFigure

    Other Reactions of Imines

    Palladium(II) and rhodium(I) catalysts and chiral disphosphane ligands allow addition of phenylboronic acid, and of phenylboroxine, to N-tosylimines, in up to 99% ee.⁷⁵ UnFigure

    Azomethine imines (24) undergo 1,3-dipolar cycloaddition to homoallylic alcohols, giving trans-pyrazolidines (25) with excellent regio-, diastereo-, and enantio-selectivities and good yields.⁷⁶ UnFigure A tartrate auxiliary and a Grignard in excess complete the protocol, with generation of the chloromagnesium salt of the homoallylic alcohol being essential to the mechanism.

    UnFigure

    An unexpected reaction of aromatic aldimines (26) with a difluoroenoxysilane gives access to 2,2-difluoro-3-hydroxy-1-ones (28)—the Mukaiyama aldol-type product—via an amine (27).⁷⁷ Zinc triflate promotes the reaction, and ¹⁸O-labelling and other experiments suggest that water is required to form the product (28).

    UnFigure

    3,4-Dihydroisoquinoline (29) undergoes aza-Henry reaction with excess nitromethane at ambient temperature to give the corresponding 1-(nitromethyl)tetrahydroisoquinoline (30), an unstable species that is trapped by acylation or alkylation, leading to Reissert-like products via an overall one-pot three-component reaction.⁷⁸ Evidence for reaction via the methyleneazinic acid tautomer of nitromethane (31) is presented.

    UnFigure

    A vinylogous imine intermediate (33), generated in situ from an arylsulfonyl indole (32), undergoes enantioselective Michael addition to malonitrile, using a chiral thiourea catalyst, to give useful 3-indolyl derivatives (34).⁷⁹ UnFigure

    UnFigure

    DFT has been used to study aziridination of diazoacetate with syn- and anti-imines in the presence of a chiral bisoxazoline-copper(I) catalyst.⁸⁰ UnFigure

    trans-2,3-Disubstituted aziridines (36) have been prepared from N-sulfinylaldimines (37) and 2-(para-tolylsulfinyl)benzyl iodide (35) in high ee/de. Whether the intermediates are benzyl halocarbenoids or benzyl carbanions has been examined using DFT.⁸¹ UnFigure UnFigure

    UnFigure

    The previously reported reaction of diarylmethyl imines with diazoacetates to give cis-aziridines (using chiral VANOL or VAPOL ligands) has now been complemented by conversion of diazoacetamides to the corresponding trans-aziridines, again with high de, ee, and yield.⁸² UnFigure UnFigure

    Systematic investigation of aziridination of benzhydryl-type imines, R UnFigure CH UnFigure N UnFigure CHAr2, has been undertaken, varying the imine aryls and using VANOL- and VAPOL-derived chiral boroxinates.⁸³ UnFigure UnFigure Typical ees of 96–97% were obtained using 2,4-dimethyl-3-methoxy as the Ar groups, and for these substrates their high activity allowed the conventional diazoacetate ester reagent to be replaced by a diazoacetamide, an option that is not really viable for simple benzhydryls (i.e. Ar = Ph). While varying the aryls varies the aziridine products, the latter are easily converted to N–H aziridines.

    2-Methylazaarenes such as 2,6-lutidine (38) undergo palladium-catalysed benzylic addition with N-sulfonyl aldimines, showing a powerful C UnFigure H activation effect and giving access to heteroarylethylamines (39); a stereoselective version is being explored.⁸⁴

    UnFigure

    Organocatalytic asymmetric Strecker reactions have been reviewed.⁸⁵ UnFigure

    Chiral BINOLs and amino alcohols have both been used as enantioselective catalysts for Strecker reaction of achiral N-phosphinoyl imines with diethylaluminium cyanide.⁸⁶ UnFigure

    Enantioselective titanium-catalysed cyanation of imines has been carried out rapidly at room temperature.⁸⁷ UnFigure

    Chiral mono- and di-meric manganese(III) salen complexes catalyse Strecker addition of TMSCN to N-benzylimines at −55 °C in the presence of 4-phenylpyridine-N-oxide.⁸⁸ UnFigure The dimeric auxiliary is more effective (ee > 99%), and the catalysts are recyclable.

    Hydrolysis of the Schiff base, N-salicylidene-meta-chloroaniline, has been studied from pH 3 to 12 at 303 K and also at other temperatures to yield thermodynamic parameters.⁸⁹

    Chiral phosphoric acids catalyse asymmetric peroxidation of imines, R²–CH UnFigure N UnFigure R¹, to give amine-peroxides with the chiral centre between the functional groups (40), using organic hydroperoxides, R³–OOH.⁹⁰ UnFigure

    Recent interest in the intermolecular carbon radical addition to the C N double bond of imines, hydrazones, and oxime ethers has been reviewed, including stereoselective approaches.⁹¹ UnFigure UnFigure

    A catalytic asymmetric exo′-selective [3 + 2] cycloaddition of iminoesters (41) to nitroalkenes yields highly functionalized proline esters (42).⁹² UnFigure UnFigure

    UnFigure

    For a homocoupling of aromatic imines, see under ‘Benzoin Condensation and Pinacol Coupling’ below. For a nucleophilic perfluoroalkylation of imines, see under ‘Addition of Organozincs’ below.

    Oximes, Hydrazones, and Related Species

    FT-ICR mass spectrometry has been used to measure gas-phase acidities of ring-substituted (E)-acetophenone oximes.⁹³ Substituent trends are the same as in DMSO solution, indicating that solvation stabilization has a consistent effect, but that there is no specific solvent effect on any particular substituent.

    The use of O-substituted hydroxylamines and oximes as electrophilic amino-transfer agents has been reviewed.⁹⁴

    2-Isoxazolines have been prepared enantioselectively by conjugate addition of oximes to α,β-unsaturated aldehydes, with anilinium catalysis.⁹⁵ UnFigure

    (O)-2-(Acyl)vinylketoximes (43) have been made as their (E)-isomers by regio- and stereo-specific addition of ketoximes (R¹R²C NOH) to acylacetylenes (Ph–C C UnFigure COR³) under mild conditions (DCM/r.t./10 mol% Ph3P).⁹⁶ Slow build-up of the (Z)-material over time indicates that the (E)-isomer is a kinetic product.

    A gold complex catalyses cyclization of O-propioloyl oximes (44), giving good yields of 4-benzylideneisoxazol-5(4H)-ones (45) after transfer of the arylidene group, but crossover experiments indicate that the arylidene ‘migration’ is in fact intermolecular.⁹⁷

    UnFigure

    Triphenylphosphine and carbon tetrachloride, together with catalytic DBU and Bu4NI, effect oxime ether formation (from oxime and alcohol) in refluxing acetonitrile.⁹⁸

    Among reports involving Beckmann rearrangement, N-imidoylbenzotriazoles (46) have been prepared in one pot in high yield from ketoximes, R¹–C(R²)—NOH, by reaction with mesyl chloride in the presence of a base and subsequent addition of benzotriazole.⁹⁹ A kinetic study of the rearrangement of cyclohexanone oxime to ε-caprolactam in aprotic solvents has been carried out, using trifluoroacetic acid as catalyst.¹⁰⁰ Bromodimethylsulfonium bromide (Me2S+Br Br−) catalyses rearrangement of ketoximes in refluxing acetonitrile, in the presence of zinc chloride.¹⁰¹ Rates of rearrangement of cyclohexanone oxime para-toluenesulfonate in eleven solvents have been described by a three-parameter linear correlation involving polarizability, electrophilicity, and solvent molar volume.¹⁰² Rearrangement of cyclododecanone oxime into ω-laurolactam has been followed by an ‘in situ’ multinuclear solid-state NMR method, and in a batch reactor process, using IL media.¹⁰³

    NiCl2 · 6H2O catalyses coupling of aldoximes with amines to give amides; the oxime can be prepared in situ from the corresponding aldehyde. ¹⁸O-Labelling studies have been used to probe the mechanism: a label in the oxime is not incorporated into the amide.¹⁰⁴

    The combination of triflic anhydride and a 30% excess of triphenylphosphine dehydrates aldoximes to nitriles at 0 °C in high yield in minutes, using 2 equiv. of triethylamine base in DCM. ¹H, ¹³C, ¹⁹F, and ³¹P NMR studies indicate that the reagent combination equilibrates to a mixture of (Ph3P+) OTf Tf− and (Ph3P+)2O·2Tf−, with the former acting as oxygen activation and dehydration reagent.¹⁰⁵

    Indium trichloride catalyses hydration of nitriles to amides: in refluxing toluene, acetaldoxime can be used as a water surrogate.¹⁰⁶ The by-product—acetonitrile—is already known to be required for some amide-to-nitrile protocols.

    Reports of oxidative deoximation back to carbonyl include an account of the kinetics of deoximation of a series of oximes of 3-alkyl-2,6-diphenylpiperidin-4-one (47) by pyridinium fluorochromate, which indicate steric crowding as the major influence.¹⁰⁷ Rates of deoximation of aldoximes and ketoximes by morpholinium chlorochromate have been measured in DMSO, showing first-order dependence on both substrate and oxidant; for acetaldoxime, 19 solvents were examined.¹⁰⁸ Quinolinium fluorochromate deoximates ketoximes in aqueous acetic acid, with a first-order dependence on both substrate and oxidant.¹⁰⁹ Oximes have also been deoximated by aerial oxidation, using manganese(I) porphyrins as catalysts and benzaldehyde as oxygen acceptor, in toluene at 50 °C. A radical trap stops the reaction, and the presence of a manganese-oxo porphyrin was confirmed by UV–vis spectra. The oximes of 2-nitrobenzaldehyde and pyridine-2-carboxaldehyde gave nitrile product; that is, ‘benzaldehydes’ with electron-withdrawing groups in the ortho-position divert in this way.¹¹⁰

    UnFigure

    Organoceriums have been added diastereoselectively to chiral aldehyde hydrazones derived from 1-aminoprolines; resulting hydrazines can be cleaved to give enantiomerically enriched amines in protected form.¹¹¹ UnFigure UnFigure The advantages of organoceriums over Grignards or organolithiums are discussed.

    Chiral N-amino cyclic carbonate hydrazones (‘ACC’ hydrazones, e.g. (48), with a rigid carbamate derived from camphor) undergo α-alkylation via deprotonation by LDA.¹¹² UnFigure UnFigure DFT has identified the features of the azaenolate intermediate that give rise to stereoselectivity. The calculations predict higher stereoselectivity than previously reported by experiment, and a modified experimental method has now yielded the higher values.

    Indium and a chiral ammonium catalyse allylation of N-benzoylhydrazones to give homoallylic amines in high yield and up to 99% ee, at room temperature in methanol.¹¹³ UnFigure

    Tetrasubstituted alkenes (49) have been accessed by coupling of N-arylsulfonylhydrazones with aryl halides, using palladium(II) catalysis.¹¹⁴

    Arylation of α-chiral ketones has been achieved by converting them to tosylhydrazones, then cross-coupling them with aryl halides, using palladium(0).¹¹⁵ UnFigure Enantiopurity is maintained, avoiding the epimerization problems found with many other approaches.

    Chiral α-hydrazino acids (50) have been accessed by asymmetric hydrocyanation of hydrazones with TMSCN; an O-silylated BINOL-phosphate formed in situ acts as auxiliary, giving α-hydrazinonitriles in a Strecker-like process, with subsequent acid hydrolysis yielding (50).¹¹⁶ UnFigure

    UnFigure

    A range of α-amido-α-aminonitrones (51) can react to form three classes of products—1,2,5-oxadiazin-4-ones, amidines, and dibenzo[d, f][1,3]diazepines—all of which retain the core structure. The products were identified by X-ray crystallography, which also pointed out unusual features, such as an exceptionally long single bond (arrowed), up to 1.54 Å, and a very high ‘trigonal’ angle of 131° for , as well as NH···O and NH···N intramolecular hydrogen-bond-like interactions. These features, together with DFT calculations, have been used to help elucidate the operative mechanisms.¹¹⁷

    For oxime formation from carbohydrates, see under ‘Reactions of Glucosides’ above.

    C–C Bond Formation and Fission: Aldol and Related Reactions

    Reviews of Organocatalysts

    General reviews include coverage of chemoselectivity in reactions involving asymmetric aminocatalysis,¹¹⁸ UnFigure the roots of asymmetric aminocatalysis over the past century, championing the seminal contributions of Knoevenagel in the 1890s,¹¹⁹ UnFigure UnFigure current approaches to improving asymmetric organocatalysts via supramolecular interactions,¹²⁰ UnFigure UnFigure and recent developments in aldolase-type organocatalytic direct reactions in water.¹²¹ UnFigure

    Chiral BINOL-phosphoric acid catalysis has been reviewed,¹²² UnFigure as has the emerging field of chiral phosphine oxides as organocatalysts of, for example, reductive aldols.¹²³ UnFigure UnFigure

    The use of NHC catalysts in aldehyde reactions has been reviewed,¹²⁴ as has been the regio- and stereo-chemistry of the aldol, with a survey of methodologies up to the present.¹²⁵ UnFigure UnFigure

    No Barrier Theory and Marcus Theory have been applied to the rates of aldol addition reactions of representative aldehydes and ketones.¹²⁶ The use of kinetic isotope effects in probing the mechanisms of stereoselective reactions has been surveyed (84 references).¹²⁷ UnFigure

    Many slow reactions not considered suitable for continuous flow processing techniques are now being reassessed under high-temperature/pressure conditions.¹²⁸

    Asymmetric Aldols Catalysed by Proline and its Derivatives

    Reviews of asymmetric aldol reactions include an account of those proceeding via enamines using organocatalysts,¹²⁹ UnFigure UnFigure their application to total synthesis of natural products in the last 5 years,¹³⁰ UnFigure and a survey of direct asymmetric aldols (357 references), which covers both organocatalytic and metal-based catalysts, noting the still low reactivity of many of the catalysts developed to date.¹³¹ UnFigure

    In reports of proline-catalysed aldol reactions, the central role of enamine intermediates has been underlined by their direct observation by NMR. E-Configured s-trans enamines (52) are detected: in DMSO, EXSY-NMR shows them arising from oxazolidinones rather than from iminium-type intermediates. The oxazolidinone-enamine equilibrium is not affected by additional water (in small amounts) or by the amount of catalyst.¹³² A computational study has compared the enamine (Houk–List) and oxazolidinone (Seebach) mechanisms, with the latter being found to be inadequate for predicting the stereochemical outcome.¹³³ UnFigure UnFigure Another DFT study has focussed on the scope for oxazolidinone intermediates,¹³⁴ UnFigure and this method has also been used to investigate further the enamine mechanism for reactions involving acetone.¹³⁵ UnFigure A coherent mechanistic rationale has been put forward for differences in kinetic behaviour in enamine reactions such as aldol, amination, and aminoxylation, with a particular focus on auto-inductive effects and on the catalytic effects of additives.¹³⁶ UnFigure DFT has also been used to identify the origin of the enantioselectivity in the aldol reaction of benzaldehyde and acetone as catalysed both by proline derivatives and by 2-azetidine carboxylic acid.¹³⁷ UnFigure

    UnFigure

    New prolinamide catalysts of the aldol reaction of para-nitrobenzaldehyde with acetone have been reported.¹³⁸ UnFigure Calix[4]arene-prolinamide organocatalysts give yields/ee/de up to 99/97/70% in direct aldols of aromatic aldehydes with cyclohexanone.¹³⁹ UnFigure UnFigure

    List's proline-catalysed stereoselective intramolecular aldols of 1,7-dicarbonyl compounds have been studied by DFT, with a polarizable continuum model employed for solvent effects. The enantioselectivity is found to arise from a key electrostatic contact between the forming alkoxide and the proline. The origin of the diastereoselectivity is typically more complex, especially for dialdehydes.¹⁴⁰ UnFigure UnFigure

    The application of reaction progress kinetic analysis to the proline-catalysed aldol has been described.¹⁴¹ UnFigure

    The possible roles of imidazolidinone intermediates or by-products in aldol reactions catalysed by prolinamides (53; R = H, NO2) has been studied by NMR and X-ray characterization of these species.¹⁴² UnFigure

    Four prolinamides (54) have been designed with enhanced acidity (EWG = Ac, Ms, Tf, and Ts) and the potential for multiple N UnFigure H···O hydrogen bonding. The mesylate gave the best performance in terms of yield/de/ee in a test aldol: 94/94/>98%, while the tosylate may involve an aryl-stacking stabilization of the transition state.¹⁴³ UnFigure UnFigure

    Two new catalysts (alcohol 55, and the corresponding ketone) have been developed for direct aldol addition in the presence of water.¹⁴⁴ UnFigure UnFigure Prepared from trans-4-hydroxy-l-proline and the steroid isosteviol, the strategy involves a hydrophilic catalytic group (the acid of proline), a lipophilic pocket (the isosteviol skeleton), and an assisting functional group (the remote alcohol/ketone). With only 1 mol% loading, yield/de/ee of up to 99/98/99% has been achieved for a cyclohexanone–araldehyde aldol at room temperature. Effects of solvent, water, temperature, and substrate structure have been studied.

    UnFigure

    Ethylene and propylene carbonate, readily prepared from epoxides and carbon dioxide, are effective solvents for proline-catalysed aldols, giving yields/de/ee up to 99/100/99%. Choice of carbonate solvent and whether or not to use water co-solvent has to be matched to substrates, and in particular to their polarity.¹⁴⁵ UnFigure UnFigure

    Intramolecular aldols of cyclic diketones are catalysed by proline, and List's studies of the effect of incorporation of a 4-fluoro substituent in the cis- or trans-position has been studied by DFT. It finds that fluorine changes pathways as well as transition states: a low energy epimerization (after the C UnFigure C bond forming process) affects product distribution.¹⁴⁶ UnFigure UnFigure

    N-(para-Dodecylphenylsulfinyl)-2-pyrrolidinecarboxamide (56) is one of the best anti-aldol catalysts to date, with yields/ee/de up to 98/99/98%, low catalyst loading, mild conditions, and convenient solvents (or none). A DFT study has now identified the origins of the diastereoselectivity in non-classical hydrogen bonds between the sulfonamide, the electrophile, and the catalyst enamine that favour the anti-Re aldol transition state.¹⁴⁷ UnFigure UnFigure

    An l-prolinethioamide catalyses aldols in water at 0 °C, with yields/ee up to 98/99%.¹⁴⁸ UnFigure

    Strong non-linear effects are observed in proline-catalysed aldols when an achiral thiourea catalyst is also employed in non-polar solvents: with an ee as low as 5% for the proline, 40% ee and 94% de are observed in the products.¹⁴⁹ UnFigure UnFigure The role of the thiourea co-catalyst in such reactions has been investigated. Examining the reaction of acetone with 4-substituted benzaldehydes, non-linear effects are observed (%eealdol versus %eeproline), but these are markedly dependent on the nature of the aromatic substituent. Results from ¹H-NMR and ESI-MS suggest that the main role of the thiourea is not that of producing a soluble proline-thiourea hydrogen-bonded complex.¹⁵⁰ UnFigure

    IL-tagged amino acid derivatives—1,2,3-triazolium salts linked to lysine or proline—give high yields/ee/de in direct aldols: the lysine surprisingly outperformed the proline.¹⁵¹ UnFigure UnFigure

    (S)-Prolinamides with a trans-4-ester moiety bearing an IL group give excellent yields, des and ees in aldol reactions in water.¹⁵² UnFigure UnFigure

    A chiral solvent effect has been seen in proline-catalysed aldols in aqueous propylene carbonate: when enantiopure (R)-propylene carbonate (57) is used with (R)-proline, they constitute a ‘matched pair’ with high de/ee, whereas (S)-proline/(57) is a mismatch.¹⁵³ UnFigure UnFigure

    UnFigure

    Racemic α-acylphosphinates undergo cross-aldol reaction with acetone to give diastereomeric α-hydroxyphosphinates (58), because of the phosphorous chiral centre. Using proline catalyst, high ees and des have been achieved.¹⁵⁴

    Efficient direct α-hydroxymethylation of ketones in homogenous aqueous solvents has been reported: a bis-prolinamide-zinc complex promotes aldol reaction with aqueous formaldehyde in good yield and up to 94% ee.¹⁵⁵ UnFigure

    Other Asymmetric Aldols

    A virtual screening method has UnFigure been demonstrated as a rapid computational tool for prediction of potential asymmetric aldol organocatalysts, throwing up several new classes such as β-amino acids and hydrazides for testing.¹⁵⁶ Amino amide catalysts that exploit a double hydrogen-bonding activation of carbonyls give high ees.¹⁵⁷ UnFigure UnFigure l-Tryptophan catalyses reaction between cyclohexanone and aldehydes in water; DFT has been used to identify the precise role of the indole substituent in stabilizing the transition state.¹⁵⁸ UnFigure UnFigure

    A siloxy-serine facilitates syn-selective direct aldols in an IL: the recyclable catalyst gives de/ee up to 88/94% under mild conditions.¹⁵⁹ UnFigure UnFigure A simple chiral diamine—picolylamine (59)—is an excellent organocatalyst for aldol reactions in water.¹⁶⁰ UnFigure UnFigure

    The combination of a primary–tertiary diamine and a Brønsted acid enables syn-selectivity in cross-aldols of aldehydes: de = 92%. A chiral diamine (60), with triflic acid, renders it enantioselective too: ee = 87%, and it works for glycolaldehyde donors.¹⁶¹ UnFigure UnFigure

    UnFigure

    Direct addition of enolizable aldehydes to α-halo thioesters gives β-hydroxy thioesters via reductive soft enolization, with syn-selectivity, whereas conventional conditions (amide bases) with esters or thioesters gives anti-product.¹⁶² UnFigure The conditions are mild (MgI2/PPh3/DCM), and the addition step is under kinetic control.

    Several stereoselective aldol reactions of β-siloxy methyl ketones with aldehydes have been developed using super-silyl stereo-directing groups such as –Si(TMS)3, including examples of both 1,5-syn- and 1,5-anti-control.¹⁶³ UnFigure DFT analysis of the influence of β-substituents has been used to explain substrate-based 1,5-syn-stereocontrol in boron-mediated aldol reactions of β-alkoxy methylketones.¹⁶⁴ UnFigure UnFigure

    Quinidine thiourea catalyses the asymmetric aldol reaction of unactivated ketones with activated carbonyl compounds via an enolate intermediate: it is suited to cases where the enamine-based organocatalysis does not work well.¹⁶⁵ UnFigure

    A chiral bifunctional thiourea catalyses aldol addition of α-isothiocyanato imides to α-ketoesters in good ee and fair de, giving access to β-hydroxy-α-amino acid derivatives.¹⁶⁶ UnFigure UnFigure

    Mukaiyama and Vinylogous Aldols

    N,N-Bis(trifluoromethanesulfonyl)squaramide (61) is a bench-stable and strong Brønsted acid. It catalyses a wide range of aldehyde reactions: Mukaiyama aldol, Mukaiyama Michael, Hosomi–Sakurai allylation, and an intramolecular carbonyl-ene reaction of a 6-enol. In reactions with silylated substrates, it appears that (61) acts to directly protonate the carbonyl compound, rather than catalysing routes involving silylated Brønsted acid.¹⁶⁷

    An (R)-BINAP platinum(II) complex catalyses enantioselective reaction of aldehydes with ketene silyl acetals, for example, Me2C C(Me)OTMS in DMF.¹⁶⁸ UnFigure UnFigure The complex undergoes a dimer/monomer equilibrium in this solvent: the monomer is apparently more catalytic.

    A new series of C2-symmetric chiral ligands (62; R UnFigure H, Me, But, etc.) has been synthesized. Complexed with europium(III), aldehydes can be activated in aqueous media, with the lanthanide still having vacant sites for hydration.¹⁶⁹ UnFigure UnFigure In addition, the lanthanide complex facilitates luminescence-decay measurements. Tested as catalysts of Mukaiyama aldols in ethanolic water, β-hydroxy carbonyl products were obtained in high yields and de/ee up to 96%, at temperatures as low as −25 °C. The use of luminescence measurements allowed binding of benzaldehyde to be observed (indirectly) via decreases in the water-coordination number of the europium cation.

    A stereoinduction model has been used to explain an unexpected syn-selectivity in the Mukaiyama aldol addition of crowded enolsilanes to α-chloroaldehydes.¹⁷⁰ UnFigure

    Pentafluorophenylammonium trifluoromethanesulfonimide, F5C6–NH3+ −NTf2, promotes Mukaiyama aldol and Mannich reactions using ketene silyl acetals with ketones and oxime ethers, respectively.¹⁷¹ ¹H-NMR and other investigations suggest in situ formation of trimethylsilyl bistriflimide, Tf2N(TMS), as the active catalyst.

    A stereoinduction model has been used to explain the unexpected syn-selectivity in the Mukaiyama aldol addition of sterically demanding enolsilanes to α-chloro aldehydes.¹⁷² UnFigure

    Vinylogous Mukaiyama aldol reactions of enals with vinylketene silyl N,O-acetals (63a/b) give 1,7- (64a) and 1,6,7- (64b) -remote asymmetric inductions in TiCl4-mediated experiments at low temperature.¹⁷³ UnFigure Specific addition of small amounts of water (but not other protic species such as alcohols or acetic acid) gives a remarkable acceleration using ent-64a, and it is found for a variety of aldehydes. Possible double activation by water and titanium(IV) is considered, or water may break up TiCl4 aggregates.

    UnFigure

    Pyrrole- and furan-based dienoxy silanes (65; X = O, N-Boc) undergo ‘uncatalysed’ vinylogous Mukaiyama aldol reaction in methanolic salty water at 40 °C, in open air, giving high des.¹⁷⁴ UnFigure The furan system is syn-selective (giving 66-O), whereas anti-product (66-N-Boc) is found for pyrroles, although this switch appears to be steric in origin (i.e. due to the bulk of the Boc), rather than being due to the change of heteroatom per se. The precise roles of water as both solvent and ‘catalyst’ are discussed in the context of the reaction not being wholly homogenous, but involving dispersed droplets of lipophilic reactants.

    UnFigure

    Other Aldol and Aldol-type Reactions

    A DFT study of the catalysis of the intramolecular aldol of acyclic keto-aldehydes by a bifunctional guanidine organocatalyst (67, 1,5,7-triazabicyclo[4.4.0]dec-5-ene, TBD) examined the model substrate 6-oxoheptanal.¹⁷⁵ UnFigure Two steps are involved: concerted proton abstraction/proton donation to enolize the substrate (with internal enolization of the ketone operative), followed by C UnFigure C bond formation concerted with proton transfer from enol to aldehyde, shuttled across the non-bridgehead nitrogens of (67). Alternative nucleophilic and enamine mechanisms have been explicitly ruled out by the calculations.

    UnFigure

    The aldol condensation of acetaldehyde in water has been studied under environmental conditions at high pH, as it may play a role in the degradation of organic matter in hyperalkaline conditions.¹⁷⁶ Analysis of the kinetics suggests that the reaction is first order in substrate, hydroxide, and carbonate, in contrast to earlier studies suggesting a second-order base dependence, with the authors claiming that they have better avoided interference by a competing Cannizzaro process.

    Benzamidine catalysis of an aldol reaction can be switched on and off reversibly with carbon dioxide, without affecting substrate or products.¹⁷⁷

    Carbanions of 3-chloropropyl phenyl sulfones containing carbonyl and imino groups (e.g. 68) in the ortho-position add intramolecularly to these groups to give aldol-type anions. Subsequent intramolecular 1,5-substitution of chlorine gives tricyclic tetrahydrofurans, pyrrolidines, and cyclopentanes.¹⁷⁸

    Halogenotin hydrides, Bu2SnXH (X = Cl, I), catalyse a reductive aldol reaction of enones to give β-hydroxyketones in good de, using a Ph2SiH2/alcohol promoter system.¹⁷⁹ UnFigure

    A strategy for controlling enantio- and diastereo-inductions in a sequential hydroformylation-aldol process involves selection of an appropriate combination of a chiral metal catalyst and a chiral organocatalyst.¹⁸⁰ UnFigure UnFigure

    A silver(I)-BINAP complex catalyses asymmetric aldol reaction of alkenyl trichloroacetates with α-keto esters.¹⁸¹ UnFigure The reaction is also promoted by dibutyltin dimethoxide, Bu2Sn(OMe)2, a species that can be regenerated by addition of methanol. The catalysts also work for the reaction of diketene and methyl benzoylformate.

    Enolizable aldehydes, R¹R²CHCHO, undergo an asymmetric Meerwein–Ponndorf–Verley–Aldol etherification reaction in methanol, giving highly functionalized products (69) with defined configurations at adjacent quaternary and tertiary centres. (−)-Menthyl-TMS is used as auxiliary, and trifluoroacetic acid is required as a catalyst.¹⁸² UnFigure UnFigure

    Aldehyde ‘dimerization’ to Tishchenko esters is catalysed by sodium hydride. While NaH is usually considered a base, it can reduce aldehydes to sodium alcoholates, and this is proposed as the first step; detection of alcohol by-product supports such a mechanism.¹⁸³

    Heteroaryl aldehydes undergo Evans–Tishchenko coupling with β-hydroxyketones using a samarium catalyst at −15 °C: high yields and de are obtained.¹⁸⁴ UnFigure At ambient temperature, a retro-aldol aldol Tishchenko process competes.

    An asymmetric direct vinylogous aldol reaction of unactivated γ-butenolide (70) with aldehydes gives the corresponding 5-(1′-hydroxy) derivatives in high yield/ee (93/83%), using a cinchona-alkaloid-based thiourea organocatalyst.¹⁸⁵ UnFigure

    A model reaction of an enal (71) and an enone (72) to give stereoselective synthesis of a trans-cyclopentene (73), catalysed by an NHC, was studied by DFT methods.¹⁸⁶ UnFigure The complex mechanism involves an initial Breslow intermediate attacking the enone to give an enol-enolate, the point where the trans-stereochemistry of (73) is determined. An intramolecular aldol condensation, extrusion of the NHC, and elimination of carbon dioxide feature in the later steps.

    UnFigure

    Asymmetric homoaldols have been reviewed.¹⁸⁷ UnFigure

    Homodimerization of 2-cyclohexanone, catalysed by l-proline, proceeds via a two-step imine/enamine addition or concerted Diels–Alder cycloaddition: the former is preferred.¹⁸⁸ UnFigure

    A silver(I) complex of a chiral quinoxaline-diphosphine gives ees up to 99% in a nitroso aldol of alkenyl trichloroacetates to give α-amino-oxy ketones.¹⁸⁹ UnFigure A tin methoxide co-catalyst is also required, presumably to convert the substrate into a tin enolate, which then adds nitrosobenzene.

    The O-nitroso aldol reaction of nitrosobenzene with enolizable aldehydes is promoted by the TMS ether of diphenylprolinol, using para-nitrobenzoic acid as a Brønsted acid co-catalyst, with ee of about 100%. The α-oxyaldehyde adducts produced are readily converted in situ to α-oxyimines, and thence to 1,2-aminoalcohols via treatment with Grignards, the latter process exhibiting des > 90%.¹⁹⁰ UnFigure UnFigure

    The Henry (Nitroaldol) Reaction

    A supramolecular chiral host, per-6-amino-β-cyclodextrin, gives ‘all-99s’ performance (yield, de, ee) for a Henry reaction in aqueous acetonitrile, and is readily recyclable without loss of activity.¹⁹¹ UnFigure UnFigure Thiourea, flanked by proline and cinchonidine substituents, gives up to 96% de and ee in conjugate addition of ketones/aldehydes to nitroalkenes.¹⁹² UnFigure UnFigure

    Copper catalysis is widely used. A high level of stereocontrol of three contiguous stereogenic centres has been achieved using a complex of copper(I) chloride and a chiral sulfonyldiamine in a Henry reaction of (R)-2-phenylpropanal and nitroethane.¹⁹³ UnFigure UnFigure Copper(II) complexes of chiral secondary diamines derived from 1,2-diaminocyclohexane catalyse reaction in 2-propanol at −30 °C in the presence of Hunig's base, i-Pr2NEt: examples of high yield/ee/de are recorded.¹⁹⁴ UnFigure UnFigure Combining diamine and bis(sulfonamide) auxiliary strategies, ligand (74)—in combination with copper(II)—gives yields/ee up to 99/99%.¹⁹⁵ UnFigure High syn-selectivity with a copper(II)-bisimidazoline has been rationalized in terms of the chiral environment around the metal, as seen in the X-ray structure.¹⁹⁶ UnFigure UnFigure

    UnFigure

    Tetramethylenediamine (TMEDA) catalyses the nitroaldol of a range of aldehyde types under mild, solvent-free conditions.¹⁹⁷ Two reviews cover advances in the asymmetric Henry reaction, focussing on organocatalysts¹⁹⁸ UnFigure and copper catalysts with chiral ligands.¹⁹⁹ UnFigure

    For other references to the Henry reaction, see under ‘Other Reactions of Imines’ above.

    The Baylis–Hillman Reaction and its Morita-variant

    The titanium-tetrachloride-promoted Baylis-Hillman reaction of methyl vinyl ketone and acetaldehyde in the absence of base has been studied by DFT, carefully dissecting the alternatives at each of the three main steps: chloride transfer to give a chloro-enolate, titanium-mediated aldol, and elimination of HCl or HOTiCl3.²⁰⁰ UnFigure

    All other reports deal with the Morita–Baylis–Hillman (MBH) reaction.

    An amino-acid-derived phosphino-thiourea catalyses an intramolecular reaction in up to 84% ee, converting ω-formyl-α,β-unsaturated carbonyl compounds to cyclic adducts.²⁰¹ UnFigure

    Brucine N-oxide and proline have been developed as a dual-catalyst system for asymmetric MBH reactions of vinyl ketones: the former activates the vinyl ketones to provide enolates via conjugate addition, while the proline forms iminium intermediates with electron-deficient aryl aldehydes.²⁰² UnFigure

    Enantioselective MBH reactions and their aza-variants have been carried out with trifunctional organocatalysts featuring a Brønsted acid and base and a Lewis base; counterion effects are significant.²⁰³ UnFigure

    The mechanism and stereoselectivity of the reaction between formaldehyde and methyl vinyl ketone has been investigated by DFT for N-methylprolinol (75), a bifunctional catalyst. Of the two steps—C UnFigure C bond formation and hydrogen migration—the latter is accelerated by water, leaving the former determining the rate and the stereochemistry.²⁰⁴ UnFigure

    A detailed mechanistic investigation highlights key deficiencies in the use of B3LYP calculations for this reaction and substitutes the MO6-2X DFT computational method.²⁰⁵ The failure to accelerate the reaction with higher temperatures has been explained by VT (variable temperature) experiments and MP2 calculations: the equilibrium shifts towards the reactants even with moderate increases in temperature. The authors also examine two key alternative mechanisms for proton transfer: Aggarwal's protic route and McQuade's aprotic one. They are found to be typically in competition, with the mechanistic balance depending on both the amount of protic species present and on the reaction progress (early or late stage). Phenol- and auto-catalysis are also accurately modelled.

    Enjoying the preview?
    Page 1 of 1