11
This journal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev., 2011, 40, 3405–3415 3405 Cite this: Chem. Soc. Rev., 2011, 40, 3405–3415 Metal-catalysed approaches to amide bond formation C. Liana Allen and Jonathan M. J. Williams* Received 30th November 2010 DOI: 10.1039/c0cs00196a Amongst the many ways of constructing the amide bond, there has been a growing interest in the use of metal-catalysed methods for preparing this important functional group. In this tutorial review, highlights of the recent literature have been presented covering the key areas where metal catalysts have been used in amide bond formation. Acids and esters have been used in coupling reactions with amines, but aldehydes and alcohols have also been used in oxidative couplings. The use of nitriles and oximes as starting materials for amide formation are also emerging areas of interest. The use of carbon monoxide in the transition metal catalysed coupling of amines has led to a powerful methodology for amide bond formation and this is complemented by the addition of an aryl or alkenyl group to an amide typically using palladium or copper catalysts. 1. Introduction The amide bond is one of the most important functional groups in contemporary chemistry. It is essential to sustain life, making up the peptide bonds in proteins such as enzymes. It is found in numerous natural products and it is also one of the most prolific moieties in modern pharmaceutical molecules. 1 Despite their obvious importance, the majority of amide bond syntheses involve the use of stoichiometric amounts of coupling reagents, making them generally expensive and wasteful procedures. 2 This discrepancy has encouraged efforts towards the identification and development of more atom-efficient, catalytic methods for amide bond formation, as evidenced by the increasing number of publications in the area in recent years. Currently, the most popular industrial methods of amide synthesis rely on activation of a carboxylic acid (using a coupling reagent such as a carbodiimide) and subsequent coupling of the activated species with an amine (Scheme 1). Although a huge amount of development has been devoted to fine tuning these coupling reagents for more efficient amide synthesis, this methodology still suffers from the inherent drawback of producing a stoichiometric amount of waste product along with the desired amide. 3 With this come the Scheme 1 Activation of a carboxylic acid. Department of Chemistry, University of Bath, Claverton Down, Bath, BA2 7AY, UK. E-mail: [email protected]; Fax: +44 (0)1225 386231; Tel: +44 (0)1225 383942 C. Liana Allen Liana Allen was born near Manchester, UK. She received her Masters in Chemistry for Drug Discovery at the University of Bath and is currently a PhD candidate at the same university under the supervision of Prof. Jonathan Williams. Liana’s current research interests are developing novel, efficient, Lewis acid catalysed syntheses of amide bonds and applying them to the synthesis of pharmaceutical molecules. Jonathan M. J. Williams Jonathan Williams was born in Stourbridge, England in 1964. He received a BSc from University of York, a DPhil. from University of Oxford (with Prof. S G Davies), and was then a post-doctoral fellow at Harvard with Prof. D. A. Evans (1989–1991). He was appointed to a Lectureship in Organic Chemistry at Loughborough University in 1991, and was then appointed as a Professor of Organic Chemistry at the University of Bath in 1996, where his research has mainly involved the use of transition metals for the catalysis of organic reactions. Chem Soc Rev Dynamic Article Links www.rsc.org/csr TUTORIAL REVIEW Downloaded by University of Oxford on 22 June 2011 Published on 17 March 2011 on http://pubs.rsc.org | doi:10.1039/C0CS00196A View Online

Citethis:Chem. Soc. Re.2011 0 ,34053415 TUTORIAL REVIEW€¦ · his ournal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev.,2011,40,34053415 3405 Citethis:Chem. Soc. Re.2011

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Citethis:Chem. Soc. Re.2011 0 ,34053415 TUTORIAL REVIEW€¦ · his ournal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev.,2011,40,34053415 3405 Citethis:Chem. Soc. Re.2011

This journal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev., 2011, 40, 3405–3415 3405

Cite this: Chem. Soc. Rev., 2011, 40, 3405–3415

Metal-catalysed approaches to amide bond formation

C. Liana Allen and Jonathan M. J. Williams*

Received 30th November 2010

DOI: 10.1039/c0cs00196a

Amongst the many ways of constructing the amide bond, there has been a growing interest in the

use of metal-catalysed methods for preparing this important functional group. In this tutorial

review, highlights of the recent literature have been presented covering the key areas where metal

catalysts have been used in amide bond formation. Acids and esters have been used in coupling

reactions with amines, but aldehydes and alcohols have also been used in oxidative couplings. The

use of nitriles and oximes as starting materials for amide formation are also emerging areas of

interest. The use of carbon monoxide in the transition metal catalysed coupling of amines has led

to a powerful methodology for amide bond formation and this is complemented by the addition

of an aryl or alkenyl group to an amide typically using palladium or copper catalysts.

1. Introduction

The amide bond is one of the most important functional

groups in contemporary chemistry. It is essential to sustain

life, making up the peptide bonds in proteins such as enzymes.

It is found in numerous natural products and it is also one of

the most prolific moieties in modern pharmaceutical molecules.1

Despite their obvious importance, the majority of amide bond

syntheses involve the use of stoichiometric amounts of

coupling reagents, making them generally expensive and wasteful

procedures.2 This discrepancy has encouraged efforts towards

the identification and development of more atom-efficient,

catalytic methods for amide bond formation, as evidenced

by the increasing number of publications in the area in

recent years.

Currently, the most popular industrial methods of amide

synthesis rely on activation of a carboxylic acid (using a

coupling reagent such as a carbodiimide) and subsequent

coupling of the activated species with an amine (Scheme 1).

Although a huge amount of development has been devoted to

fine tuning these coupling reagents for more efficient amide

synthesis, this methodology still suffers from the inherent

drawback of producing a stoichiometric amount of waste

product along with the desired amide.3 With this come the

Scheme 1 Activation of a carboxylic acid.

Department of Chemistry, University of Bath, Claverton Down, Bath,BA2 7AY, UK. E-mail: [email protected];Fax: +44 (0)1225 386231; Tel: +44 (0)1225 383942

C. Liana Allen

Liana Allen was born nearManchester, UK. She receivedher Masters in Chemistryfor Drug Discovery at theUniversity of Bath and iscurrently a PhD candidate atthe same university under thesupervision of Prof. JonathanWilliams. Liana’s currentresearch interests aredeveloping novel, efficient,Lewis acid catalysed synthesesof amide bonds and applyingthem to the synthesis ofpharmaceutical molecules. Jonathan M. J. Williams

Jonathan Williams was bornin Stourbridge, England in1964. He received a BSc fromUniversity of York, a DPhil.from University of Oxford(with Prof. S G Davies), andwas then a post-doctoralfellow at Harvard with Prof.D. A. Evans (1989–1991). Hewas appointed to a Lectureshipin Organic Chemistry atLoughborough University in1991, and was then appointedas a Professor of OrganicChemistry at the Universityof Bath in 1996, where his

research has mainly involved the use of transition metals forthe catalysis of organic reactions.

Chem Soc Rev Dynamic Article Links

www.rsc.org/csr TUTORIAL REVIEW

Dow

nloa

ded

by U

nive

rsity

of

Oxf

ord

on 2

2 Ju

ne 2

011

Publ

ishe

d on

17

Mar

ch 2

011

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C0C

S001

96A

View Online

Page 2: Citethis:Chem. Soc. Re.2011 0 ,34053415 TUTORIAL REVIEW€¦ · his ournal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev.,2011,40,34053415 3405 Citethis:Chem. Soc. Re.2011

3406 Chem. Soc. Rev., 2011, 40, 3405–3415 This journal is c The Royal Society of Chemistry 2011

further problems of increased costs and significant waste

by-products which need to be removed from the reaction

mixture and cannot be recycled. Enzymatic methods are also

available, although high isolation costs and somewhat limited

substrate ranges can be problematic.4

In the search for an alternative to coupling reagents and

enzymes in amide bond synthesis, non-metal catalysts such as

organocatalysts and boron reagents5 have been reported,

although these still often suffer from low atom efficiency and

difficult isolations. A possible solution to these drawbacks lies

with metal catalysis. Increasing attention is now being devoted

to developing such amide bond syntheses which are not only

atom-economical but also low cost and more environmentally

friendly. Employing metal catalysis in amide syntheses also

creates the possibility to start from substrates other than

carboxylic acids, opening up previously unavailable synthetic

routes to target molecules.

2. Amides from carboxylic acids

With the abundance of coupling reagents available for the

formation of amide bonds from carboxylic acids and amines,

there has been little development in the area of metal catalysts

for this reaction. The natural equilibrium of a carboxylic acid

and an amine is heavily towards the salt formation (except at

high temperatures), making a catalytic reaction between them

challenging.

However, the N-formylation of amines using formic acid as

the formylating agent has recently been reported to proceed

under catalytic conditions. Hosseini-Sarvari and Sharghi

published the first example of this in their highly efficient

reaction using ZnO as a catalyst under solvent-free conditions

at 70 1C, achieving some excellent yields in short reaction

times (Scheme 2).6 They also demonstrated the reusability of

the ZnO catalyst, incurring only a small decrease in yield of

amide after the third use.

Rao and co-workers later published their investigation into

a range of Lewis acid catalysts for the same reaction, reporting

dichloride complexes of zinc, tin, lanthanum, iron, aluminium

and nickel to give yields in the range of 80–100%.7 They found

the best results were obtained when using ZnCl2 as a catalyst

under the same solvent-free, 70 1C conditions that Hosseini-

Sarvari and Sharghi had used. Although the catalyst loading

could now be reduced from 50 mol% (ZnO) to just 10 mol%

(ZnCl2), there was no report of possible recovery and reuse of

Rao’s catalyst.

Kim and Jang have reported the use of indium metal as a

catalyst for the N-formylation of amines with formic acid,

again under solvent-free conditions at 70 1C.8 They found the

reaction to be efficiently catalysed by 10 mol% of the indium

metal, which they presume reacts with the formic acid to form

In(O2CH)3 and acts as a Lewis acid in the reaction.

3. Amides from esters

As the preparation of amides from carboxylic acids is difficult

to achieve in a catalytic manner, their derivatives, particularly

esters, have been explored as an alternative in catalytic amide

forming reactions.

In 2003, a simple procedure was published by Ranu and

Dutta, using a catalytic amount of indium triiodide and an

excess of the amine.9 The elimination of toxic reagents and

operational simplicity made this reaction a good alternative to

the methods known at that time. Several excellent yields were

reported for a range of amides containing functional groups

using their conditions, but the reaction was not successful with

secondary amines, making it unsuitable for the synthesis of

tertiary amides.

The same transformation was reported in 2005 by Gupta

et al., using zinc dust as a reusable catalyst under either

microwave or conventional heating.10 A modest range of

amides was synthesised, using only aromatic esters and amines

and again no demonstration of a tertiary amide synthesis.

Despite the reported substrate range being limited, their

procedure had the advantages of the zinc dust being able to

be reused up to six times (after washing with dilute HCl) with

only slight decreases in yield and a very short reaction time

(2–8 minutes) when microwave heating was used.

The same year, Porco and co-workers reported their

findings of group (IV) metal alkoxide complexes which, in

conjunction with an activator, could be used for the formation

of amides from esters and amines (Scheme 3).11 They demon-

strated an impressively varied substrate range, including

the formation of amides 1–3 along with an intramolecular

example giving 4.

The wide range of structurally diverse amines and esters

successfully coupled and the excellent yields attained under

their reaction conditions make this a valuable methodology.

A detailed mechanistic study was also carried out by the

group. Using X-ray crystallography and NMR studies, they

were able to determine the structures of key intermediates

Scheme 2 N-Formylation of an amine with formic acid.

Scheme 3 Zirconium catalysed coupling of esters and amines with

selected examples (HOAt = 1-hydroxy-7-azabenzotriazole, HOBt =

1-hydroxy-1H-benzotriazole, HYP = L-hydroxyproline).

Dow

nloa

ded

by U

nive

rsity

of

Oxf

ord

on 2

2 Ju

ne 2

011

Publ

ishe

d on

17

Mar

ch 2

011

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C0C

S001

96A

View Online

Page 3: Citethis:Chem. Soc. Re.2011 0 ,34053415 TUTORIAL REVIEW€¦ · his ournal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev.,2011,40,34053415 3405 Citethis:Chem. Soc. Re.2011

This journal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev., 2011, 40, 3405–3415 3407

along the reaction pathway and deduce the active catalyst to

be a dimeric zirconium complex (Scheme 4).

Dimeric species 5 in which both zirconium centres are

hexacoordinate is formed in the presence of amines. Coordi-

nation of an ester to one of the zirconium centres results in

formation of either 6 or 7 by breaking of one bridging Zr–O

bonds. Nucleophilic attack of the amine onto the ester then

proceeds via a six-membered (path A) or four-membered (path B)

transition state.

4. Amides from aldehydes or alcohols

Aldehydes and alcohols are desirable starting materials in

amide synthesis due to their ready availability and non-toxic

nature. In the last ten years, catalytic systems to effect this

transformation have been substantially developed into useful

organic reactions.

4.1 Amides from aldehydes [not via an oxime]

Oxidative amidation of aldehydes into amides has been known

since the early 1980s. The general mechanism of this process is

based on the reaction of an aldehyde with an amine to form a

hemiaminal intermediate and subsequent oxidation to the

amide product (Scheme 5). Loss of water from the hemiaminal

to form the imine, then hydrogenation of the imine to form an

amine is a potential side reaction in this process, one which has

been exploited in the ‘borrowing hydrogen’ reaction where

alcohols are used as alkylating agents for amines.12

The first catalyst system to be used for this transformation

was Pd(OAc)2 (5 mol%), triphenylphosphine (15 mol%),

potassium carbonate and an aryl bromide as the oxidant.13

Several aldehydes were successfully coupled with morpholine

to form the corresponding amides under these conditions.

Improved reaction conditions using another palladium catalyst

were reported by Torisawa and co-workers in 2008.14 Their

use of H2O2 in urea as the oxidant combined with PdCl2(2.5 mol%) and xantphos (2.5 mol%) allowed the reaction

temperature and required time to be reduced, as well as greatly

expanding the range of amides that could be synthesised in this

reaction.

A copper catalysed oxidative amidation has been reported

(Scheme 6).13 tert-Butyl hydroperoxide solution in water

serves as the oxidising agent and the use of amine hydro-

chloride salts minimises the competing reaction, oxidation of

the amine. Their yields decreased when an aliphatic or

electron-poor arylaldehyde was used, but when the reaction

was applied to an enantiomerically pure amine, no racemisation

occurred.

Recently, several lanthanide catalysts have been reported to

catalyse the oxidative coupling of aldehydes and amines. These

catalysts are generally capable of facilitating the reaction at

room temperature, which is a useful advantage. Additionally,

no external oxidant is required, as the aldehyde is presumed to

act as a hydrogen acceptor in the proposed catalytic cycle.

Marks and Seos’ lanthanide-amido complex La[N(TMS)2]3achieved varied yields of amides at room temperature in

deuterated benzene.13 Due to the aldehyde substrate also

acting as a hydrogen acceptor, a threefold excess was required.

Shen and co-workers have reported several bimetallic lanthanide

complexes used to catalyse this reaction. Their 2010 paper15

details the first structurally characterised complex of lanthanide

and lithium metals with dianionic guanidinate ligands,

with the Nd demonstrating its effectiveness for amidation of

aldehydes with amines at just 1 mol% catalyst loading at room

temperature.

4.2 Amides from alcohols [via an aldehyde]

The direct catalytic conversion of alcohols and amines into

amides and dihydrogen is a particularly desirable reaction due

to its high atom efficiency and widely available starting

materials. Murahasi and Naota reported the synthesis of

lactams from an intramolecular reaction of amino alcohols

in 1991,16 but the intermolecular reaction was realised by

Milstein and co-workers in their breakthrough report from

2007.17 Catalysed by a ruthenium pincer complex with molecular

hydrogen being the only by-product, this exceptionally clean

and simple reaction has been emulated by several other groups

since Milstein’s original publication (Scheme 8).

The general mechanism of this reaction is the same as that

of oxidative amidation involving aldehydes and amine, but

with an additional oxidation step at the start to convert the

alcohol into the aldehyde (Scheme 7).

Milstein’s PNN pincer complex 8 undergoes an aromatisation/

dearomatisation catalytic cycle, where initial addition of the

Scheme 4 Formation of key intermediate dimeric zirconium

complexes.

Scheme 5 General mechanism of amide formation from aldehydes

and amines.

Scheme 6 Copper catalysed oxidative amidation.

Dow

nloa

ded

by U

nive

rsity

of

Oxf

ord

on 2

2 Ju

ne 2

011

Publ

ishe

d on

17

Mar

ch 2

011

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C0C

S001

96A

View Online

Page 4: Citethis:Chem. Soc. Re.2011 0 ,34053415 TUTORIAL REVIEW€¦ · his ournal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev.,2011,40,34053415 3405 Citethis:Chem. Soc. Re.2011

3408 Chem. Soc. Rev., 2011, 40, 3405–3415 This journal is c The Royal Society of Chemistry 2011

alcohol leads to aromatisation giving the pyridine complex 9.

Subsequent loss of an aldehyde generates the known trans

ruthenium dihydride complex 10. Elimination of dihydrogen

regenerates the catalyst 8 and enables the catalytic cycle to

continue (Scheme 9). The aldehyde forms an aminol by

reaction with the amine and a similar cycle oxidises this to

the amide product.

Alternative ruthenium catalyst systems employing ruthenium

precursors in combination with N-heterocyclic carbenes have

been found to give exclusively the amide product (as opposed

to the amine) as have been reported by the groups of Madsen18

and Hong.19 Although this was the first step towards the use of

a simple, commercially available catalyst in this reaction, there

was no real improvement in terms of yields or scope of

reagents on those reported by Milstein. The first commercially

available catalyst system for formation of amides from alcohols

and amines was reported by our own group (although

Milstein’s catalyst is now commercially available).20 The use

of [Ru(p-cymene)Cl2]2 in combination with dppb, a base and a

hydrogen acceptor produced the amide products in reasonable

to good yields, however required increased reaction times

compared with alternative catalysts.

An example of primary amide formation in this reaction was

demonstrated by the group of Grutzmacher.21 Using a

rhodium catalyst and ammonia as the nitrogen source,

primary amides were obtained in excellent yields in just four

hours and at a temperature of �30 1C to 25 1C.

These catalysts, though efficient and highly chemoselective

in favour of amide formation, can be expensive, difficult to

handle and do not tolerate secondary amines well. Some of

these issues were addressed by Satsuma and co-workers in

their reported g-alumina supported silver cluster.22 This

re-usable, heterogeneous, easily prepared catalyst is a more

economic alternative to the homogenous ruthenium and

rhodium catalysts reported. In addition to this, secondary

amines could be used in the reaction, giving tertiary amides

in very good yields.

5. Amides from nitriles

Nitriles are well recognised as important substrates in organic

chemistry due to their chemical versatility, allowing addition

to the CRN triple bond by nucleophiles or electrophiles to

lead to new C–N, C–C and C–O bonds. Their use in the

synthesis of amides has, however, been somewhat limited to

the three reactions discussed below.

5.1 Hydration of nitriles into primary amides

A simple and efficient method of synthesising primary amides

is the hydration of nitriles. Many metal catalysts have

been reported to facilitate this transformation efficiently, an

excellent and in depth review of which was pub lished by

Kukushkin and Pombeiro in 2005.23 More recently, catalysts

which allow the hydration of organonitriles under ambient

conditions24 or in reaction times as short as 2 hours using

microwave radiation25 have been published (Scheme 10).

5.2 Coupling of nitriles with amines

A little known reaction which yields amides is the hydrolytic

amidation of nitriles with amines. This was first published in

1986 by Murahashi and co-workers using the ruthenium

catalyst RuH2(PPh3)4 (Scheme 11).26 They demonstrated the

wide scope of this reaction with the efficient synthesis of

several drug precursors (including 11) as well as lactams from

the intramolecular reaction and diamide 12 from cinnamonitrile

and a triamine.

Development of this reaction has been sporadic; in 2000 a

platinum catalyst was found by de Vries and co-workers to

Scheme 7 General mechanism of amide formation from alcohols and

amines.

Scheme 8 Milstein’s catalyst for conversion of alcohols into amides.

Scheme 9 Catalytic cycle for Milstein’s catalyst. Scheme 10 Rhodium and gold catalysed nitrile hydration.

Dow

nloa

ded

by U

nive

rsity

of

Oxf

ord

on 2

2 Ju

ne 2

011

Publ

ishe

d on

17

Mar

ch 2

011

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C0C

S001

96A

View Online

Page 5: Citethis:Chem. Soc. Re.2011 0 ,34053415 TUTORIAL REVIEW€¦ · his ournal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev.,2011,40,34053415 3405 Citethis:Chem. Soc. Re.2011

This journal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev., 2011, 40, 3405–3415 3409

perform the coupling,27 then in 2009 an iron catalysed version

was published by our group.28 The former decreased the

catalyst loading from 3 mol% with the ruthenium catalyst to

levels as low as 0.1 mol% with the platinum(II) complex, but

the reaction temperature remained high at 160 1C and the

yields were moderate. Our own iron catalysed reaction allowed

the reaction to be performed at 125 1C under solvent-free

conditions, but a higher catalyst loading (10 mol%) was

required as well as an excess of one of the reagents.

Brief mechanistic investigations suggest that the involvement

of an amidine intermediate formed from nucleophilic attack of

the amine on the nitrile seems likely, as opposed to initial

hydration of the nitrile to the primary amide thenN-alkylation

(Scheme 12).

De Vries and co-workers found that when the reaction is

run in the absence of water, the major product isolated is the

amidine. The possibility of the reaction proceeding through

initial hydration of the nitrile to the primary amide then

subsequent N-acylation of the amine was ruled out after we

found only a 21% conversion into secondary amide in the

reaction between butyramide and benzylamine.

5.3 Coupling of nitriles with alcohols

Nitriles can also be coupled with alcohols to form amides in

the Ritter reaction. As an alternative to sulfuric acid, the

Ritter reaction can be catalysed by metal complexes. One of

the first examples of this was the use of bismuth triflate by

Barrett and co-workers who published a wide range of amides

synthesised from coupling various nitriles and tertiary alcohols.29

A later example by Cossy’s group described an iron-catalysed

Ritter reaction with a wider substrate range than those

previously reported, but the reaction conditions are less

desirable (Scheme 13).30

6. Amides from oximes

Oximes have been used in organic synthesis since the 19th

century in a diverse range of reactions. Their applications

in metal catalysed amide bond synthesis have been well

documented in the areas of aldoxime rearrangement into a

primary amide and the Beckmann rearrangement of ketoximes.

Recently, aldoximes have also been shown to react with

amines to form secondary and tertiary amides.

6.1 Rearrangement of aldoximes into primary amides

The rearrangement of aldoximes into primary amides has been

shown to be catalysed by several metal complexes. This highly

atom-efficient reaction also offers the option to start from an

aldehyde and hydroxylamine (which forms the aldoxime

in situ) (Scheme 14), or even the alcohol oxidation state.

The examples of catalysts suitable for this rearrangement

are largely based on precious metals. Chang31 and Mizuno32

have independently published reports of rhodium complexes,

the latter being an example of a supported, reusable catalyst.

Ruthenium catalysts have also been reported by both

Crabtree33 and our group to perform this rearrangement.34

Another report by our group demonstrates the potential to

start from the alcohol with the first step of the mechanism then

being oxidation of the alcohol to the aldehyde, then condensation

with hydroxylamine and subsequent rearrangement to the

primary amide (Scheme 15).35 The oxidation step was

conveniently catalysed by the same iridium catalyst used for

the rearrangement and use of styrene as a sacrificial hydrogen

acceptor. These current conditions are not particularly desirable,

Scheme 11 Murahashi’s ruthenium-catalysed hydrolytic coupling of

nitriles with amines.

Scheme 12 Proposed mechanism pathways for the addition of amines

to nitriles.

Scheme 13 Iron-catalysed Ritter reaction.

Scheme 15 Formation of primary amides from alcohols via

aldoximes.

Scheme 14 Rearrangement of aldoximes into primary amides starting

from an aldehyde.

Dow

nloa

ded

by U

nive

rsity

of

Oxf

ord

on 2

2 Ju

ne 2

011

Publ

ishe

d on

17

Mar

ch 2

011

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C0C

S001

96A

View Online

Page 6: Citethis:Chem. Soc. Re.2011 0 ,34053415 TUTORIAL REVIEW€¦ · his ournal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev.,2011,40,34053415 3405 Citethis:Chem. Soc. Re.2011

3410 Chem. Soc. Rev., 2011, 40, 3405–3415 This journal is c The Royal Society of Chemistry 2011

but the novel concept of transformation of an alcohol into a

primary amide via an oxime is clearly demonstrated.

The most recent advances on this reaction have introduced

more catalysts; a gold/silver co-catalysed system (Nolan et al.),36

a palladium acetate complex (Ali and Punniyamurthy)37 and

most recently the use of simple metal salts InCl3 or Zn(NO3)2at low catalyst loadings (0.1 mol% and 10 mol% respectively)

have been published by our own group, representing the most

cost-efficient catalysts reported so far.38

6.2 Beckmann rearrangement of ketoximes into secondary

amides

The Beckmann rearrangement of ketoximes into amides is a

powerful methodology in organic synthesis. Traditionally this

reaction requires harsh conditions such as high temperatures

and strong acids, although recently, metal catalysed Beckmann

rearrangements have been published, including conditions

such as using an ionic liquid medium, or performing the

reaction in the vapour phase. A notable metal catalysed

varient has been reported by Ramalingan and Park and details

a mercury(II) chloride catalysed reaction, run in acetonitrile at

80 1C (Scheme 16).39 A wide range of ketoximes was transformed

into the corresponding secondary amides under their conditions,

including several halogen substituted amides including 13 and a

cyclic ketoxime yielding caprolactam 14 as the product.

A further example of a metal catalysed Beckmann

rearrangement uses a novel heterobimetallic catalyst with a

cobalt centre and Lewis acidic zinc site on the periphery.40

Reasonable yields were achieved in a reaction time of 2 hours

at 5 mol% catalyst loading.

6.3 Coupling of aldoximes and amines

Our own recent paper on amide bond syntheses from aldoximes

and amines describes the novel synthesis of amides from the

coupling of aldoximes and amines.41 This reaction can also be

run with the aldehyde and hydroxylamine forming the oxime

in situ, catalysed by the simple metal salt NiCl2, present

at 5 mol% in the reaction mixture (Scheme 17). As an additional

oxidising agent is not required, this method of coupling

aldehydes and amines is a more atom-efficient method

compared with other conditions reported for this coupling.

Based on the knowledge that a nitrile is an intermediate in

the rearrangement of aldoximes into primary amides and

amines and nitriles can couple to form secondary and tertiary

amides, we proposed a reaction mechanism whereby the nitrile

intermediate was intercepted by an amine.Further mechanistic investigations using 18O labelled

oximes suggested that a bimolecular mechanism was operating

(Scheme 18). A nickel metallocycle has been suggested to be a

key intermediate in the mechanism, susceptible to nucleophilic

attack in the presence of an amine.

7. Aminocarbonylation

The combination of amines with carbon monoxide has been

used in the catalytic aminocarbonylation of a variety of

substrates to give amide products.42 Schoenberg and Heck

published the first examples of the palladium-catalysed

aminocarbonylation of aryl halides and vinyl halides over

35 years ago.43 Murahashi and co-workers later reported aScheme 16 Mercury catalysed Beckmann rearrangement of

ketoximes.

Scheme 17 Nickel-catalysed formation of amides from aldehydes via

aldoximes.

Scheme 18 Proposed mechanism of aldoxime rearrangement into

primary amide.

Dow

nloa

ded

by U

nive

rsity

of

Oxf

ord

on 2

2 Ju

ne 2

011

Publ

ishe

d on

17

Mar

ch 2

011

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C0C

S001

96A

View Online

Page 7: Citethis:Chem. Soc. Re.2011 0 ,34053415 TUTORIAL REVIEW€¦ · his ournal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev.,2011,40,34053415 3405 Citethis:Chem. Soc. Re.2011

This journal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev., 2011, 40, 3405–3415 3411

palladium catalysed carbonylation of allylamines to give

b,g-unsaturated amides, avoiding the use of a halide species

in the reaction. They proposed a p-allylpalladium complex,

formed from oxidative addition of allylamines to a co-

ordinatively unsaturated palladium–phosphine species, as a

key intermediate in the mechanism of their reaction.44

More recent reports include the use of a Pd(OAc)2/xantphos

catalyst developed by Buchwald and co-workers.45 For

example, aryl bromides such as 1-bromonaphthalene 15 and

3-bromothiophene 16 were converted into the corresponding

amide 17 or Weinreb amide 18 using the Pd(OAc)2/xantphos

catalyst (Scheme 19). These coupling reactions involve the

oxidative addition of the active Pd(0) catalyst into the aryl

halide, ligand combination with CO to give an acylated

palladium species which is then converted into amide by the

addition of amine either directly or by prior co-ordination to

the metal.

Amongst the many other reports of aminocarbonylation of

aryl halides, Beller and co-workers have prepared CNS active

amphetamine derivatives 19 by the three component coupling

of unprotected 5-bromoindole, carbon monoxide and a

piperazine (Scheme 20).46 Aminocarbonylation reactions

have also been performed in water using Pd(OAc)2 under

ligand-free conditions.47

The direct synthesis of primary amides by aminocarbonylation

can be problematic due to the handling difficulties of ammonia,

its low nucleophilicity and ability to complex strongly with

palladium. Several indirect approaches for the formation of

primary amides have been developed, including the use of

tert-butylamine as an ammonia equivalent followed by

deprotection.48 Using this procedure, iodobenzene was

converted into benzamide after removal of the t-Bu group

with TBDMSOTf (tert-butyldimethylsilyl triflate) as shown in

Scheme 21. Formamide has also been used as an ammonia

equivalent for the formation of primary aromatic amides

from aryl halides.49 However, Beller and co-workers have

demonstrated that careful selection of a phosphine ligand does

provide a catalytic system capable of converting aryl bromides

with ammonia and CO directly into the corresponding

primary amides.50 The use of Pd(OAc)2 in combination with

dppf was effective for a range of aryl bromides and, under

more forcing conditions, aryl chlorides. For example,

4-bromotoluene was converted into toluamide in good yield

(Scheme 21).

Aryl chlorides are usually cheaper alternatives to other aryl

halides although they require more forcing conditions to

undergo aminocarbonylation. Milstein and co-workers

reported the first examples of the use of aryl chlorides in

aminocarbonylation reactions where electron-rich bidentate

phosphines were found to improve the reactivity of the

palladium catalyst.51 Buchwald and co-workers have

developed an interesting and practical approach to the amino-

carbonylation of aryl chlorides which proceeds via an inter-

mediate phenyl ester.52 The intermediate palladium acyl

complex 20 reacts with sodium phenoxide which is a better

nucleophile than the amine. The so-formed phenyl ester 21

then reacts with the amine in a subsequent step which does not

involve the palladium catalyst. This approach was used for the

conversion of a range of aryl chlorides into secondary and

tertiary amines. In one example, the reaction of aryl chloride

22 with morpholine gave the tertiary amide 23 (Scheme 22).

Mo(CO)6 has been used as a solid source of CO for the

aminocarbonylation reactions of aryl triflates using a palla-

dium catalyst, with DMAP (N,N-dimethylaminopyridine) as a

co-catalyst which is believed to catalyse acyl transfer from the

intermediate acylpalladium species.53 N,N-Dimethylformamide

has also been used as a substrate to allow aminoformylation

reactions to take place without the need for CO.54 The

N,N-dimethylformamide is activated by the addition of POCl3and then used to convert aryl iodides into the N,N-dimethyl-

amide products (Scheme 23).

Scheme 19 Pd-catalysed aminocarbonylation of aryl bromides.

Scheme 20 Aminocarbonylation of an unprotected indole. Scheme 21 Synthesis of primary amides.

Dow

nloa

ded

by U

nive

rsity

of

Oxf

ord

on 2

2 Ju

ne 2

011

Publ

ishe

d on

17

Mar

ch 2

011

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C0C

S001

96A

View Online

Page 8: Citethis:Chem. Soc. Re.2011 0 ,34053415 TUTORIAL REVIEW€¦ · his ournal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev.,2011,40,34053415 3405 Citethis:Chem. Soc. Re.2011

3412 Chem. Soc. Rev., 2011, 40, 3405–3415 This journal is c The Royal Society of Chemistry 2011

Aryl substrates which do not contain a halide or other

suitable leaving group are able to undergo palladium-

catalysed coupling reactions by C–H insertion, but usually

require the presence of an oxidant in order to allow for

catalyst recycling. Orito and co-workers have demonstrated

that substrates such as the amine-containing arene 24 can

undergo direct aromatic carbonylation to give the benzolactam

product 25.55 The reaction is catalysed by Pd(OAc)2 in the

presence of Cu(OAc)2 and air which re-oxidises the catalyst

(Scheme 24).

Alkynes can also undergo aminocarbonylation reactions

and this methodology is completely atom-efficient. In one

example, Matteoli and co-workers have coupled phenylacetylene

with aniline and CO to give the acrylamide product 26 with

almost complete regioselectivity for the branched product

(Scheme 25).56 Lu and Alper have reported a related example

using a supported palladium catalyst to cyclise an aniline onto

a pendant alkyne.57

The aminocarbonylation of alkenes has been less widely

explored, although the three component coupling of an alkene,

an amine and CO has considerable potential as an atom

efficient synthesis of saturated amides. Chung and co-workers

have used a heterogeneous cobalt catalyst for the amino-

carbonylation of 1-pentene with aniline to give the amide 27

(Scheme 26).58 However, high temperatures and long

reaction times were required and some alkenes led to lower

conversions. An interesting double aminocarbonylation of

alkynes using a rhodium catalyst has been reported by Huang

and Hua.59 Using phenylacetylene as the substrate, the

reaction was believed to occur by initial aminocarbonylation

of the alkyne with pyrrolidine to give a branched acrylamide

followed by a second aminocarbonylation leading to the

1,4-diamide 28 shown in Scheme 26. The reaction was also

successful for aliphatic alkynes.

Using isocyanates in place of the amine and CO, Schleicher

and Jamison used a nickel catalyst complexed to the

N-heterocyclic carbene ligand IPr to convert simple alkenes

into the corresponding acrylamides.60 Although some alkene/

isocyanate combinations led to mixtures of branched and

unbranched products, the use of vinylcyclohexane with

tert-butylisocyanate gave the branched acrylamide 29 with

complete selectivity and good isolated yield (Scheme 27).

8. N-Arylation and N-alkenylation of amides

The cross coupling of amides with aryl and alkenyl halides is

an important process that has been widely used for the

preparation of amides of pharmaceutical interest.61 These

reactions have mainly been achieved using palladium or

copper catalysts, although other metals have been shown to

be effective, including simple iron salts.62

8.1 Palladium-catalysed reactions

Buchwald and co-workers have developed a palladium

catalyst that has shown broad scope for the N-arylation of

amides wth aryl chlorides, aryl tosylates and aryl nonaflates

(ArONf = ArOSO2CF2CF2CF2CF3).63 Typical examples

Scheme 22 Aminocarbonylation of aryl chlorides.

Scheme 23 Use of DMF in aminocarbonylation.

Scheme 24 Direct aromatic carbonylation.

Scheme 25 Aminocarbonylation of an alkyne.

Scheme 26 Cobalt and rhodium catalysed aminocarbonylation.

Scheme 27 Aminocarbonylation using an isocyanate, IPr = 1,3-

bis(2,6-diisopropylphenyl)imidazol-2-ylidene.

Dow

nloa

ded

by U

nive

rsity

of

Oxf

ord

on 2

2 Ju

ne 2

011

Publ

ishe

d on

17

Mar

ch 2

011

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C0C

S001

96A

View Online

Page 9: Citethis:Chem. Soc. Re.2011 0 ,34053415 TUTORIAL REVIEW€¦ · his ournal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev.,2011,40,34053415 3405 Citethis:Chem. Soc. Re.2011

This journal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev., 2011, 40, 3405–3415 3413

include the use of [Pd(allyl)Cl]2 with JackiePhos for the

coupling of aryl nonaflate 30 with amide 31 and the aryl

chloride 33 with amide 34 (Scheme 28). The electron-with-

drawing nature of the trifluoromethyl groups in JackiePhos

was important for achieving full conversion under the reaction

conditions.

The more readily available ligand, xantphos (see

Scheme 19), has also been shown to be useful for other

palladium-catalysed cross coupling reactions of amides.

Wallace and co-workers have used the Pd(0)/xantphos

combination for the coupling of vinyl triflates with amides.64

Vinyl triflate 36 was coupled with acetamide to give the

product 37 with good isolated yield (Scheme 29). An interesting

tandem process involving N-alkenylation and N-arylation was

designed by Willis and co-workers for the synthesis of

indoles.65 When substrate 38 was reacted with benzamide, a

double cross coupling led to the formation of the N-benzoy-

lindole 39 (Scheme 29).

The functional group tolerance of the palladium-catalysed

arylation of amides is apparent from the work of a team of

chemists from GlaxoSmithKline who reported the coupling of

amide 40 with a range of heterocyclic triflates and bromides

including triflate 41 to give the arylated product 42

(Scheme 30).66

An alternative approach to the formation of amides by

palladium-catalysed cross coupling involving the use of

anilines was developed by Xu and co-workers.67 The gem-

dibromoalkene 43 underwent palladium-catalysed amination

where the initially-formed bromoenamine is readily hydrolysed

to liberate the amide 44 which was isolated in good yield

(Scheme 31).

8.2 Copper-catalysed reactions

The copper-catalysed arylation of amides offers a cheaper

alternative to the palladium-catalysed reactions. Typically, a

copper(I) salt is used in combination with a diamine, such as 45

or 46, or a related ligand.68 Buchwald and co-workers have

reported several related CuI/diamine catalysts which are effective

for the N-arylation of amides with aryl halides. Representative

examples for the formation of amides 47 and 48 are given in

Scheme 28 Palladium-catalysed N-arylation reactions.

Scheme 29 Use of palladium/xantphos for N-arylation.

Scheme 30 Synthesis of potential anti-bacterial agents.

Scheme 31 Amide formation by amination/hydrolysis.

Scheme 32 Copper-catalysed N-arylation of aryl halides.

Dow

nloa

ded

by U

nive

rsity

of

Oxf

ord

on 2

2 Ju

ne 2

011

Publ

ishe

d on

17

Mar

ch 2

011

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C0C

S001

96A

View Online

Page 10: Citethis:Chem. Soc. Re.2011 0 ,34053415 TUTORIAL REVIEW€¦ · his ournal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev.,2011,40,34053415 3405 Citethis:Chem. Soc. Re.2011

3414 Chem. Soc. Rev., 2011, 40, 3405–3415 This journal is c The Royal Society of Chemistry 2011

Scheme 32. The N-arylation of amides could also be achieved

using aryl chlorides, although a four-fold excess of the chloride

was required under solvent-free conditions.

The copper-catalysed coupling of vinyl halides with amides

is also known,69,70 which has been exploited in an interesting

example reported by the group of Li.71 N-Arylpyrroles were

formed by the tandem coupling of an amide with dienyl-

diiodide 49. Using pentanamide, the acylpyrrole 50 was

isolated in 95% yield (Scheme 33), although yields were lower

in other reported examples.

Potassium alkenyltrifluoroborate salts have also been used

for the N-alkenylation of amides, as reported by Bolshan and

Batey.72 These reactions occur under oxidative conditions with

lower temperatures than those needed by vinyl halides. For

example, various amides undergo copper-catalysed coupling

with potassium hexenyltrifluoroborate at 40 1C (Scheme 34).

9. Enamide formation by addition reactions

The transition metal catalysed addition of amides to alkynes

provides a useful approach to the preparation of enamides.

Gooßen and co-workers identified Ru(methallyl)(cod) with

n-Bu3P and DMAP as an efficient catalyst for the selective

formation of the (E)-enamide 51 from the coupling partners

1-hexyne and N-methylformamide.73 None of the branched

product arising from Markovnikov addition to the alkyne was

observed. The reaction was mainly used for the addition of

lactams to alkynes. RuCl3 could be used as an alternative

ruthenium source under otherwise similar reaction conditions,74

while the use of the rhenium catalyst Re2(CO)10 was also

found to be highly regio- and stereoselective (Scheme 35).75

The oxidative coupling of amides with conjugated alkenes

under Wacker-type conditions has been found to lead to the

selective formation of (Z)-enamides.76 Using a palladium

catalyst which required a copper co-catalyst for re-oxidation,

benzamide reacted with ethyl acrylate to give the (Z)-enamide

52. The selective formation of the (Z)-enamide was attributed

to hydrogen bonding between the ester carbonyl oxygen and

the amide NH in the intermediate (Scheme 36).

10. Conclusions

In conclusion, there has been impressive progress in the use of

metal catalysts for the formation of amides. The range of

available methods allows the synthetic chemist to choose from

a variety of starting materials for the construction of the

amide bond.

Acknowledgements

We thank the Engineering and Physical Sciences Research

Council for financial support (for C. L. A.) through the

Doctoral Training Account.

References

1 J. S. Carey, D. Laffan, C. Thomson and M. T. Williams, Org.Biomol. Chem., 2006, 4, 2337–2347.

2 D. J. C. Constable, P. J. Dunn, J. D. Hayler, G. R. Humphrey,J. L. Leazer, R. J. Linderman, K. Lorenz, J. Manley,B. A. Pearlman, A. Wells, A. Zaks and T. Y. Zhang, Green Chem.,2007, 9, 411–420.

3 E. Valeur and M. Bradley, Chem. Soc. Rev., 2009, 38, 606–631.4 M. X. Wang, Top. Catal., 2005, 35, 117–130.5 J. E. Anderson, R. Davis, R. N. Fitzgerald and J. M. Haberman,Synth. Commun., 2006, 36, 2129–2133.

6 M. Hosseini-Sarvari and H. Sharghi, J. Org. Chem., 2006, 71,6652–6654.

7 A. C. Shekhar, A. R. Kumar, G. Sathaiah, V. L. Paul, M. Sridharand P. S. Rao, Tetrahedron Lett., 2009, 50, 7099–7101.

8 J.-G. Kim and D. O. Jang, Synlett, 2000, 1231–1234.9 B. C. Ranu and P. Dutta, Synth. Commun., 2003, 33, 297–301.10 R. Arora, S. Paul and R. Gupta, Can. J. Chem., 2005, 83,

1137–1140.11 C. Han, J. P. Lee, E. Lobkovsky and J. A. Porco, J. Am. Chem.

Soc., 2005, 127, 10039–10044.12 M. Hamid, C. L. Allen, G. W. Lamb, A. C. Maxwell,

H. C. Maytum, A. J. A. Watson and J. M. J. Williams, J. Am.Chem. Soc., 2009, 131, 1766–1774.

13 K. Ekoue-Kovi and C. Wolf, Chem.–Eur. J., 2008, 14, 6302–6315and references therein.

14 Y. Suto, N. Yamagiwa and Y. Torisawa, Tetrahedron Lett., 2008,49, 5732–5735.

15 C. Qian, X. Zhang, Y. Zhang and Q. Shen, J. Organomet. Chem.,2010, 695, 747–752 and references therein.

16 T. Naota and S.-I. Murahasi, Synlett, 1991, 693–694.17 C. Gunanathan, Y. Ben-David and D. Milstein, Science, 2007, 317,

790–792.18 L. U. Nordstrøm, H. Vogt and R. Madsen, J. Am. Chem. Soc.,

2008, 130, 17672–17674.19 Y. Zhang, C. Chen, S. C. Ghosh, Y. Li and S. H. Hong, Organo-

metallics, 2010, 29, 1374–1378.20 A. J. A. Watson, A. C. Maxwell and J. M. J. Williams, Org. Lett.,

2009, 11, 2667–2670.21 T. Zweifel, J.-V. Naubron and H. Grutzmacher, Angew. Chem.,

Int. Ed., 2009, 48, 559–563.

Scheme 33 Copper-catalysed N-arylation of aryl halides.

Scheme 34 Copper-catalysed N-arylation of aryl halides.

Scheme 35 Ru-catalysed addition of an amide to an alkyne.

Scheme 36 Oxidative coupling of an amide and alkene.

Dow

nloa

ded

by U

nive

rsity

of

Oxf

ord

on 2

2 Ju

ne 2

011

Publ

ishe

d on

17

Mar

ch 2

011

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C0C

S001

96A

View Online

Page 11: Citethis:Chem. Soc. Re.2011 0 ,34053415 TUTORIAL REVIEW€¦ · his ournal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev.,2011,40,34053415 3405 Citethis:Chem. Soc. Re.2011

This journal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev., 2011, 40, 3405–3415 3415

22 K. Shimizu, K. Ohshima and A. Satsuma, Chem.–Eur. J., 2009, 15,9977–9980.

23 V. Y. Kukushkin and A. J. L. Pombeiro, Inorg. Chim. Acta, 2005,358, 1–21.

24 A. Goto, K. Endo and S. Saito, Angew. Chem., Int. Ed., 2008, 47,3607–3609.

25 R. S. Ramon, N. Marion and S. P. Nolan, Chem.–Eur. J., 2009, 15,8695–8697 and ref. 8–18 therein.

26 S.-I. Murahashi, T. Naota and E. Saito, J. Am. Chem. Soc., 1986,108, 7846–7847.

27 C. J. Cobley, M. van den Heuvel, A. Abbadi and J. G. de Vries,Tetrahedron Lett., 2000, 41, 2467–2470.

28 C. L. Allen, A. A. Lapkin and J. M. J. Williams, Tetrahedron Lett.,2009, 50, 4262–4264.

29 E. Callens, A. J. Burton and A. G. M. Barrett, Tetrahedron Lett.,2006, 47, 8699–8701.

30 B. Anxionnat, A. Guerinot, S. Reymond and J. Cossy, TetrahedronLett., 2009, 50, 3470–3473.

31 M. Kim, J. Lee, H.-Y. Lee and S. Chang, Adv. Synth. Catal., 2009,351, 1807–1812.

32 H. Fujiwara, Y. Ogasawara, M. Kotani, K. Yamaguchi andN. Mizuno, Chem.–Asian J., 2008, 3, 1715–1721.

33 D. Gnanamgari and R. H. Crabtree, Organometallics, 2009, 28,922–924.

34 N. A. Owston, A. J. Parker and J. M. J. Williams, Org. Lett., 2007,9, 3599–3601.

35 N. A. Owston, A. J. Parker and J. M. J. Williams, Org. Lett., 2007,9, 73–75.

36 R. S. Ramon, J. Bosson, S. Diez-Gonzalez, N. Marion andS. P. Nolan, J. Org. Chem., 2010, 75, 1197–1202.

37 M. A. Ali and T. Punniyamurthy, Adv. Synth. Catal., 2010, 352,288–292.

38 C. L. Allen, C. Burel and J. M. J. Williams, Tetrahedron Lett.,2010, 51, 2724–2726.

39 C. Ramalingan and Y.-T. Park, J. Org. Chem., 2007, 72,4536–4538.

40 A. Mishra, A. Ali, S. Upreti and R. Gupta, Inorg. Chem., 2008, 47,154–161 and ref. 26,27 therein.

41 C. L. Allen, S. Davulcu and J. M. J. Williams, Org. Lett., 2010, 12,5096–5099.

42 A. Brennfuhrer, H. Neumann and M. Beller, Angew. Chem., Int.Ed., 2009, 48, 4114–4133.

43 A. Schoenberg and R. F. Heck, J. Org. Chem., 1974, 39,3327–3331.

44 S.-I. Murahashi, Y. Imada and K. Nishimura, J. Chem. Soc.,Chem. Commun., 1988, 1578–1579.

45 J. R. Martinelli, D. A. Watson, D. M.M. Freckmann, T. E. Barderand S. L. Buchwald, J. Org. Chem., 2008, 73, 7102–7107.

46 K. Kumar, A. Zapf, D. Michalik, A. Tillack, T. Heinrich,H. Bottcher, M. Arlt and M. Beller, Org. Lett., 2004, 6,7–10.

47 P. J. Tambade, Y. P. Patil, M. J. Bhanushali and B. M. Bhanage,Synthesis, 2008, 2347.

48 E. Takacs, C. Varga, R. Skoda-Foldes and L. Kollar, TetrahedronLett., 2007, 48, 2453–2456.

49 A. Schnyder, M. Beller, G. Mehltretter, T. Nsenda, M. Studer andA. F. Indolese, J. Org. Chem., 2001, 66, 4311–4315.

50 X.-F. Wu, H. Neumann and M. Beller, Chem.–Asian J., 2010, 5,2168–2172.

51 Y. Ben-David, M. Portnoy and D. Milstein, J. Am. Chem. Soc.,1989, 111, 8742–8744.

52 J. R. Martinelli, T. P. Clark, D. A. Watson, R. H. Munday andS. L. Buchwald, Angew. Chem., Int. Ed., 2007, 46, 8460–8463.

53 L. R. Odell, J. Savmaker and M. Larhed, Tetrahedron Lett., 2008,49, 6115–6118.

54 K. Hosoi, K. Nozaki and T. Hiyama, Org. Lett., 2002, 4,2849–2851.

55 K. Orito, A. Horibata, T. Nakamura, H. Ushito, H. Nagasaki,M. Yuguchi, S. Yamashita and M. Tokuda, J. Am. Chem. Soc.,2004, 126, 14342–14343.

56 U. Matteoli, A. Scrivanti and V. Beghetto, J. Mol. Catal. A:Chem., 2004, 213, 183–186.

57 S.-M. Lu and H. Alper, J. Am. Chem. Soc., 2008, 130, 6451–6455.58 S. I. Lee, S. U. Son and Y. K. Chung, Chem. Commun., 2002,

1310–1311.59 Q. Huang and R. Hua, Adv. Synth. Catal., 2007, 349, 849–852.60 K. D. Schleicher and T. F. Jamison, Org. Lett., 2007, 9, 875–878.61 B. P. Fors, K. Dooleweerdt, Q. Zeng and S. L. Buchwald, Tetra-

hedron, 2009, 65, 6576–6583 and references therein.62 A. Correa, S. Elmore and C. Bolm, Chem.–Eur. J., 2008, 14,

3527–3529.63 J. D. Hicks, A. M. Hyde, A. Martinez Cuezva and S. L. Buchwald,

J. Am. Chem. Soc., 2009, 131, 16720–16734.64 D. J. Wallace, D. J. Klauber, C.-y. Chen and R. P. Volante, Org.

Lett., 2003, 5, 4749–4752.65 M. C. Willis, G. N. Brace and I. P. Holmes, Angew. Chem., Int.

Ed., 2005, 44, 403–406.66 C. Barfoot, G. Brooks, P. Brown, S. Dabbs, D. T. Davies,

I. Giordano, A. Hennessy, G. Jones, R. Markwell, T. Miles,N. Pearson and C. A. Smethurst, Tetrahedron Lett., 2010, 51,2685–2689.

67 W. Ye, J. Mo, T. Zhao and B. Xu, Chem. Commun., 2009,3246–3248.

68 D. S. Surry and S. L. Buchwald, Chem. Sci., 2010, 1, 13–31 andreferences therein.

69 X. Pan, Q. Cai and D. Ma, Org. Lett., 2004, 6, 1809–1812.70 L. Jiang, G. E. Job, A. Klapars and S. L. Buchwald, Org. Lett.,

2003, 5, 3667–3669.71 X. Yuan, X. Xu, X. Zhou, J. Yuan, L. Mai and Y. Li, J. Org.

Chem., 2007, 72, 1510–1513.72 Y. Bolshan and R. A. Batey, Tetrahedron, 2010, 66, 5283–5294.73 L. J. Gooßen, J. E. Rauhaus and G. Deng, Angew. Chem., Int. Ed.,

2005, 44, 4042–4045.74 L. J. Gooßen, M. Arndt, M. Blanchot, F. Rudolphi, F. Menges

and G. Nieder-Schatteburg, Adv. Synth. Catal., 2008, 350,2701–2707.

75 S. Yudha, Y. Kuninobu and K. Takai, Org. Lett., 2007, 9,5609–5611.

76 J. M. Lee, D.-S. Ahn, D. Y. Jung, J. Lee, Y. Do, S. K. Kim andS. Chang, J. Am. Chem. Soc., 2006, 128, 12954–12962.

Dow

nloa

ded

by U

nive

rsity

of

Oxf

ord

on 2

2 Ju

ne 2

011

Publ

ishe

d on

17

Mar

ch 2

011

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C0C

S001

96A

View Online