27.06.2013 Views

Palladium- and Copper-Catalyzed Aryl Halide Amination ...

Palladium- and Copper-Catalyzed Aryl Halide Amination ...

Palladium- and Copper-Catalyzed Aryl Halide Amination ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

REVIEW 1<br />

<strong>Palladium</strong>- <strong>and</strong> <strong>Copper</strong>-<strong>Catalyzed</strong> <strong>Aryl</strong> <strong>Halide</strong> <strong>Amination</strong>, Etherification <strong>and</strong><br />

Thioetherification Reactions in the Synthesis of Aromatic Heterocycles<br />

<strong>Palladium</strong>- Jessie <strong>and</strong> <strong>Copper</strong>-<strong>Catalyzed</strong> Heterocycle Synthesis E. R. Sadig, Michael C. Willis*<br />

Department of Chemistry, University of Oxford, Chemistry Research Laboratory, Mansfield Road, Oxford, OX1 3TA, UK<br />

Fax +44(1865)285002; E-mail: michael.willis@chem.ox.ac.uk<br />

Received 4 August 2010; revised 13 August 2010<br />

Abstract: This article reviews the use of palladium- <strong>and</strong> coppercatalyzed<br />

aryl halide amination, etherification <strong>and</strong> thioetherification<br />

processes in the synthesis of heteroaromatic molecules. The review<br />

is structured by the nature of the key C–X bond being formed,<br />

<strong>and</strong> then by heterocycle type. Where applicable individual heterocycles<br />

are further divided into syntheses based on intermolecular,<br />

intramolecular <strong>and</strong> cascade processes. In order to limit the length of<br />

the article, processes that do not deliver an aromatic heterocycle<br />

from the key C–X bond-forming event are excluded. Processes for<br />

the functionalization of intact heteroaromatics are also not included.<br />

1 Introduction<br />

2 Carbon–Nitrogen Bond Formation<br />

2.1 Indoles<br />

2.2 Carbazoles<br />

2.3 Benzimidazoles <strong>and</strong> Benzimidazolones<br />

2.4 Indazoles <strong>and</strong> Indazolones<br />

2.5 Pyrroles<br />

2.6 Pyrazoles<br />

2.7 Oxazoles<br />

2.8 Quinolones<br />

2.9 Quinazolines, Quinazolinones <strong>and</strong> Quinazolinediones<br />

2.10 Phenazines<br />

2.11 Cinnolines<br />

3 Carbon–Oxygen Bond Formation<br />

3.1 Benzofurans<br />

3.2 Benzoxazoles<br />

3.3 Isocoumarins<br />

4 Carbon–Sulfur Bond Formation<br />

4.1 Benzothiophenes<br />

4.2 Benzothiazoles<br />

4.3 Oxathioles<br />

5 Conclusion<br />

Key words: palladium catalysis, copper catalysis, aromatic heterocycles,<br />

amination, etherification<br />

1 Introduction<br />

Given the numerous applications of aromatic heterocycles<br />

in medicine, agriculture <strong>and</strong> materials, it is not surprising<br />

that a whole host of methods have been developed for<br />

their preparation. Prominent amongst these are routes<br />

based on transition metal catalyzed transformations. 1–8 Indeed,<br />

many transition metal catalyzed processes have<br />

been developed with the explicit goal of delivering new<br />

synthetic routes to heteroaromatics. This forging of new<br />

SYNTHESIS 2011, No. 1, pp 0001–0022xx.xx.2010<br />

Advanced online publication: 12.10.2010<br />

DOI: 10.1055/s-0030-1258294; Art ID: E27810SS<br />

© Georg Thieme Verlag Stuttgart · New York<br />

routes, allowing the introduction of new classes of starting<br />

materials, or access to alternative substitution patterns, is<br />

one of the key advantages offered by transition metal catalysis.<br />

The last fifteen years has seen the development of<br />

efficient <strong>and</strong> user-friendly methods, based on both palladium<br />

<strong>and</strong> copper catalysis, for the formation of carbon–<br />

nitrogen, carbon–oxygen <strong>and</strong> carbon–sulfur bonds using<br />

aryl halide substrates. Collectively, these processes<br />

present almost ideal tools for aromatic heterocycle synthesis,<br />

as witnessed by the rapidly increasing number of<br />

applications that have been published during this time.<br />

For example, reference to the first edition of Li <strong>and</strong><br />

Gribble’s excellent treatise on the use of palladium catalysis<br />

in heterocyclic chemistry, 1 published in 2000, shows<br />

only a h<strong>and</strong>ful of examples of palladium-catalyzed aryl<br />

amination reactions being employed in the synthesis of aromatic<br />

heterocycles. This is certainly not the case today.<br />

Migita described the palladium-catalyzed coupling of<br />

aminostannanes with aryl halides as early as 1983; 9 however,<br />

it was not until the report of tin-free catalytic amination<br />

reactions, by Buchwald10 <strong>and</strong> Hartwig, 11 that the<br />

synthetic potential of these processes began to be realized.<br />

<strong>Copper</strong>-based aryl halide amination (<strong>and</strong> amidation)<br />

chemistry has an even longer history, dating back to the<br />

original reports by Ullmann12 <strong>and</strong> Goldberg, 13 but again,<br />

it was not until the development of mild catalytic variants<br />

of these reactions that the majority of applications began<br />

to be developed. In recent years both processes have undergone<br />

enormous development <strong>and</strong> now encompass a<br />

myriad of different nitrogen nucleophiles <strong>and</strong> aryl halide<br />

(<strong>and</strong> equivalent) coupling partners. Advances to carbon–<br />

oxygen <strong>and</strong> carbon–sulfur bond-forming variants have<br />

also been achieved. It is beyond the scope of this review<br />

to examine the development <strong>and</strong> mechanistic details of<br />

these underpinning catalytic methods, but extensive reviews<br />

of both the palladium14–17 <strong>and</strong> copper18–23 chemistries<br />

exist. The following discussion is divided by the key<br />

catalytic bond construction – carbon–nitrogen, carbon–<br />

oxygen or carbon–sulfur – <strong>and</strong> then by heterocycle type.<br />

Where appropriate, individual heterocycles are then subdivided<br />

into intermolecular, intramolecular or cascade<br />

processes. The review is focused on the formation of aromatic<br />

heterocyles, <strong>and</strong> accordingly we have not included<br />

reports of the functionalization of intact heterocyclic<br />

cores, nor processes that lead to non-aromatic systems.


2 J. E. R. Sadig, M. C. Willis REVIEW<br />

2 Carbon–Nitrogen Bond Formation<br />

2.1 Indoles<br />

A number of new indole syntheses based on aryl halide<br />

amination have been developed; however, one of the first<br />

heterocycle syntheses to be reported using palladiumcatalyzed<br />

amination chemistry was concerned with intercepting<br />

reaction intermediates from a very well established<br />

route to indoles. 24 In 1998 Buchwald demonstrated<br />

that a variety of aryl halides could be combined with benzophenone<br />

hydrazone using intermolecular palladiumcatalyzed<br />

amination reactions to generate the corresponding<br />

N-arylhydrazones (1, Scheme 1). 25 A palladium(II)<br />

acetate/XantPhos (2) catalyst system, in combination with<br />

sodium tert-butoxide, was found to be optimal. The Narylhydrazones<br />

could either be isolated or, after simple<br />

filtration through a silica plug, treated directly with an<br />

enolizable ketone under acid hydrolysis conditions. The<br />

ensuing Fischer cyclization provided the corresponding<br />

indoles in good to excellent yields. Variations to access either<br />

N-alkyl- or N-arylindoles were also developed, although<br />

in these cases functionalization of the isolated Narylhydrazones<br />

was necessary.<br />

A limitation of the benzophenone hydrazone methodology<br />

is the difficulty in removing benzophenone from the<br />

final indole products, particularly in large-scale<br />

applications. Cho <strong>and</strong> Lim reported a related procedure, in<br />

which N-Boc arylhydrazines were employed in Fischer<br />

cyclizations. 26 The required N-Boc arylhydrazines were<br />

prepared using palladium-catalyzed amination chemistry,<br />

but in these cases were isolated before indole formation.<br />

The use of intramolecular carbon–nitrogen bond formation<br />

onto an aryl halide has proved to be a popular method<br />

to achieve indole synthesis. In one of the first routes based<br />

Biographical Sketches<br />

Jessie Sadig received her<br />

MChem degree from the<br />

University of Oxford in July<br />

2008, where she carried out<br />

her final year project under<br />

Michael Willis received his<br />

undergraduate education at<br />

Imperial College London,<br />

<strong>and</strong> his PhD from the University<br />

of Cambridge working<br />

with Prof. Steven V.<br />

Ley, FRS. After a postdoctoral<br />

stay with Prof. David<br />

A. Evans at Harvard University,<br />

as a NATO/Royal<br />

Synthesis 2011, No. 1, 1–22 © Thieme Stuttgart · New York<br />

on this approach, Watanabe et al. demonstrated that N,Ndimethylhydrazones<br />

derived from o-chloroarylacetaldehydes<br />

(4) underwent cyclization under the action of palladium<br />

catalysis to provide the corresponding N-aminoindoles<br />

(Scheme 2). 27 Tri(tert-butyl)phosphine <strong>and</strong> the<br />

ferrocene derivative 5 were found to be the optimal<br />

lig<strong>and</strong>s. The authors went on to demonstrate that if an appropriately<br />

functionalized substrate was employed, then a<br />

second palladium-catalyzed transformation could be<br />

achieved in the reactions. For example, dichloride 6 could<br />

N<br />

Ph Ph<br />

NH2<br />

+<br />

MeO<br />

MeO<br />

Scheme 1<br />

65%<br />

Br<br />

Me<br />

N<br />

H<br />

Me Me<br />

PPh 2<br />

the supervision of Dr. Jeremy<br />

Robertson. She then<br />

joined the Willis group <strong>and</strong><br />

is currently working towards<br />

her DPhil degree. Her<br />

Society Research Fellow, he<br />

was appointed to a lectureship<br />

at the University of<br />

Bath in November 1997. In<br />

January 2007 he moved to<br />

the University of Oxford,<br />

where he is a University<br />

Lecturer <strong>and</strong> Fellow of<br />

Lincoln College. He was<br />

awarded an EPSRC Ad-<br />

O<br />

i) Pd(OAc) 2 (0.1 mol%)<br />

XantPhos (0.11 mol%)<br />

NaOt-Bu, toluene, 80 °C<br />

ii) TsOH.<br />

H2O, EtOH<br />

reflux<br />

O<br />

Me<br />

Me<br />

Pent<br />

Ph<br />

Ph<br />

N<br />

HN<br />

Me<br />

1<br />

PPh 2<br />

N<br />

H<br />

Ph Ph<br />

79%<br />

Pent<br />

N<br />

H<br />

Pent<br />

N<br />

H<br />

Me<br />

Me<br />

70% 70%<br />

(BINAP used as lig<strong>and</strong>)<br />

XantPhos (2) (rac)-BINAP (3)<br />

PPh2<br />

PPh2<br />

research focuses on palladium-<br />

<strong>and</strong> copper-catalyzed<br />

cascade processes for heterocycle<br />

synthesis.<br />

vanced Research Fellowship<br />

in 2005 <strong>and</strong> the 2008<br />

AstraZeneca Research<br />

Award for Organic Chemistry.<br />

His group’s research interests<br />

are based on the<br />

development <strong>and</strong> application<br />

of new catalytic processes<br />

for organic synthesis.


REVIEW <strong>Palladium</strong>- <strong>and</strong> <strong>Copper</strong>-<strong>Catalyzed</strong> Heterocycle Synthesis 3<br />

be reacted under the st<strong>and</strong>ard conditions but with the addition<br />

of phenyl boronic acid, to provide 4-phenylindole<br />

7, resulting from t<strong>and</strong>em carbon–carbon <strong>and</strong> carbon–<br />

nitrogen bond formation. In addition to aryl boronic acids<br />

being introduced by Suzuki couplings, a range of azoles<br />

<strong>and</strong> amines could also be effectively incorporated via a<br />

second carbon–nitrogen bond-forming process.<br />

F<br />

Cl<br />

4<br />

6<br />

Scheme 2<br />

Doye <strong>and</strong> co-workers demonstrated that N-alkyl imines<br />

corresponding to hydrazones 4 can also be effectively cyclized<br />

under the action of palladium catalysis to provide<br />

N-alkylindoles. 28 In their approach, the imine substrates<br />

were prepared from the corresponding alkynes using a<br />

titanium-catalyzed hydroamination process. In this onepot<br />

protocol, the imines were then subjected to palladium<br />

catalysis to deliver the indole products. Scheme 3 shows<br />

an example in which alkyne 8 was converted in a one-pot<br />

process into indole 9. N-Heterocyclic carbene (NHC)<br />

lig<strong>and</strong> 10 was most effective. Substrates containing a tethered<br />

amine nucleophile could also be included, leading to<br />

the formation of N–C2 annulated indoles.<br />

Scheme 3<br />

H<br />

N<br />

Cl NMe2<br />

H<br />

N<br />

Cl NMe2<br />

Pd(dba) 2 (3 mol%)<br />

5 (4.5 mol%)<br />

NaOt-Bu<br />

o-xylene, 120 °C<br />

PhB(OH) 2<br />

Pd(dba)2 (5 mol%)<br />

5 (7.5 mol%)<br />

Cs 2CO 3<br />

o-xylene, 120 °C<br />

Fe<br />

NMe2 Pt-Bu2<br />

5<br />

N<br />

60% NMe2 N<br />

NMe2 7, 56%<br />

10<br />

8<br />

MeO<br />

Pr i) [Cp2TiMe2] (5 mol%)<br />

toluene<br />

110 °C, 24 h<br />

MeO<br />

+ H2N Me<br />

ii) Pd2(dba)3 (5 mol%)<br />

Cl Me<br />

Me<br />

(10 mol%)<br />

KOt-Bu, dioxane 9, 65%<br />

N<br />

Pr<br />

Me<br />

110 °C<br />

Me Me<br />

Me<br />

Me<br />

Isolated dehydrohalophenylalinate derivatives have been<br />

converted into indoles using related cyclizations. For example,<br />

Brown was the first to report the cyclization of<br />

enamines such as 11 (Scheme 4) into the corresponding<br />

indole-2-carboxylates. In this report from 2000, a simple<br />

dppf-derived catalyst was employed in combination with<br />

F<br />

Ph<br />

Me Me Me<br />

Cl<br />

N N<br />

Me<br />

–<br />

10<br />

+<br />

potassium acetate as base. 29 Kondo <strong>and</strong> co-workers subsequently<br />

showed that polymer-supported substrates corresponding<br />

to 11 can also be cyclized effectively. 30<br />

O 2N<br />

Scheme 4<br />

11<br />

HN<br />

I<br />

Br<br />

CO2Et<br />

PdCl 2(dppf)<br />

(5 mol%)<br />

KOAc, DMF<br />

90 °C<br />

PPh 2<br />

Fe dppf (12)<br />

PPh2<br />

O2N 83%<br />

Br<br />

Lautens <strong>and</strong> co-workers established gem-dihalovinylanilines<br />

as versatile substrates for indole synthesis based<br />

on a series of t<strong>and</strong>em metal-catalyzed processes. The substrates<br />

were readily accessed from the relevant o-nitrobenzaldehyde<br />

derivatives via Ramirez olefinations<br />

followed by reduction of the nitro group. The early chemistry<br />

focused on t<strong>and</strong>em palladium-catalyzed intramolecular<br />

amination reactions <strong>and</strong> intermolecular Suzuki<br />

couplings to deliver a series of variously substituted indoles.<br />

31a,b For example, reaction of gem-dibromoaniline<br />

13 with thienyl-3-boronic acid delivered the expected<br />

indole in 86% yield (Scheme 5). The electron-rich<br />

biphenyl-based phosphine SPhos (14), developed by<br />

Buchwald, proved to be optimal. Significant variation in<br />

the substitution pattern of the substrates <strong>and</strong> in the type of<br />

organoboron coupling partner was possible. For example,<br />

it was possible to prepare indoles with individual substituents<br />

at positions C2–C7; N-aryl substrates could also be<br />

employed. <strong>Aryl</strong>, alkenyl <strong>and</strong> alkyl boron reagents were all<br />

used successfully. For many of the examples it was possible<br />

to use palladium loadings of only 1 mol%. Although<br />

no reaction intermediates were detected, a brief mechanistic<br />

investigation suggested that the intramolecular amination<br />

reaction preceded the intermolecular Suzuki<br />

coupling. It is interesting to note that the amination reactions<br />

took place on alkenyl halides, 32 as opposed to the<br />

usual aryl halide substrates. To demonstrate the synthetic<br />

utility of the method, indole 15, prepared in 86% yield<br />

from the corresponding gem-dibromovinylaniline <strong>and</strong> 2methoxyquinoline<br />

boronic acid, was utilized in a short<br />

synthesis of KDR kinase inhibitors. 31c Bisseret <strong>and</strong> coworkers<br />

also reported a single example of the synthesis of<br />

a 2-arylindole using a similar strategy. 33<br />

In an elegant extension of this chemistry the same group<br />

was able to demonstrate that the corresponding pyridinederived<br />

substrates could be utilized to access azaindoles.<br />

31d The example shown in Scheme 6 illustrates the<br />

preparation of a 7-azaindole using reaction conditions almost<br />

identical to those for the parent indole series. An important<br />

modification from the parent system, needed to<br />

achieve high yields, was the use of a nitrogen-protecting<br />

group. It was also possible to extend the chemistry to the<br />

preparation of 6-azaindoles; however, the synthesis of the<br />

Synthesis 2011, No. 1, 1–22 © Thieme Stuttgart · New York<br />

N<br />

CO2Et


4 J. E. R. Sadig, M. C. Willis REVIEW<br />

BnO<br />

13<br />

+<br />

(HO) 2B<br />

MeO<br />

O<br />

Scheme 5<br />

N<br />

Scheme 6<br />

Br<br />

NH2<br />

S<br />

Br<br />

Pd(OAc) 2 (1 mol%)<br />

SPhos (2 mol%) S<br />

N<br />

H<br />

Ph<br />

NH<br />

Me<br />

Me<br />

N<br />

86% 73%<br />

77%<br />

15, 86%<br />

(2 mol% Pd)<br />

Br<br />

NH<br />

Bn<br />

Br<br />

+<br />

N<br />

H<br />

MeO<br />

(HO) 2B Ph<br />

K3PO .<br />

4 H2O toluene, 90 °C<br />

regioisomeric 5- <strong>and</strong> 4-azaindoles was more challenging<br />

<strong>and</strong> required the use of N-oxide substrates. Thienopyrroles<br />

could also be prepared starting from the corresponding<br />

thiophene-derived substrates.<br />

The same gem-dihalovinylaniline substrates have been<br />

utilized in a number of different t<strong>and</strong>em processes. For example,<br />

as Scheme 7 illustrates, the initial intramolecular<br />

amination reactions have been partnered with a number of<br />

alternative reactions, including Heck olefinations (16 →<br />

17), 31e carbonylations (16 → 18) 34 <strong>and</strong> Sonogashira couplings<br />

(16 → 19), 31f to deliver acrylate-, ester- <strong>and</strong> alkynesubstituted<br />

indoles, respectively. Bisseret <strong>and</strong> co-workers<br />

used the same substrates in t<strong>and</strong>em palladium-catalyzed<br />

amination/phosphonation sequences to deliver 2-phosphonate-substituted<br />

indoles. 33 Tethered substrates were<br />

used in the Heck chemistry to provide polycyclic products<br />

via an intramolecular second step. 31e Related tethered substrates,<br />

invoking a second intramolecular amination reaction,<br />

31g or an intramolecular direct arylation reaction, 31h<br />

were also developed to access alternative polycyclic scaffolds.<br />

It is interesting to note that the double amination<br />

processes were achieved using copper catalysis. Lautens<br />

<strong>and</strong> co-workers also demonstrated that the reactivity of<br />

gem-dibromovinylanilines can be controlled to allow access<br />

to 2-bromoindoles by a single (as opposed to a t<strong>and</strong>em)<br />

palladium-catalyzed intramolecular amination<br />

reaction. 31i<br />

The Willis research group has developed cascade catalytic<br />

amination strategies to access a range of indole deriva-<br />

Synthesis 2011, No. 1, 1–22 © Thieme Stuttgart · New York<br />

N<br />

Me<br />

Pd(OAc) 2 (3 mol%)<br />

SPhos (14) (6 mol%)<br />

K3PO .<br />

4 H2O toluene, 100 °C<br />

N<br />

H<br />

Me<br />

86%<br />

(dichloro substrate)<br />

Cy2P OMe<br />

MeO<br />

SPhos (14)<br />

N<br />

N<br />

Bn<br />

74%<br />

Ph<br />

t-BuO 2C<br />

Scheme 7<br />

Br<br />

NH<br />

16<br />

R<br />

SiMe 3<br />

Pd(OAc) 2 (4 mol%)<br />

Me4NCl (100 mol%)<br />

K3PO4 .<br />

H2O, Et3N<br />

toluene, reflux<br />

Br<br />

PdCl2(PPh3) 2<br />

Ph3P (10 mol%)<br />

CO (10 atm)<br />

DIPEA, THF<br />

MeOH, 110 °C<br />

P(p-MeOC6H4)3 (8 mol%)<br />

Pd/C (2 mol%)<br />

CuI (4 mol%)<br />

i-Pr2NH, toluene, 100 °C<br />

tives. They focused on the use of 2-(2-haloalkenyl)aryl<br />

halides, together with the corresponding alkenyl triflates,<br />

as indole substrates. 35a,b Scheme 8 shows examples of<br />

both the aryl halide/alkenyl triflate (20) <strong>and</strong> the aryl halide/alkenyl<br />

halide (21) substrates in palladium-catalyzed<br />

indole-forming reactions. Both reactions feature an initial<br />

intermolecular amination reaction followed by an intramolecular<br />

amination as the second step of the cascade,<br />

to deliver N-functionalized indole products. The authors<br />

noted that it was possible to employ either Z- or E-configured<br />

alkene isomers in the process, a consequence of the<br />

initial amination reaction taking place at the alkenyl halide<br />

<strong>and</strong> generating a configurationally unstable enamine<br />

intermediate. Reactions employing the triflate substrates<br />

were best achieved using DPEPhos (22) or XantPhos (2)<br />

lig<strong>and</strong>s, while the dihalide substrates delivered best yields<br />

with the Buchwald diphenyl lig<strong>and</strong>s, such as SPhos (14).<br />

By selecting the appropriate class of substrate it was possible<br />

to access a number of indole substitution patterns.<br />

Significant variation of the nitrogen coupling partner was<br />

possible, with examples of aniline, amine, amide, hydrazine<br />

<strong>and</strong> sulfonamide nucleophiles all being reported.<br />

Sterically dem<strong>and</strong>ing nitrogen nucleophiles, such as tertbutylamine,<br />

could also be introduced; 35c the ability to access<br />

indoles bearing bulky N-substituents was exploited<br />

in a synthesis of the natural product demethylasterriquinone<br />

A, in which N-(reverse prenyl)indole 23 was utilized as<br />

a key intermediate. A recent extension of the chemistry<br />

has seen trihalogenated substrates employed, allowing access<br />

to 4-, 5-, 6- <strong>and</strong> 7-chloroindoles. 35d The Li research<br />

group has adapted the method to encompass trifluoromethyl-substituted<br />

substrates in order to prepare a series<br />

of N-aryl-2-trifluoromethyl-substituted indoles. 36 A related<br />

intramolecular amination reaction between an aniline <strong>and</strong><br />

an alkenyl triflate was employed by Smith <strong>and</strong> co-workers<br />

in their synthesis of the nodulisporic acids tetracycle. 37<br />

The Willis research group also demonstrated that similar<br />

palladium-catalyzed cascade processes can be achieved<br />

using pyridine-derived substrates to provide access to the<br />

corresponding azaindole products. 35e For example, reaction<br />

of dihalopyridine substrate 24 with p-anisidine, using<br />

a DPEPhos-derived catalyst, delivered the corresponding<br />

N<br />

Bn<br />

17, 79%<br />

CO2t-Bu<br />

CO2Me N<br />

18, 70%<br />

Bn<br />

N<br />

H<br />

SiMe3<br />

19, 57%


REVIEW <strong>Palladium</strong>- <strong>and</strong> <strong>Copper</strong>-<strong>Catalyzed</strong> Heterocycle Synthesis 5<br />

Br OTf<br />

Cl<br />

+<br />

20 H2N<br />

N<br />

71%<br />

Ph<br />

Scheme 8<br />

7-azaindole in good yield (Scheme 9). The process was<br />

shown to work well for 7-azaindoles; however, access to<br />

the remaining regioisomers was less successful.<br />

Scheme 9<br />

The 2-(2-haloalkenyl)aryl halide substrates have also<br />

been shown to undergo cascade copper-catalyzed amination<br />

reactions to deliver N-functionalized indoles. 35f An<br />

example featuring a carbamate coupling partner is shown<br />

in Scheme 10. Although some overlap in scope with the<br />

palladium-catalyzed version of the process was established,<br />

there were also significant differences in reactivity.<br />

In general, the copper-catalyzed variant was more<br />

limited, with chloro-substituted substrates performing<br />

poorly; however, greater success was achieved with<br />

amide nucleophiles, relative to the palladium system.<br />

Scheme 10<br />

Pd2(dba)3 (2.5 mol%)<br />

DPEPhos (6 mol%)<br />

Cs2CO3 toluene, 100 °C<br />

Br Cl<br />

Ph<br />

21, E/Z 2:7<br />

+ H2N<br />

Ph<br />

Pd2(dba)3 (2.5 mol%)<br />

SPhos (14) (7.5 mol%)<br />

NaOt-Bu<br />

toluene, 80 °C<br />

N Cl Br<br />

+<br />

O<br />

OEt<br />

24<br />

H2N OMe<br />

O<br />

PPh2 PPh2<br />

DPEPhos (22)<br />

N<br />

Ph 81%<br />

OMe<br />

Pd2(dba) 3 (5 mol%)<br />

DPEPhos (22) (12 mol%)<br />

Cs2CO3 toluene, 110 °C<br />

Br Br<br />

25<br />

MeO<br />

CuOAc (10 mol%)<br />

MeO<br />

MeO<br />

H2N<br />

+<br />

Ot-Bu<br />

(20 mol%)<br />

Cs2CO3<br />

toluene, 110 °C<br />

MeO<br />

O<br />

MeN NMe<br />

H H 25<br />

75%<br />

94%<br />

N<br />

N<br />

Ph<br />

N<br />

23, 77%<br />

Me Me<br />

N<br />

83%<br />

N<br />

t-BuO<br />

82%<br />

OEt<br />

O<br />

OMe<br />

N<br />

O<br />

Cl<br />

Ackermann <strong>and</strong> co-workers developed efficient indole<br />

syntheses based on a cascade amination process starting<br />

from o-alkynylhaloarenes. 38 In his original report,<br />

Ackermann described both palladium- <strong>and</strong> copper-catalyzed<br />

variants of the process (Scheme 11). 38a For example,<br />

the combination of o-alkynylchloroarene 26 with<br />

benzylamine under the action of an NHC-derived palladium<br />

catalyst delivered indole 27 in 66% yield. The original<br />

copper-catalyzed protocol simply employed copper(I) iodide<br />

to combine o-alkynylchoroarenes with anilines furnishing<br />

the corresponding N-arylindoles in good yields.<br />

The reactions proceed via initial intermolecular N-arylation,<br />

followed by cyclization onto the alkyne. The two<br />

different metal systems were shown to display differing,<br />

<strong>and</strong> complementary, reactivities; for example, the palladium-catalyzed<br />

methods were effective for N-alkyl <strong>and</strong> Naryl<br />

nucleophiles, while the copper system tolerated Naryl<br />

<strong>and</strong> N-acyl substrates. 38b The N-acyl systems required<br />

the use of diamine lig<strong>and</strong> 25. Scheme 11 shows several<br />

examples of products obtained using the copper methodology,<br />

including an N-acyl indole, as well as an azaindole.<br />

An indole corresponding to the Chek1/KDR kinase inhibitor<br />

pharmacophore was also prepared. The palladiumcatalyzed<br />

version of the chemistry was shown to be effective<br />

for the preparation of indoles bearing sterically dem<strong>and</strong>ing<br />

substituents; for example, the adamantylsubstituted<br />

indole 28 was obtained in 94% yield. 38c,d The<br />

Hu 39 <strong>and</strong> Sanz 40 research groups have reported related<br />

indole-forming chemistries.<br />

F3C<br />

Scheme 11<br />

29<br />

26<br />

Hex<br />

Pd(OAc)2 (5 mol%)<br />

(5 mol%)<br />

+<br />

H2N<br />

Cl<br />

Ph<br />

K3PO4<br />

toluene, 105 °C<br />

Bu<br />

i-Pr i-Pr<br />

Cl<br />

+<br />

N N<br />

–<br />

F 3C<br />

N<br />

27, 66% Bn<br />

Hex<br />

H2N<br />

Cl<br />

+ OMe<br />

CuI (10 mol%)<br />

KOt-Bu<br />

toluene, 105 °C<br />

N<br />

Bu<br />

OMe<br />

69%<br />

t-BuO<br />

61%<br />

N<br />

O<br />

S<br />

N<br />

N<br />

t-BuO<br />

86%<br />

O<br />

i-Pr i-Pr<br />

29<br />

Ph<br />

Synthesis 2011, No. 1, 1–22 © Thieme Stuttgart · New York<br />

N<br />

28, 94%<br />

F


6 J. E. R. Sadig, M. C. Willis REVIEW<br />

The Hsung research group employed related cascade processes<br />

for the synthesis of 2-aminoindoles. 41 Scheme 12<br />

shows an example in which ynamide 30 was combined<br />

with p-toluamine to deliver indole 31 in 85% yield. An<br />

XPhos-derived palladium catalyst was found to be optimal,<br />

<strong>and</strong> a variety of 2-carbamate-substituted indoles<br />

were prepared in good yields. The ynamide substrates<br />

were prepared, in a separate operation, by a copper-catalyzed<br />

amidation of the corresponding bromo-alkynes.<br />

30<br />

Br<br />

Scheme 12<br />

Ackermann <strong>and</strong> colleagues were able to extend their original<br />

methodology to a three-component system, featuring<br />

the in situ formation of the key o-alkynylhaloarenes. 38a,e<br />

The process involved the initial combination of an ochloroiodobenzene<br />

with an alkyne in the presence of a<br />

mixed palladium/copper catalyst system (Scheme 13); after<br />

two hours the required nitrogen nucleophile was introduced<br />

along with additional base, resulting in arylamination<br />

followed by a 5-endo ring closure. A variety of<br />

1,2-disubstituted indoles were obtained in good yields. A<br />

possible limitation of the method is the poor commercial<br />

availability of alternative o-chloroiodobenzene derivatives.<br />

Scheme 13<br />

O<br />

N<br />

O Pd2(dba) 3 (2.5 mol%)<br />

XPhos (5 mol%)<br />

Cs2CO3 +<br />

Ph toluene, 110 °C<br />

H2N p-Tol<br />

Cy2P i-Pr<br />

i-Pr<br />

XPhos (32)<br />

N N<br />

p-Tol<br />

Ph<br />

31, 85%<br />

The Barluenga research group developed a successful cascade<br />

process, based on palladium-catalyzed aza-enolate<br />

a-arylation followed by intramolecular N-arylation, for<br />

the synthesis of a variety of indole derivatives. 42 The process<br />

is outlined in Scheme 14: Initial palladium-catalyzed<br />

aza-enolate arylation joins N-phenylimine 33 with 1,2-dibromobenzene,<br />

to provide imine 34. <strong>Palladium</strong>-catalyzed<br />

intramolecular amination, presumably via enamine intermediate<br />

35, then delivers the expected indole in an excel-<br />

i-Pr<br />

29<br />

I i) Pd(OAc) 2 (10 mol%)<br />

CuI (10 mol%), (10 mol%)<br />

Cs2CO3 Cl toluene, 105 °C<br />

H<br />

+<br />

Ph<br />

ii) KOt-Bu<br />

H2N p-Tol<br />

Ph<br />

65%<br />

Cl<br />

Ph<br />

N<br />

p-Tol<br />

Synthesis 2011, No. 1, 1–22 © Thieme Stuttgart · New York<br />

O<br />

O<br />

lent 86% yield. The same XPhos-derived catalyst proved<br />

optimal for both steps of the cascade. A wide range of<br />

imine derivatives could be incorporated. The use of mixed<br />

halogen systems, such as 1-bromo-2-iodobenzenes, allowed<br />

the regioselective synthesis of substituted indoles<br />

by exploiting the greater reactivity of the aryl iodide substituent.<br />

The authors were also able to show that mixed<br />

halide/sulfonate substrates were efficient reaction components,<br />

with both triflate <strong>and</strong> nonaflate systems being successfully<br />

employed. The ability to generate the required<br />

sulfonate derivatives from the parent o-chlorophenols significantly<br />

widened the scope of the process <strong>and</strong> allowed<br />

unusual substitution patterns to be accessed, such as the<br />

2,4,6-trisubstituted example 36. 42b Although an N-aryl or<br />

N-alkyl substituent was a requirement of the imine components,<br />

the authors demonstrated that N-tert-butylsubstituted<br />

indoles could be deprotected under a variety of<br />

conditions. For example, the N-H indole derived from Ntert-butylindole<br />

37 was obtained in 97% yield for the indole<br />

formation <strong>and</strong> deprotection (AlCl 3) sequence. The<br />

group also reported a three-component variant of the<br />

methodology, in which an initial palladium-catalyzed alkenyl<br />

halide amination was used to prepare the imine reaction<br />

components.<br />

Scheme 14<br />

33 86%<br />

Br<br />

+<br />

Br Me<br />

N<br />

Ph<br />

Ph<br />

Pd2(dba)3 (2 mol%)<br />

XPhos (32) (4 mol%)<br />

NaOt-Bu, dioxane<br />

110 °C<br />

N<br />

Ph<br />

N<br />

86% Ph<br />

Br Ph<br />

N<br />

2.2 Carbazoles<br />

Ph<br />

Br Ph<br />

HN<br />

34 35<br />

OMe<br />

Catalytic amination chemistry has a number of applications<br />

in carbazole synthesis. Nozaki was one of the first to<br />

exploit cascade amination processes for the preparation of<br />

a wide range of carbazole architectures. 43 An example of<br />

the general process developed by the Nozaki research<br />

group is shown in Scheme 15; the key heterocycle-forming<br />

reaction is a coupling between a nitrogen nucleophile<br />

<strong>and</strong> a doubly activated biphenyl. In this particular example,<br />

ditriflate 37 was coupled with tert-butylcarbamate using<br />

a XantPhos-derived catalyst to deliver carbazole 38 in<br />

70% yield. 43b Deprotection of the Boc group from carbazole<br />

38 revealed the natural product mukinone. The same<br />

group has exploited their methodology in the synthesis of<br />

heteroacenes, 43d,e chiral carbazole variants, 43a as well as<br />

Ph<br />

Ph<br />

36, 74%<br />

N<br />

Ph<br />

Ph<br />

N<br />

37, 91%<br />

Me<br />

Ph<br />

Me<br />

Me<br />

(from Cl/ONf substrate) (from I/Cl substrate)<br />

Cl


REVIEW <strong>Palladium</strong>- <strong>and</strong> <strong>Copper</strong>-<strong>Catalyzed</strong> Heterocycle Synthesis 7<br />

helicenes. 43b The second example shown in Scheme 15<br />

was reported by Chida <strong>and</strong> colleagues, <strong>and</strong> shows that related<br />

strategies developed using dibromobiphenyls are<br />

also effective in carbazole synthesis. 44a The example<br />

shows the preparation of a key intermediate in the synthesis<br />

of the natural product murrayazoline. 44b In order to<br />

achieve an efficient transformation using the sterically dem<strong>and</strong>ing<br />

amine 39, it was necessary to employ a high catalyst<br />

loading (20 mol%). A copper-catalyzed version of<br />

these transformations has also been reported. 45 The cascade<br />

coupling of nitrogen nucleophiles with doubly activated<br />

bi-aromatics has proven to be a powerful method for<br />

carbazole synthesis <strong>and</strong> has been exploited by several other<br />

groups. For example, Samyn <strong>and</strong> co-workers utilized<br />

dibromo-2,2¢-bithiophene substrates to prepare<br />

dithienopyrroles, 46 as did the Barlow research group. 47<br />

MeO<br />

Me<br />

O<br />

37<br />

OMOM<br />

Scheme 15<br />

O<br />

OMe<br />

OTf<br />

OTf<br />

Pd2(dba)3.CHCl3<br />

(5 mol%)<br />

XantPhos (2) (10 mol%)<br />

K3PO4<br />

o-xylene, 100 °C<br />

Br<br />

Pd2(dba) 3 (20 mol%)<br />

XPhos (32) (60 mol%)<br />

Br<br />

H2N +<br />

Me<br />

Me<br />

NaOt-Bu<br />

toluene, 130 °C<br />

O<br />

+<br />

t-BuO<br />

NH2<br />

39<br />

O<br />

MeO2C<br />

Bedford et al. developed a cascade sequence based on intermolecular<br />

aryl-amination followed by intramolecular<br />

direct C–H arylation as a route to carbazole derivatives;<br />

several examples are shown in Scheme 16. 48 The process<br />

combines o-chloroanilines with aryl bromides using palladium<br />

catalysis. Use of the bulky electron-rich tri(tertbutyl)phosphine<br />

lig<strong>and</strong> in combination with palladium(II)<br />

acetate was the optimal catalyst system <strong>and</strong> allowed the<br />

preparation of a range of carbazoles in good yields. Significant<br />

substitution could be tolerated on the coupling<br />

partners, as illustrated by the preparation of the natural<br />

product clausine P (40).<br />

Ackermann et al. developed a related aryl-amination <strong>and</strong><br />

direct C–H arylation sequence for carbazole synthesis. 49<br />

In the Ackermann approach, 1,2-dihaloaromatics were<br />

combined with anilines to deliver carbazole products<br />

(Scheme 17). A combination of palladium(II) acetate <strong>and</strong><br />

Me<br />

N<br />

OMe<br />

Boc<br />

70%<br />

N-Boc mukonine (38)<br />

MOMO<br />

murrayazoline<br />

59%<br />

O O<br />

Me<br />

N<br />

Me<br />

F 3C<br />

Me<br />

Scheme 16<br />

tricyclohexylphosphine was found to be optimal <strong>and</strong> allowed<br />

a wide range of carbazoles to be prepared. Using<br />

this catalyst system it was possible to employ inexpensive<br />

1,2-dichloroarenes as coupling partners, as well as heterocyclic<br />

derivatives, illustrated by the preparation of pyrazine<br />

<strong>and</strong> pyridine derivatives 41 <strong>and</strong> 42. The authors<br />

exploited the methodology in a synthesis of the natural<br />

product murrayafoline A (43). 49b Although less accessible,<br />

1,2-dihaloalkenes could also be employed as substrates,<br />

allowing the same cascade sequence to be applied<br />

to indole synthesis.<br />

+<br />

+<br />

N<br />

H<br />

43%<br />

Cl<br />

H<br />

N<br />

N<br />

Cl<br />

Scheme 17<br />

Ph<br />

N<br />

Ph<br />

41, 93%<br />

Bn<br />

NH<br />

Cl<br />

Br<br />

N<br />

Pd(OAc)2 (4 mol%)<br />

t-Bu3P (5 mol%)<br />

NaOt-Bu<br />

toluene, reflux<br />

OMe<br />

Pd(OAc)2 (5 mol%)<br />

Cy3P (10 mol%)<br />

NaOt-Bu<br />

NMP, 130 °C<br />

N<br />

Ph<br />

42, 93%<br />

N<br />

69% Bn<br />

Kan <strong>and</strong> co-workers reported a palladium-catalyzed cascade<br />

route to carbazoles based on initial intermolecular<br />

Suzuki coupling followed by an intramolecular aryl amination.<br />

A small scoping study was described, although a<br />

possible limitation is the availability of suitably functionalized<br />

boronic acids. 50 Fujii <strong>and</strong> Ohno have also described<br />

a cascade route to carbazoles. In their approach, an intermolecular<br />

aryl amination between an aryl triflate <strong>and</strong> an<br />

aniline was used to construct a diphenylamine which then<br />

underwent an oxidative coupling to generate a carbazole<br />

product. The efficiency of the oxidative step was shown to<br />

be dependent on the substitution pattern of the diphenylamine<br />

intermediates. 51<br />

N<br />

F 3C<br />

MeO<br />

N<br />

H<br />

69%<br />

clausine P (40)<br />

CF 3<br />

Synthesis 2011, No. 1, 1–22 © Thieme Stuttgart · New York<br />

Me<br />

Me<br />

N<br />

Ph<br />

85%<br />

OMe<br />

N<br />

MeO H<br />

murrayafoline A (43)<br />

74% from Br/Cl-Ar<br />

72% from Cl/Cl-Ar<br />

Me<br />

Me


8 J. E. R. Sadig, M. C. Willis REVIEW<br />

2.3 Benzimidazoles <strong>and</strong> Benzimidazolones<br />

A host of approaches to benzimidazoles <strong>and</strong> related derivatives,<br />

featuring inter- <strong>and</strong> intramolecular <strong>and</strong> cascade reactions,<br />

as well as both palladium <strong>and</strong> copper catalysts,<br />

have been reported. Ma <strong>and</strong> co-workers developed a number<br />

of copper-catalyzed routes to these important heterocycles.<br />

52 In 2007, they demonstrated that o-haloacetanilides<br />

could be combined with primary amines to deliver<br />

benzimidazoles (Scheme 18). For example, a copper(I)<br />

iodide/proline catalyst system was effective for the union<br />

of aryl iodide 44 with allylamine, to furnish benzimidazole<br />

45 in 94% yield. 52a When more sterically dem<strong>and</strong>ing<br />

amines, or less activated aryl units, were employed it was<br />

necessary to use either heat or an acid/heat combination to<br />

achieve cyclization to the aromatic system; for example,<br />

the formation of cyclohexyl-substituted benzimidazole 46<br />

required the addition of acid. A broad range of amines <strong>and</strong><br />

aryl units, including pyridine derivatives, could be included<br />

in the method, delivering benzimidazoles in good to<br />

excellent yields. Although Scheme 18 shows only aryl iodide<br />

substrates, aryl bromides were also used. When the<br />

starting substrates were changed to o-iodoarylcarbamates,<br />

the same reaction system was used to access benzimidazolones<br />

(47 → 48). 52b Aqueous ammonia could also be<br />

employed as the nitrogen nucleophile in both reaction<br />

pathways, leading to the corresponding N–H derivatives.<br />

52c<br />

44<br />

+<br />

H2N +<br />

H2N 47<br />

+<br />

H2N<br />

H<br />

N<br />

Scheme 18<br />

I<br />

H<br />

N<br />

I<br />

H<br />

N<br />

I<br />

O<br />

O<br />

O<br />

CF3<br />

O<br />

OMe<br />

N Boc<br />

CuI (10 mol%)<br />

L-proline (20 mol%)<br />

K 2CO 3, DMSO, r.t.<br />

i) CuI (10 mol%)<br />

L-proline (20 mol%)<br />

K2CO3, DMSO, 40 °C<br />

ii) AcOH, 140 °C<br />

CuI (10 mol%)<br />

L-proline (20 mol%)<br />

K 2CO 3, DMSO<br />

50 °C then 130 °C<br />

OH<br />

N<br />

H H<br />

O<br />

L-proline<br />

N<br />

45, 94%<br />

46, 73%<br />

48, 74%<br />

Buchwald <strong>and</strong> co-workers developed a related palladiumcatalyzed<br />

route to aryl-substituted benzimidazoles. For<br />

example, the coupling of o-bromacetanilide 49 with otoluamine<br />

using an XPhos-derived catalyst provided the<br />

corresponding N-arylbenzimidazole in 94% yield<br />

Synthesis 2011, No. 1, 1–22 © Thieme Stuttgart · New York<br />

N<br />

N<br />

N<br />

H<br />

N<br />

N<br />

N<br />

CF3<br />

O<br />

Boc<br />

O<br />

(Scheme 19). 53a A good range of anilines <strong>and</strong> aryl substrates<br />

(both Br <strong>and</strong> Cl) were described; Scheme 19 also<br />

shows an aza-example, derived from the corresponding<br />

pyridine substrate, as well as a cyclopropyl-substituted<br />

product. Some variation of lig<strong>and</strong> between XPhos (32)<br />

<strong>and</strong> RuPhos (50) was needed for certain substrates. In<br />

2006, Scott reported a palladium-catalyzed synthesis of<br />

imidazopyridinones (aza-benzimadazolones) in a process<br />

analogous to the conversion of 47 into 48 shown in<br />

Scheme 18. 54<br />

Me<br />

N<br />

H2N<br />

Me<br />

Scheme 19<br />

Buchwald <strong>and</strong> Zheng also described a complementary<br />

copper-catalyzed protocol, based on aryl halide amidation<br />

(as opposed to amination), for benzimidazole synthesis. 53b<br />

The basic process is shown in Scheme 20; copper(I)<br />

iodide/diamine-catalyzed coupling of hexamide with oiodo-N-alkylaniline<br />

51, followed by base-promoted cyclization,<br />

delivered the desired N-alkylbenzimidazole 52<br />

in 87% yield. Alkyl, alkenyl <strong>and</strong> aryl amides could all be<br />

incorporated effectively.<br />

Scheme 20<br />

N<br />

N<br />

(from ArCl using lig<strong>and</strong> 50)<br />

F 3C<br />

49<br />

Br<br />

H<br />

N Me<br />

Pd2(dba) 3 (1 mol%)<br />

+<br />

O<br />

Br<br />

XPhos (32) (8 mol%)<br />

K3PO4 t-BuOH, 110 °C<br />

88%<br />

H2N<br />

51<br />

+<br />

O<br />

H<br />

N<br />

I<br />

Pent<br />

N<br />

Me<br />

Me<br />

Me<br />

N<br />

Me<br />

Me<br />

83%<br />

(using lig<strong>and</strong> 50)<br />

i) CuI (5 mol%)<br />

25 (20 mol%)<br />

Cs2CO3, dioxane, 90 °C<br />

ii) K3PO4<br />

t-BuOH, 110 °C<br />

F3C<br />

94%<br />

Intramolecular aryl amidation using urea substrates has<br />

been achieved using both palladium <strong>and</strong> copper catalysis<br />

<strong>and</strong> provides a further route to benzimidazolones<br />

(Scheme 21). The cyclization of urea 53, using a palladium(II)<br />

acetate/XPhos catalyst system, was described by<br />

the process group at Merck. 55 Chloro-, bromo- <strong>and</strong> iodoaryl<br />

derivatives could all be employed as substrates, as<br />

could chloropyridines. <strong>Copper</strong>-catalyzed variants of similar<br />

cyclizations have also been reported, 56 including a<br />

polymer-supported example; 57 the second reaction shown<br />

in Scheme 21, reported by SanMartin, Domínguez <strong>and</strong> co-<br />

Me<br />

Br<br />

N<br />

N<br />

Cy 2P Oi-Pr<br />

i-PrO<br />

RuPhos (50)<br />

52, 87%<br />

N<br />

N<br />

Pent<br />

Me<br />

Me<br />

Me


REVIEW <strong>Palladium</strong>- <strong>and</strong> <strong>Copper</strong>-<strong>Catalyzed</strong> Heterocycle Synthesis 9<br />

MeO<br />

Scheme 21<br />

workers, 58 illustrates a copper(I) iodide catalyzed cyclization<br />

using water as the solvent.<br />

Intramolecular reactions to access benzimidazoles can be<br />

achieved using amidines as substrates. An example, reported<br />

by Brain et al., 59 is shown in Scheme 22 in which<br />

amidine 54 was converted into the corresponding benzimidazole<br />

under the action of a Pd2(dba) 3/triphenylphosphine<br />

catalyst. The authors were also able to show that use<br />

of a ‘catch-<strong>and</strong>-release’ purification strategy, 59b involving<br />

Amberlyst resin, allowed the benzimidazoles to be obtained<br />

in pure form without recourse to chromatography.<br />

Related cyclizations in which the amidine is embedded<br />

within a heterocycle have also been described. 60,61<br />

De Meijere <strong>and</strong> Lygin reported that amidines, generated<br />

in situ from an isocyanide <strong>and</strong> a primary amine, can also<br />

be cyclized using copper(I) conditions to access N-alkylbenzimidazoles.<br />

62<br />

Me<br />

53<br />

54<br />

Scheme 22<br />

PMB PMB<br />

Pd(OAc) 2 (1 mol%)<br />

N O XPhos (32) (3 mol%)<br />

N<br />

NH2<br />

Cl<br />

Bn<br />

N<br />

O<br />

NH2 Br<br />

NaHCO3<br />

i-PrOH, 83 °C<br />

CuI (8.5 mol%)<br />

TMEDA (3.5 equiv)<br />

H 2O, 120 °C<br />

MeO<br />

92%<br />

92%<br />

Batey <strong>and</strong> Evindar used related cyclizations employing<br />

guanidine substrates to access 2-aminobenzimidazoles. 63<br />

Efficient reactions were achieved using either palladium<br />

or copper catalysis; Scheme 23 gives an example of the<br />

reaction conditions needed for each metal. In general, the<br />

copper conditions were found to be superior, providing<br />

higher yields <strong>and</strong> more selective reactions. Both aryl iodides<br />

<strong>and</strong> aryl bromides could be employed as substrates,<br />

allowing a broad range of 2-aminobenzimidazoles to be<br />

Me<br />

Br<br />

HN<br />

Bn<br />

N<br />

Scheme 23<br />

N<br />

N Me<br />

NHMe<br />

Br<br />

Pd2(dba)3, (1.5 mol%)<br />

Ph3P (12 mol%)<br />

NaOH, H 2O–DME<br />

160 °C, MW<br />

Pd(PPh3)4 (10 mol%)<br />

Cs2CO3, DME, 80 °C<br />

or<br />

CuI (5 mol%)<br />

1,10-Phen (10 mol%)<br />

Cs2CO3, DME, 80 °C<br />

Me<br />

Me<br />

82%<br />

N<br />

N<br />

Bn N<br />

Pd: 66%<br />

Cu: 90%<br />

N<br />

N<br />

N<br />

H<br />

Bn<br />

N<br />

N<br />

H<br />

Me<br />

O<br />

O<br />

Me<br />

prepared in good yields. Szczepankiewicz et al. applied<br />

this type of copper-catalyzed cyclization to substrates<br />

based on uracil templates to prepare purine <strong>and</strong> related<br />

fused imidazole systems. 64 2-Mercaptobenzimidazoles<br />

can similarly be accessed using copper-catalyzed cyclization<br />

of in situ generated isothioureas. 65<br />

The research groups of both Zhang66<strong>and</strong> Wu67 showed<br />

that o-halophenylimidoyl chlorides are effective substrates<br />

for copper-catalyzed benzimidazole synthesis.<br />

Scheme 24 presents an example from the Zhang group,<br />

<strong>and</strong> shows how imidoyl chloride 55 can be combined with<br />

benzylamine using a copper(I) iodide catalyst, to deliver<br />

the expected benzimidazole in excellent yield. Both alkyl<strong>and</strong><br />

arylamines could be employed. In both reports it was<br />

necessary to include an electron-withdrawing substituent<br />

on the imidoyl chloride substrate; for example, the trifluoromethyl<br />

substituent on imidoyl chloride 55.<br />

N<br />

Scheme 24<br />

Maes <strong>and</strong> co-workers explored a range of cascade amination<br />

strategies to access a variety of benzo-fused benzimidazole<br />

systems. 68 In their lead publication, they were able<br />

to combine 2-chloro-3-iodopyridine with 2-picoline to<br />

generate dipyridoimidazole 56 in 96% yield (Scheme 25).<br />

The example shown employs a palladium(II) acetate/<br />

BINAP catalyst, although XantPhos (2) was also shown to<br />

be an effective lig<strong>and</strong>. 68a The chemistry has been extended<br />

to the preparation of a number of benzo-fused <strong>and</strong> aza analogues,<br />

68a–c,69 <strong>and</strong> in an interesting application a temperature/halide<br />

dependent regioselectivity switch was<br />

developed. 68d<br />

Scheme 25<br />

I<br />

H2N Ph<br />

N<br />

55<br />

+<br />

I<br />

Cl<br />

+<br />

CF3<br />

Cl N NH 2<br />

CuI (10 mol%)<br />

TMEDA (20 mol%)<br />

Cs 2CO 3<br />

toluene, 110 °C<br />

Pd(OAc) 2 (3 mol%)<br />

BINAP (3) (3 mol%)<br />

Cs2CO3<br />

toluene, reflux<br />

98%<br />

Batey <strong>and</strong> co-workers described a copper-catalyzed cascade<br />

route to benzimidazoles from the combination of a<br />

1,2-dihalobenzene with an amidine, although only a single<br />

example was reported. 70 Deng et al. reported a related<br />

cascade in which amidines were exchanged for<br />

guanidines, leading to the synthesis of 2-aminobenzimidazoles<br />

(Scheme 26). The majority of examples delivered<br />

N–H products, although N-substitution could also be introduced.<br />

71<br />

Synthesis 2011, No. 1, 1–22 © Thieme Stuttgart · New York<br />

N<br />

N<br />

N<br />

CF3<br />

Ph<br />

N<br />

N<br />

56, 96%


10 J. E. R. Sadig, M. C. Willis REVIEW<br />

Me<br />

HN<br />

H2N<br />

+<br />

N<br />

Scheme 26<br />

2.4 Indazoles <strong>and</strong> Indazolones<br />

Indazoles have proven to be popular targets for amination<br />

chemistry. A number of groups have described the cyclization<br />

of appropriately substituted arylhydrazones.<br />

Scheme 27 illustrates an intramolecular coupling of bromo-substituted<br />

arylhydrazone 57 to deliver 1H-indazole<br />

58 in 85% yield. 72 The DPEPhos-derived catalyst system<br />

was effective for a wide range of substrates, although aryl<br />

chloride substrates performed poorly. Analogous N-tosylhydrazones<br />

were also established as effective indazole<br />

precursors, 73a <strong>and</strong> were utilized in a synthesis of the natural<br />

product nigellicine. 73b 3-Amino-1H-indazoles were<br />

also prepared by similar palladium-catalyzed cyclizations.<br />

74 Song <strong>and</strong> Yee demonstrated that appropriately<br />

substituted hydrazines are also useful indazole precursors.<br />

For example, palladium-catalyzed ring closure using hydrazine<br />

59 delivered the corresponding aromatic 1H-indazole<br />

directly in 87% yield, following intramolecular<br />

amination <strong>and</strong> spontaneous aromatization. 75 The mechanism<br />

of aromatization was not established. The authors<br />

noted the instability of certain hydrazine substrates to<br />

long-term storage, <strong>and</strong> as alternatives established that the<br />

corresponding N-triphenylphosphonium bromide salts<br />

provided convenient stable precursors that could be cyclized<br />

under identical reaction conditions.<br />

MeO<br />

57<br />

Scheme 27<br />

Br<br />

I<br />

Br<br />

N<br />

O<br />

CuI (15 mol%)<br />

25 (30 mol%)<br />

Cs2CO3<br />

DMA, 165 °C<br />

N<br />

N O<br />

N<br />

H<br />

53%<br />

There are a number of reports of halo-substituted hydrazones<br />

being formed in situ <strong>and</strong> then cyclized to yield 1Hindazoles;<br />

Scheme 28 presents examples using both palladium<br />

<strong>and</strong> copper catalysis. Cho et al. were able to show<br />

that o-bromobenzaldehydes could be combined with phenylhydrazine<br />

using a palladium(II) chloride/dppp catalyst<br />

system to furnish the corresponding indazoles in good<br />

yields (60 → 61). 76 The copper example, reported by<br />

Synthesis 2011, No. 1, 1–22 © Thieme Stuttgart · New York<br />

Me<br />

Me<br />

H<br />

N<br />

Pd(dba)2 (2 mol%<br />

N<br />

DPEPhos (22) (2 mol%)<br />

N<br />

Me<br />

NH<br />

K 3PO 4<br />

toluene, 110 °C<br />

MeO<br />

Pd(OAc) 2 (5 mol%)<br />

dppf (12) (7.5 mol%)<br />

Me<br />

58, 85%<br />

Br HN<br />

NaOt-Bu<br />

toluene, 90 59 °C<br />

87%<br />

N<br />

N<br />

Me<br />

Pabba et al., required a two-step one-pot approach in<br />

which a ten-minute microwave reaction was used to form<br />

the hydrazone before a copper(I)/diamine catalyst was<br />

added to the system (62 → 63). 77 Del Olmo <strong>and</strong> co-workers<br />

reported a related copper-catalyzed process which also<br />

allowed the use of aryl carboxylic acid substrates to deliver<br />

1-hydroxy-1H-indazoles. 78<br />

MeO<br />

MeO<br />

F<br />

Scheme 28<br />

Guillaumet <strong>and</strong> co-workers reported an intermolecular<br />

copper-catalyzed amination method for the preparation of<br />

pyrazolopyridines (azaindazoles). 79 3-Cyano-2-chloropyridine<br />

was combined with a range of hydrazines using a<br />

copper(I) iodide/phenanthroline catalyst to deliver 3-amino-1H-azaindazoles<br />

in good yields (Scheme 29). The 3amino<br />

products were converted into the corresponding 3iodo<br />

derivatives by way of their diazonium salts, <strong>and</strong> were<br />

employed in a range of palladium-catalyzed coupling processes<br />

including Stille, Heck <strong>and</strong> Suzuki reactions.<br />

N<br />

CN<br />

Cl<br />

Scheme 29<br />

O<br />

60<br />

O<br />

+<br />

62<br />

H<br />

+ H2N<br />

Br<br />

H<br />

+ H2N<br />

Br<br />

Et<br />

H<br />

N NH2<br />

NH<br />

Ph<br />

NH<br />

Ph<br />

PdCl 2 (2 mol%)<br />

dppp (3 mol%) MeO<br />

NaOt-Bu<br />

toluene, 100 °C<br />

i) NMP, 160 °C<br />

10 min, MW<br />

ii) CuI (5 mol%)<br />

64 (10 mol%)<br />

K2CO 3, 160 °C<br />

10 min, MW<br />

NHMe<br />

NHMe<br />

MeO<br />

61, 65%<br />

N<br />

N<br />

N<br />

N<br />

The less thermodynamically stable 2H-indazole isomers<br />

can also be accessed using amination chemistry. In an approach<br />

mirroring their route to the 1H-isomers (see<br />

Scheme 27), Song <strong>and</strong> Yee employed a palladium-catalyzed<br />

cyclization of appropriately substituted hydrazines.<br />

80 For example, N-alkyl-N-arylhydrazine 65 was<br />

converted into 2-aryl-2H-indazole 66 in 60% yield<br />

(Scheme 30). Katayama <strong>and</strong> co-workers showed that N2–<br />

C3-fused examples can also be prepared using similar<br />

chemistry. 81<br />

64<br />

CuI (5 mol%)<br />

1,10-phenanthroline<br />

(10 mol%)<br />

Cs 2CO 3<br />

DMF, 60 °C<br />

N N<br />

1,10-phenanthroline<br />

F<br />

63, 84%<br />

NH 2<br />

N<br />

N N<br />

86% Et


REVIEW <strong>Palladium</strong>- <strong>and</strong> <strong>Copper</strong>-<strong>Catalyzed</strong> Heterocycle Synthesis 11<br />

N Ph<br />

MeO<br />

Pd(OAc) 2 (5 mol%)<br />

MeO<br />

dppf (12) (7.5 mol%)<br />

NH2 Br<br />

NaOt-Bu<br />

toluene, 90 °C<br />

65<br />

Scheme 30<br />

Hall<strong>and</strong>, Lindenschmidt <strong>and</strong> co-workers reported an alternative<br />

route to the 2H-indazole isomers employing the<br />

same o-alkynylhaloarene substrates used successfully by<br />

others to access indoles (see Schemes 11 <strong>and</strong> 12). The reactions<br />

proceeded via an initial regioselective amination<br />

reaction using a monosubstituted hydrazine to generate an<br />

N,N¢-disubstituted hydrazine, which then underwent intramolecular<br />

hydroamination to form a dihydroindazole<br />

intermediate (67, Scheme 31). 82 Isomerization from these<br />

intermediates to the aromatic 2H-indazoles occurred<br />

spontaneously under the reaction conditions of palladium(II)<br />

chloride/tri(tert-butyl)phosphine with cesium carbonate<br />

in N,N-dimethylformamide. Good functional<br />

group tolerance was demonstrated <strong>and</strong> an extensive range<br />

of substituted products was described; three examples are<br />

shown in Scheme 31.<br />

67<br />

Scheme 31<br />

Indazolones can be prepared by the copper-catalyzed cyclization<br />

of o-halobenzohydrazides. For example, treatment<br />

of hydrazide 68 with a copper(I) iodide/proline<br />

catalyst system delivered indazolone 69 in 60% yield<br />

(Scheme 32). 83 Iodo substrates delivered the most efficient<br />

reactions; the bromo <strong>and</strong> chloro substrates were also<br />

shown to deliver the desired products, albeit in reduced<br />

yields.<br />

Scheme 32<br />

Ph<br />

+ H2N<br />

Cl<br />

O<br />

I<br />

Ph<br />

N Ph<br />

N<br />

H<br />

N<br />

H<br />

H<br />

N<br />

2.5 Pyrroles<br />

PdCl2 (5 mol%)<br />

Pt-Bu3 HBF4 (10 mol%)<br />

NH<br />

Cs2CO<br />

Ph<br />

3<br />

DMF, 110–130 °C<br />

.<br />

93%<br />

CO2t-Bu<br />

N Ph<br />

N<br />

CuI (10 mol%)<br />

L-proline (20 mol%)<br />

K 2CO 3, DMSO<br />

r.t. to 70 °C<br />

68 69, 60%<br />

N Ph<br />

N<br />

66, 60%<br />

N Ph<br />

N<br />

79%<br />

OEt<br />

The Buchwald <strong>and</strong> Li research groups both reported cascade<br />

copper-catalyzed alkenyl amidation routes to pyr-<br />

Ph<br />

N Ph<br />

N<br />

55%<br />

O<br />

OEt<br />

NH<br />

N<br />

roles. The Buchwald group utilized a copper(I) iodide/<br />

diamine 25 catalyst system to combine carbamates (<strong>and</strong> a<br />

limited number of amides) with 1,4-diiodo-1,3-dienes to<br />

generate highly substituted pyrrole products (70 → 71,<br />

Scheme 33). 84 The methodology displayed excellent<br />

functional group tolerance <strong>and</strong> was applied to the synthesis<br />

of a wide range of pyrroles, including tetrasubstituted<br />

examples. The method was also applicable to the synthesis<br />

of heteroarylpyrroles, such as thienopyrrole 72. The Li<br />

approach exploited similar diene substrates in combination<br />

with a range of amide coupling partners. 45,85 For example,<br />

diiododiene 73 was combined with phenylacetamide<br />

using a copper(I) iodide/diamine 64 catalyst to<br />

deliver the expected pyrrole in 86% yield (Scheme 33).<br />

Pr<br />

S<br />

Bu<br />

Bu<br />

SiMe3<br />

Pr<br />

70<br />

73<br />

Scheme 33<br />

I O<br />

I<br />

+<br />

H2N Ot-Bu<br />

Me<br />

Pr<br />

CuI (5 mol%)<br />

25 (20 mol%)<br />

Cs 2CO 3<br />

THF, 80 °C<br />

CuI (20 mol%)<br />

64 (20 mol%)<br />

Cs 2CO 3<br />

dioxane, 100 °C<br />

Pr<br />

N<br />

Boc<br />

71, 86%<br />

N<br />

SiMe3<br />

Pr<br />

N<br />

SiMe3<br />

N<br />

Boc<br />

Boc<br />

Boc<br />

98% 80%<br />

72, 83%<br />

I O<br />

+<br />

I<br />

H2N Ph<br />

Buchwald <strong>and</strong> colleagues reported a complementary stepwise<br />

approach to pyrroles also based on copper-catalyzed<br />

alkenylation reactions. 86 N,N¢-Di(Boc)-protected alkenylhydrazides<br />

74, themselves prepared by copper-catalyzed<br />

alkenylation reactions, were coupled with a second alkenyl<br />

iodide to generate bis(ene)hydrazide 75 (Scheme 34).<br />

Thermolysis triggered a [3,3] rearrangement to generate<br />

bis-imine 76, cyclization of which provided the pyrrole<br />

Pr<br />

Pr<br />

Pr<br />

Boc<br />

Oct<br />

N<br />

Boc<br />

N<br />

N<br />

H<br />

Boc<br />

74<br />

+<br />

Boc<br />

N<br />

Pr Oct<br />

I<br />

[3,3]<br />

Bu<br />

(i) CuI (10 mol%)<br />

1,10-phenanthroline<br />

(20 mol%)<br />

Cs2CO3, DMF<br />

80 °C, 30 h<br />

ii) o-xylene, 140 °C, 30 h<br />

iii) p-TsOH, r.t.<br />

Boc Boc<br />

N N<br />

Pr<br />

Pr Oct<br />

Synthesis 2011, No. 1, 1–22 © Thieme Stuttgart · New York<br />

Pr<br />

S<br />

Bu<br />

N<br />

Bu<br />

O<br />

86% Ph<br />

75 76 77<br />

Scheme 34<br />

Pr<br />

Pr<br />

Pr Oct<br />

N<br />

68% Boc<br />

Boc<br />

N<br />

Pr Oct<br />

SiMe3<br />

H<br />

NBoc


12 J. E. R. Sadig, M. C. Willis REVIEW<br />

via intermediate 77. For certain substrates it was necessary<br />

to add acid to achieve complete conversion into the<br />

aromatic system. The bis(ene)hydrazides could be isolated,<br />

or more conveniently used directly in the next step of<br />

the sequence with no purification. The overall method<br />

represents a modified Piloty–Robinson reaction.<br />

The final pyrrole synthesis considered here comprises a<br />

copper-catalyzed cascade alkenyl amidation/hydroamination<br />

sequence. The process is essentially a pyrroleforming<br />

version of the indole synthesis described in<br />

Scheme 11. In this approach, described by Buchwald <strong>and</strong><br />

co-workers, haloenynes such as 78 replace the alkynylhaloarenes<br />

used for indole synthesis, <strong>and</strong>, when combined<br />

with tert-butyl carbamate <strong>and</strong> a copper(I) iodide/diamine<br />

catalyst, provide the corresponding pyrroles in good<br />

yields (Scheme 35). 87 A broad range of di- <strong>and</strong> trisubstituted<br />

pyrroles were prepared, <strong>and</strong> although the study focused<br />

on the use of iodoenynes, it was also possible to<br />

employ bromoenynes. A brief mechanistic study established<br />

that the reactions likely proceed via initial intermolecular<br />

aryl carbon–nitrogen bond formation followed by<br />

a 5-endo-dig intramolecular hydroamination.<br />

Pent<br />

78<br />

OTIPS<br />

I<br />

+<br />

t-BuO NH2<br />

Scheme 35<br />

O<br />

2.6 Pyrazoles<br />

25<br />

CuI (5 mol%)<br />

(20 mol%) OTIPS<br />

Cs2CO3 THF, 80 °C<br />

Pent<br />

N<br />

Boc<br />

83%<br />

Buchwald <strong>and</strong> co-workers utilized haloenyne substrates<br />

in a t<strong>and</strong>em copper-catalyzed pyrazole synthesis. 87 For<br />

example, combination of iodoenyne 78 with bis(Boc)hydrazine<br />

using a copper(I) iodide/diamine catalyst provided<br />

pyrazole 79 in 78% yield after Boc-deprotection with<br />

trifluoroacetic acid (Scheme 36). As in the related pyrrole<br />

syntheses, a good range of di- <strong>and</strong> trisubstituted aromatics<br />

were prepared. The mechanism was again established as<br />

proceeding via intermolecular aryl carbon–nitrogen bond<br />

formation followed by intramolecular hydroamination, although<br />

the cyclizations were in this case 5-exo-dig processes.<br />

Cho <strong>and</strong> Patel reported a pyrazole synthesis based on the<br />

palladium-catalyzed combination of b-bromovinyl aldehydes<br />

with hydrazines. 88 For example, treatment of<br />

Pent<br />

25 78<br />

(i) CuI (5 mol%)<br />

(20 mol%)<br />

Pent<br />

I<br />

+<br />

H<br />

N<br />

Boc NH<br />

Boc<br />

OTIPS<br />

Cs2CO3 THF, 80 °C<br />

ii) TFA, CH2Cl2, r.t.<br />

Scheme 36<br />

N<br />

N<br />

H<br />

79, 78%<br />

Synthesis 2011, No. 1, 1–22 © Thieme Stuttgart · New York<br />

OTIPS<br />

bromo-enal 80 with phenylhydrazine in the presence of a<br />

dppf-derived catalyst yielded pyrazole 81 in 77%<br />

(Scheme 37). Both cyclic <strong>and</strong> acyclic substrates could be<br />

employed, although no ketone-derived substrates were reported.<br />

Haddad <strong>and</strong> co-workers developed a pyrazole synthesis<br />

in which aryl benzophenone hydrazones, prepared<br />

by the palladium-catalyzed coupling of benzophenone hydrazone<br />

with aryl halides, were treated with a range of<br />

1,3-bifunctional substrates under acidic conditions. 89 1,3-<br />

Diketones, keto esters <strong>and</strong> ester/acid chlorides could be<br />

combined with the hydrazones to provide a range of substituted<br />

pyrazole products.<br />

80<br />

O<br />

H<br />

+<br />

Br<br />

Scheme 37<br />

H2N Pd(OAc) 2 (5 mol%)<br />

dppf (12) (7.5 mol%)<br />

NH NaOt-Bu, toluene<br />

Ph 125 °C<br />

N<br />

N<br />

Ph<br />

81, 77%<br />

2.7 Oxazoles<br />

Buchwald <strong>and</strong> co-workers utilized copper-catalyzed alkenyl<br />

iodide amidation reactions as the key step in a route to<br />

oxazoles. 90 For example, enamides such as 82 could be<br />

treated with iodine <strong>and</strong> base to provide the expected<br />

trisubstituted oxazoles in good yields (Scheme 38). The<br />

required enamides were prepared from the corresponding<br />

alkenyl bromides using a copper(I) iodide/diamine-catalyzed<br />

coupling with the appropriate amide; both aryl <strong>and</strong><br />

alkyl amides could be used. Depending on the substitution<br />

pattern of the oxazole product, certain examples required<br />

the addition of p-toluenesulfonic acid to achieve complete<br />

conversion into the aromatic molecule. Attempts to prepare<br />

mono- <strong>and</strong> disubstituted oxazoles using this method<br />

resulted in complex reaction mixtures, <strong>and</strong> an alternative<br />

process, based on the use of 1,2-dihaloalkene substrates,<br />

was developed.<br />

Ph Br<br />

+<br />

O<br />

Ph<br />

H2N Ph<br />

Scheme 38<br />

Ph<br />

Ph<br />

H<br />

N<br />

82<br />

O<br />

2.8 Quinolones<br />

i) CuI (5 mol%)<br />

(20 mol%)<br />

25<br />

Cs2CO3, THF, 80 °C<br />

ii) I2, DBU<br />

r.t. to 80 °C<br />

Ph<br />

Ph<br />

Ph<br />

Manley <strong>and</strong> Bilodeau used palladium-catalyzed intermolecular<br />

aryl halide amidation followed by an in situ aldol<br />

condensation to prepare 2-quinolones. 91 In this way o-bromobenzaldehydes<br />

were combined with a range of enolizable<br />

amides to deliver the quinolone products. Scheme 39<br />

I<br />

H<br />

N<br />

O<br />

Ph<br />

Ph<br />

Ph<br />

N<br />

O<br />

77%<br />

Ph


REVIEW <strong>Palladium</strong>- <strong>and</strong> <strong>Copper</strong>-<strong>Catalyzed</strong> Heterocycle Synthesis 13<br />

illustrates the combination of the parent aldehyde (83)<br />

with phenylacetamide using a XantPhos-derived catalyst<br />

in combination with cesium carbonate as base. A variety<br />

of substituted amides could be employed, as could ketonebased<br />

substrates, allowing the synthesis of 4-substituted<br />

products such as quinolone 84. Pyridine derivatives could<br />

also be employed, thus providing a route to naphthyridinones.<br />

Ester- <strong>and</strong> nitrile-substituted aryl bromides were<br />

also examined <strong>and</strong> allowed access to hydroxy- <strong>and</strong> aminosubstituted<br />

products, respectively. For a small number of<br />

amide coupling partners, the researchers were able to employ<br />

a copper(I) iodide/diamine catalyst system.<br />

83<br />

Scheme 39<br />

O<br />

O<br />

H<br />

+ H2N<br />

Br<br />

N<br />

H<br />

94%<br />

O<br />

N<br />

Ph<br />

Pd 2(dba)3 (1 mol%)<br />

XantPhos (2) (3 mol%)<br />

Cs2CO3<br />

toluene, 100 °C<br />

N<br />

H<br />

84, 55%<br />

N<br />

H<br />

94%<br />

Huang et al. utilized the related o-acetylbromoarenes as<br />

substrates in a palladium-catalyzed synthesis of 4-quinolones.<br />

The simplest version of the process employed formamide<br />

as the nitrogen nucleophile <strong>and</strong>, when coupled<br />

with bromide 85, delivered quinolone product 86 in 77%<br />

yield (Scheme 40). 92 The reactions proceeded via initial<br />

palladium-catalyzed amidation followed by a base-promoted<br />

intramolecular condensation step. A XantPhosderived<br />

catalyst in combination with cesium carbonate as<br />

base was optimal for the amidation step, <strong>and</strong> the addition<br />

of sodium tert-butoxide as a second base was found to be<br />

necessary to achieve high yields for the combined process.<br />

As can be seen from the examples presented in<br />

Scheme 40, a range of substituted amides could be employed,<br />

including lactams, which allowed the synthesis of<br />

N-fused products such as quinolone 87. The Buchwald research<br />

group reported a related process based on coppercatalyzed<br />

amidation, although in their case it was neces-<br />

MeO<br />

N<br />

H<br />

Scheme 40<br />

85<br />

O<br />

Ph<br />

Ph<br />

O<br />

Me<br />

Br H O<br />

Pd2(dba) 3 (1 mol%)<br />

XantPhos (2) (2.5 mol%)<br />

+<br />

2N H<br />

Cs2CO3, dioxane<br />

100 °C, 2–48 h then<br />

NaOt-Bu, 100 °C<br />

MeO<br />

O<br />

N<br />

H<br />

O<br />

N<br />

N<br />

H<br />

75%<br />

Ph<br />

O<br />

Ph<br />

O<br />

O<br />

N<br />

H<br />

86, 77%<br />

S<br />

82% 91% 87, 85%<br />

Me<br />

O<br />

N<br />

sary to isolate the initial amidation products before cyclization.<br />

93<br />

A t<strong>and</strong>em Heck/intramolecular amidation strategy was reported<br />

by Cacchi <strong>and</strong> co-workers as a route to 4-aryl-2quinolones.<br />

94 Using molten tetrabutylammonium acetate/<br />

tetrabutylammonium bromide as the reaction medium <strong>and</strong><br />

a simple palladium(II) acetate catalyst, the combination of<br />

o-bromophenylacrylamides (88) <strong>and</strong> aryl iodides provided<br />

the quinolone products in moderate to good yields<br />

(Scheme 41). Although only a single acrylamide substrate<br />

was employed, a good range of aryl iodide coupling partners<br />

could be incorporated. A brief mechanistic investigation<br />

supported the Heck followed by intramolecular<br />

amidation pathway.<br />

88<br />

+<br />

Br<br />

Scheme 41<br />

Willis <strong>and</strong> co-workers exploited 2-(2-haloalkenyl)aryl halide<br />

substrates, previously employed in indole syntheses<br />

(see Scheme 8), in the preparation of 2-quinolones. 95 A<br />

cascade palladium-catalyzed carbonylation/intramolecular<br />

amidation sequence was employed to access a range of<br />

quinolone products. For example, combination of the simple<br />

dibromide 89 <strong>and</strong> p-methoxybenzylamine under a balloon<br />

pressure of carbon monoxide delivered quinolone 90<br />

in 80% yield (Scheme 42). Although all of the products<br />

shown in Scheme 42 were obtained using a dppp-derived<br />

catalyst, the researchers found that lig<strong>and</strong> variation was<br />

needed for particular substrate/amine combinations. In<br />

addition, purging the reaction of carbon monoxide was<br />

shown to benefit the efficiency of certain amidation reactions.<br />

By delaying the introduction of the carbon monoxide<br />

<strong>and</strong> running the reaction in a two-stage process, it was<br />

also possible access the regioisomeric isoquinolone products,<br />

although in these cases competing indole formation<br />

was problematic.<br />

O<br />

Scheme 42<br />

O<br />

I<br />

NH 2<br />

+<br />

H2N<br />

Br<br />

Br<br />

89<br />

Pd(OAc)2 (5 mol%)<br />

n-Bu4NOAc, n-Bu4NBr<br />

120 °C<br />

CO (balloon)<br />

Pd2(dba)3 (3 mol%)<br />

dppp (6 mol%)<br />

Cs2CO3, toluene<br />

100 °C<br />

OMe<br />

N<br />

H<br />

75%<br />

Synthesis 2011, No. 1, 1–22 © Thieme Stuttgart · New York<br />

O<br />

N O<br />

PMB<br />

90, 80%<br />

O N O<br />

MeO<br />

N O N N O<br />

Oct<br />

O Oct<br />

Oct<br />

69% 65% 73%


14 J. E. R. Sadig, M. C. Willis REVIEW<br />

2.9 Quinazolines, Quinazolinones <strong>and</strong><br />

Quinazolindiones<br />

In 2006 Willis <strong>and</strong> co-workers reported the palladiumcatalyzed<br />

coupling of o-bromobenzoate esters with<br />

monosubstituted ureas as a route to 3-alkylated quinazolinediones.<br />

96 The reactions proceeded via initial intermolecular<br />

carbon–nitrogen bond formation followed by<br />

intramolecular, base-promoted, amidation. A XantPhosderived<br />

catalyst in combination with cesium carbonate allowed<br />

efficient <strong>and</strong> regioselective processes; for example,<br />

bromobenzoate 91 was combined with N-butyl urea to<br />

provide quinazolinedione 92 in 77% yield as a single regioisomer<br />

(Scheme 43). A variety of substituents were<br />

tolerated on both the aryl bromide <strong>and</strong> urea coupling partners.<br />

The observed regiocontrol originates from the initial<br />

aryl carbon–nitrogen bond formation taking place at the<br />

least hindered nitrogen of the urea nucleophile.<br />

Cl<br />

91<br />

Scheme 43<br />

O<br />

OMe<br />

Br<br />

+<br />

Bu<br />

N<br />

H<br />

O<br />

Pd 2(dba) 3 (2.5 mol%)<br />

XantPhos (2) (5 mol%)<br />

NH2<br />

Cs 2CO 3<br />

dioxane, 100 °C<br />

N<br />

H<br />

92, 77%<br />

The Fu research group reported a related strategy for the<br />

synthesis of quinazolinones <strong>and</strong> quinazolines. In their<br />

original report, they exploited a copper-catalyzed coupling<br />

of amidines with o-bromobenzoic acids to access<br />

quinazolinones. For example, acid 93 was combined with<br />

acetimidamide 94 to provide quinazolinone 95 in 81%<br />

yield (Scheme 44). 97a The reaction conditions consisted of<br />

copper(I) iodide without any added lig<strong>and</strong>, in N,N-dimethylformamide<br />

at room temperature. The ability to use<br />

such low-temperature conditions was attributed to the formation<br />

of a chelated intermediate involving the oxygen<br />

atom of the ortho-positioned carboxylic acid. The reaction<br />

was applicable to a broad range of amidines <strong>and</strong> benzoic<br />

acids. The researchers next extended the chemistry to include<br />

guanidines as the nitrogen nucleophiles, resulting in<br />

the formation of 3-aminoquinazolinone products such as<br />

96. 97b The carboxylic acid substrates could also be replaced;<br />

the use of the related ketones in combination with<br />

guanidines resulted in the synthesis of 3-aminoquinazolines<br />

such as 97. Ding <strong>and</strong> co-workers reported an alternative<br />

copper-catalyzed route to quinazolinones based on<br />

the use of o-iodobenzamide substrates. 98 A typical reaction<br />

is shown in Scheme 44: coupling of benzamide 98<br />

with formimidamide, using copper(I) iodide as catalyst,<br />

provided quinazolinone 99 in 77% yield. The method was<br />

shown to tolerate reasonable variation of both reaction<br />

components.<br />

Li <strong>and</strong> co-workers reported a cascade process involving in<br />

situ amidine formation followed by palladium-catalyzed<br />

Synthesis 2011, No. 1, 1–22 © Thieme Stuttgart · New York<br />

Cl<br />

O<br />

N<br />

Bu<br />

O<br />

O<br />

93 95, 81%<br />

94<br />

OH<br />

CuI (20 mol%)<br />

NH<br />

+<br />

H2N<br />

Br<br />

NH.HCl Cs2CO3 DMF, r.t. N Me<br />

Me<br />

75%<br />

O<br />

N<br />

cyclization as an entry into ring-fused quinazolinone derivatives.<br />

The key amidine intermediates were generated<br />

from intramolecular amide addition to a nitrile; the nitrile<br />

group also being introduced by a palladium-catalyzed<br />

process. 99 Scheme 45 outlines the overall conversion,<br />

with aryl iodide 100 being transformed into quinazolinone<br />

101 in 91% yield. A DPEPhos-derived catalyst was effective<br />

for both the initial cyanation reaction to generate nitrile<br />

102, <strong>and</strong> then to achieve the final ring closure from<br />

the proposed amidine intermediate to give quinazolinone<br />

101. Using a bromoquinoline-derived substrate, the authors<br />

were able to apply the methodology to a synthesis of<br />

the natural product luotonin (103).<br />

Scheme 45<br />

NH<br />

2.10 Phenazines<br />

O<br />

N<br />

96, 81%<br />

NH<br />

The Kamikawa <strong>and</strong> Beifuss groups have both reported intramolecular<br />

palladium-catalyzed aryl-amination routes<br />

to phenazines. Both explored the ring-closure of 2-amino-<br />

2¢-bromodiphenylamines to access the target systems.<br />

Scheme 46 shows an example from Beifuss <strong>and</strong> co-workers,<br />

in which a JohnPhos-derived catalyst was used to convert<br />

aniline 104 into phenazine 105 in 76% yield. 100 The<br />

Kamikawa research group utilized BINAP-derived cata-<br />

N<br />

O<br />

O<br />

Ph<br />

N<br />

97, 56%<br />

98<br />

O<br />

O<br />

I<br />

Scheme 44<br />

Ph<br />

N<br />

H<br />

+ H2N NH<br />

AcOH<br />

CuI (10 mol%)<br />

K2CO3 DMF, 80 °C<br />

N<br />

99, 77%<br />

N<br />

Ph<br />

O<br />

NH<br />

Pd(OAc)2 (5 mol%)<br />

DPEPhos (22) (10 mol%)<br />

O<br />

I<br />

Br<br />

+ KCN<br />

dioxane, reflux<br />

then dppf (10 mol%)<br />

K2CO3<br />

N<br />

N<br />

100 101, 91%<br />

N<br />

H<br />

O<br />

CN Br<br />

102<br />

N<br />

N<br />

O<br />

N<br />

N<br />

91%<br />

luotonin (103)<br />

N


REVIEW <strong>Palladium</strong>- <strong>and</strong> <strong>Copper</strong>-<strong>Catalyzed</strong> Heterocycle Synthesis 15<br />

MeO<br />

Scheme 46<br />

lysts. 101 In both cases the phenazine ring system was isolated<br />

directly from the amination reactions.<br />

2.11 Cinnolines<br />

An intermolecular copper-catalyzed aryl amination was<br />

used by Nishida <strong>and</strong> co-workers to access cinnoline derivatives.<br />

In the example shown in Scheme 47, hydrazonesubstituted<br />

aryl iodide 107 was converted into N-acyldihydrocinnoline<br />

108 using a copper(I) iodide/diamine<br />

catalyst. 102 The use of superstoichiometric amounts of catalyst<br />

led to mixtures of N-acyl product 108 together with<br />

smaller amounts (up to 40%) of the aromatic cinnoline being<br />

obtained. Cyclization of hydrazines derived from hydrazone<br />

107 allowed access to 1-aminoindoles.<br />

I<br />

NH2<br />

107<br />

Scheme 47<br />

H<br />

N<br />

Br<br />

104<br />

OTBS<br />

CO2t-Bu HN<br />

Ac<br />

N<br />

3 Carbon–Oxygen Bond Formation<br />

Initially, the development of catalytic carbon–oxygen<br />

bond-forming processes using aryl halide substrates<br />

lagged behind the corresponding carbon–nitrogen forming<br />

reactions; however, efficient methods, using both palladium<br />

<strong>and</strong> copper catalysts, are now well established.<br />

3.1 Benzofurans<br />

Pd 2(dba) 3 (3 mol%)<br />

JohnPhos (6 mol%)<br />

NaOt-Bu<br />

toluene, 100 °C MeO<br />

P(t-Bu)2<br />

JohnPhos (106)<br />

CuI (10 mol%)<br />

25 (10 mol%)<br />

Cs 2CO 3<br />

DMSO, r.t.<br />

N N<br />

Ac<br />

108, 89%<br />

Few examples of benzofuran syntheses that proceed via a<br />

metal-catalyzed intermolecular (aryl)carbon–oxygen<br />

bond-forming reaction exist. In one example, Buchwald<br />

<strong>and</strong> co-workers were able to apply their palladium-catalyzed<br />

phenol synthesis to the preparation of benzofurans.<br />

103 The chemistry was based on the use of<br />

potassium hydroxide as a nucleophile in the palladiumcatalyzed<br />

hydroxylation of aryl halides to provide phenols.<br />

When applied to benzofuran synthesis, o-chloroarylalkyne<br />

substrates reacted with potassium hydroxide in the<br />

presence of t-Bu-XPhos as catalyst, to give o-hydroxyalkynylarenes,<br />

which, as previously shown, 104 undergo<br />

N<br />

N<br />

105, 76%<br />

OTBS<br />

CO2t-Bu cyclization to the required benzofurans (109 → 110,<br />

Scheme 48). You’s research group went on to develop a<br />

copper-catalyzed version of the hydroxylation reaction<br />

<strong>and</strong> also demonstrated its use in benzofuran synthesis, in<br />

this case from an o-iodoarylalkyne to generate benzofuran<br />

112. 105<br />

F 3C<br />

Cl<br />

109<br />

Scheme 48<br />

A greater number of research groups have utilized intramolecular<br />

carbon–oxygen bond formation as the key<br />

step in benzofuran syntheses. In 2004 Willis et al. demonstrated<br />

the use of a-(o-haloaryl) ketones as precursors to<br />

the required oxygen heterocycles via an enolization/palladium-catalyzed<br />

intramolecular O-arylation reaction, with<br />

a Pd 2(dba) 3/DPEPhos catalyst system proving optimum<br />

for the process (Scheme 49). 106 The starting ketones were<br />

themselves formed by a palladium-catalyzed ketone arylation;<br />

however, attempts to achieve a one-pot combination<br />

of these processes was not straightforward, <strong>and</strong> after<br />

optimization only a single high-yielding example of the<br />

cascade could be achieved. Kotschy <strong>and</strong> co-workers<br />

showed that the same cyclization of o-bromobenzyl ketones,<br />

which they accessed from aromatic aldehydes <strong>and</strong><br />

2-bromobenzyl bromide using dithiane chemistry, is possible<br />

using a palladium–NHC catalyst system. 107<br />

O<br />

Scheme 49<br />

Cl<br />

I<br />

O<br />

Br<br />

Me<br />

86%<br />

(NaOt-Bu)<br />

Ph<br />

S<br />

KOH<br />

F3C<br />

Pd2(dba) 3 (2 mol%)<br />

t-BuXPhos (8 mol%)<br />

H2O, dioxane<br />

100 °C<br />

KOH<br />

Ph CuI (10 mol%)<br />

1,10-phenathroline<br />

(20 mol%)<br />

H 2O, DMSO<br />

100 °C<br />

t-Bu2P i-Pr<br />

i-Pr<br />

t-BuXPhos (111)<br />

Cs2CO3<br />

toluene, 100 °C<br />

Synthesis 2011, No. 1, 1–22 © Thieme Stuttgart · New York<br />

Cl<br />

i-Pr<br />

Pd 2(dba) 3 (2.5 mol%)<br />

DPEPhos (22) (6 mol%)<br />

N<br />

O<br />

86%<br />

(NaOt-Bu)<br />

(chloro substrate)<br />

F<br />

O<br />

110, 87%<br />

O<br />

112, 86%<br />

O<br />

95%<br />

O<br />

74%<br />

(NaOt-Bu)<br />

Ph<br />

S


16 J. E. R. Sadig, M. C. Willis REVIEW<br />

Chen <strong>and</strong> Dormer reported a copper(I) iodide catalyzed,<br />

lig<strong>and</strong>-free modification of this benzofuran synthesis using<br />

o-iodo- or o-bromoaromatic ketones. 108 A range of<br />

substituents at the resulting 2- <strong>and</strong> 3-positions of the product<br />

benzofuran were tolerated, <strong>and</strong> the first use of an aldehyde<br />

substrate was demonstrated, yielding 3benzylbenzo[b]furan<br />

113 in 92% yield (Scheme 50). An<br />

on-water variant of this protocol was described by the<br />

SanMartin <strong>and</strong> Domínguez group, with the use of a diamine<br />

lig<strong>and</strong> being required. 109 Ackermann <strong>and</strong> Kaspar<br />

also utilized a related copper-catalyzed cyclization in<br />

combination with alkyne hydration chemistry to access<br />

benzofurans. 110<br />

O<br />

Br<br />

Scheme 50<br />

Willis <strong>and</strong> co-workers developed a second strategy to<br />

access benzofurans, again based on an intramolecular<br />

carbon–oxygen bond-forming reaction. Scheme 8 highlighted<br />

the use of aryl halide/alkenyl triflate substrates in<br />

the synthesis of indole derivatives; the research group was<br />

able to further demonstrate the utility of these substrates<br />

as general heterocycle precursors, through their use in a<br />

copper-catalyzed benzofuran synthesis. 111 Coupling of the<br />

same substrates with potassium hydroxide yielded benzofurans<br />

via presumed enolate intermediates.<br />

Lautens <strong>and</strong> co-workers reported an approach to 2-bromobenzofurans<br />

using an intramolecular carbon–oxygen<br />

coupling of gem-dibromovinyl phenols. 31i As in the related<br />

indole chemistry (see Scheme 5), the dibromovinyl<br />

phenol substrates were synthesized using a Ramirez olefination<br />

process. As shown in Scheme 51, a lig<strong>and</strong>-free<br />

copper(I) iodide catalyst was found to be effective, providing<br />

the benzofurans in excellent yields.<br />

Br<br />

Scheme 51<br />

H<br />

Br<br />

OH<br />

Br<br />

Cl<br />

CuI (10 mol%)<br />

K3PO4 DMF, 105 °C<br />

CuI (5 mol%)<br />

K3PO4<br />

THF, 80 °C<br />

H<br />

O<br />

113, 92%<br />

In 1999 Miura <strong>and</strong> co-workers developed a t<strong>and</strong>em palladium-catalyzed<br />

intermolecular carbon–carbon/intramolecular<br />

carbon–oxygen bond-forming reaction of benzyl<br />

phenyl ketones with o-dibromoarenes to yield benzofurans.<br />

112 A palladium(II) acetate/triphenylphosphine catalyzed<br />

system was found to be optimal, with high reaction<br />

temperatures also being used (Scheme 52). The reactions<br />

proceeded via initial palladium-catalyzed enolate C-arylation,<br />

followed by palladium-catalyzed O-arylation. The<br />

protocol was extended to the use of phenol coupling part-<br />

Synthesis 2011, No. 1, 1–22 © Thieme Stuttgart · New York<br />

Br<br />

O<br />

96%<br />

Br<br />

Cl<br />

ners in place of ketones, to give dibenzofuran products<br />

(114 → 115). SanMartin, Domínguez <strong>and</strong> co-workers reported<br />

a similar method for the synthesis of benzofurans;<br />

113 in their account they compared the use of<br />

homogeneous <strong>and</strong> heterogeneous polymer-supported palladium<br />

catalysis, with the latter affording the heterocycles<br />

in slightly inferior yields.<br />

MeO<br />

MeO<br />

t-Bu<br />

OH<br />

114<br />

Scheme 52<br />

Ma <strong>and</strong> co-workers reported a related copper-catalyzed<br />

cascade route to benzofurans. In the optimized system, bketo<br />

esters were combined with 1-bromo-2-iodobenzenes<br />

to provide benzofurans via initial intermolecular carbon–<br />

carbon bond formation followed by intramolecular formation<br />

of the carbon–oxygen bond (Scheme 53). 114 Substitution<br />

on both the aryl halide <strong>and</strong> keto ester was well<br />

tolerated, providing benzofurans in good yields.<br />

Ph<br />

Cl<br />

O<br />

Scheme 53<br />

+<br />

+<br />

Br<br />

Br<br />

O<br />

Ph<br />

3.2 Benzoxazoles<br />

+<br />

O<br />

I<br />

Br<br />

OEt<br />

Pd(OAc) 2 (5 mol%)<br />

Ph3P (20 mol%)<br />

Cs2CO3 o-xylene, 160 °C<br />

Br<br />

Pd(OAc)2–4Ph3P<br />

(5 mol%)<br />

O<br />

78%<br />

CsCO3<br />

Br o-xylene, 160 °C<br />

O<br />

t-Bu<br />

115, 66%<br />

CuI (10 mol%)<br />

K2CO3 THF, 100 °C<br />

The use of catalytic intramolecular carbon–oxygen bondforming<br />

reactions has proved to be a popular route to<br />

benzoxazoles. In 2006 Batey <strong>and</strong> Evindar developed a<br />

copper-catalyzed cyclization of o-halobenzanilides to<br />

generate a variety of alkyl, aryl, benzyl, alkenyl, dienyl<br />

<strong>and</strong> heterocyclic 2-substituted benzoxazoles (as well as a<br />

h<strong>and</strong>ful of benzothiazoles – see section 4.2). As shown in<br />

Scheme 54 the optimal catalyst was a copper(I) iodide/<br />

phenanthroline combination. 115 The majority of examples<br />

employed aryl bromide substrates, although the iodo derivatives<br />

also performed well. A single aryl chloride example<br />

was included. SanMartin, Domínguez <strong>and</strong> coworkers<br />

reported a similar copper-catalyzed cyclization to<br />

access benzoxazoles. They developed two catalyst systems,<br />

using either copper(I) chloride or copper(II) triflate<br />

in combination with N,N,N¢,N¢-tetramethylethylenediamine<br />

(TMEDA), on water, to yield the desired heterocy-<br />

Cl<br />

O<br />

78%<br />

Ph<br />

CO2Et<br />

OMe<br />

OMe<br />

Ph


REVIEW <strong>Palladium</strong>- <strong>and</strong> <strong>Copper</strong>-<strong>Catalyzed</strong> Heterocycle Synthesis 17<br />

cles in good yields from either the o-iodo, o-bromo- or ochlorobenzanilides.<br />

116 Similar cyclizations have also been<br />

reported by Jiang <strong>and</strong> Ma 117 <strong>and</strong> Kantam <strong>and</strong> co-workers.<br />

118<br />

H<br />

N<br />

Br<br />

N<br />

Scheme 54<br />

In 2009 Sun <strong>and</strong> co-workers extended this type of cyclization<br />

to the synthesis of 2-substituted oxazolopyridines. 119<br />

As shown in Scheme 55, a copper(I) iodide/diamine or<br />

copper(I) iodide/phenanthroline catalyst system proved to<br />

be optimal, yielding the desired heterocycles from the cyclization<br />

of o-chloro- or o-bromopyridylamides, respectively.<br />

Me<br />

N<br />

Scheme 55<br />

CuI (5 mol%)<br />

1,10-phenanthroline<br />

(10 mol%)<br />

O OMe Cs 2CO 3<br />

DME, relfux<br />

O Ph<br />

89%<br />

N N<br />

O<br />

90%<br />

Cbz<br />

25<br />

H<br />

N<br />

CuI (5 mol%)<br />

(10 mol%)<br />

N<br />

O<br />

Cl<br />

K2CO3 toluene, reflux<br />

N O<br />

91%<br />

N<br />

H<br />

N<br />

O<br />

Br<br />

Ph<br />

Cl<br />

CuI (5 mol%)<br />

1,10-phenanthroline<br />

(10 mol%)<br />

Cs2CO3<br />

THF, reflux<br />

MeO<br />

N<br />

The Batey research group developed a further synthesis of<br />

benzoxazoles involving intramolecular carbon–oxygen<br />

bond formation. The substrates were again o-halobenzanilides;<br />

however, these were formed in situ from the reaction<br />

of o-bromoanilines <strong>and</strong> acyl chlorides. 70 A copper(I)<br />

iodide/1,10-phenanthroline combination was found to be<br />

the optimal catalyst system for this one-pot strategy. In<br />

addition, the use of microwave irradiation gave substantially<br />

better results than conventional heating. A library of<br />

24 benzoxazoles was prepared, examples of which are<br />

shown in Scheme 56.<br />

A t<strong>and</strong>em approach to benzoxazoles was reported by<br />

Glorius <strong>and</strong> Altenhoff, in which reaction of o-dihalobenzenes<br />

with amides underwent copper(I) iodide/diamine<br />

catalyzed carbon–nitrogen followed by carbon–oxygen<br />

cross-couplings to yield the desired heterocycles. 120 Examples<br />

of diiodo-, dibromo- <strong>and</strong> mixed dihalobenzenes,<br />

as well as dihalopyridines (Br <strong>and</strong> Cl) were successfully<br />

coupled to benzamide affording 2-phenylbenzoxazoles<br />

(Scheme 57). The dibromobenzene substrate was utilized<br />

to demonstrate variation of the amide partner, giving aryl,<br />

Me<br />

N<br />

O<br />

99%<br />

N<br />

O<br />

90%<br />

97%<br />

N<br />

O<br />

Ph<br />

Ph<br />

Cl<br />

Me<br />

S<br />

O<br />

+<br />

Scheme 56<br />

alkyl, vinyl <strong>and</strong> heterocyclic 2-substituted benzoxazoles.<br />

The use of o-bromochlorobenzenes allowed the regioselective<br />

synthesis of substituted benzoxazoles, with the reaction<br />

proceeding via initial amidation at the aryl bromide<br />

position. Batey <strong>and</strong> co-workers had previously reported<br />

on investigations of related processes, but had not<br />

achieved an efficient system. 70<br />

Scheme 57<br />

Cl<br />

O<br />

NH 2<br />

Br<br />

N<br />

3.3 Isocoumarins<br />

CuI (10 mol%)<br />

1,10-phenanthroline<br />

(20 mol%)<br />

Cs 2CO3, MeCN<br />

210 °C, 15 min, MW<br />

O<br />

91%<br />

In 1999 Shen <strong>and</strong> Wang described the synthesis of isocoumarins<br />

from a palladium-catalyzed reaction of gem-dibromovinyl<br />

benzoates with an organostannane. 121 For<br />

example, reaction of benzoate 114 with phenyltrimethylstannane<br />

using a trifurylphosphine-derived palladium catalyst<br />

delivered isocoumarin 115 in 92% yield (Scheme<br />

58). The t<strong>and</strong>em process is believed to proceed via an initial<br />

Stille reaction of the ‘E’ bromide with the stannane,<br />

which is followed by an intramolecular carbon–oxygen<br />

bond-forming cyclization <strong>and</strong> ensuing elimination of methyl<br />

bromide. Examples using phenyl, furyl, thienyl <strong>and</strong><br />

vinyl tin reagents gave the 3-substituted isocoumarins in<br />

mostly excellent yields, <strong>and</strong> both esters <strong>and</strong> methoxy<br />

groups could be tolerated on the aromatic ring. Willis <strong>and</strong><br />

co-workers reported a palladium-catalyzed carbonylative<br />

isocoumarin synthesis, commencing from the same a-(o-<br />

Synthesis 2011, No. 1, 1–22 © Thieme Stuttgart · New York<br />

Me<br />

O O Ph<br />

F3C<br />

O<br />

66%<br />

85%<br />

Br<br />

+<br />

Br<br />

O OMe<br />

114<br />

85% O<br />

Scheme 58<br />

O<br />

H 2N Ph<br />

Br<br />

Br<br />

+<br />

PhSnMe3 O<br />

O<br />

CuI (5 mol%)<br />

25 (10 mol%)<br />

K2CO 3<br />

toluene, 110 °C<br />

80%<br />

Pd2(dba)3 (2.5 mol%)<br />

P(2-furyl)3 (15 mol%)<br />

toluene, 100 °C<br />

O<br />

O<br />

S<br />

MeO<br />

N<br />

MeO<br />

N<br />

N<br />

O<br />

90%<br />

115, 92%<br />

81%<br />

O<br />

O<br />

S<br />

Ph<br />

O<br />

O<br />

Ph<br />

Ph


18 J. E. R. Sadig, M. C. Willis REVIEW<br />

haloaryl) ketone substrates previously used in benzofuran<br />

synthesis (see Scheme 49). 122<br />

4 Carbon–Sulfur Bond Formation<br />

Relative to the carbon–nitrogen <strong>and</strong> carbon–oxygen<br />

bond-forming reactions discussed above, there are far<br />

fewer examples of catalytic aryl carbon–sulfur bondforming<br />

processes in the literature. However, reactions<br />

are beginning to be developed that exploit the highly nucleophilic<br />

character of thiols (<strong>and</strong> related functional<br />

groups) <strong>and</strong> synthetically useful methods with very low<br />

catalyst loadings are being reported. 123 The number of applications<br />

of these reactions to the synthesis of heterocycles<br />

is also growing.<br />

4.1 Benzothiophenes<br />

Although a number of metal-catalyzed benzothiophene<br />

syntheses exist, 1,2 few of these involve a key carbon–sulfur<br />

bond formation using an aryl halide substrate. As discussed<br />

in section 3.1, Willis et al. demonstrated the use of<br />

a palladium-catalyzed intramolecular O-enolate arylation<br />

in the synthesis of benzofurans (see Scheme 49). 106 Similarly,<br />

the same group showed that thio ketones, derived<br />

from the same a-(o-haloaryl) ketone substrates using<br />

phosphorus pentasulfide, could also undergo this enolization–cyclization,<br />

again using a DPEPhos catalyst system<br />

to afford benzothiophenes. 106 Scheme 59 shows an aryl<br />

bromide example, although the corresponding aryl chloride<br />

also underwent cyclization, albeit in a reduced 44%<br />

yield.<br />

Br S<br />

Scheme 59<br />

Pd2(dba)3 (2.5 mol%)<br />

DPEPhos (22) (6 mol%)<br />

Cs2CO3 S<br />

toluene, 100 °C 74%<br />

In 2009 Lautens <strong>and</strong> co-workers extended the use of gemdihalovinylanilines<br />

in indole synthesis (see Scheme 5) to<br />

establish similar thiophenols as precursors for benzothiophenes.<br />

124 The combination of a palladium-catalyzed<br />

carbon–sulfur bond-forming reaction with a second<br />

cross-coupling process, such as a Suzuki–Miyaura, Heck<br />

or Sonogashira reaction, yielded diversely functionalized<br />

benzothiophenes. For example, combination of thiophenol<br />

116 with thiophene-3-boronic acid using an SPhos-derived<br />

catalyst delivered benzothiophene 117 in 99% yield<br />

(Scheme 60). The majority of examples reported involved<br />

Suzuki chemistry; a broad range of boronic acids, as well<br />

as other boron reagents, were readily included <strong>and</strong> allowed<br />

the introduction of aryl, alkenyl <strong>and</strong> alkyl C2 substituents.<br />

Application of the methodology to a variety of<br />

thiophenol backbones afforded the required heterocycles<br />

in mostly excellent yields.<br />

Synthesis 2011, No. 1, 1–22 © Thieme Stuttgart · New York<br />

S<br />

Scheme 60<br />

4.2 Benzothiazoles<br />

Two research groups have established benzothiazole syntheses<br />

based on a key catalytic intermolecular carbon–sulfur<br />

bond-forming step. In the first approach, Itoh <strong>and</strong><br />

Mase utilized a palladium-catalyzed thioetherification of<br />

o-bromoanilides using a thiol surrogate coupling partner.<br />

125 For example, reaction between aryl bromide 118<br />

<strong>and</strong> thiol 119, an odorless thiol surrogate, using a<br />

XantPhos-derived catalyst, ultimately delivered benzothiophene<br />

120 in 75% yield (Scheme 61). The reaction<br />

proceeded via intermediate sulfide 121, which was<br />

cleaved under basic conditions <strong>and</strong> then cyclized in the<br />

presence of acid to generate the aromatic product. p-<br />

Methoxybenzylthiol could also be employed as the thiol<br />

surrogate, allowing sulfide cleavage under acid conditions.<br />

Scheme 61<br />

Br<br />

Br<br />

SH<br />

116<br />

+<br />

B(OH) 2<br />

PdCl2 (3 mol%)<br />

SPhos (14) (3 mol%)<br />

K 3PO4, Et3N<br />

dioxane, 110 °C<br />

S<br />

117, 99%<br />

118<br />

121<br />

F<br />

+<br />

H<br />

N<br />

O<br />

Br<br />

Me<br />

Pd2(dba)3 (5 mol%)<br />

XantPhos (2) (10 mol%)<br />

i-Pr2NEt, dioxane<br />

reflux<br />

F<br />

H<br />

N<br />

S<br />

O<br />

Me<br />

O<br />

SH O<br />

OR<br />

O<br />

119<br />

Me<br />

Me<br />

NaOEt, THF, r.t.<br />

then TFA, reflux<br />

S<br />

120, 75%<br />

Rather than use a thiol surrogate, Ma <strong>and</strong> co-workers exploited<br />

metal sulfides in copper-catalyzed couplings with<br />

o-haloanilides to generate benzothiazoles. 126 They were<br />

able to show that sodium sulfide nonahydrate could be<br />

coupled with o-iodoanilides, <strong>and</strong> following acidic workup,<br />

deliver the desired benzothiazole products. For o-bromo<br />

substrates, the use of potassium sulfide was optimal.<br />

Both systems utilized a lig<strong>and</strong>less copper(I) iodide catalyst;<br />

significant variation of the substrate substituents was<br />

possible, delivering benzothiazoles in good to excellent<br />

yields (Scheme 62).<br />

The use of an intramolecular carbon–sulfur bond-forming<br />

reaction has proved more popular in the synthesis of benzothiazoles.<br />

In 1982, Bowman, Heaney <strong>and</strong> Smith reported<br />

an intramolecular, copper-catalyzed S-arylation in the<br />

synthesis of 2-alkyl- <strong>and</strong> 2-aryl-1,3-benzothiazoles from<br />

F<br />

N<br />

S<br />

Me


REVIEW <strong>Palladium</strong>- <strong>and</strong> <strong>Copper</strong>-<strong>Catalyzed</strong> Heterocycle Synthesis 19<br />

F<br />

H<br />

N<br />

I<br />

O<br />

Me<br />

Na2S<br />

CuI (10 mol%)<br />

DMF, 80 °C<br />

then HCl, r.t. F<br />

N<br />

S<br />

75%<br />

.9H2O<br />

+<br />

Scheme 62<br />

o-halothioacetanilides <strong>and</strong> o-halothiobenzanilides, respectively.<br />

127 Castillón <strong>and</strong> co-workers extended this<br />

methodology to a more efficient palladium-catalyzed cyclization<br />

of o-bromothioamides (Scheme 63). 128 It was<br />

also shown that cyclization of o-bromothioureas, prepared<br />

from o-bromophenylisothiocyanates <strong>and</strong> amines, afforded<br />

2-aminobenzothiazoles under similar reaction conditions<br />

(122 → 123).<br />

H<br />

N<br />

S<br />

Br<br />

122<br />

Scheme 63<br />

Batey <strong>and</strong> co-workers reported a comparison of the copper-<br />

<strong>and</strong> palladium-catalyzed syntheses of 2-aminobenzothiazoles<br />

based on this same cyclization of obromothioureas.<br />

129 The use of copper catalysis generally<br />

led to higher yields <strong>and</strong> conversions. Both metals were<br />

also shown to catalyze an example of a one-pot thiourea<br />

formation–cyclization reaction in the same excellent<br />

yield. Batey’s research group also demonstrated that cyclization<br />

of o-bromothioamides under the same copper(I)<br />

iodide/1,10-phenanthroline catalyst system afforded benzothiazoles<br />

in excellent yields; 115 a representative example<br />

is shown in Scheme 64. Jiang <strong>and</strong> Ma reported a<br />

related copper-catalyzed cyclization employing oxazolidin-2-one<br />

as the lig<strong>and</strong>; a number of aryl chloride substrates<br />

were included in their study, <strong>and</strong> effectively<br />

converted into benzothiazoles. 117 Pan <strong>and</strong> co-workers reported<br />

a related preparation of 2-aminobenzothiazoles using<br />

a copper(I) iodide/oxazoline catalyst system. 130<br />

Scheme 64<br />

H<br />

N Ph<br />

+<br />

O<br />

Br<br />

Me Me<br />

Me<br />

H<br />

N NMe 2<br />

S<br />

Br<br />

H<br />

N<br />

S<br />

Br<br />

K 2S<br />

CuI (10 mol%)<br />

DMF, 140 °C<br />

then HCl, r.t.<br />

Pd2(dba)3 (5 mol%)<br />

JohnPhos (106)<br />

(5.5 mol%)<br />

Cs2CO 3<br />

dioxane, 80 °C<br />

Pd 2(dba) 3 (5 mol%)<br />

P(t-Bu)3 (5.5 mol%)<br />

OMe<br />

Cs2CO 3<br />

dioxane, 80 °C<br />

CuI (5 mol%)<br />

1,10-phenanthroline<br />

(10 mol%)<br />

Cs 2CO3, reflux<br />

88%<br />

N<br />

S<br />

Me<br />

Ph<br />

N Me<br />

Me<br />

S Me<br />

88%<br />

N<br />

S<br />

123, 92%<br />

N<br />

S<br />

93%<br />

NMe 2<br />

OMe<br />

The Wu research group described the synthesis of 2-aminobenzothiazoles<br />

by the reaction of o-iodobenzamines<br />

with isothiocyanates using a copper(I) iodide/phenanthroline<br />

catalyst system (Scheme 65). 131 The method was useful<br />

as it eliminated the need to generate an ohalobenzothiourea<br />

cyclization precursor in a separate<br />

step, with the addition/carbon–sulfur coupling reaction<br />

occurring in one pot. The authors exploited the method in<br />

the preparation of an 18-membered library.<br />

F<br />

Scheme 65<br />

A number of benzothiazole syntheses that involve t<strong>and</strong>em<br />

processes have also appeared in the literature. Vera <strong>and</strong><br />

Pelletier utilized t<strong>and</strong>em palladium-catalyzed carbon–<br />

sulfur <strong>and</strong> carbon–nitrogen arylation reactions to prepare<br />

a series of aminobenzothiazoles. 132 For example, the<br />

combination of dibromothiobenzamide 124 <strong>and</strong> isopropylamine,<br />

using a JohnPhos catalyst, delivered 4-aminobenzothiazole<br />

125 in 47% yield (Scheme 66). As can be seen<br />

from the remainder of the examples in Scheme 66, it was<br />

possible to alter the position of the second bromine substituent,<br />

to generate 5-, 6-, <strong>and</strong> 7-amino-substituted products.<br />

Scheme 66<br />

NH 2<br />

I<br />

+<br />

SCN<br />

CuI (10 mol%)<br />

1,10-phenanthroline<br />

(20 mol%)<br />

DABCO<br />

toluene, 50 °C<br />

99%<br />

Br NHi-Pr<br />

124<br />

i-PrHN<br />

H<br />

N Ph<br />

+<br />

H2N(i-Pr)<br />

S<br />

Br<br />

16%<br />

N<br />

S<br />

Ph<br />

Pd 2(dba) 3 (10 mol%)<br />

JohnPhos (106) (20 mol%)<br />

i-PrHN<br />

Patel <strong>and</strong> co-workers showed that 2-arylthiobenzothiazoles<br />

can be accessed from cascade intra- <strong>and</strong> intermolecular<br />

carbon–sulfur bond-forming reactions using a single<br />

catalytic system. 133 A combination of o-iodo- or o-bromodithiocarbamates<br />

<strong>and</strong> iodoarenes were subjected to a<br />

copper(I) iodide/diamine catalyst system yielding the substituted<br />

benzothiazoles in mostly excellent yields<br />

(Scheme 67). The methodology was successfully applied<br />

to the synthesis of a cathespin-D inhibitor analogue.<br />

The same research group developed a cascade protocol for<br />

the synthesis of 2-thio- or 2-oxa-benzothiazoles by the<br />

copper-catalyzed reaction of o-iodo- or o-bromoarylisothiocyanates<br />

with a sulfur or oxygen nucleophile, respectively.<br />

134 The required dithiocarbamates or<br />

Synthesis 2011, No. 1, 1–22 © Thieme Stuttgart · New York<br />

F<br />

Cs 2CO 3, NaOt-Bu<br />

toluene-dioxane<br />

80 °C<br />

50%<br />

N<br />

S<br />

Ph<br />

N<br />

S<br />

NH<br />

N<br />

S<br />

125, 47%<br />

N<br />

S<br />

NHi-Pr 39%<br />

Ph<br />

Ph


20 J. E. R. Sadig, M. C. Willis REVIEW<br />

Me<br />

I<br />

Scheme 67<br />

thiocarbamates, which were formed in situ under basic<br />

conditions, readily underwent copper-catalyzed intramolecular<br />

carbon–sulfur bond formation. Thiophenols <strong>and</strong><br />

phenols showed the greatest reactivity; however, the corresponding<br />

alkyl series could also be employed, albeit to<br />

generate the products in reduced yields (Scheme 68).<br />

73%<br />

+<br />

Scheme 68<br />

4.3 Oxathioles<br />

Bao <strong>and</strong> co-workers reported a novel one-pot synthesis of<br />

2-iminobenzo-1,3-oxathioles using a cascade addition/<br />

intramolecular carbon–sulfur coupling process. 135 o-Iodophenols<br />

<strong>and</strong> isothiocyanates were combined using a<br />

copper(I) iodide/phenanthroline catalyst system to afford<br />

the desired heterocycles in good to excellent yields. A<br />

representative example is shown in Scheme 69.<br />

Scheme 69<br />

H<br />

N<br />

5 Conclusion<br />

I<br />

S<br />

NCS<br />

+<br />

Br<br />

HS Ph<br />

N<br />

S<br />

OPh<br />

S HNEt3<br />

CuI (5 mol%)<br />

(10 mol%)<br />

126<br />

OMe<br />

K 2CO 3<br />

DMSO, 90 °C<br />

126<br />

NH2<br />

NH2<br />

CuI (5 mol%)<br />

1,10-phenanthroline<br />

(10 mol%)<br />

K2CO 3<br />

dioxane, 90 °C<br />

N<br />

SBn<br />

S<br />

69%<br />

(iodo substrate)<br />

By definition, palladium- <strong>and</strong> copper-catalyzed aryl amination,<br />

aryl etherification <strong>and</strong> aryl thioetherification reactions<br />

are transformations designed to fashion bonds<br />

between heteroatoms <strong>and</strong> aromatic rings. It is perhaps not<br />

surprising that these reactions have enjoyed considerable<br />

success when applied to the synthesis of aromatic heterocycles.<br />

The examples presented above show how these re-<br />

Synthesis 2011, No. 1, 1–22 © Thieme Stuttgart · New York<br />

Me<br />

OH<br />

CuI (10 mol%)<br />

1,10-phenanthroline<br />

O<br />

SCN<br />

+<br />

I<br />

OMe<br />

(20 mol%)<br />

Cs2CO3<br />

toluene, 70–90 °C<br />

S<br />

86%<br />

Cl<br />

72%<br />

N<br />

S<br />

Ar = 4-MeOC 6H 4<br />

N<br />

S<br />

76%<br />

N<br />

S<br />

54%<br />

N<br />

O<br />

SPh<br />

SAr<br />

Ph<br />

OMe<br />

actions have been exploited towards a wide range of<br />

heterocyclic targets. They also show these reactions being<br />

used to provide new entries to existing, classic synthetic<br />

routes, as well as in the formulation of completely new<br />

disconnections. As advances in the underpinning transformations<br />

continue to develop – new coupling partners,<br />

more active catalysts <strong>and</strong> milder reaction conditions – the<br />

number of applications will undoubtedly continue to<br />

grow. The importance of heteroaromatic molecules virtually<br />

assures it.<br />

References<br />

(1) <strong>Palladium</strong> in Heterocyclic Chemistry, 1st ed.; Li, J. J.;<br />

Gribble, G. W., Eds.; Elsevier: Oxford, 2000.<br />

(2) Nakamura, I.; Yamamoto, Y. Chem. Rev. 2004, 104, 2127.<br />

(3) Zeni, G.; Larock, R. C. Chem. Rev. 2004, 104, 2285.<br />

(4) Cacchi, S.; Fabrizi, G. Chem. Rev. 2005, 105, 2873.<br />

(5) Zeni, G.; Larock, R. C. Chem. Rev. 2006, 106, 4644.<br />

(6) <strong>Palladium</strong> in Heterocyclic Chemistry, 2nd ed.; Li, J. J.;<br />

Gribble, G. W., Eds.; Elsevier: Oxford, 2007.<br />

(7) Patil, N. T.; Yamamoto, Y. Chem. Rev. 2008, 108, 3395.<br />

(8) Willis, M. C. Tetrahedron 2009, 65, 8907.<br />

(9) Kosugi, M.; Kameyama, M.; Migita, T. Chem. Lett. 1983,<br />

927.<br />

(10) (a) Guram, A. S.; Buchwald, S. L. J. Am. Chem. Soc. 1994,<br />

116, 7901. (b) Guram, A. S.; Rennels, R. A.; Buchwald, S.<br />

L. Angew. Chem. Int. Ed. 1995, 34, 1348. (c) Wolfe, J. P.;<br />

Buchwald, S. L. J. Org. Chem. 1996, 61, 1133. (d) Wolfe,<br />

J. P.; Wagaw, S.; Buchwald, S. L. J. Am. Chem. Soc. 1996,<br />

118, 7215.<br />

(11) (a) Paul, F.; Patt, J.; Hartwig, J. F. J. Am. Chem. Soc. 1994,<br />

116, 5969. (b) Louie, J.; Hartwig, J. F. Tetrahedron Lett.<br />

1995, 36, 3609. (c) Hartwig, J. F.; Richards, S.; Barañano,<br />

D.; Paul, F. J. Am. Chem. Soc. 1996, 118, 3626. (d) Driver,<br />

M. S.; Hartwig, J. F. J. Am. Chem. Soc. 1996, 118, 7217.<br />

(12) Ullmann, F. Ber. Dtsch. Chem. Ges. 1903, 36, 2382.<br />

(13) Goldberg, I. Ber. Dtsch. Chem. Ges. 1906, 39, 1691.<br />

(14) Hartwig, J. F. In H<strong>and</strong>book of Organopalladium Chemistry<br />

for Organic Synthesis, Vol. 1; Negishi, E. I., Ed.; Wiley-<br />

Interscience: New York, 2002, 1051–1096.<br />

(15) Jiang, L.; Buchwald, S. L. In Metal-<strong>Catalyzed</strong> Cross-<br />

Coupling Reactions, 2nd ed.; de Meijere, A.; Diederich, F.,<br />

Eds.; Wiley-VCH: Weinheim, 2004.<br />

(16) Surry, D. S.; Buchwald, S. L. Angew. Chem. Int. Ed. 2008,<br />

47, 6338.<br />

(17) Hartwig, J. F. Acc. Chem. Res. 2008, 41, 1534.<br />

(18) Ley, S. V.; Thomas, A. W. Angew. Chem. Int. Ed. 2003, 42,<br />

5400; Corrigendum: Angew. Chem. Int. Ed. 2004, 43, 1043.<br />

(19) Beletskaya, I. P.; Cheprakov, A. V. Coord. Chem. Rev. 2004,<br />

248, 2337.<br />

(20) Evano, G.; Blanchard, N.; Toumi, M. Chem. Rev. 2008, 108,<br />

3054.<br />

(21) Ma, D.; Cai, Q. Acc. Chem. Res. 2008, 41, 1450.<br />

(22) Monnier, F.; Taillefer, M. Angew. Chem. Int. Ed. 2009, 48,<br />

6954.<br />

(23) Surry, D. S.; Buchwald, S. L. Chem. Sci. 2010, 1, 13.<br />

(24) Although the majority of examples described in this<br />

approach do not lead directly to an aromatic heterocycle, but<br />

to an intermediate, it is included because of its historical<br />

relevance to the development of palladium-catalyzed aryl<br />

halide amination chemistry in heterocycle synthesis.<br />

(25) (a) Wagaw, A.; Yang, B. H.; Buchwald, S. L. J. Am. Chem.<br />

Soc. 1998, 120, 6621. (b) Wagaw, A.; Yang, B. H.;<br />

Buchwald, S. L. J. Am. Chem. Soc. 1999, 121, 10251.


REVIEW <strong>Palladium</strong>- <strong>and</strong> <strong>Copper</strong>-<strong>Catalyzed</strong> Heterocycle Synthesis 21<br />

(26) Lim, Y.-K.; Cho, C.-G. Tetrahedron Lett. 2004, 45, 1857.<br />

(27) Watanabe, M.; Yamamoto, T.; Nishiyama, M. Angew.<br />

Chem. Int. Ed. 2000, 39, 2501.<br />

(28) Siebeneicher, H.; Bytschkov, I.; Doye, S. Angew. Chem. Int.<br />

Ed. 2003, 42, 3042.<br />

(29) Brown, J. A. Tetrahedron Lett. 2000, 41, 1623.<br />

(30) (a) Yamazaki, K.; Nakamura, Y.; Kondo, Y. J. Chem. Soc.,<br />

Perkin Trans. 1 2002, 2137. (b) Yamazaki, K.; Nakamura,<br />

Y.; Kondo, Y. J. Org. Chem. 2003, 68, 6011.<br />

(31) (a) Fang, Y.-Q.; Lautens, M. Org. Lett. 2005, 7, 3549.<br />

(b) Fang, Y.-Q.; Lautens, M. J. Org. Chem. 2008, 73, 538.<br />

(c) Fang, Y.-Q.; Karisch, R.; Lautens, M. J. Org. Chem.<br />

2007, 72, 1341. (d) Fang, Y.-Q.; Yuen, J.; Lautens, M.<br />

J. Org. Chem. 2007, 72, 5152. (e) Fayol, A.; Fang, Y.-Q.;<br />

Lautens, M. Org. Lett. 2006, 8, 4203. (f) Nagamochi, M.;<br />

Fang, Y.-Q.; Lautens, M. Org. Lett. 2007, 9, 2955.<br />

(g) Yuen, J.; Fang, Y.-Q.; Lautens, M. Org. Lett. 2006, 8,<br />

653. (h) Bryan, C. S.; Lautens, M. Org. Lett. 2008, 10, 4633.<br />

(i) Newman, S. G.; Aureggi, V.; Bryan, C. S.; Lautens, M.<br />

Chem. Commun. 2009, 5236.<br />

(32) (a) Willis, M. C.; Brace, G. N. Tetrahedron Lett. 2002, 43,<br />

9085. (b) Barluenga, J.; Fernández, M. A.; Aznar, F.;<br />

Valdés, C. Chem. Commun. 2002, 2362. (c) Wallace, D. J.;<br />

Klauber, D. J.; Chen, C.-y.; Volante, R. P. Org. Lett. 2003,<br />

5, 4759. (d) Barluenga, J.; Fernández, M. A.; Aznar, F.;<br />

Valdés, C. Chem. Eur. J. 2004, 10, 494. (e) Klapers, A.;<br />

Campos, K. R.; Chen, C.-y.; Volante, R. P. Org. Lett. 2005,<br />

7, 1185. (f) Willis, M. C.; Chauhan, J.; Whittingham, W. G.<br />

Org. Biomol. Chem. 2005, 3094. (g) Willis, M. C.; Brace,<br />

G. N.; Holmes, I. P. Synthesis 2005, 3229.<br />

(33) Thielges, S.; Meddah, E.; Bisseret, P.; Eustache, J.<br />

Tetrahedron Lett. 2004, 45, 907.<br />

(34) (a) Vieira, T. O.; Meaney, L. A.; Shi, Y.-L.; Alper, H. Org.<br />

Lett. 2008, 10, 4899. (b) Arthuis, M.; Pontikis, R.; Florent,<br />

J.-C. Org. Lett. 2009, 11, 4608.<br />

(35) (a) Willis, M. C.; Brace, G. N.; Holmes, I. P. Angew. Chem.<br />

Int. Ed. 2005, 44, 403. (b) Willis, M. C.; Brace, G. N.;<br />

Findlay, T. J. K.; Holmes, I. P. Adv. Synth. Catal. 2006, 348,<br />

851. (c) Fletcher, A. J.; Bax, M. N.; Willis, M. C. Chem.<br />

Commun. 2007, 4764. (d) Henderson, L. C.; Lindon, M. J.;<br />

Willis, M. C. Tetrahedron 2010, 66, 6632. (e) Hodgkinson,<br />

R. C.; Schulz, J.; Willis, M. C. Tetrahedron 2009, 65, 8940.<br />

(f) Hodgkinson, R. C.; Schulz, J.; Willis, M. C. Org. Biomol.<br />

Chem. 2009, 7, 432.<br />

(36) Dong, S.-X.; Zhang, X.-G.; Liu, Q.; Tang, R.-Y.; Zhong, P.;<br />

Li, J.-H. Synthesis 2010, 1521.<br />

(37) Smith, A. B. III.; Kürti, L.; Davulcu, A. H. Org. Lett. 2006,<br />

8, 2167.<br />

(38) (a) Ackermann, L. Org. Lett. 2005, 7, 439. (b) Ackermann,<br />

L.; Barfüsser, S.; Potukuchi, H. K. Adv. Synth. Catal. 2009,<br />

351, 1064. (c) Ackermann, L.; S<strong>and</strong>mann, R.; Kondrashov,<br />

M. V. Synlett 2009, 1219. (d) Ackermann, L.; S<strong>and</strong>mann,<br />

R.; Schinkel, M.; Kondrashov, M. V. Tetrahedron 2009, 65,<br />

8930. (e) Kaspar, L. T.; Ackermann, L. Tetrahedron 2005,<br />

61, 11311.<br />

(39) Tang, Z.-Y.; Hu, Q.-S. Adv. Synth. Catal. 2006, 348, 846.<br />

(40) Sanz, R.; Castroviejo, M. P.; Guilarte, V.; Pérez, A.;<br />

Fañanás, F. J. J. Org. Chem. 2007, 72, 5113.<br />

(41) Yao, P.-Y.; Zhang, Y.; Hsung, R. P.; Zhao, K. Org. Lett.<br />

2008, 10, 4275.<br />

(42) (a) Barluenga, J.; Jiménez-Aquino, A.; Valdés, C.; Aznar, F.<br />

Angew. Chem. Int. Ed. 2007, 46, 1529. (b) Barluenga, J.;<br />

Jiménez-Aquino, A.; Aznar, F.; Valdés, C. J. Am. Chem.<br />

Soc. 2009, 131, 4031.<br />

(43) (a) Nozaki, K.; Takahashi, K.; Nakano, K.; Hiyama, T.;<br />

Tang, H.-Z.; Fujiki, M.; Yamaguchi, S.; Tamao, K. Angew.<br />

Chem. Int. Ed. 2003, 42, 2051. (b) Kuwahara, A.; Nakano,<br />

K.; Nozaki, K. J. Org. Chem. 2005, 70, 413. (c) Nakano,<br />

K.; Hidehira, Y.; Takahashi, K.; Hiyama, T.; Nozaki, K.<br />

Angew. Chem. Int. Ed. 2005, 44, 7136. (d) Kawaguchi, K.;<br />

Nakano, K.; Nozaki, K. J. Org. Chem. 2007, 72, 5119.<br />

(e) Kawaguchi, K.; Nakano, K.; Nozaki, K. Org. Lett. 2008,<br />

10, 1199.<br />

(44) (a) Kitawaki, T.; Hayashi, Y.; Ueno, A.; Chida, N.<br />

Tetrahedron 2006, 62, 6792. (b) Ueno, A.; Kitawaki, T.;<br />

Chida, N. Org. Lett. 2008, 10, 1999.<br />

(45) Li, E.; Xu, X.; Li, H.; Zhang, H.; Xu, X.; Yuan, X.; Li, Y.<br />

Tetrahedron 2009, 65, 8961.<br />

(46) Koeckelberghs, G.; Cremer, L. D.; Vanormelingen, W.;<br />

Dehaen, W.; Verbiest, T.; Persoons, A.; Samyn, C.<br />

Tetrahedron 2005, 61, 687.<br />

(47) Odom, S. A.; Lancaster, K.; Bevrina, L.; Lefler, K. M.;<br />

Thompson, N. J.; Coropceanu, V.; Brédas, J.-L.; Marder, S.<br />

R.; Barlow, S. Chem. Eur. J. 2007, 13, 9637.<br />

(48) (a) Bedford, R. B.; Cazin, C. S. J. Chem. Commun. 2002,<br />

2310. (b) Bedford, R. B.; Betham, M. J. Org. Chem. 2006,<br />

71, 9403. (c) Bedford, R. B.; Betham, M.; Charmant, J. P.<br />

H.; Weeks, A. L. Tetrahedron 2008, 64, 6038.<br />

(49) (a) Ackermann, L.; Althammer, A. Angew. Chem. Int. Ed.<br />

2007, 46, 1627. (b) Ackermann, L.; Althammer, A.; Mayer,<br />

P. Synthesis 2009, 3493.<br />

(50) Kitamura, Y.; Yoshikawa, S.; Furuta, T.; Kan, T. Synlett<br />

2008, 377.<br />

(51) (a) Watanabe, T.; Ueda, A.; Inuki, S.; Oishi, S.; Fujii, N.;<br />

Ohno, H. Chem. Commun. 2007, 4516. (b) Watanabe, T.;<br />

Oishi, S.; Fujii, N.; Ohno, H. J. Org. Chem. 2009, 74, 4720.<br />

(52) (a) Zou, B.; Yuan, Q.; Ma, D. Angew. Chem. Int. Ed. 2007,<br />

46, 2598. (b) Zou, B.; Yuan, Q.; Ma, D. Org. Lett. 2007, 9,<br />

4291. (c) Diao, X.; Wang, Y.; Jiang, Y.; Ma, D. J. Org.<br />

Chem. 2009, 74, 7974.<br />

(53) (a) Zheng, N.; Anderson, K. W.; Xiaohua, X.; Nguyen, H.<br />

N.; Buchwald, S. L. Angew. Chem. Int. Ed. 2007, 46, 7509.<br />

(b) Zheng, N.; Buchwald, S. L. Org. Lett. 2007, 9, 4749.<br />

(54) Scott, J. P. Synlett 2006, 2083.<br />

(55) McLaughlin, M.; Palucki, M.; Davies, I. W. Org. Lett. 2006,<br />

8, 3311.<br />

(56) Li, Z.; Sun, H.; Jiang, H.; Liu, H. Org. Lett. 2008, 10, 3263.<br />

(57) Xu, X.-J.; Zong, Y.-X. Tetrahedron Lett. 2007, 48, 129.<br />

(58) Barbero, N.; Carril, M.; SanMartin, R.; Domínguez, E.<br />

Tetrahedron 2008, 64, 7283.<br />

(59) (a) Brain, C. T.; Brunton, S. A. Tetrahedron Lett. 2002, 43,<br />

1893. (b) Brain, C. T.; Steer, J. T. J. Org. Chem. 2003, 68,<br />

6814.<br />

(60) (a) Ventatesh, C.; Sundaram, G. S. M.; Ila, H.; Junjappa, H.<br />

J. Org. Chem. 2006, 71, 1280. (b) Kumar, S.; Ila, H.;<br />

Junjappa, H. J. Org. Chem. 2009, 74, 7046.<br />

(61) Alen, J.; Robeyns, K.; De Borggraeve, W. M.;<br />

Van Meervelt, L.; Compernolle, F. Tetrahedron 2008, 64,<br />

8128.<br />

(62) Lygin, A. V.; De Meijere, A. Eur. J. Org. Chem. 2009, 5138.<br />

(63) Evindar, G.; Batey, R. A. Org. Lett. 2003, 5, 133.<br />

(64) Szczepankiewicz, B. G.; Rohde, J. J.; Kurukulasuriya, R.<br />

Org. Lett. 2005, 7, 1833.<br />

(65) Murru, S.; Patel, B. K.; Le Bras, J.; Muzart, J. J. Org. Chem.<br />

2009, 74, 2217.<br />

(66) Chen, M.-W.; Zhang, X.-G.; Zhong, P.; Hu, M.-L. Synthesis<br />

2009, 1431.<br />

(67) Zhu, J.; Xie, H.; Chen, Z.; Li, S.; Wu, Y. Chem. Commun.<br />

2009, 2338.<br />

(68) (a) Loones, K. T. J.; Maes, B. U. W.; Dommisse, R. A.;<br />

Lemiére, G. L. F. Chem. Commun. 2004, 2466. (b) Loones,<br />

K. T. J.; Maes, B. U. W.; Meyers, C.; Deruytter, J. J. Org.<br />

Chem. 2006, 71, 260. (c) Loones, K. T. J.; Maes, B. U. W.;<br />

Herrebout, W. A.; Dommisse, R. A.; Lemiére, G. L. F.;<br />

Synthesis 2011, No. 1, 1–22 © Thieme Stuttgart · New York


22 J. E. R. Sadig, M. C. Willis REVIEW<br />

Van der Veken, B. Tetrahedron 2007, 63, 3818.<br />

(d) Loones, K. T. J.; Maes, B. U. W.; Dommisse, R. A.<br />

Tetrahedron 2007, 63, 8954.<br />

(69) Bogányi, B.; Kámán, J. J. Heterocycl. Chem. 2009, 46, 33.<br />

(70) Viirre, R. D.; Evindar, G.; Batey, R. A. J. Org. Chem. 2008,<br />

73, 3452.<br />

(71) Deng, X.; McAllister, H.; Mani, N. J. Org. Chem. 2009, 74,<br />

5742.<br />

(72) Lebedev, A. Y.; Khartulyari, A. S.; Voskoboynikov, A. Z.<br />

J. Org. Chem. 2005, 70, 596.<br />

(73) (a) Inamoto, K.; Katsuno, M.; Yoshino, T.; Suzuki, I.;<br />

Hiroya, K.; Sakamoto, T. Chem. Lett. 2004, 33, 1026.<br />

(b) Inamoto, K.; Katsuno, M.; Yoshino, T.; Arai, Y.; Hiroya,<br />

K.; Sakamoto, T. Tetrahedron 2007, 63, 2695.<br />

(74) Suryakiran, N.; Prabhakar, P.; Venkateswarlu, Y. Chem.<br />

Lett. 2007, 36, 1370.<br />

(75) Song, J. J.; Yee, N. K. Tetrahedron Lett. 2001, 42, 2937.<br />

(76) Cho, C. S.; Lim, D. K.; Heo, N. H.; Kim, T.-J.; Shim, S. C.<br />

Chem. Commun. 2004, 104.<br />

(77) Pabba, C.; Wang, H.-J.; Mulligan, S. R.; Chen, Z.-J.; Stark,<br />

T. M.; Gregg, B. T. Tetrahedron Lett. 2005, 46, 7553.<br />

(78) Viña, D.; del Olmo, E.; López-Pérez, J. L.; San Feliciano, A.<br />

Org. Lett. 2007, 9, 525.<br />

(79) Levecchia, G.; Berteina-Raboin, S.; Guillaumet, G.<br />

Tetrahedron Lett. 2004, 45, 2389.<br />

(80) Song, J. J.; Yee, N. K. Org. Lett. 2000, 2, 519.<br />

(81) Zhu, Y.-m.; Kiryu, Y.; Katayama, H. Tetrahedron Lett.<br />

2002, 43, 3577.<br />

(82) Hall<strong>and</strong>, N.; Nazaré, M.; R’kyek, O.; Alonso, J.; Urmann,<br />

M.; Lindenschmidt, A. Angew. Chem. Int. Ed. 2009, 48,<br />

6879.<br />

(83) Tanimori, S.; Ozaki, Y.; Iesaki, Y.; Kirihata, M. Synlett<br />

2008, 1973.<br />

(84) Martin, R.; Larsen, C. H.; Cuenca, A.; Buchwald, S. L. Org.<br />

Lett. 2007, 9, 3379.<br />

(85) Yuan, X.; Xu, X.; Zhou, X.; Yuan, J.; Mai, L.; Li, Y. J. Org.<br />

Chem. 2007, 72, 1510.<br />

(86) Rodriguez Rivero, M.; Buchwald, S. L. Org. Lett. 2007, 9,<br />

973.<br />

(87) Martin, R.; Rodriguez Rivero, M.; Buchwald, S. L. Angew.<br />

Chem. Int. Ed. 2006, 45, 7079.<br />

(88) Cho, C. S.; Patel, D. B. Tetrahedron 2006, 62, 6388.<br />

(89) (a) Haddad, N.; Baron, J. Tetrahedron Lett. 2002, 43, 2171.<br />

(b) Haddad, N.; Salvagno, A.; Busacca, C. Tetrahedron Lett.<br />

2004, 45, 5935.<br />

(90) Martin, R.; Cuenca, A.; Buchwald, S. L. Org. Lett. 2007, 9,<br />

5521.<br />

(91) Manley, P. J.; Bilodeau, M. T. Org. Lett. 2004, 6, 2433.<br />

(92) Huang, J.; Chen, Y.; King, A. O.; Dilmeghani, M.; Larsen,<br />

R. D.; Faul, M. M. Org. Lett. 2008, 10, 2609.<br />

(93) Jones, C. P.; Anderson, K. W.; Buchwald, S. L. J. Org.<br />

Chem. 2007, 72, 7968.<br />

(94) Battistuzzi, G.; Bernini, R.; Cacchi, S.; De Salve, I.; Fabrizi,<br />

G. Adv. Synth. Catal. 2007, 349, 297.<br />

(95) Tadd, A. C.; Matsuno, A.; Fielding, M. R.; Willis, M. C.<br />

Org. Lett. 2009, 11, 583.<br />

(96) Willis, M. C.; Snell, R. H.; Fletcher, A. J.; Woodward, R. L.<br />

Org. Lett. 2006, 8, 5089.<br />

(97) (a) Liu, X.; Fu, H.; Jiang, Y.; Zhao, Y. Angew. Chem. Int. Ed.<br />

2009, 48, 348. (b) Huang, X.; Yang, H.; Fu, H.; Qiao, R.;<br />

Zhao, Y. Synthesis 2009, 2679.<br />

(98) Zhou, J.; Fu, L.; Lv, M.; Liu, J.; Pei, D.; Ding, K. Synthesis<br />

2008, 3974.<br />

(99) Ju, Y.; Liu, F.; Li, C. Org. Lett. 2009, 11, 3582.<br />

(100) Tietze, M.; Iglesias, A.; Merisor, E.; Conrad, J.; Klaiber, I.;<br />

Beifuss, U. Org. Lett. 2005, 7, 1549.<br />

Synthesis 2011, No. 1, 1–22 © Thieme Stuttgart · New York<br />

(101) Emoto, T.; Kubosaki, N.; Yamagiwa, Y.; Kamikawa, T.<br />

Tetrahedron Lett. 2000, 41, 355.<br />

(102) Hasegawa, K.; Kimura, N.; Arai, S.; Nishida, A. J. Org.<br />

Chem. 2008, 73, 6363.<br />

(103) Anderson, K. W.; Ikawa, T.; Tundel, R. E.; Buchwald, S. L.<br />

J. Am. Chem. Soc. 2006, 128, 10694.<br />

(104) Liao, Y.; Smith, J.; Fathi, R.; Yan, Z. Org. Lett. 2005, 7,<br />

2707.<br />

(105) Zhao, D.; Wu, N.; Zhang, S.; Xi, P.; Su, X.; Lan, J.; You, J.<br />

Angew. Chem. Int. Ed. 2009, 48, 8729.<br />

(106) (a) Willis, M. C.; Taylor, D.; Gillmore, A. T. Org. Lett.<br />

2004, 6, 4755. (b) Willis, M. C.; Taylor, D.; Gillmore, A. T.<br />

Tetrahedron 2006, 62, 11513.<br />

(107) Faragó, J.; Kotschy, A. Synthesis 2009, 85.<br />

(108) Chen, C.-y.; Dormer, P. G. J. Org. Chem. 2005, 70, 6964.<br />

(109) Carril, M.; SanMartin, R.; Tellitu, I.; Domínguez, E. Org.<br />

Lett. 2006, 8, 1467.<br />

(110) Ackermann, L.; Kaspar, L. T. J. Org. Chem. 2007, 72, 6149.<br />

(111) Tadd, A. C.; Fielding, M. R.; Willis, M. C. Tetrahedron Lett.<br />

2007, 48, 7578.<br />

(112) Terao, Y.; Satoh, T.; Miura, M.; Nomura, M. Bull. Chem.<br />

Soc. Jpn. 1999, 72, 2345.<br />

(113) Churruca, F.; SanMartin, R.; Tellitu, I.; Domínguez, E. Eur.<br />

J. Org. Chem. 2005, 2481.<br />

(114) Lu, B.; Wang, B.; Zhang, Y.; Ma, D. J. Org. Chem. 2007, 72,<br />

5337.<br />

(115) Evindar, G.; Batey, R. A. J. Org. Chem. 2006, 71, 1802.<br />

(116) Barbero, N.; Carril, M.; SanMartin, R.; Domínguez, E.<br />

Tetrahedron 2007, 63, 10425.<br />

(117) Ma, H. C.; Jiang, X. Z. Synlett 2008, 1335.<br />

(118) Kantam, M. L.; Venkanna, G. T.; Kumar, K. B. S.;<br />

Balasubramanyam, V.; Bhargava, S. Synlett 2009, 1753.<br />

(119) Xu, D.; Xu, X.; Liu, Z.; Sun, L.-P.; You, Q. Synlett 2009,<br />

1172.<br />

(120) Altenhoff, G.; Glorius, F. Adv. Synth. Catal. 2004, 346,<br />

1661.<br />

(121) Wang, L.; Shen, W. Tetrahedron Lett. 1998, 39, 7625.<br />

(122) Tadd, A. C.; Fielding, M. R.; Willis, M. C. Chem. Commun.<br />

2009, 6744.<br />

(123) (a) Fernández-Rodríguez, M. A.; Hartwig, J. F. J. Org.<br />

Chem. 2009, 74, 1663. (b) Eichman, C. C.; Stambuli, J. P.<br />

J. Org. Chem. 2009, 74, 4005. (c) Lee, J.-Y.; Lee, P. H.<br />

J. Org. Chem. 2008, 73, 7413. (d) Fernández-Rodríguez,<br />

M. A.; Hartwig, J. F. Chem.Eur. J. 2006, 12, 7782.<br />

(e) Murata, M.; Buchwald, S. L. Tetrahedron 2004, 60,<br />

7397.<br />

(124) Bryan, C. A.; Braunger, J. A.; Lautens, M. Angew. Chem.<br />

Int. Ed. 2009, 48, 7064.<br />

(125) Itoh, T.; Mase, T. Org. Lett. 2007, 9, 3687.<br />

(126) Ma, D.; Xie, S.; Xue, P.; Zhang, X.; Dong, J.; Jiang, Y.<br />

Angew. Chem. Int. Ed. 2009, 48, 4222.<br />

(127) Bowman, W. R.; Heaney, H.; Smith, P. H. G. Tetrahedron<br />

Lett. 1982, 23, 5093.<br />

(128) Bendi, C.; Bravo, F.; Uriz, P.; Fernández, E.; Claver, C.;<br />

Castillón, S. Tetrahedron Lett. 2003, 44, 6073.<br />

(129) Joyce, L. L.; Evindar, G.; Batey, R. A. Chem. Commun.<br />

2004, 446.<br />

(130) Wang, J.; Peng, F.; Jiang, J.-l.; Lu, Z.-j.; Wang, L.-y.; Bai,<br />

J.; Pan, Y. Tetrahedron Lett. 2008, 49, 467.<br />

(131) Ding, Q.; He, X.; Wu, J. J. Comb. Chem. 2009, 11, 587.<br />

(132) Vera, M. D.; Pelletier, J. C. J. Comb. Chem. 2007, 9, 569.<br />

(133) Murru, S.; Ghosh, H.; Sahoo, S. K.; Patel, B. K. Org. Lett.<br />

2009, 11, 4254.<br />

(134) Murru, S.; Mondal, P.; Yella, R.; Patel, B. K. Eur. J. Org.<br />

Chem. 2009, 5406.<br />

(135) Lv, X.; Liu, Y.; Qian, W.; Bao, W. Adv. Synth. Catal. 2008,<br />

350, 2507.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!