palladium-catalyzed cross-coupling reactions in total synthesis.pdf

(2901 KB) Pobierz
Reviews
K. C. Nicolaou et al.
Synthetic Methods
Palladium-Catalyzed Cross-Coupling Reactions in Total
Synthesis
K. C. Nicolaou,* Paul G. Bulger, and David Sarlah
Keywords:
C C coupling · cross-coupling ·
palladium catalysis · synthetic
methods · total synthesis
Dedicated to Richard F. Heck
on the occasion of his 74th birthday
Angewandte
Chemie
4442
DOI: 10.1002/anie.200500368
2005 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Angew. Chem. Int. Ed. 2005, 44, 4442–4489
1081964094.026.png 1081964094.027.png
 
Angewandte
Chemie
C C Coupling
I n studying the evolution of organic chemistry and grasping its
essence, one comes quickly to the conclusion that no other type of
reaction plays as large a role in shaping this domain of science
than carbon–carbon bond-forming reactions. The Grignard,
Diels–Alder, and Wittig reactions are but three prominent exam-
ples of such processes, and are among those which have undeni-
ably exercised decisive roles in the last century in the emergence of
chemical synthesis as we know it today. In the last quarter of the
20th century, a new family of carbon–carbon bond-forming
reactions based on transition-metal catalysts evolved as powerful
tools in synthesis. Among them, the palladium-catalyzed cross-
coupling reactions are the most prominent. In this Review, high-
lights of a number of selected syntheses are discussed. The
examples chosen demonstrate the enormous power of these
processes in the art of total synthesis and underscore their future
potential in chemical synthesis.
From the Contents
1. Introduction
4443
2. The Heck Reaction
4445
3. The Stille Reaction
4452
4. The Suzuki Reaction
4458
5. The Sonogashira Reaction
4468
6. The Tsuji–Trost Reaction
4473
7. The Negishi Reaction
4478
8. Summary and Outlook
4481
1. Introduction
formally result from the substitution of a hydrogen atom in
the alkene coupling partner. The first examples of this
reaction as we would recognize it today were reported
independently by Mizoroki (1971) [2] and, in an improved
form, byHeck (1972). [3] However, it wouldprove to bemore
than a decade before the broader applicability of this
transformation began to be investigated by the wider
synthetic organic community. The development of catalytic
asymmetric Heck reactions in the late 1980s led to a further
resurgence of interest in this field. [4] The Heck reaction now
stands as a remarkably robust and efficient method for
carbon–carbonbondformation,particularlyinthegeneration
of tertiary and quaternary stereocenters and intramolecular
ring formation, and remains a flourishing area of research.
Significantly, it inspired important variations that, with time,
have assumed their own names, identities, and place in total
synthesis.
The palladium-catalyzed cross-coupling of organic elec-
trophiles with vinyl organotin compounds is today known as
the Stille reaction (Scheme1), [5] after the late Professor J.K.
Stille who pioneered (1978) [6] and subsequently developed [7]
this reaction, although the seeds of discovery were sown
earlier by Kosugi and his group, who published the first
reports of transition-metal-catalyzed carbon-carbon bond-
forming reactions with organotin compounds a year ear-
lier. [8,9] Nearly30yearslatertheStillereactionremainsoneof
Ever since the first laboratory construction of a carbon–
carbonbondbyKolbeinhishistoricsynthesisofaceticacidin
1845, carbon–carbon bond-forming reactions have played an
enormously decisive and important role in shaping chemical
synthesis. Aldol- and Grignard-type reactions, the Diels–
Alder and related pericyclic processes, and the Wittig and
related reactions are but a few examples of such processes
that have advanced our ability to construct increasingly
complexcarbonframeworksand,thus,enabledthesyntheses
of a myriad of organic compounds. In the last quarter of the
20th century, a new paradigm for carbon–carbon bond
formation has emerged that has enhanced considerably the
prowess of synthetic organic chemists to assemble complex
molecular frameworks and has changed the way we think
about synthesis. Based on transition-metal catalysis, this
newlyacquiredabilitytoforgecarbon–carbonbondsbetween
or within functionalized and sensitive substrates provided
new opportunities, particularly in total synthesis but also in
medicinal and process chemistry as well as in chemical
biology and nanotechnology.
Prominent among these processes are the palladium-
catalyzed carbon–carbon bond-forming reactions. Because
the historical, mechanistic, theoretical, and practical aspects
of these processes have been amply discussed, [1] in this
Review we focus only on selected applications of the most
commonlyappliedpalladium-catalyzedcarbon–carbonbond-
forming reactions in total synthesis, namely, the Heck, Stille,
Suzuki, Sonogashira, Tsuji–Trost, and the Negishi reactions,
withparticularemphasisonthepioneeringaswellassomeof
themostrecentandexcitingexamples.Indoingso,wehopeto
illustrate the tremendous enabling ability of these modern
synthetic tools.
The Heck reaction can be broadly defined as the
palladium-catalyzed coupling of alkenyl or aryl (sp 2 ) halides
or triflates with alkenes (Scheme1) to yield products which
[*] Prof. Dr. K. C. Nicolaou, Dr. P. G. Bulger, D. Sarlah
Department of Chemistry
and The Skaggs Institute for Chemical Biology
The Scripps Research Institute
10550 North Torrey Pines Road, La Jolla, CA 92037 (USA)
Fax: ( + 1)858-784-2469
E-mail: kcn@scripps.edu
and
Department of Chemistry and Biochemistry
University of California San Diego
9500 Gilman Drive, La Jolla, CA 92093 (USA)
4443
DOI: 10.1002/anie.200500368
Angew. Chem. Int. Ed. 2005, 44, 4442 –4489
2005 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
1081964094.028.png 1081964094.001.png 1081964094.002.png 1081964094.003.png 1081964094.004.png 1081964094.005.png 1081964094.006.png 1081964094.007.png 1081964094.008.png 1081964094.009.png
 
Reviews
K. C. Nicolaou et al.
the most widely applied palladium-catalyzed carbon–carbon
bond-forming reactions, in large part due to typically mild
reactionconditions,theeaseofpreparationofawiderangeof
coupling partners, and the tolerance of a wide variety of
sensitive functionalities in this transformation. In particular,
the number of ingenious and daring applications of Stille
couplingsinthechallengingprovinggroundoftotalsynthesis
bears testament to the faith placed in the robustness and
versatilityofthisreactionbypractitionersofthisart.Notably,
the Stille reaction can be viewed as a variation of the Heck
reactioninwhichahydrogenatomisreplacedbyatin-bearing
substituent. [10]
Another extraordinarily useful palladium-catalyzed
carbon–carbon bond-forming reaction involves the palla-
dium-mediatedcouplingoforganicelectrophiles,suchasaryl
oralkenylhalidesandtriflates,withorganoboroncompounds
in the presence of a base (Scheme1), [11] a process known
today as the Suzuki reaction. The first examples of this
protocol were reported by the Suzuki group in 1979 [12]
K. C. Nicolaou was born in Cyprus and edu-
cated in the UK and USA. He is Chairman
of the Department of Chemistry at The
Scripps Research Institute where he holds
the Darlene Shiley Chair in Chemistry and
the Aline W. and L. S. Skaggs Professorship
in Chemical Biology. He is also Professor of
Chemistry at the University of California,
San Diego. His impact on chemistry, biology
and medicine flows from contributions to
chemical synthesis, which are described in
numerous publications and patents.
Paul G. Bulger was born in London (UK) in
1978. He received his M.Chem in 2000
from the University of Oxford, where he
completed his Part II project under Dr.
Mark G. Moloney. He obtained his D.Phil in
chemistry in 2003 for research conducted
under Professor Sir Jack E. Baldwin. In the
fall of 2003, he joined Professor K. C. Nico-
laou’s group as a postdoctoral researcher.
His research interests encompass reaction
mechanism and design and their application
to complex natural product synthesis and
chemical biology.
Scheme 1. The most commonly utilized palladium-catalyzed
cross-coupling reactions.
David Sarlah was born in Celje, Slovenia in
1983. He is currently student in the Faculty
of Chemistry and Chemical Technology, Uni-
versity of Ljubljana (Slovenia). Since 2001,
he has been a research assistant at the Lab-
oratory of Organic and Medicinal Chemistry
at the National Institute of Chemistry (Slov-
enia) where he carried out research on
asymmetric catalysis under the direction of
Dr. B. Mohar. During the summer of 2004,
he was engaged in total synthesis endeavors
as a member of the azaspiracid team under
Professor K. C. Nicolaou.
although, again, the inspiration or the seeds of this develop-
ment can be found in earlier work by others, in this case the
groups of Heck (1973) [13] and Negishi (1977). [14] The ensuing
quarter of a century saw remarkable developments in the
field. Amongst its manifold applications, the Suzuki reaction
is particularly useful as a method for the construction of
conjugateddienesandhigherpolyenesystemsofhighstereo-
isomeric purity, as well as of biaryl and related systems.
Furthermore, tremendous progress has been made in the
development of Suzuki coupling reactions of unactivated
4444
2005 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Angew. Chem. Int. Ed. 2005, 44, 4442–4489
1081964094.010.png 1081964094.011.png 1081964094.012.png 1081964094.013.png 1081964094.014.png 1081964094.015.png 1081964094.016.png 1081964094.017.png 1081964094.018.png 1081964094.019.png 1081964094.020.png 1081964094.021.png
Angewandte
Chemie
C C Coupling
alkyl halides, enabling C(sp 2 )–C(sp 3 ) and even C(sp 3 )–C(sp 3 )
bond-forming processes. [15,16] The ease of preparation of
organoboron compounds (e.g. aryl, vinyl, alkyl) and their
relativestabilitytoairandwater,combinedwiththerelatively
mild conditions for the reaction as well as the formation of
nontoxic by-products, makes the Suzuki reaction a valuable
addition to the armory of the synthetic organic chemist.
Indeed, it has become one of the most reliable and widely
applied palladium-catalyzed cross-coupling reactions in total
synthesis, where it has found a prominent role. [17] It is, again,
worthmentioningthattheSuzukireactionmaybeconsidered
as a variation of the Heck reaction, in which a boron-
containing group replaces a hydrogen atom in the olefinic
partner of the cross-coupling.
The palladium-catalyzed coupling of terminal alkynes
withvinylorarylhalideswasfirstreportedindependentlyand
simultaneously by the groups of Cassar [18] and Heck [19] in
1975. A few months later, Sonogashira and co-workers
demonstratedthat,inmanycases,thiscross-couplingreaction
couldbeacceleratedbytheadditionofcocatalyticCu I saltsto
the reaction mixture. [20,21] This protocol, which has become
knownastheSonogashirareaction,canbeviewedasbothan
alkyne version of the Heck reaction and an application of
palladium catalysis to the venerable Stephens–Castro reac-
tion (the coupling of vinyl or aryl halides with stoichiometric
amountsofcopper( i )acetylides). [22] TheSonogashirareaction
provides a valuable method for the synthesis of conjugated
acetylenic systems, which are used in a diverse array of
important applications from natural products and pharma-
ceuticals to designed molecules of interest in biotechnology
and nanotechnology. Interestingly, the utility of the “copper-
free” Sonogashira protocol (i.e. the original Cassar–Heck
version of this reaction) has subsequently been “rediscov-
ered” independently by a number of other researchers in
recent years. [23]
The palladium-catalyzed nucleophilic substitution of
allylic compounds, known as the Tsuji–Trost reaction
(Scheme1), is arguably one of the most synthetically useful
carbon–carbon bond-forming reactions to emerge in the last
quarterofthepreviouscentury. [24] Allylacetatesarebyfarthe
mostcommonlyemployedelectrophiles,andsoftanionssuch
as those derived from b-dicarbonyl compounds are most
routinelyusedasthenucleophiliccouplingpartner.However,
a wide variety of substrate combinations is possible, which
givesthereactionanexceptionallybroadscope.Thedevelop-
mentofthepalladium-catalyzedasymmetricallylicalkylation
reaction over the last decade has considerably increased the
enablingnatureofthistransformation. [25] Anotablefeatureof
the allylic alkylation reaction is its net reaction at sp 3 -
hybridizedcarbonatomcenters,afeaturecommonamongthe
palladium-catalyzed coupling reactions only to the Fu mod-
ification of the Suzuki reaction. The asymmetric allylic
alkylation reaction now provides a powerful method for
ring formation, 1,3-chirality transfer, desymmetrization of
meso substrates, the resolution of racemic compounds, and a
plethora of other applications.
Historically, the use of organozinc reagents as the
nucleophilic component in palladium-catalyzed cross-coup-
ling reactions, known as the Negishi coupling (Scheme1),
actually predates the development of both the organostan-
nane- (Stille, 1978) [6] and organoborane-based (Suzuki,
1979) [12] procedures, with the first such examples being
reported in 1977. [26] Nevertheless, the rapid and widespread
embracementofthelattertwoprotocolsbysyntheticchemists
duringthe1980sledtothepotentialoforganozincreagentsin
cross-coupling processes being relatively underappreciated
and underutilized during this time, particularly in total
synthesis. However, recent years have witnessed a resurgent
interest in the development and application of organozinc-
mediated cross-couplings, largely fueled by the recognition
that these reagents offer complementary modes of reactivity
to those of the less electropositive metals species (e.g. B and
Sn). Organozinc reagents exhibit a very high intrinsic
reactivity in palladium-catalyzed cross-coupling reactions,
which combined with the availability of a number of
procedures for their preparation and their relatively low
toxicity, makes the Negishi coupling an exceedingly useful
alternative to other cross-coupling procedures, as well as
constituting an important method for carbon–carbon bond
formation in its own right. [27]
In the following sections, we discuss the contributions of
thesepalladium-catalyzedcarbon–carbonbond-formingreac-
tions to the art and science of total synthesis and the new
thinking that they have precipitated in the field.
2. The Heck Reaction
Total synthesis has benefited enormously from the Heck
reaction, which has been widely applied in both its intermo-
lecular and intramolecular variants. [28–30] The enabling attrib-
utes of this remarkable reaction manifest themselves in a
plethora of ways, including appendage attachments, polyene
construction, fragment couplings, and ring-closure reactions.
Inthissection,wehighlightafewexamplesthatdemonstrate
theeleganceandeffectivenessofstrategiesbasedontheHeck
reaction as the key step.
Among alkaloid total syntheses employing the Heck
reaction, that of dehydrotubifoline (3) by Rawal and co-
workers stands out (Scheme2). [31] In this instance the
palladium-catalyzed process was used to forge the final
carbon–carbon bond and cast the final ring of the polycyclic
structure of the target in 79% yield (1 ! 3, Scheme2). This
effectiveoperation,involvinga6- exo -palladationfollowedby
b-hydride elimination and tautomerization of the resulting
enamine species, would seem simple enough and certainly
predictable, until one considers the different outcome
observedwiththecorresponding N -carbomethoxycyclization
substrate 2.Inthatinitialattempt,theresearchersnoticedthe
exclusive formation of the unexpected compound 7, the
apparent product of a 7- endo -cyclopalladation reaction, in
84%yield.Moreover,closescrutinyofthespectroscopicdata
of pentacyclic compound 7 revealed that inversion of
geometry of the exocyclic double bond had occurred, an
outcome inconsistent with a direct 7- endo mode of ring
closure. The proposed mechanistic explanation for the
formation of 7 is both intriguing and illuminating. Thus,
under the Jeffery modification [32] of the Heck conditions
4445
Angew. Chem. Int. Ed. 2005, 44, 4442 –4489
2005 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
1081964094.022.png
Reviews
K. C. Nicolaou et al.
ence of a b-hydrogen atom. [33] Instead, this intermediate is
sufficientlylong-livedtoundergoasecondcyclopalladationto
form the cyclopropylmethyl palladium complex 5, which is
forced by steric congestion to undergo a 1208 rotation about
the s bond. This then allows the proper alignment required
for fragmentation of the other cyclopropane bond to give
palladium complex 6, which is no longer stabilized by
carbamate complexation and undergoes the anticipated b-
hydrideeliminationtoprovidetheobservedproduct 7.Based
on these mechanistic considerations, the Rawal group sub-
jected the “carbamate-free” substrate 1 to the same reaction
conditions realizing, much to their delight, the formation of
thenaturalproductdehydrotubifoline(3).Thiscaseservesto
illustrate the fact that the “normal” mechanistic pathways of
metal-catalyzed processes may be diverted, in certain cases,
by the judicious placementof coordinating groupswithin the
employed substrates. [34]
Palladium-catalyzed reactions abound in the spectacular
synthesis of quadrigemineC (13, Scheme3), a tetrameric
member of the polypyrrolidinoindoline alkaloid family, by
Overman and co-workers. [35] Noting that the quaternary C3
andC3’’’ stereocenterswithinthetargetmolecule 13 havethe
same absolute configuration, these researchers applied cata-
lytic asymmetric Heck reactions [4] to effect the desymmetri-
zationofanadvanced meso intermediate 10,formationofthe
two peripheral indoline residues, and installation of the final
twoquaternarystereocentersinasinglestep.Thisremarkable
double cyclization (10 ! 11 ! 12) was preceded by another
highly effective palladium-catalyzed carbon–carbon bond-
formingreaction,namelyaStillecoupling,whichwasusedto
assemble the required precursor 10 from its constituent
Scheme 2. Intramolecular Heck reactions in the total synthesis of
()-dehydrotubifoline (3) (Rawal et al., 1993). [31]
(Pd(OAc) 2 cat.,K 2 CO 3 , n Bu 4 NCl,DMF,608C),theexpected
6- endo cyclization occurs to yield initially s-alkyl–palladium
species 4 which, due to stabilization by intramolecular
carbamate complexation, is prevented from undergoing a
normally facile syn -b-hydride elimination, despite the pres-
Scheme 3. Sequential tandem Stille couplings and asymmetric intramolecular Heck reactions in the enantioselective synthesis of
()-quadrigemine C (13) (Overman et al., 2002). [35]
4446
2005 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Angew. Chem. Int. Ed. 2005, 44, 4442–4489
1081964094.023.png 1081964094.024.png 1081964094.025.png
Zgłoś jeśli naruszono regulamin