Tải bản đầy đủ (.pdf) (79 trang)

The Art and Science of Total Synthesis

Bạn đang xem bản rút gọn của tài liệu. Xem và tải ngay bản đầy đủ của tài liệu tại đây (2.85 MB, 79 trang )

The Art and Science of
Total Synthesis


REVIEWS

The Art and Science of Total Synthesis at the Dawn
of the Twenty-First Century**
K. C. Nicolaou,* Dionisios Vourloumis, Nicolas Winssinger, and Phil S. Baran
Dedicated to Professor E. J. Corey for his outstanding contributions to organic synthesis

At the dawn of the twenty-first century, the state of the art and science of
total synthesis is as healthy and vigorous as ever. The birth of this exhilarating, multifaceted, and boundless science is marked by Wöhlers synthesis
of urea in 1828. This milestone eventÐ
as trivial as it may seem by todays
standardsÐcontributed to a ªdemystification of natureº and illuminated the
entrance to a path which subsequently
led to great heights and countless rich
dividends for humankind. Being both a
precise science and a fine art, this
discipline has been driven by the constant flow of beautiful molecular architectures from nature and serves as the
engine that drives the more general
field of organic synthesis forward.
Organic synthesis is considered, to a
large extent, to be responsible for some

of the most exciting and important
discoveries of the twentieth century in
chemistry, biology, and medicine, and
continues to fuel the drug discovery
and development process with myriad


processes and compounds for new
biomedical breakthroughs and applications. In this review, we will chronicle the past, evaluate the present, and
project to the future of the art and
science of total synthesis. The gradual
sharpening of this tool is demonstrated
by considering its history along the
lines of pre-World War II, the Woodward and Corey eras, and the 1990s,
and by accounting major accomplishments along the way. Today, natural
product total synthesis is associated
with prudent and tasteful selection of
challenging and preferably biologically
important target molecules; the dis-

1. Prologue
ªYour Majesty, Your Royal Highnesses, Ladies and Gentlemen.
In our days, the chemistry of natural products attracts a very
lively interest. New substances, more or less complicated,
[*] K. C. Nicolaou, D. Vourloumis, N. Winssinger, P. S. Baran
Department of Chemistry
and The Skaggs Institute for Chemical Biology
The Scripps Research Institute
10550 North Torrey Pines Road, La Jolla, CA 92037 (USA)
and
Department of Chemistry and Biochemistry
University of California, San Diego
9500 Gilman Drive, La Jolla, CA 92093 (USA)
Fax: (‡ 1) 858-784-2469
E-mail:
[**] A list of abbreviations can be found at the end of the article.
Angew. Chem. Int. Ed. 2000, 39, 44 ± 122


covery and invention of new synthetic
strategies and technologies; and explorations in chemical biology through
molecular design and mechanistic
studies. Future strides in the field are
likely to be aided by advances in the
isolation and characterization of novel
molecular targets from nature, the
availability of new reagents and synthetic methods, and information and
automation technologies. Such advances are destined to bring the power of
organic synthesis closer to, or even
beyond, the boundaries defined by
nature, which, at present, and despite
our many advantages, still look so far
away.
Keywords: drug research ´ natural
products ´ synthetic methods ´ total
synthesis

more or less useful, are constantly discovered and investigated. For the determination of the structure, the architecture
of the molecule, we have today very powerful tools, often
borrowed from Physical Chemistry. The organic chemists of
the year 1900 would have been greatly amazed if they had
heard of the methods now at hand. However, one cannot say
that the work is easier; the steadily improving methods make
it possible to attack more and more difficult problems and the
ability of Nature to build up complicated substances has, as it
seems, no limits.
In the course of the investigation of a complicated
substance, the investigator is sooner or later confronted by

the problem of synthesis, of the preparation of the substance
by chemical methods. He can have various motives. Perhaps
he wants to check the correctness of the structure he has
found. Perhaps he wants to improve our knowledge of the
reactions and the chemical properties of the molecule. If the

 WILEY-VCH Verlag GmbH, D-69451 Weinheim, 2000

1433-7851/00/3901-0045 $ 17.50+.50/0

45


REVIEWS

K. C. Nicolaou et al.

substance is of practical importance, he may hope that the
synthetic compound will be less expensive or more easily
accessible than the natural product. It can also be desirable to
modify some details in the molecular structure. An antibiotic
substance of medical importance is often first isolated from a
microorganism, perhaps a mould or a germ. There ought to
exist a number of related compounds with similar effects; they
may be more or less potent, some may perhaps have
undesirable secondary effects. It is by no means, or even
probable, that the compound produced by the microorganismÐmost likely as a weapon in the struggle for existenceÐis
the very best from the medicinal point of view. If it is possible
to synthesize the compound, it will also be possible to modify
the details of the structure and to find the most effective

remedies.

K. C. Nicolaou

D. Vourloumis

The synthesis of a complicated molecule is, however, a very
difficult task; every group, every atom must be placed
in its proper position and this should be taken in its most
literal sense. It is sometimes said that organic synthesis
is at the same time an exact science and a fine art. Here
nature is the uncontested master, but I dare say that
the prize-winner of this year, Professor Woodward, is a good
second.º[1]
With these elegant words Professor A. Fredga, a member of
the Nobel Prize Committee for Chemistry of the Royal
Swedish Academy of Sciences, proceeded to introduce R. B.
Woodward at the Nobel ceremonies in 1965, the year in which
Woodward received the prize for the art of organic synthesis.
Twenty-five years later Professor S. Gronowitz, then a member of the Nobel Prize Committee for Chemistry, concluded

N. Winssinger

P. S. Baran

K.C. Nicolaou, born in Cyprus and educated in England and the US, is currently Chairman of the Department of Chemistry
at The Scripps Research Institute, La Jolla, California, where he holds the Darlene Shiley Chair in Chemistry and the
Aline W. and L. S. Skaggs Professorship in Chemical Biology as well as Professor of Chemistry at the University of
California, San Diego. His impact on chemistry, biology, and medicine flows from his works in organic synthesis described
in nearly 500 publications and 70 patents as well as his dedication to chemical education, as evidenced by his training of

over 250 graduate students and postdoctoral fellows. His recent book titled ªClassics in Total Synthesisº,[3] which he coauthored with Erik J. Sorensen, is used around the world as a teaching tool and source of inspiration for students and
practitioners of organic synthesis.
Dionisios Vourloumis, born in 1966 in Athens, Greece, received his B.Sc. degree from the University of Athens and his
Ph.D. from West Virginia University under the direction of Professor P. A. Magriotis, in 1994, working on the synthesis of
novel enediyne antibiotics. He joined Professor K. C. Nicolaous group in 1996, and was involved in the total synthesis of
epothilones A and B, eleutherobin, sarcodictyins A and B, and analogues thereof. He joined Glaxo Wellcome in early 1999
and is currently working with the Combichem Technology Team in Research Triangle Park, North Carolina.
Nicolas Winssinger was born in Belgium in 1970. He received his B.Sc. degree in chemistry from Tufts University after
conducting research in the laboratory of Professor M. DAlarcao. Before joining The Scripps Research Institute as a
graduate student in chemistry in 1995, he worked for two years under the direction of Dr. M. P. Pavia at Sphinx
Pharmaceuticals in the area of molecular diversity focusing on combinatorial chemistry. At Scripps, he joined the
laboratory of Professor K. C. Nicolaou, where he has been working on methodologies for solid-phase chemistry and
combinatorial synthesis. His research interests include natural products synthesis, molecular diversity, molecular evolution,
and their application to chemical biology.
Phil S. Baran was born in Denville, New Jersey in 1977. He received his B.Sc. degree in chemistry from New York University
while conducting research under the guidance of Professors D. I. Schuster and S. R. Wilson, exploring new realms in
fullerene science. Upon entering The Scripps Research Institute in 1997 as a graduate student in chemistry, he joined the
laboratory of Professor K. C. Nicolaou where he embarked on the total synthesis of the CP molecules. His primary research
interest involves natural product synthesis as an enabling endeavor for the discovery of new fundamental processes and
concepts in chemistry and their application to chemical biology.

46

Angew. Chem. Int. Ed. 2000, 39, 44 ± 122


REVIEWS

Natural Products Synthesis
his introduction of E. J. Corey, the 1990 Nobel prize winner,

with the following words:
ª...Corey has thus been awarded with the Prize for three
intimately connected contributions, which form a whole.
Through retrosynthetic analysis and introduction of new
synthetic reactions, he has succeeded in preparing biologically
important natural products, previously thought impossible to
achieve. Coreys contributions have turned the art of synthesis
into a science...º[2]
This description and praise for total synthesis resonates
today with equal validity and appeal; most likely, it will be
valid for some time to come. Indeed, unlike many one-time
discoveries or inventions, the endeavor of total synthesis[3±6] is
in a constant state of effervescence and flux. It has been on the
move and center stage throughout the twentieth century and
continues to provide fertile ground for new discoveries and
inventions. Its central role and importance within chemistry
will undoubtedly ensure its present preeminence into the
future. The practice of total synthesis demands the following
virtues from, and cultivates the best in, those who practice it:
ingenuity, artistic taste, experimental skill, persistence, and
character. In turn, the practitioner is often rewarded with
discoveries and inventions that impact, in major ways, not
only other areas of chemistry, but most significantly material
science, biology, and medicine. The harvest of chemical
synthesis touches upon our everyday lives in myriad ways:
medicines, high-tech materials for computers, communication
and transportation equipment, nutritional products, vitamins,
cosmetics, plastics, clothing, and tools for biology and
physics.[7]
But why is it that total synthesis has such a lasting value as a

discipline within chemistry? There must be several reasons for
this phenomenon. To be sure, its dual nature as a precise
science and a fine art provides excitement and rewards of rare
heights. Most significantly, the discipline is continually being
challenged by new structural types isolated from natures
seemingly unlimited library of molecular architectures. Happily, the practice of total synthesis is being enriched constantly
by new tools such as new reagents and catalysts as well as
analytical instrumentation for the rapid purification and
characterization of compounds.
Thus, the original goal of total synthesis during the first part
of the twentieth century to confirm the structure of a natural
product has been replaced slowly but surely with objectives
related more to the exploration and discovery of new
chemistry along the pathway to the target molecule. More
recently, issues of biology have become extremely important
components of programs in total synthesis. It is now clear that
as we enter the twenty-first century both exploration and
discovery of new chemistry and chemical biology will be
facilitated by developments in total synthesis.
In this article, and following a short historical perspective of
total synthesis in the nineteenth century, we will attempt to
review the art and science of total synthesis during the
twentieth century. This period can be divided into the preWorld War II Era, the Woodward Era, the Corey Era, and the
1990s. There are clearly overlaps in the last three eras and
many more practitioners deserve credit for contributing to the
evolution of the science during these periods than are
Angew. Chem. Int. Ed. 2000, 39, 44 ± 122

mentioned. The labeling of these eras is arbitraryÐnot
withstanding the tremendous impact Woodward and Corey

had in shaping the discipline of total synthesis during their
time. As in any review of this kind, omissions are inevitable
and we apologize profusely, and in advance, to those
whose brilliant works were omitted as a result of space
limitations.

2. Total Synthesis in the Nineteenth Century
The birth of total synthesis occurred in the nineteenth
century. The first conscious total synthesis of a natural product
was that of urea (Figure 1) in 1828 by Wöhler.[8] Significantly,
this event also marks the beginning of organic synthesis and
OH
O
NH2

O
NH2

Me

OH

HO
HO

O
OH
OH

urea


acetic acid

glucose

[Wöhler, 1828][8]

[Kolbe, 1845][9]

[Fischer, 1890][12]

Figure 1. Selected nineteenth century landmark total syntheses of natural
products.

the first instance in which an inorganic substance
(NH4CNO:ammonium cyanate) was converted into an organic substance. The synthesis of acetic acid from elemental
carbon by Kolbe in 1845[9] is the second major achievement in
the history of total synthesis. It is historically significant that,
in his 1845 publication, Kolbe used the word ªsynthesisº for
the first time to describe the process of assembling a chemical
compound from other substances. The total syntheses of
alizarin (1869) by Graebe and Liebermann[10] and indigo
(1878) by Baeyer[11] spurred the legendary German dye
industry and represent landmark accomplishments in the
field. But perhaps, after urea, the most spectacular total
synthesis of the nineteenth century was that of (‡)-glucose
(Figure 1) by E. Fischer.[12] This total synthesis is remarkable
not only for the complexity of the target, which included, for
the first time, stereochemical elements, but also for the
considerable stereochemical control that accompanied it.

With its oxygen-containing monocyclic structure (pyranose)
and five stereogenic centers (four controllable), glucose
represented the state-of-the-art in terms of target molecules
at the end of the nineteenth century. E. Fischer became the
second winner of the Nobel Prize for chemistry (1902), after
J. H. vant Hoff (1901).[13]

3. Total Synthesis in the Twentieth Century
The twentieth century has been an age of enormous
scientific advancement and technological progress. To be
sure, we now stand at the highest point of human accomplishment in science and technology, and the twenty-first century
promises to be even more revealing and rewarding. Advances
47


REVIEWS
in medicine, computer science, communication, and transportation have dramatically changed the way we live and the
way we interact with the world around us. An enormous
amount of wealth has been created and opportunities for new
enterprises abound. It is clear that at the heart of this
technological revolution has been science, and one cannot
deny that basic research has provided the foundation for this
to occur.
Chemistry has played a central and decisive role in shaping
the twentieth century. Oil, for example, has reached its
potential only after chemistry allowed its analysis, fractionation, and transformation into myriad of useful products such
as kerosene and other fuels. Synthetic organic chemistry is
perhaps the most expressive branch of the science of
chemistry in view of its creative power and unlimited scope.
To appreciate its impact on modern humanity one only has to

look around and recognize that this science is a pillar behind
pharmaceuticals, high-tech materials, polymers, fertilizers,
pesticides, cosmetics, and clothing.[7] The engine that drives
forward and sharpens our ability to create such molecules
through chemical synthesis (from which we can pick and
choose the most appropriate for each application) is total
synthesis. In its quest to construct the most complex and
challenging of natures products, this endeavorÐperhaps
more that any otherÐbecomes the prime driving force for
the advancement of the art and science of organic synthesis.
Thus, its value as a research discipline extends beyond
providing a test for the state-of-the-art. It offers the opportunity to discover and invent new science in chemistry and
related disciplines, as well as to train, in a most rigorous way,
young practitioners whose expertise may feed many peripheral areas of science and technology.[6]

3.1. The Pre-World War II Era
The syntheses of the nineteenth century were relatively
simple and, with a few exceptions, were directed towards
benzenoid compounds. The starting materials for these target
molecules were other benzenoid compounds, chosen for their
resemblance to the targeted substance and the ease by which
the synthetic chemist could connect them by simple functionalization chemistry. The twentieth century was destined to
bring dramatic advances in the field of total synthesis. The
pre-World War II Era began with impressive strides and with
increasing molecular complexity and sophistication in strategy design. Some of the most notable examples of total
synthesis of this era are a-terpineol (Perkin, 1904),[14]
camphor (Komppa, 1903; Perkin, 1904),[15] tropinone (Robinson, 1917; Willstätter, 1901),[16±17] haemin (H. Fischer,
1929),[18] pyridoxine hydrochloride (Folkers, 1939),[19±20] and
equilenin (Bachmann, 1939)[21] (Figure 2). Particularly impressive were Robinsons one-step synthesis of tropinone
(1917)[16] from succindialdehyde, methylamine, and acetone

dicarboxylic acid and H. Fischers synthesis of haemin[18]
(1929). These total syntheses are among those which will be
highlighted below. Both men went on to win a Nobel Prize for
Chemistry (Fischer, 1929; Robinson, 1947).[13]
48

K. C. Nicolaou et al.

Figure 2. Selected landmark total syntheses of natural products from 1901
to 1939.

3.2. The Woodward Era
In 1937 and at the age of 20 R. B. Woodward became an
assistant professor in the Department of Chemistry at
Harvard University where he remained for the rest of his
life. Since that time, total synthesis and organic chemistry
would never be the same. A quantum leap forward was about
to be taken, and total synthesis would be elevated to a
powerful science and a fine art. Woodwards climactic
contributions to total synthesis included the conquest of some
of the most fearsome molecular architectures of the time. One
after another, diverse structures of unprecedented complexity
succumbed to synthesis in the face of his ingenuity and
resourcefulness. The following structures (some are shown in
Figure 3) are amongst his most spectacular synthetic achievements: quinine (1944),[22] patulin (1950),[23] cholesterol and
cortisone (1951),[24] lanosterol (1954),[25] lysergic acid (1954),[26]
strychnine (1954),[27] reserpine (1958),[28] chlorophyll a (1960),[29]
colchicine (1965),[286] cephalosporin C (1966),[30] prostaglandin F2a (1973),[31] vitamin B12 (with A. Eschenmoser) (1973),[32]
and erythromycin A (1981).[33] Some of these accomplishments
will be briefly presented in Section 3.5.

Woodward brought his towering intellect to bear on these
daunting problems of the 1940s, 1950s, and 1960s with
distinctive style and unprecedented glamour. His spectacular
successes were often accompanied by appropriate media
coverage and his lectures and seminars remained legendary
for their intellectual content, precise delivery, and mesmerizing style, not to mention their colorful nature and length!
What distinguished him from his predecessors was not just his
powerful intellect, but the mechanistic rationale and stereochemical control he brought to the field. If Robinson
introduced the curved arrow to organic chemistry (on paper),
Woodward elevated it to the sharp tool that it became for
teaching and mechanistic understanding, and used it to
explain his science and predict the outcome of chemical
reactions. He was not only a General but, most importantly, a
generalist and could generalize observations into useful
theories. He was master not only of the art of total synthesis,
but also of structure determination, an endeavor he cherished
Angew. Chem. Int. Ed. 2000, 39, 44 ± 122


REVIEWS

Natural Products Synthesis
H
H
O
H

O

N


HO

OH

H

OH

patulin (1950)[23]

N

H
H
MeO2C H

reserpine (1958)[28]

N

O

O
H
OMe

Me
H


N

O Me
Me
H
H2N

N

OH

NH2

O

Me

Me

O

O H
O P
O
O
H

H

O


MeO

NH2

HO

H OH

HO

PGF2α (1973)[31]

H
O

marasmic acid (1976)[288]
O

O

cephalosporin C (1966)[30]

isolongistrobine (1973)[287]

OHC

OMe OMe

O


illudalic acid

CO2H

(1977)[289]

illudinine

(1977)[289]

HO

H
N

R
O

MeO

N

Me

Me

Me

OMe


O

H

OHC

N

OAc

N

OH

HO

NHAc

colchicine (1965)[286]
CO2H

N

O

O

illudacetalic acid


penems
(1977)[289]

H

HO
S

N
O

H
OH

MeO

O

Me

OH

CO2H

O

vitamin B12 (1973)[32]
[with A. Eschenmoser]

OH

O

O

Me

N

O

OMe

O
CO2

N
HO

O

O

H H H
N
S

H3N

Me O


Me

MeO
OH

HO

N

Me
HO

O H

NMe2

O

NH2

N

OH

O

chlorophyll a (1960)[29]

O
H

N

N

H

6-demethyl-6-deoxytetracycline (1962)[285]

OMe

Me

CN

lanosterol (1954)[25]

NH2
N

OMe

Co

NH

HN

lysergic acid (1954)[26]

strychnine (1954)[27]


Me MeH

N

MeO2C

H2N

Me
HO

Mg

MeO

Me

NMe

O
H

H

O

H2N

O


H

H

N
N
H

H

cortisone (1951)[24]

quinine (1944)

Me

H

O

[22]

H

H

H
N


O

CO2H

N

Me

O
Me

MeO

Me

OH

O

R'

OH
Me
HO
O

OH

Me


Me
O

O
CO2H

(1978)[290]

O

O

OMe

NMe2
Me

Me

Me
OH
O
Me
erythromycin A (1981)[33]

Figure 3. Selected syntheses by the Woodward Group (1944 ± 1981).

throughout his career. He clearly influenced the careers of not
only his students, but also of his peers and colleagues, for
example, J. Wilkinson (sandwich structure of ferrocene), K.

Block (steroid biosynthesis), R. Hoffmann (Woodward and
Hoffmann rules), all of whom won the Nobel Prize for
chemistry.[13]
His brilliant use of rings to install and control stereochemical centers and to unravel functionality by rupturing
them is an unmistakable feature of his syntheses. This theme
appears in his first total synthesis, that of quinine,[22] and
appears over and over again as in the total synthesis of
reserpine,[28] vitamin B12 ,[3, 32] and, remarkably, in his last
synthesis, that of erythromycin.[33] Woodwards mark was that
of an artist, treating each target individually with total
mastery as he moved from one structural type to another.
He exercised an amazing intuition in devising strategies
toward his targets, magically connecting them to suitable
starting materials through elegant, almost balletlike, maneuvers.
However, the avalanche of new natural products appearing
on the scene as a consequence of the advent and development
of new analytical techniques demanded a new and more
systematic approach to strategy design. A new school of
thought was appearing on the horizon which promised to take
the field of total synthesis, and that of organic synthesis in
general, to its next level of sophistication.

3.3. The Corey Era
In 1959 and at the age of 31 E. J. Corey arrived at Harvard
as a full professor of chemistry from the University of Illinois,
Angew. Chem. Int. Ed. 2000, 39, 44 ± 122

Urbana-Champaign. His dynamism and brilliance were to
make him the natural recipient of the total synthesis baton
from R. B. Woodward, even though the two men overlapped

for two decades at Harvard. Coreys pursuit of total synthesis
was marked by two distinctive elements, retrosynthetic
analysis and the development of new synthetic methods as
an integral part of the endeavor, even though Woodward
(consciously or unconsciously) must have been engaged in
such practices. It was Coreys 1961 synthesis of longifolene[34]
that marked the official introduction of the principles of
retrosynthetic analysis.[4] He practiced and spread this concept
throughout the world of total synthesis, which became a much
more rational and systematic endeavor. Students could now
be taught the ªlogicº of chemical synthesis[4] by learning how
to analyze complex target molecules and devise possible
synthetic strategies for their construction. New synthetic
methods are often incorporated into the synthetic schemes
towards the target and the exercise of the total synthesis
becomes an opportunity for the invention and discovery of
new chemistry. Combining his systematic and brilliant approaches to total synthesis with the new tools of organic
synthesis and analytical chemistry, Corey synthesized hundreds of natural and designed products within the thirty year
period stretching between 1960 and 1990 (Figure 4)Ðthe year
of his Nobel Prize.
Corey brought a highly organized and systematic approach
to the field of total synthesis by identifying unsolved and
important structural types and pursuing them until they fell.
The benefits and spin-offs from his endeavors were even more
impressive: the theory of retrosynthetic analysis, new synthetic methods, asymmetric synthesis, mechanistic proposals,
and important contributions to biology and medicine. Some of
49


REVIEWS


K. C. Nicolaou et al.

Figure 4. Selected syntheses by the Corey Group (1961 ± 1999).

50

Angew. Chem. Int. Ed. 2000, 39, 44 ± 122


REVIEWS

Natural Products Synthesis
his most notable accomplishments in the field are highlighted
in Section 3.5.
The period of 1950 ± 1990 was an era during which total
synthesis underwent explosive growth as evidenced by
inspection of the primary chemical literature. In addition to
the Woodward and Corey schools, a number of other groups
contributed notably to this rich period for total synthesis[3±5]
and some continue to do so today. Indeed, throughout the
second half of the twentieth century a number of great
synthetic chemists made significant contributions to the field,
as natural products became opportunities to initiate and focus
major research programs and served as ports of entry for
adventures and rewarding voyages.
Among these great chemists are G. Stork, A. Eschenmoser,
and Sir D. H. R. Barton, whose sweeping contributions began
with the Woodward era and spanned over half a century. The
Stork ± Eschenmoser hypothesis[35] for the stereospecific

course of biomimetic ± cation cyclizations, such as the conversion of squalene into steroidal structures, stimulated much
synthetic work (for example, the total synthesis of progesterone by W. S. Johnson, 1971).[36] Storks elegant total syntheses
(for example, steroids, prostaglandins, tetracyclins)[37±39] decorate beautifully the chemical literature and his useful
methodologies (for example, enamine chemistry, anionic ring
closures, radical chemistry, tethering devices)[40±43] have found
important and widespread use in many laboratories and
industrial settings.
Similarly, Eschenmosers beautiful total syntheses (for
example, colchicine, corrins, vitamin B12 , designed nucleic
acids)[44±47] are often accompanied by profound mechanistic
insights and synthetic designs of such admirable clarity and
deep thought. His exquisite total synthesis of vitamin B12
(with Woodward), in particular, is an extraordinary achievement and will always remain a classic[3] in the annals of
organic synthesis. The work of D. H. R. Barton,[48] starting
with his contributions to conformational analysis and biogenetic theory and continuing with brilliant contributions
both in total synthesis and synthetic methodology, was
instrumental in shaping the art and science of natural products
synthesis as we know it today. Among his most significant
contributions are the Barton reaction, which involves the
photocleavage of nitrite esters[49] and its application to the
synthesis of aldosterone-21-acetate,[50] and his deoxygenation
reactions and related radical chemistry,[51] which has found
numerous applications in organic and natural product synthesis.
It seemed for a moment, in 1990, that the efforts of the
synthetic chemists had conquerred most of the known
structural types of secondary metabolites: prostaglandins,
steroids, b-lactams, macrolides, polyene macrolides, polyethers, alkaloids, porphyrinoids, endiandric acids, palitoxin
carboxyclic acid, and gingkolide; all fell as a result of the
awesome power of total synthesis. Tempted by the lure of
other unexplored and promising fields, some researchers even

thought that total synthesis was dead, and declared it so. They
were wrong. To the astute eye, a number of challenging and
beautiful architectures remained standing, daring the synthetic chemists of the time and inviting them to a feast of
discovery and invention. Furthermore, several new structures
were soon to be discovered from nature that offered
Angew. Chem. Int. Ed. 2000, 39, 44 ± 122

unprecedented challenges and opportunities. To be sure, the
final decade of the twentieth century proved to be a most
exciting and rewarding period in the history of total synthesis.

3.4. The 1990s Era
The climactic productivity of the 1980s in total synthesis
boded well for the future of the science, and the seeds were
already sown for continued breakthroughs and a new
explosion of the field. Entirely new types of structures were
on the minds of synthetic chemists, challenging and presenting
them with new opportunities. These luring architectures
included the enediynes such as calicheamicin and dynemicin,
the polyether neurotoxins exemplified by brevetoxins A and
B, the immunosuppressants cyclosporin, FK506, rapamycin,
and sanglifehrin A, taxol and other tubulin binding agents,
such as the epothilones eleutherobin and the sarcodictyins,
ecteinascidin, the manzamines, the glycopeptide antibiotics
such as vancomycin, the CP molecules, and everninomicin
13,384-1 (see Section 3.5).
Most significantly, total synthesis assumed a more serious
role in biology and medicine. The more aggressive incorporation of this new dimension to the enterprise was aided and
encouraged by combinatorial chemistry and the new challenges posed by discoveries in genomics. Thus, new fields of
investigation in chemical biology were established by synthetic chemists taking advantage of the novel molecular

architectures and biological action of certain natural products.
Besides culminating in the total synthesis of the targeted
natural products, some of these new programs expanded into
the development of new synthetic methods as in the past, but
also into the areas of chemical biology, solid phase chemistry,
and combinatorial synthesis. Synthetic chemists were moving
deeper into biology, particularly as they recognized the
timeliness of using their powerful tools to probe biological
phenomena and make contributions to chemical and functional genomics. Biologists, in turn, realized the tremendous
benefits that chemical synthesis could bring to their science
and adopted it, primarily through interdisciplinary collaborations with synthetic chemists. A new philosophy for total
synthesis as an important component of chemical biology
began to take hold, and natural products continued to be in
the center of it all. In the next section we briefly discuss a
number of selected total syntheses of the twentieth century.

3.5. Selected Examples of Total Syntheses
The chemical literature of the twentieth century is adorned
with beautiful total syntheses of natural products.[3±5] We have
chosen to highlight a few here as illustrative examples of
structural types and synthetic strategies.
Tropinone (1917)
Perhaps the first example of a strikingly beautiful total
synthesis is that of the alkaloid (Æ)-tropinone (1 in Scheme 1)
reported as early as 1917 by Sir R. Robinson.[5, 16] In this
elegant synthesisÐcalled biomimetic because of its resem51


REVIEWS


K. C. Nicolaou et al.

a)
Me
N

Mannich reaction

CO2H

CHO
NMe
O

+

O

CO2H

2

H
O

N

H2NMe
CHO


O

CHO

Mannich reaction

1: tropinone

b)

+

H2NMe

3

4

Me

- H2 O

Equilenin (1939)

NMe
NMe

+ H2O

CHO

H

2: succin-dialdehyde

5

O

OH

6

7
O

[intermolecular Mannich reaction]
Me

Me

O

1

CO2H

CO2H

10


O

H
CO2

CO2H

O H

Me O

Dieckmann
cyclization

CO2H

CO2H

9
[intramolecular Mannich reaction]

The first sex hormone to be constructed in the laboratory by
total synthesis was equilenin (1 in Scheme 3). The total
synthesis of this first steroidal structure was accomplished in
a)

N

N


HCl
-2 CO2

O2 C

Me

Me
N

N

prior to elimination of the latter functionalities. In contrast to
the rather brutal reagents and conditions used in this
porphyrins synthesis, the tools of the ªtradeº when Woodward faced chlorophyll a, approximately thirty years later,
were much sharper and selective.

O

HO
O
H CO2H

8

H
MeO

HO


1: equilenin

4: Butenandt's ketone

Scheme 1. a) Strategic bond disconnecions and retrosynthetic analysis of
(Æ)-tropinone and b) total synthesis (Robinson, 1917).[16]
Me

blance to the way nature synthesizes tropinoneÐRobinson
utilized a tandem sequence in which one molecule of
succindialdehyde, methylamine, and either acetone dicarboxylic acid (or dicarboxylate) react together to afford the natural
substance in a simple one-pot procedure. Two consecutive
Mannich reactions are involved in this synthesis, the first one
in an inter- and the second one in an intramolecular fashion.
In a way, the total synthesis of (Æ)-tropinone by Robinson was
quite ahead of its time both in terms of elegance and logic.
With this synthesis Robinson introduced aesthetics into total
synthesis, and art became part of the endeavor. It was left,
however, to R. B. Woodward to elevate it to the artistic status
that it achieved in the 1950s and to E. J. Corey to make it into
the precise science that it became in the following decades.

Me

CO2Me

HO

Haemin (1 in Scheme 2), the red pigment of blood and the
carrier of oxygen within the human body, belongs to the

porphyrin class of compounds. Both its structure and total
synthesis were established by H. Fischer.[5, 18] This combined
program of structural determination through chemical synthesis is exemplary of the early days of total synthesis. Such
practices were particularly useful for structural elucidation in
the absence of todays physical methods such as NMR
spectroscopy, mass spectrometry, and X-ray crystallography.
In the case of haemin, the molecule was degraded into smaller
fragments, which chemical synthesis confirmed to be substituted pyrroles. The assembly of the pieces by exploiting the
greater nucleophilicity of pyrroles 2-position, relative to that
of the 3-position, led to haemins framework into which the
iron cation was implanted in the final step. Among the most
remarkable features of Fischers total synthesis of haemin are
the fusion of the two dipyrrole components in succinic acid at
180 ± 190 8C to form the cyclic porphyrin skeleton in a single
step by two CÀC bond-forming reactions, and the unusual way
in which the carbonyl groups were reduced to hydroxyl groups
52

CO2H

CO2Me

H

HO

2

3


Arndt-Eistert reaction

Reformatsky
reaction

a. (CO2Me)2, MeONa
b. 180 °C, glass

b)

MeO

Me

CO2Me

(90%)

O

MeI, MeONa
(92%)

O
MeO

CO2Me
O

MeO


4

5

6

[Reformatsky reaction] a. BrZnCH2CO2Me
[dehydration] b. SOCl2, py
[saponification] c. KOH, MeOH
a. CH2N2
Me
Me
CO2Me
CO2H
b. NaOH
d. Na-Hg
c. SOCl

Me

CO2H

2

H

H
O


MeO

Cl

MeO

8

Haemin (1929)

CO2H

H

Me

Me

H
MeO

O

9

7

[Dieckmann cyclizationdecarboxylation sequence]

CO2Me


H

:

CO2Me

a. MeONa
b. HCl, AcOH
CO2Me

MeO

10

CO2H

MeO

3a
(84% overall)
a. CH2N2
b. Ag2O, MeOH [-N2]

[Arndt-Eistert
reaction]

(39% overall)
CO2H


Me O

H

HO

(92%)

1: equilenin

Scheme 3. a) Strategic bond disconnections and retrosynthetic analysis of
equilenin and b) total synthesis (Bachmann et al., 1939).[21]

1939 by Bachmann and his group at the University of
Michigan.[21, 52] This synthesis featured relatively simple
chemistry as characteristically pointed out by the authors:
ªThe reactions which were used are fairly obvious ones...º[21]
Specifically, the sequence involves enolate-type chemistry, a
Reformatsky reaction, a sodium amalgam reduction, an
Arndt ± Eistert homologation, and a Dieckmann cyclization ± decarboxylation process to fuse the required cyclopentanone ring onto the pre-existing tricyclic system of the
starting material. As the last pre-World War II synthesis of
note, this example was destined to mark the end of an era; A
new epoch was about to begin in the 1940s with R. B.
Woodward and his school of chemistry at the helm.
Angew. Chem. Int. Ed. 2000, 39, 44 ± 122


REVIEWS

Natural Products Synthesis

a)

Me

Me

Me
Me
Me
Me

N

Me

N

Br

Me

Me

1: haemin

HO2C

b)

Me


H

HO

Me

Me

Me

CO2H

H

Me

5
CO2H

NH HN

Me

3

HO2C
H

HO


Me

N
H

4
Br

Me

OHC

N
H

2

Fe
N

Me

NH HN

N

CO2Et

N

H

CO2H

6
Me

Me

Me

Me

HBr
Me

N
H

4

Me

Me

N
H

O


5

H

Me

H
NH HN
Me

Me

7

Me

a. H2SO4

HCO2H

b. ∆
Me

HCl

CO2Et

N
H


Me

OHC

Me

CO2Et

N
H

11

Me

12

CO2H

Me

piperidine
[Knoevenagel]

CO2Et

N
H

Me


N
H

Me

Me

Me

Me

H

CO2Et

N
H

6

Me

CO2Et

N
H

17


16
HO2C

HO2C

Me

2

HO2C

H

CO2Et

N
H

NH HN
Me

H

10

HO2C

HO2C

H2 O


Me

NH HN
Me

HBr,
Br2

Me

15
HO2C

Me
HO

9 Br

Na/Hg

CO2Et

N
H

13

HO2C


Me

HO2C

HO2C

14
Me

Me

NH HN
Me

Me

8

CO2H

EtO2C

Me

NH HN
Me

Me

H Me


Br

CO2Et

CO2Et

N
H

22

CO2Et

N
H

21

Br

CO2Et

N
H

δ

19


20

CO2Et

N
H

δ-

+

Br

Br

18
O

HO2C

EtO2C
Me

Me

Me

CO2Et
NH HN


Me

HO2C

HO

N
H

CO2Et

N
H

22

CO2Et

H
CO2H

23

H
H
Me

Me

NH HN


H

δ- Br Br δ+ Me

Me

Me

CO2H

HO2C

Me
Me

Me

Me

Me

Me

NH HN
Me

Br

Me


NH HN

CO2H

Br

2

Br
NH HN

Me

Me [fusion in succinic acid]

27
HO2C

CO2H

26
Me

NH HN
Me Br

CO2H

CO2H


NH HN

Me

CO2H

25

28

29
HO2C

Me

H

Br

– [CO2]
HO2C

NH HN

NH HN

O

HO2C


O

24

Br

H

NH HN

NH HN

Me

Me
Me

HO2C

CO2H

3

CO2H

CO2H

HO2C


[oxidation]
O

Me
Me

N

a. Fe3
b. Ac2O, AlCl3

HN

Me

N

HN

c. H
Me

NH

N

Me

NH


HO2C

31

30
CO2H

Me
O

KOH,EtOH, ∆

Me

Me

N

HN

OH

[reduction]

[Friedel-Crafts acylation]
Me

HO

Me


N

Me
CO2H

Me

NH

O2C

32

N

a. ∆/H
[dehydration]

Me

Me

CO2

N

N

Cl

Me

1: haemin
HO2C

HO2C

N
Fe

b. Fe3
Me

N

CO2H

Scheme 2. a) Strategic bond disconnections and retrosynthetic analysis of haemin and b) total synthesis (Fisher, 1929).[18]

Before we close this era of total synthesis and enter into a
new one, the following considerations might be instructive in
atempting to understand the way of thinking of the pre-World
War II chemists as opposed to those who followed them. The
rather straightforward synthesis of equilenin is representative
of the total syntheses of pre-World War II eraÐwith the
exception of Robinsons unique tropinone synthesis. In
contemplating a strategy towards equilenin, Bachmann must
have considered several possible starting materials before
recognizing the resemblance of his target molecule to
Butenands ketone (4 in Scheme 3). After all, three of

equilenins rings are present in 4 and all he needed to do
was fuse the extra ring and introduce a methyl group onto the
Angew. Chem. Int. Ed. 2000, 39, 44 ± 122

cyclohexane system in order to accomplish his goal. The issue
of stereochemistry of the two stereocenters was probably left
open to chance in contrast to the rational approaches towards
such matters of the later periods. Connecting the chosen
starting material 4 with the target molecule 1 was apparently
obvious to Bachmann, who explicitly stated the known nature
of the reactions he used to accomplish the synthesis.
Since the motivations for total synthesis were strongly tied
to the proof of structure, one needed a high degree of
confidence that the proposed transformations did indeed lead
to the proposed structure. Furthermore, the limited arsenal of
chemical transformations did not entice much creative deviation from the most straightforward course. This high degree of
53


REVIEWS
confidence that synthetic chemists had in their designed
strategies was soon to decrease as the complexity of newly
discovered natural products increased, thus catalyzing the
development of novel strategies and new chemistry in
subsequent years. In addition, advances in theoretical and
mechanistic organic chemistry as well as new synthetic tools
were to allow much longer sequences to be planned with a
heightened measure of confidence and considerable flexibility
for redesign along the way.
Strychnine (1954)

As the most notorious poison[53] of the Strychnos plant
species, strychnine (1 in Scheme 4) occupied the minds of

K. C. Nicolaou et al.
structural chemists for a rather long time. Its gross structure
was revealed in 1946[54] and was subsequently confirmed by
X-ray crystallographic analysis.[55] In 1952, Sir Robert Robinson commented that strychnine: ªFor its molecular size it is
the most complex substance known.º[56] This estimation had
not, apparently, escaped R. B. Woodwards attention who had
already been fully engaged in strychnines total synthesis. In
1948 Woodward put forth the idea that oxidative cleavage of
electron-rich aromatic rings might be relevant in the biogenesis of the strychnos alkaloids.[57] This provocative idea was
implemented in his 1954[27] synthesis of strychnine, which
established Woodward as the undisputed master of the art at
the time. The total synthesis of (À)-strychnine by Woodward
(Scheme 4) ushered in a golden era of total synthesis and

Scheme 4. a) Strategic bond disconnections and retrosynthetic analysis of (À)-strychnine and b) total synthesis (Woodward et al., 1954).[27]

54

Angew. Chem. Int. Ed. 2000, 39, 44 ± 122


REVIEWS

Natural Products Synthesis
installed unprecedented confidence in, and respect for, the
science of organic synthesis. Although several of its steps were
beautifully designed and executed, perhaps the most striking

feature is its reliance on only the simplest of reagents to carry
out what seemed to be rather complex chemical transformations. With its challenging molecular structure, the molecule
of strychnine continued to occupy the minds of several
subsequent practitioners of the art and several other total
syntheses have since appeared in the literature.[58, 59]
Penicillin (1957)
Few discoveries of the twentieth century can claim higher
notoriety than that of penicillin (1 in Scheme 5). Discovered
in 1928 by Alexander Fleming[60] in the secretion of the mold
Penicillium notatum, penicillin was later shown to possess
remarkable antibacterial properties by Chain and Florey.[61]
Following a massive development effort known as the
Anglo ± American penicillin project[62, 63] the substance was

introduced as a drug during World War II and saved countless
lives. Its molecular structure containing the unique and
strained b-lactam ring was under the cloud of some controversy until Dorothy Crowfoot-Hodgkin confirmed it by X-ray
crystallographic analysis.[64]
Not surprisingly, penicillin immediately became a highly
priced synthetic target attracting the attention of major
players in total synthesis of the time. Finally, it was Sheehan
and Henery-Logan[65] at the Massachusetts Institute of
Technology who delivered synthetic penicillin V by total
synthesis of the ªenchantedº molecule, as Sheehan later
called it.[66] Their synthesis, reported in 1957 and summarized
in Scheme 5, was accompanied by the development of the
phthalimide and tert-butyl ester protecting groups and the
introduction of an aliphatic carbodiimide as a condensing
agent to form amide bonds andÐin the eventÐpenicillins
fragile b-lactam ring. With this milestone, another class of

natural products was now open to chemical manipulation and
a new chapter in total synthesis had begun.
Reserpine (1958)

a)
Ring formation

Amide formation
H
N

PhO

H
S

O

N

Me

Me

O

Me

Me


O

OH

Me

HCl•H2N

CO2H

3

2
Me

NH2

HN

5: valine

6

4
O
O

Me

(75%)


O

O

Me

Ac2O, 60 °C

CO2H

Me

(72-80%)

Me

HS

+

CO2H

ClCH2COCl

CO2H

Me

S


tBuO2C HN

O

CO2H
Lactamization
1:penicillin V

b)

PhtN

H

PhtN

Me

H

N

7

Cl

Me

O

Cl

OAc
SH

O

Me

O

Me

[isomerization]

Me

Me

O

Me

Me

11

Me

O

H

Me

O

N

N

O

Me
O

O
N

N

10

Cl

9

Cl

8


(75%) H2S, NaOMe [Michael addition]
O

Me Me

Me O OMe
OMe Me

HS

O

HS

O

N

Me

N

12

13

Me

O


Me

HS
O

H

N

CHO

Me
HCl•H2N

tBuO2C

NaOAc
a. N2H4

H

N

S

tBuO2C HN

Me

HCl•H2N


(82%)

S

Me

H
N

H
S

PhO

tBuO2C HN

CO2H

16

PhOCH2COCl,
Et3N
(70%) O

H

b. HCl, H2O

Me


Me

tBuO2C HN

Me
CO2H

2

H
S
N

O
O

Me
Me

a. KOH (1.0 equiv)
b. DCC, H2O, dioxane
(12%)

O

H
N

(100%)

H
S

PhO
HO2C HN

Me
Me

CO2H

CO2K

[potassium salt of 1]

Me

19
a. HCl
b. py, acetone, H2O

H
N

Me

CO2H

18


PhO

CO2H

O
H

3a

O

Me

S
N

CO2H

4: D-penicillamine
hydrochloride

tBuO2C

17

O

HCO2H
Ac2O


(74%)
a. brucine
b. resolution
Me
c. HCl, H2O
Me
d. HCl

O

O

Me
CO2H

15

Me

HS

+

Me

S
N
H

CO2Me (100%)


14

O

N

Me
Me

b. Me2CO

N
H

Me

a. tBuONa
b. tBuOCHO

a. HCl, H2O

Me

20

Scheme 5. a) Strategic bond disconnections and retrosynthetic analysis of
penicillin V and b) total synthesis (Sheehan et al., 1957).[65]

Angew. Chem. Int. Ed. 2000, 39, 44 ± 122


Reserpine (1 in Scheme 6), a constituent of the Indian
snakeroot Rauwolfia serpentina Benth., is an alkaloid substance with curative properties[67] for the treatment of hypertension, as well as nervous and mental disorders.[68] Reserpine
was isolated in 1952 and yielded to structural elucidation in
1955 (Schlittler and co-workers)[69] and to total synthesis in
1958 (Woodward et al.).[28] The first total synthesis of reserpine (Scheme 6), considered by some as one of Woodwards
greatest contributions to synthesis, inspires admiration and
respect by the manner in which it exploits molecular
conformation to arrive at certain desired synthetic objectives.
During this synthesis, Woodward demonstrated brilliantly the
power of the venerable Diels ± Alder reaction to construct a
highly functionalized 6-membered ring, to control stereochemistry around the periphery of such a ring, and most
importantly, to induce a desired epimerization by constraining
the molecule into an unfavorable conformation by intramolecular tethering. All in all, Woodwards total synthesis of
reserpine remains as brilliant in strategy as admirable in
execution. It was to be followed by several others.[70]
The synthesis of reserpine appropriately represents Woodwards approach to total synthesis. Even though Woodward
did not talk about retrosynthetic analysis, he must have
practiced it subconsciously. In his mind, reserpine consisted of
three parts: the indole (the AB unit, see Scheme 6), the
trimethoxybenzene system, and the highly substituted E-ring
cyclohexane. Given the simplicity of the first two fragments
and their obvious attachment to fragment 3, Woodward
concerned himself primarily with the stereoselective construction of 3 and the stereochemical problem encountered in
completing the architecture of the CD ring system. He
brilliantly solved the first problem by employing the Diels ±
Alder reaction to generate a cyclic template onto which he
installed the required functionality by taking advantage of the
special effects of ring systems on the stereochemical outcomes
of reactions. He addressed the second issue, that of the last

stereocenter to be set at the junction of rings C and D, by
55


REVIEWS

K. C. Nicolaou et al.

a)
MeO

Lactamization

A

MeO

B
N
H

C-C bond
formation

C

A
B
N O
H


N
H

D

H

H
MeO2C

Imine
formation

N
H

D

O

E

H

OMe

O

E


MeO2C

OMe

OAc
OMe

OMe

1: reserpine

OMe

2

Esterification

O

O
H

O
H

+

Diels-Alder
reaction


O
H
MeO2C

CO2Me

O

MeO2C

H

E
OAc

MeO2C
H

6

5

3

4

OMe

[Meerwein-Pondorff-Verley

reduction]

b)
O

[Diels-Alder
reaction]
O

O

H

H

AcOH

OH
H

HO

H

H

13

H


3

B
C
N
N
H H
D

H

H

A

OAc

C

B
N
H

NaBH4

MeO2C

OAc

POCl3


N

H
OAc

MeO2C

OMe

N
H

OAc

15

OAc
R
HN

MeO

R = CO2Me

17

H
N H


H
O

MeO

HN
O

18

19

tBuCO2H, ∆
[isomerization]

OMe

OMe

A
B
C
N
N
H H
D
H

H
N


NaOMe, MeOH, ∆

H

H
O

E
OH

21

H
HN

MeO

MeO2C

O

OMe

20

O

py


Cl

OMe
OMe

MeO

OMe

B
C
N
N
H H
D
H

H
O

E

MeO2C

a. MeOH/CHCl3
(+)-CSA
b. resolution
c. 1 N NaOH
OMe


O
OMe

23

OMe

[esterification]

22

A

OMe
OMe

Chlorophyll a (1 in Scheme 7), the green pigment of plants
and the essential molecule of photosynthesis, is distinguished
from its cousin molecule haemin by the presence of two extra
hydrogen atoms (and, therefore, two chiral centers) in one of
its pyrrole rings, the presence of the phytyl side chain, and the
encapsulation of a magnesium cation rather than an iron
cation. Its total synthesis by R. B. Woodward et al. in 1960[29]
represents a beautiful example of bold planning and exquisite
execution. This synthesis includes improvements over Fischers routes to porphyrin building blocks and, most importantly, a number of clever maneuvers for the installment of the
three stereocenters and the extra five-membered ring residing
on the periphery of the chlorin system of chlorophyll a. The
chemical synthesis of chlorophyll a is a significant advance
over Fischers total synthesis of haemin,[18] and must have
given Woodward the confidence, and prepared the ground, for

his daring venture towards vitamin B12 in which he was to be
joined by A. Eschenmoser (see p. 61).

OMe

a. KOH, MeOH
b. DCC, py

N H

H

Chlorophyll a (1960)

E

16

N

MeO

H

D

OMe

H


MeO

OAc

2

B
N Cl
H

E

H

OMe
OAc

E

A

H

D

H

H

OMe


N

H

17

R

N

D

MeO2C

MeO

A

OMe

H

A

B
N
H

MeO


OMe

9

NH2

OMe

H

H

O

B
O
N
H

E

MeO2C

H

14
NaBH4, MeOH
[reductive amination-lactamization]


MeO

A

OH

OMe

E

MeO2C

H

O

10

OMe

MeO

H2O

MeO

OH

HO2C


H
H H2SO4 H

O
H
MeO2C

E

NaOMe
MeOH

H

H

11

a. CH2N2
b. Ac2O
c. OsO4
d. HIO4
e. CH2N2

Br

O

O


OMe

Br

8

O H

H

H

H

H

O

NBS

H2Cr2O7 H
AcOH

O

H

HO
H


O

H

O

H

OH

OMe

12

H

H

[elimination-conjugate addition]

Zn

H

O

7

Br
H


H
OH

O

Zn

O

Br2

H

4

Br

H

H

OH

OH
H

5+6 CO2Me

O


H

H

MeO2C

H

Zn

H

O Al(OiPr)3,
H
iPrOH



O

A
B
C
N
N
H H
D
H


H
O

E

OMe

O

MeO2C
OMe

1: (–)-reserpine

OMe
OMe

Scheme 6. a) Strategic bond disconnections and retrosynthetic analysis of
reserpine and b) total synthesis (Woodward et al., 1958).[28]

56

cleverly coaxing his polycycle into an unfavorable conformation (through intramolecular tethering), which forced an
isomerization to give the desired stereochemistry.
These maneuvers clearly constituted unprecedented sophistication and rational thinking in chemical synthesis
design. While this rational thinking was to be further
advanced and formalized by Coreys concepts on retrosynthetic analysis, the stereocontrol strategies of this era were to
dominate synthetic planning for some time before being
complemented and, to a large degree, eclipsed by acyclic
stereoselection and asymmetric synthesis advances which

emerged towards the end of the century.

Longifolene (1961)
The publication of the total synthesis of longifolene (1 in
Scheme 8) in 1961 by Corey et al.[34] is of historical significance in that in it Corey laid out the foundation of his
systematic approach to retrosynthetic analysis. Our thinking
about synthetic design has been profoundly affected and
shaped by the principles of retrosynthetic analysis ever since,
and the theory is sure to survive for a long time to come.
Coreys longifolene synthesis[34] exemplifies the identification
and mental disconnection of strategic bonds for the purposes
of simplifying the target structure. The process of retrosynthetic analysis unravels a retrosynthetic tree with possible
pathways and intermediates from which the synthetic chemist
can choose the most likely to succeed and/or most elegant
strategies. The total synthesis of longifolene itself, shown in
Scheme 8, involves a Wittig reaction, an osmium tetroxidemediated dihydroxylation of a double bond, a ring expansion,
and an intramolecular Michael-type alkylation to construct
the longifolene skeleton. This synthesis remains a landmark in
the evolution of the art and science of total synthesis.
Angew. Chem. Int. Ed. 2000, 39, 44 ± 122


REVIEWS

Natural Products Synthesis
a)

Hofmann elimination reaction
Me
H2N

Me

N

Me
OHC

N

Me
HN

Mg
Me

N

Me

N

NH

Me

H
H

Dieckmann cyclization
H


O

O

Me

Me

3

H2N
NC

Me

NH

NH

HN

HBr

NH

Me

NH


Me

HN

AcOH
b. H2S

Me

a. MeO2C
Me b. NaOH
c. CH2N2

HN

Me a. EtNH2,

NaBH4

HN

MeO2C

2

9

CO2Me

CO2Me


CN
Me

Me

Me

CO2Et

HN

HN

HCl

Me

CO2Et

formation]

6

NC

HN

HN


Me

CO2Et [thioaldehyde
MeO2C

Me
Cl

O

O

5

CN

O

OHC

SHC

MeO2C

NH

Me

Me
NH


Me

Me

Me

4: phytol

CO2Me

2

H3N

Me

Me

Me

Me

O

MeO2C

OHC

H2N


Me

CO2Et

1: chlorophyll a

b)

OH

+
Me

Me
Me

Ester
formation

Me

HN

NH

Me

O
CO2Me

Me

+

Cl

CO2Et

8

3

Me

7

HCl

AcHN

H3N
N
Me

Me

AcHN

Me


NH HN

Me

NH HN

Me

N

Me a. I2 [oxidation]

Me
Me

NH

Me

Me

NH HN

Me

Me

Me

AcOH, ∆


(50% overall)

Me

Me

NH HN

Me

b. Ac2O, py
NH HN

AcHN

Me

Me

NH

Me

N

NH HN

Me


Me

CO2Me

H

CO2Me

CO2Me

10

H

CO2Me

CO2Me

CO2Me

CO2Me

11

NH

CO2Me

13


12

Me

NH

Me

N

NH

Me a. resolution

Me

N

Me

N

HN

Me

H
H

CHO CO2Me


CO2Me

b. CH2N2 Me

N

b. NaOH, H2O Me

HN

Me

H
HO2C

H

19

CO2Me

14

HO

O

NH


N

N

HN

Me

O2, hv

Me

H
O CHO

CO2Me

NH

N

N

HN

CO2Me

MeO2C
MeO2C


17

Me

[Hofmann
elimination] Me
Me a. HCl,
MeOH
b. Me2SO4, Me
Me
NaOH
H

H

[photooxygenation]

MeO2C
MeO2C

O

18

Me

Me

N


[highly specific
photochemical
Me
Me cleavage of the
cyclopentadiene ring]

AcOH/∆

[hydrolysis]

Me

NH

N

N

HN

Me
CO2Me

MeO2C
MeO2C

16

Me


15

[cyanohydrin lactone
formation]

HCN, Et3N

Me

NH

a. KOH, MeOH

with quinine
Me

Me

Me

Me
CO2Me

CO2Me

AcHN
Me

Me


N

CO2Me

CO2Me

CO2Me

NH

[oxidation] Me

Me

O
CO2Me

N

air

N

Me

HN

Me

H

H
NC
CO2Me

O

Me

Me

a. Zn, AcOH
b. CH2N2
c. MeOH, HCl

[reduction]
[methylation]
[methanolysis]

O

20

Me

Me

NH

N


N

HN

H
H
MeO2C
CO2Me

Me
Me

Me

[Dieckmann
cyclization]

Me

b. H ,

N

Me

N

HN

Me


H
H

CO2Me

21

NH

NaOH, py

Me

a. NaOH, H2O
4

Me

N

c. Mg(OEt)2

[ester exchangemagnesium
insertion sequence]

Me

N


N

H

Me

Me

H

H

MeO2C
CO2Me

N
Mg

H
O

22

O

O

O
CO2Me
Me


1: chlorophyll a
Me

Me

Me
Me

Scheme 7. a) Strategic bond disconnections and retrosynthetic analysis of chlorophyll a and b) total synthesis (Woodward et al., 1960).[29] The locking of 2
and 9 together through formation of a schiff base forces the cyclization to proceed with the desired regioselectivity.

Lycopodine (1968)
Lycopodine (1 in Scheme 9), first isolated in 1881, is the
most wildly distributed alkaloid from the genus lycopodium.[71] In addition to the great challenge of synthesizing this
novel polycyclic framework in a stereocontrolled manner, one
must effectively address the challenge posed by the C13
quaternary center, which is common to all four rings. Gilbert
Stork was one of the first to successfully complete the total
synthesis of lycopodine.[72] This masterfully executed synthesis
Angew. Chem. Int. Ed. 2000, 39, 44 ± 122

features a unique ªaza-annulationº strategy which utilizes the
Stork enamine methodology[73] (a generally useful strategy to
generate and trap enolates regiospecifically) to construct
quinolone systems, a stereospecific cationic cyclization to
establish the C13 quaternary center, and a series of functional
group manipulations to elaborate the resulting aromatic ring
into ring D. Several syntheses of lycopodine have since
appeared,[74] each featuring a unique strategy complementary

to Storks beautiful synthesis.
57


REVIEWS
a)

K. C. Nicolaou et al.
Alkylation

Me

Me

Michael Me
addition

a)

Me
O

Allylic

H

Me

H oxidation


Me

O

OMe

H

H
O
H

Olefination

Me

H

1: longifolene

2

O

N
N

H

Me


O

Me

O

H

3

Cl3C

Lactamization
O

O

Me

O

pinacol
rearrangement

Me

1: lycopodine

O


Me

5: Wieland-Miescher
ketone

OH
OTs

4

Ozonolysis
CO2Me

O

H

Me

Me

N
H

Cationic
cyclization
2

Stork enamine


3

O

Me
O

CHO
O

OMe
MeO

3

OMe

CO2Et

6

Conjugate
addition

+

b)

O


HO

a.

OH

O

pTsOH, ∆
b. Ph3P=CHMe

Me

O

a. OsO4, py
b. pTsCl, py

Me

O Me

7
O

6

[pinacol
(31% overall) rearrangement]


4 H
OTs
LiClO4,
CaCO3

b)
MeO

HO

O
O

O

OH

O Me

O

Et3N , ∆

O

3

a.


Me
O
O

H

HS

7

Me Me
S

SH

BF3•Et2O
b. LiAlH4, ∆

S

Me

Me

O

9

H
Me

Me

H

a. Na, NH2NH2, ∆
b. CrO3, AcOH

O
H

Me

Me

O

H
Me

N
H

[Stork
enamine]
Ar

O

3


H2N
N

Me

(55%) O

N

O

N

Me

4

H

c. KOtBu
d. O
OMe

12

11

Cl

O


H

Me

a. LiAlH4
b. Li-NH3
[Birch
reduction]

H

N
H

OMe

Scheme 8. a) Strategic bond disconnections and retrosynthetic analysis of
longifolene and b) total synthesis (Corey et al., 1961).[34]

H
H

H

(49% overall)

Me

10


H3PO4:HCO2H (1:1)
H

Me

a. MeLi, ∆
b. SOCl2, py

H

H

N
O

O

CCl3

13

OMe

Cl3C

O3, MeOH
H

Me


Me

H

Cephalosporin C (1966)
Cephalosporin C (1 in Scheme 10) was isolated from
Cephalosporium acremonium in the mid-1950s[75] and was
structurally elucidated by X-ray crystallographic analysis in
1961.[76] Reminiscent of the penicillins, the cephalosporins
represent the second subclass of b-lactams, several of which
became legendary antibiotics in the latter part of the
twentieth century. Having missed the opportunity to deliver
penicillin, the Woodward group became at once interested in
the synthesis of cephalosporin C and, by 1965, they completed
the first total synthesis of the molecule.[30]
This total synthesis of cephalosporin C was the sole topic of
Woodwards 1965 Nobel lecture in Stockholm. Indeed, in a
move that broke tradition, R. B. Woodward described on that
occasion for the first time, and in a breathtaking fashion, the
elegant synthesis of cephalosporin C. Highlights of this synthesis, which is summarized in Scheme 10, include the
development of the azodicarboxylate-mediated functionalization of the methylene group adjacent to the sulfur atom of
l-cysteine, the aluminum-mediated closure of the aminoester
to the b-lactam functionality, the brilliant formation of
cephalosporins sulfur-containing ring, and the use of the
b,b,b-trichloroethyloxy moiety to protect the hydroxyl group.
This total synthesis stands as a milestone accomplishment in
the field of natural product synthesis.

Ar


O

(20-25%
of
desired
isomer)

Me

[cationic
cyclization]

2

10

H
N



Me

9

H2N

1: longifolene


58

O

OH

H
N

H

Me

OH
H

[conjugate
addition]
(90%)

OMe

Me

Ph3CNa; MeI (60%)
Me

a. LiAlH4
b. MeMgBr,
CuCl2


8

Me
Me 8

OMe

OMe

(36%)

O

(10-20%)

CH3

4

b. K2CO3, H2O
[decarboxylation]

O Me

2 N HCl, ∆

N

5


[Isomerization;
Michael addition]
a. NaOEt, 7

CO2Et

6

Me

H

O

O

Me

O

5

EtO2C

H

Me

H


H
H

X

O
O

N
O
Cl3C

N
O

O

16

CO2Me

O
Cl3C

15

O

CO2Me


Cl3C

H
O
O

N
OH

O

H

Me

OH

O

14

SeO2
N
H2O2
(30%
O
overall)O

CO2Me


Cl3C

2

CHO

CO2Me

a. NaOMe [formate methanolysis]
b. Zn, MeOH [deprotection of amine]
Me

H

Me

H

N
H

O

Me

N
O

H

H

O

a. LiAlH4
b. CrO3-H2SO4

N

O
H

H

17
MeO

H
H

1: lycopodine

18

O

Scheme 9. a) Strategic bond disconnections and retrosynthetic analysis of
(Æ)-lycopodine and b) total synthesis (G. Stork et al., 1968).[72]

Prostaglandins F2a and E2 (1969)

The prostaglandins were discovered by von Euler in the
1930s[77] and their structures became known in the mid-1960s
primarily as a result of the pioneering work of Bergström and
his group.[78] With their potent and important biological
activities and their potential applications in medicine,[79] these
scarce substances elicited intense efforts directed at their
chemical synthesis. By 1969 Corey had devised and completed
his first total synthesis of prostaglandins F2a (1 in Scheme 11)
and E2 .[80] These syntheses amplified brilliantly Coreys
Angew. Chem. Int. Ed. 2000, 39, 44 ± 122


REVIEWS

Natural Products Synthesis
a)

CO2H

Conjugate
addition

O
H

NH H

H2N
H


Cyclization
O

CH2OAc

N

H

S

MeO2C
H

H2N

Me
CO2Me

N N

MeO2C
H

8

S

4


Me

H

tBuOC(O)N

S
Me

7

N3
H

tBuOC(O)N

Me

5

MeO2C

Me

6

MeO2C
H

S


Me

a. acetone
MeO2C
H
b. tBuOCOCl
N
c.
CH
2
2
SH
tBuOC(O)N

HO2C
H

O

3

OAc
H

tBuOC(O)N

6: L-(+)-cysteine

b)


H
Me

2

HO2C
H
SH

+

S

Me

O

1: cephalosporin C

H2N

CHO

H

tBuOC(O)N

Amide bond formation
CO2H


CO2CH2CCl3

NH

N CO Me
2

S N

Me

Me CO2Me

9

OAc

CO2Me

MeO2C H
H
S

CO2Me

MeO2C H
H

N


tBuOC(O)N

S

Me

tBuOC(O)N

S

[oxidation]

Me

12

H

N

tBuOC(O)N
Me

MeO2C
H

Pd(OAc)4

N


N
S

Me

13

CO2Me

tBuOC(O)N

CO2Me

Me

tBuOC(O)N

CO2Me

Me

MeO2C
H

NH

Me

11


MeO2C
H

OAc

N

Me

CO2Me
N

10

Me CO2Me

MeO2C OAc
N
H
N

OAc

N

NH

S N


tBuOC(O)N

CO2Me

Me

S

Me

14

CO2Me

Me

15

OAc

[-N2]
MeO2C
H

OAc
H

tBuOC(O)N
Me


MeO2C
H

+

S

tBuOC(O)N

Me

Me

5: major product

MeO2C
H

OAc
H

OAc

tBuOC(O)N

S
Me

S


Me

17: minor product

Me

16

AcO, MeOH
O
MeO2C
OH a. MeSO2Cl
H
H b. NaN3
tBuOC(O)N
S

MeO2C
H

tBuOC(O)N
Me

H

Me

Me

18


Me

[lactamization]

CHO

H
NHCO2CH2CCl3
H

N H
H

O
CO2CH2CCl3

a.

27

O

CO2CH2CCl3
O
N

TFA

H

tBuOC(O)N

S

Me

N H
H

S

CO2CH2CCl3

H

py

[equilibration]

H

O

S

H

Me

CH2OAc


CO2H
O

CH2OAc

N
H

N H
H

O
CO2CH2CCl3

28

O

23

N

H
NHCO2CH2CCl3

O

3


CHO

O

CH2OAc

N

O
CO2CH2CCl3

H

H

O

H

25
24
b. CCl3CH2OH, DCC

26

H

H

NHCO2CH2CCl3 H2N

H
CO2H
CO2H

CO2CH2CCl3
H

22

N

DCC

Me

CO2CH2CCl3

CO2CH2CCl3

a. BH3 [neutral reducing agent]
b. Ac2O,
CO2CH2CCl3
py
O
H
NHCO2CH2CCl3

O

O


CHO
S

H



21

CO2CH2CCl3
N

Me

H

NaO

20

O

H
S

2

CCl3


CHO

HO2C H

S

HO

HO CO2H a. CCl3CH2OH, pTsOH
b. NaIO4
O
O

H

19: tartaric acid

NH

H
a. Al/Hg/MeOH
tBuOC(O)N
b. iBu3Al

4

[SN2 with inversion
of configuration]

OH


[reduction of azide]

N3

S

Zn, AcOH

NH2

N H
H

H
CO2H

[reductive removal of
the protecting groups]

S

O

1: cephalosporin C

Scheme 10. a) Strategic bond disconnections and retrosynthetic analysis of
cephalosporin C and b) total synthesis (Woodward et al., 1966).[30]

retrosynthetic analysis concepts and demonstrated the utilization of the bicycloheptane system derived from a Diels ±

Alder reaction as a versatile key intermediate for the synthesis of several of the prostaglandins. A large body of
Angew. Chem. Int. Ed. 2000, 39, 44 ± 122

Scheme 11. a) Strategic bond disconnections and retrosynthetic analysis of
(Æ)-PGF2a and b) the total synthesis (Corey et al., 1969).[80]

synthetic work[81±83] followed the initial Corey synthesis and
myriad prostaglandin analogues have since been synthesized,
aiding both biology and medicine tremendously.
Coreys original strategy evolved alongside the impressive
developments in the field of asymmetric catalysis, many of
which he instigated, which culminated by the 1990s, in a
refined, highly efficient and stereocontrolled synthesis of the
prostaglandins.[84] Thus, in its original version, the Corey
synthesis of prostaglandins F2a and E2 was nonstereoselective
and delivered the racemate and as a mixture of C15 epimers.
Then, in 1975, came a major advance in the use of a chiral
auxiliary to control the stereochemical outcome of the crucial
Diels ± Alder reaction to form the bicyclo[2.2.1]heptane
system in its optically active form.[85] The theme of chiral
auxiliaries to control stereochemistry played a major role in
the development of organic and natural products synthesis in
the latter part of the century. In addition to the contributions
59


REVIEWS

K. C. Nicolaou et al.


of Corey, those of A. I. Myers,[86] D. A. Evans,[87] W. Oppolzer,[88] and H. C. Brown[89] as well as many others helped shape
the field.
Finally came the era of catalyst design and here again the
prostaglandins played a major role in providing both a driving
force and a test. In a series of papers, Corey disclosed a set of
chiral aluminum- and boron-based[90, 91] catalysts for the
Diels ± Alder reaction (and several other reactions) that
facilitated the synthesis of an enantiomerically enriched
intermediate along the route to prostaglandins. And, finally,
the problem of stereoselectivity at C15 was solved by the
introduction of the oxazaborolidine catalyst (CBS) by Corey
in 1987.[92] These catalysts not only refined the industrial
process for the production of prostaglandins, but also found
uses in many other instances both in small scale laboratory
operations and manufacturing processes of drug candidates
and pharmaceuticals. For a more in-depth analysis of the
Corey syntheses of prostaglandins F2a and E2 and other
advances on asymmetric catalysis, the reader is referred to
ref. [4] and other appropriate literature sources.

Progesterone (1971)
Progesterone (1 in Scheme 12), a hormone that prepares
the lining of the uterus for implantation of an ovum, is a
member of the steroid class of compounds that is found
ubiquitously in nature. Its linearly fused polycyclic carbon
framework is characteristic of numerous natural products of
steroidal or triterpenoid structures. A daring approach to
progesterones skeleton by W. S. Johnson[93] was inspired by
the elucidated enzyme-catalyzed conversion[94] of squalene
oxide into lanosterol or to the closely related plant triterpenoid dammaradienol. This biomimetic strategy was also

encouraged by the Stork ± Eschenmoser hypothesis, which
was proposed in 1955[35] to rationalize the stereochemical
outcome of the biosynthetic transformation of squalene oxide
to steroid. According to this postulate it was predicted that
polyunsaturated molecules with trans CˆC bonds, such as
squalene oxide, should cyclize in a stereospecific manner, to
furnish polycyclic systems with trans,anti,trans stereochemistry at the ring fusion.
This brilliant proposition was confirmed by W. S. Johnson
and his group through the biomimetic total synthesis of
progesterone (Scheme 12). A tertiary alcohol serves as the
initiator of the polyolefinic ring-closing cascade, in this
instance, but other groups have also been successfully
employed in this regard (for example, acetal, epoxide). The
methylacetylenic group performed well as a terminator of the
cascade in the original work. A number of new terminating
systems have since been successfully employed (for example,
allyl or propargyl silanes, vinyl fluoride). The work of W. S.
Johnson was complemented by that of van Tamelen[95] and
others[3, 4] who also explored such biomimetic cascades.

Tetrodotoxin (1972)
Tetrodotoxin (1 in Scheme 13) is the poisonous compound
of the Japanese puffer fish and its structure was elucidated by
60

Scheme 12. a) Strategic bond disconnections and retrosynthetic analysis of
progesterone and b) total synthesis (Johnson et al., 1971).[93]

Woodward in 1965.[96] By 1972 Kishi and his group had
published the total synthesis[97] of this highly unusual and

challenging structure. This outstanding achievement from
Japan was received at the time with great enthusiasm and
remains to this day as a classic in total synthesis. The target
molecule was reached through a series of maneuvers which
included a Diels ± Alder reaction of a quinone with butadiene,
a Beckman rearrangement to install the first nitrogen atom,
stereoselective reductions, strategic oxidations, unusual functional group manipulations, and, finally, construction of the
guanidinium system. As a highly condensed and polyfuncAngew. Chem. Int. Ed. 2000, 39, 44 ± 122


REVIEWS

Natural Products Synthesis
a)

HO

C-N Bond
formation

N O
H
HO

NH2
AcHN

O
O


1: tetrodotoxin

H
CH2OH

HN
H2N

O

OH

H

OH
OAc
CH2OAc

N

O

AcO H

OH

2

O


OAc

trans-Esterification/
epoxide opening

Orthoester
formation
Epoxide opening/
cyclization

H

H

O

HO

H

AcO
OAc
AcNH

4

3

O


N

O

H

, SnCl4

Me

Me

[Lewis acid catalyzed N
Diels-Alder
HO
reaction]

O

5

O
Me

O

a. MsCl, Et3N
b. H2O, ∆
(61%)


H

[Beckmann
rearrangement]

AcN O
H

Me

(83%)

HO

O

O

O
AcNH

Diels-Alder
reaction

CH2OAc

O
O

H


b)

Baeyer-Villiger
oxidation

Me

6

Me

7

[regio- and stereoselective reduction] a. NaBH4, MeOH
(72%)
[epoxide-mediated etherification] b. mCPBA, CSA

a. SeO2
b. NaBH4
(100%) O

H
OH
O

O

O AcN
H OAc


9
a. mCPBA
b. Ac2O, py
c. TFA, H2O;
Ac2O, py

O

O
O
AcO AcN OAc
H

10

O
AcHN

OH

H

14

H

OH

HO

H2N
AcHN

O
H

OAc

OAc
CH2OAc
H

N

O
S
H2N

OH

H

H

AcN O
H

4

O


O
OAc

a. KOAc, AcOH
b. Ac2O, CSA, ∆

CH2OAc

OAc

O

N
H
AcO

H O
H

O

13

a. H5IO6
b. NH4OH
(9% overall)

OAc


15

OH

H

OAc

OAc

O

AcO

Me
O

O
O
AcHN
O
(100%)
c. vacuum,
H O
[Baeyer-Villiger O
300 °C
AcO
oxidation]
AcO AcN OAc [acetate elimination]
H

H
O
(80%)
11
3
a. OsO4, py
(65%)
b. (MeO)2CMe2, CSA
c. Et3O BF4 , Na2CO3; AcOH

BF4 ;
Ac2O, py
b. acetamide
(50%)

O

a. BF3, TFA
b. TFA, H2O
HO

c. Al(OiPr)3, iPrOH, ∆
d. Ac2O, py
(86% overall)

H

mCPBA

H O a. Et O

3

AcO

H
HO

OAc

AcO-

CH2OAc
H

OAc
N

AcHN

O

OH

HO

8
d. (EtO)3CH, CSA; Ac2O, py [diethylketal formation]
e. ∆
[ethyl enol ether formation]
f. mCPBA, K2CO3

[epoxidation]
g. AcOH
[epoxide opening and acetylation]
(53% overall)
OAc

H

Me

O AcN
H OAc

H

O

a. CrO3, py
b. BF3•Et2O,

H

O

CH2OAc
H

OAc
CH2OAc
H


OAc
H2N

H O
a. BrCN, NaHCO3 AcO
H
b. H2S
O
(100%)
12

HO

H
H

OH
CH2OH

HN
H2N

N O
H
HO H

H
O
O


OH

1: tetrodotoxin

Scheme 13. a) Strategic bond disconnections and retrosynthetic analysis of
tetrodotoxin and b) total synthesis (Kishi et al., 1972).[97]

tional molecule, tetrodotoxin was certainly a great conquest
and elevated the status of both the art and the practitioner,
and at the same time was quite prophetic of things to come.
Vitamin B12 (1973)
The total synthesis of vitamin B12 (1 in Scheme 14),
accomplished in 1973 by a collaboration between the groups
of Woodward and Eschenmoser,[3, 32] stands as a monumental
achievement in the annals of synthetic organic chemistry.
Angew. Chem. Int. Ed. 2000, 39, 44 ± 122

Rarely before has a synthetic project yielded so much
knowledge, including: novel bond-forming reactions and
strategies, ingenious solutions to formidable synthetic problems, biogenetic considerations and hypotheses, and the seeds
of the principles of orbital symmetry conservation known as
the Woodward and Hoffmann rules.[98] The structure of
vitamin B12 was revealed in 1956 through the elegant X-ray
crystallographic work of Dorothy Crowfoot-Hodgkin.[99] The
escalation of molecular complexity from haemin to chlorophyll a to vitamin B12 is interesting not only from a structural
point of view, but also in that the total synthesis of each
molecule reflects the limits of the power of the art and science
of organic synthesis at the time of the accomplishment.
One of the most notable of the many elegant maneuvers of

the Woodward ± Eschenmoser synthesis of vitamin B12 is the
photoinduced ring closure of the corrin ring from a preorganized linear system wrapped around a metal template,
which was an exclusive achievement of the Eschenmoser
group. The convergent approach defined cobyric acid (2 in
Scheme 14) as a landmark key intermediate, which had
previously been converted into vitamin B12 by Bernhauser
et al.[100] The synthesis of vitamin B12 defined the frontier of
research in organic natural product synthesis at that time. For
an in depth discussion of this mammoth accomplishment, the
reader is referred to ref. [4].
Erythronolide B (1978)
The macrolide antibiotics, of which erythromycin is perhaps
the most celebrated, stood for a long time as seemingly
unapproachable by chemical synthesis. The origin of the
initial barriers and difficulties was encapsulated in the
following statement made by Woodward in 1956, ªErythromycin, with all our advantages, looks at present hopelessly
complex, particularly in view of its plethora of asymmetric
centers.º[101] In addition to the daunting stereochemical
problems of erythromycin and its relatives, also pending was
the issue of forming the macrocyclic ring. These challenges
gave impetus to the development of new synthetic technologies and strategies to address the stereocontrol and macrocyclization problems.
The brilliant total synthesis of erythronolide B[102] (1 in
Scheme 15), the aglycon of erythromycin B, by Corey et al.
published in 1979, symbolizes the fall of this class of natural
products in the face of the newly acquired power of organic
synthesis. Additionally, it provides further illustration of the
classical strategy for the setting of stereocenters on cyclic
templates. The synthesis began with a symmetrical aromatic
system that was molded into a fully substituted cyclohexane
ring through a short sequence of reactions in which two

bromolactonizations played important roles. A crucial Baeyer ± Villiger reaction then completed the oxygenated stereocenter at C6 and rendered the cyclic system cleavable to an
open chain for further elaboration.
As was the case in many of Coreys syntheses, the total
synthesis of erythronolide B was preceded by the invention of
a new method, namely the double activation procedure for the
formation of macrocyclic lactones employing 2-pyridinethiol
esters.[103] This landmark invention allowed the synthesis of
61


REVIEWS

K. C. Nicolaou et al.

Scheme 14. a) Strategic bond disconnections and retrosynthetic analysis of (À)-vitamin B12 , b) key synthetic methodologies developed in the course of the
total synthesis, c) and final synthetic steps in the Woodward-Eschenmoser total synthesis of vitamin B12 (Woodward ± Eschenmoser, 1973).[32]

62

Angew. Chem. Int. Ed. 2000, 39, 44 ± 122


REVIEWS

Natural Products Synthesis
a)

Me

O

Me

Me
Me

Me
O

Me
OBz
O

HO

OH

Me

OH

OTBS
O

OTBS

OBz
Me

Me


Me

Me

1: erythronolide B

2

Bromolactonization

Me

Br

Me

Bromolactonization

O

Me

Me

O

Me

O


NaOMe

(72%)

Me

10

O

O

O

O

Me

Me
O

13
a. H2, Raney-Ni
b. BzCl

12

OBz

OBz


a. LDA
Me
b. MeI
(75%
O overall) BzO
Me

Br2, KBr

Me

7

Me

(91%)

O

Me

Me

O
Me

O

6


11

[bromolactonization]

CO2H

OBz
Me

Me

BzO
Me

Me

O

O
Br

O

Me

a. KOH, H2O (98%)
b. resolution

O


(93%)

O

Me
O

Me

nBu3SnH
Me
AIBN

Me

O

O

CO2H

9

(76%)

a. LiOH
Me
b. CrO3, H2SO4
(80%)


O

OBz
Me

Me

Me

CH3CO3H

BzO
Me

(70%) BzO

O

Me

Me

CO2H

[Baeyer-Villiger
oxidation]

Me


16

14

5
a. H2O2, Na2WO4
b. resolution
c. ClCO2Et
Me d. NaBH4 OMe
e. POCl3,
(76% overall)

O
Me

18

15
a.
Li
b. Amberlyst IRC-50
c. ArSO2Cl, py
Me d. Me2CuLi
Me
I
O e. TBSCl, imid.
Me
f. LDA; MeI
OMe g. [Cp2ZrHCl]
OTBS

Me h. I2, CCl4
Me

19

3

Ph3P

17
N

S S

(65%)

N

OBz
Me

Me

BzO
Me

O O O
S

tBuLi, MgBr2


N

4

Me

(90%)

OH
Me

Me

Me

Me

a. AcOH
HO
b. LiOH
OBz
H2O2 Me

Me

BzO
Me

O O O


Me

Me

N

OH

iPr
Me

O

Me

OH

tBu
N

N
S S

23
Ph3P;

Me

iPr


PhMe, ∆
(50% )

a. MnO2
(98% )
b. H2O2, NaOH

10
OH

Me
OH
OH
OH

O

Me

24

O

a. H2, Pd/C [epoxide reduction]
b. K2CO3, MeOH
[epimerization at C-10]
c. HCl

Me


1: erythronolide B

Me

Me
O

Me

Me
OH

Me
O

Me

O
Me

O

O
Me

Me

25


Scheme 15. a) Strategic bond disconnections and retrosynthetic analysis of
erythronolide B and b) total synthesis (Corey et al., 1978).[102]

Angew. Chem. Int. Ed. 2000, 39, 44 ± 122

Monensin[105] (1 in Scheme 16), isolated from a strain of
Streptomyces cinamonensis, is perhaps the most prominent
member of the polyether class of antibiotics. Also known as
ionophores, these naturally occurring substances have the
ability to complex and transport metals across membranes,
thus exerting potent antibacterial action.[106, 107] These structures are characterized by varying numbers of tetrahydropyran, tetrahydrofuran, and/or spiroketals. Kishis total synthesis
of monensin,[108] which followed his synthesis of the simpler
ionophore lasalocid,[109] represents a milestone achievement
in organic synthesis (Scheme 16). This accomplishment demonstrates the importance of convergency in the total synthesis
of complex molecules and is one of the first examples of
stereoselective total synthesis through acyclic stereocontrol,
and elegantly marked the application of the Cram rules within
the context of natural-product synthesis. By unraveling the
spiroketal moiety of the molecule Kishi was able to adopt an
aldol-based strategy to couple monensins two segments. A
series of daring reactions (for example, hydroborations,
epoxidations) on acyclic systems with pre-existing stereocenters allowed the construction of the two heavily substituted fragments of the molecule which were then successfully
coupled and allowed to fold into the desired spiroketal upon
deprotection. Kishis beautiful synthesis of monensin also
provided a demonstration of the importance of 1,3-allylic
strain in acyclic conformational preferences, which in turn can
be exploited for the purposes of stereocontrolled reactions
(for example, epoxidation).
A second total synthesis of monensin was accomplished in
1980 by W. C. Still and his group (Scheme 17).[110] Just as

elegant as Kishis synthesis, the Still total synthesis of
monensin demonstrates a masterful application of chelationcontrolled additions to the carbonyl function. A judicious
choice of optically active starting materials as well as a highly
convergent strategy that utilized the same aldol ± spiroketalization sequence as in Kishis synthesis allowed rapid access to
monensins rather complex structure.

O
O

Me

Me

Me
O

OH
Me

O

O

O

Me

Me
Me
O


Me

22

Me

Me

Me

N

O

Me

Me Me

20

tBu

Me
OH
Me

OTBS
Me


Me

2

Me

HO

OTBS

d. Amberlyst IRC-50
e. KOH
(61% overall)

Me

Me

O

Me

Me

OH
Me

Me

Me


O

OMe

Me

Zn(BH4)2

O

OBz

21 HO
a. KOH
b. CH2N2
c. HBr,

Me
OBz

OH

Me
OH

OBz

OBz


Me

Me

O

O

Me

CO2H

O

O

HO2C

5
O

(96%)
Me

8

HO

O


O [bromolactonization]
Br
Me
Br2, KBr Me

Me

Br

Me Al/Hg

BzO
Me

6

a. BH3•THF;
Me
H2O2, NaOH
b. CrO3, H2SO4

Me

Me

Me

Me

Me


O

OH

Monensin (1979, 1980)

N

4

OBz
Br

Me

Me 7

S
Me

O

O

Me

O O O

Baeyer-Villiger

oxidation

O

O

8 Me

b)

Me

Alkylation

Me

Me

Me
BzO
Me

C-C bond formation

Lactonization

OH

3


Me
Me

OH Functionalization

O

several macrolides including erythronolide B and, most
significantly, catalyzed the development of several improvements and other new methods for addressing the macrocyclization problem.[104] Soon to follow Coreys synthesis of
erythronolide B was Woodwards total synthesis of erythromycin A.[33]

I

Me

Me

Me

OH

Me

OBz

Endiandric Acids (1982)
The endiandric acids (Scheme 18) are a fascinating group of
natural products discovered in the early 1980s in the
Australian plant Endiandra introsa (Lauraceae) by Black
et al.[111] Their intriguing structures and racemic nature gave

rise to the so called ªBlack hypothesisº for their plant origin,
which involved a series of non-enzymatic electrocyclizations
from acyclic polyunsaturated precursors (see Scheme 18).
Intrigued by these novel structures and Blacks hypothesis for
their ªbiogeneticº origin, we directed our attention towards
their total synthesis. Two approaches were followed, a
63


REVIEWS

K. C. Nicolaou et al.

a) Aldol condensation

Spiroketalization
O

HO

Me

Me

Me

O

O
Me


Me H

O

H

H

Me
CO2H

a.

OMe
Me

2

Me

Ph3P

Me

O

O

H Me


H

CO2Me

3

HO O

MeO

Me
H

O

6

BH3;
KOH, H2O2
(85%)

Me

7

OH

4


Me

BH3

O
Me

Me

8

[8:1 mixture]

MeO

a. O
(MeO)2P

a. KH, MeI
b. H2, 10% Pd/C
c. resolution
d. PCC
(77% overall)

OBn

H Me

H


HO O

OBn

b. LiAlH4
c. BnBr, KH
(66% overall)

H

O

O

CO2Et
Me

O

O

+

Me

Et

MeO

Me


H Me

1: monensin HO

b)
a. nBuLi, MeI
CN
b. KOH
c. LiAlH4
5
d. PCC

O

H

Me

Et

Me
O OH

OBn

H
HO O

OMe


O

H

Me

Et

OMe O

Me

CO2Me

OMe

Me

Me
O
Me

b. LiAlH4
(73%)

Me

O
Me


9

Me

(80%)
Me a. mCPBA

Me

Me

Me

b. KOH aq.
c. resolution

CO2H

O

OH

(35% overall)

PPh3

14

13


15

a. PhCHO, CSA
b. LiAlH4-AlCl3 (1:4)
OH c. resolution
HO

HO

CH3C(OEt)3
CH3CH2CO2H, ∆

Et
H

(93%)

OBn

a. PCC
b.

Et

HO

[Johnson
orthoester Claisen
rearrangement]


H

16

Me

Me

Me

Ar
O

Et H

[bromoetherification]

26

KO2,
[18]crown-6
DMSO

Me

Ar
H

Et H


O

H H

OH

27
Me
Me
MeO OH
OH

Ar

OH
H

O

Et H

O

Me

OH

Me


11
[12:1 mixture]

Et

H

21 OH

O
Et

Et

Ar

O

mCPBA

H

OH

(36%)

[7:2 mixture]
24

(78%)


OH

[hydroxyl-directed
epoxidation]
22

23

a. NaOMe, MeOH
b. (CH3O)3CH
MeOH, CSA

Me
Me
Me
a. Cl3CCOCl, py
Me
b. OsO4, py
Ar
OBz
O
c. BzCl, py
O
O
H
Et H
H HO
d. CrO3, H2SO4
CCl3

28
O
Me

H
Me

PPh3

[Wittig reaction]
25

20
a. pTsCl
O
b. LiAlH4
c. CSA
Ar
d. OsO4, NaIO4

Me

a. LiAlH4
Ar
b. PCC
OBn c. MeOC H MgBr
6 4
H
d. CrO3, H2SO4
e. BCl3

(31% overall)

EtO2C

OBn

OMe OH

(47%)

Me

O

15

H

Me

Me

NBS Ar
O
OH
(57%)
H
Et H

Br


H

Me

O

19

Me

Me

Me

O

18

17

EtO

BH3; H2O2

O

Et

Et


OBn

MgBr

H

a. MOMBr,
OMe OBn O
OMe OBn OMOM
a. CH2N2
PhNMe2
MeO2C
HO2C
H
b. BnBr, KH O
b. HCl
c. O3 , MeOH
Me Me Me
Me Me Me
c. PCC
12
2
(33% from 11)

Me

Me

12 steps


OH

10

Me

H

(53% overall)

Me

Me

Me

MeMgBr
O

Et H

O

H H

O

OMe


(22% from 4)

OHC

MeO O

Me

O

O

Et H

H H

Li, EtOH
NH3 (l)

OH

O

OMe

[Birch reduction]

4

Me

OH

Me

MeO

H O Et H O H H O OMe

31

OH

30

a. (CH3O)3CH
MeOH, CSA
b. O3 , MeOH
c. MgBr2

Me

Me

Me

MeO
H

O


Et H

O

O

H H

OH
OMe

29

a. O3 , MeOH
b. HCl, MeOH
c. MeLi
OH
H
Me

Me
O OH

Et
O

HO

Me


iPrNMgBr, 2

H Me

(21%, 92% based on recovered SM)

H

OBn
Me
OMe
Me

HO O

Et

Me

Me

Me

H

OH

O H
O


a. H2 , 10% Pd/C
b. CSA, H2O

Me

c. NaOH-MeOH (1:5)

Me

H Me

H

Me

3

[8:1 mixture]
MeO
32

Et
O

O

Me H

O


H

H
OMe

Me

HO O
CO2Me

MeO

HO

O

O

Me

H Me

H
HO

O

CO2Na
HO


Me

Me

1-Na: (+)-monensin sodium salt

Scheme 16. a) Strategic bond disconnections and retrosynthetic analysis of monensin and b) total synthesis (Kishi et al., 1979).[108]

stepwise (Scheme 19 b) and a direct one-step strategy
(Scheme 19 c). Both strategies involve an 8-p-electron electrocyclization, a 6-p-electron electrocyclization, and a Diels ±
Alder-type [4‡2] cycloaddition reaction to assemble the
polycyclic skeletons of endiandric acids. The total synthesis[112]
of these architecturally interesting structures demonstrated a
number of important principles of organic chemistry and
verified Blacks hypothesis for their natural origin. In
particular, the ªone-potº construction of these target mole64

cules from acyclic precursors from the endiandric acid cascade
is remarkable, particularly if one considers the stereospecific
formation of no less than four rings and eight stereogenic
centers in each final product.
Efrotomycin (1985)
Efrotomycin (1 in Scheme 20; see p. 67), the most complex
member of the elfamycin class of antibiotics[113] that includes
Angew. Chem. Int. Ed. 2000, 39, 44 ± 122


REVIEWS

Natural Products Synthesis


Scheme 17. a) Strategic bond disconnections and retrosynthetic analysis of monensin and b) total synthesis (Still et al., 1980).[110]

aurodox, was isolated from Nocardia lactamdurans. [114] Its
molecular structure, which contains nineteen stereocenters
and seven geometrical elements of stereochemistry, presented
considerable challenge to the synthetic chemists of the 1980s,
particularly in regard to the oligosaccharide domain and the
all-cis-tetrasubstituted tetrahydrofuran system. The total synAngew. Chem. Int. Ed. 2000, 39, 44 ± 122

thesis of efrotomycin, accomplished in 1985 in our laboratories,[115] addressed both problems by devising new methodologies for the stereoselective construction of glycosides and
tetrahydrofurans. Scheme 20 summarizes this total synthesis
in which the two-stage activation procedure for the synthesis
of oligosaccharides utilizing thioglycosides and glycosyl
65


REVIEWS

K. C. Nicolaou et al.

CO2R

CO2R
Ph

a

Ph


a

a CO2R

CO2R

a

Ph

Ph

CO2R

RO2C

CO2R
Ph Ph

b

H
H
HO2C

H

endiandric acid D

Ph


b

Ph

H

endiandric acid E
Ph

H
H

CO2H
HO2C

H

endiandric acid F

Diels-Alder
Ph

endiandric acid G
Diels-Alder

H

H
HO2C


H

H
H

endiandric acid A

H

H

H
H

CO2H

H

Ph

H

Diels-Alder

H

Ph

b


H
H

H
H CO H
2

Ph
Ph

RO2C

b

H

H
H

endiandric acid B

CO2H

H
H

H

Ph


H

endiandric acid C

Scheme 18. The endiandric acid cascade (Black et al., R ˆ Me, H). a) Conrotatory 8-p-electron cyclization; b) disrotatory 6-p-electron cyclization.[111]

Scheme 19. a) Strategic bond disconnections and retrosynthetic analysis of endiandric acids A ± C, b, c) total synthesis, and d) ªbiomimeticº synthesis of
endiandric acid methyl esters A ± C (Nicolaou et al., 1982).[112]

66

Angew. Chem. Int. Ed. 2000, 39, 44 ± 122


REVIEWS

Natural Products Synthesis
a)

OMe

Glycosidation
H

O
Me

O
Me


Wittig
olefination

O

OH

O

O

Me

OMe H

H
N

Me
O
Me

O

OH

OMe
OTBS


F

Me

H

O

2

OMe H

H
N

O

Me

O

Me
O

Me
OMe

O

OH


Me

H

O

O

Me

Me

O

Me

Me

O

N

Ph3P

OH

3

1: efrotomycin


O

Me

O

OH

TBSO

OH

O

Me

Me

O
O

Me

5

N
PhS

Me

HO OH

Me

Amide
formation

CCl3

OMe Me

OMe
OH

OMe

HO

Glycosidation

Me

O

Br

4

OBn


6

b)

O

O

a. LiCuMe2; TMSCl
b. O3; Me2S
c.
OMe

a. KCH2S(O)CH3
b. TBSCl, imid.

MeO2C

H

O

H

OMe H

OTBS

CuLi


MeO2C
O

O

O

8

7
Me

O

Me

OH Me
OMe
OH

a. nBu2SnO, ∆
b. BnBr
c. KH, MeI

O
Me

OMe
OBn


Me

9

(55%)

O

OMe Me

d. NBS, DAST

12

MeO

Me

a. H2, 5% Pd/C
b. TBSCl, imid.
c. PhSTMS, ZnI2

OMe Me

H

O

OTBS


O

Me

Me

5, 6

Me
TMSO

10

Me

OTMS

OMe
OMe
OTBS

O

F

a. AgClO4, SnCl2
b. NBS, DAST

O
O


O
HO

NBS, AIBN

Ph

O

(100%)

c. PhSTMS, ZnI2;
K2CO3, MeOH
(70%)

14

a. Swern [O]
b. tBuOK,

Me
O

a. [Rh], H2
b. LiAlH4
c. CSA, acetone

Me
O


O

Me

O

15

OH
Me Me

(70%)

17

18

OMe
O

(80%)

O

O

OH

O


O

AcO

3

16

Me

a. (-)-DET, Ti(iPrO)4
tBuOOH
b. BnOC(O)Cl, py
c. AlCl3; H2O
(65%)

HO

O
O

c. DIBAL-H
(85%)

Me

a. AcOH, H2O
b. NaOH/EtOH


Me Me

19

Me

O
Me

Me

(85%)

O
H

OMe

Me

(85%)

Me

OH

O
O

a. (MeO)2CMe2,

CSA
b. K2CO3, MeOH;
CSA
c. RuO2, NaIO4

O
Me Me O

20

O

a. AcOH, H2O
b. PCC

O

c.

Me

22

Me

O

Me

H


O
Me

a. AlMe3, 10
b. HF•py
c. DDQ, MeOH

OH

O

24

Me
OMe H

H
N

O
Me

OH

O

OMe
OH


Me

Me

(26% overall)

21

Me

HO
O

Me Me

Me

OMe
O

Me

P(O)Ph2

(59%)
OMe

Me
Me
CO2H


O

O
O

Me

O

OMe
OH

(86%)
Me
Me

, nBuLi

Me

Me

Me
OH

O

Me


Me

O

O

Me

O
O

O

O

HO
O

CH3CH2CH2CO2Et,
LDA

Me

O
Me

OEt
O

23

c. 16, AgClO4, SnCl2
d. K2CO3, MeOH
OMe

(86%)

Me
Me

OH

4

OH

O

OMe
OAc

Me

Me
Me

Me

OMe

OH


Me
CO2Me

O

F
Me

TBSO

Me

MeO3P(O)

O

SPh

OPMB

O

TBSO

(63%)

a. nBu3SnH, AIBN
b. TBSCl, imid.


Br
OBz

HO

13

Me

OMe

O

OMe
OTBS

OMe
OMe

OMe

N

Me

O

a. TBAF
b. Ac2O, 4-DMAP
OMe


O

Me

OMe

2

F

(66%)

Me
O

O

H

H2N

O

d. KH, MeI
e. AcOH, H2O

(90%)

11


MeO

O

Me

Me
OMe H

O

1: efrotomycin

Me

O

H

Me
HO OH

Me

O

N

O


OH

Scheme 20. a) Strategic bond disconnections and retrosynthetic analysis of efrotomycin and b) total synthesis (Nicolaou et al., 1985).[115]

fluorides[116] as well as the base-induced zip-type diepoxide
opening were highlighted as powerful methods for organic
synthesis. Numerous applications and extensions of these
synthetic technologies have since followed.[117]
Okadaic acid (1986)
Okadaic acid[118] (1 in Scheme 21) is a marine toxin isolated
from Halichondria Okadai. Besides its shellfish toxicity,
Angew. Chem. Int. Ed. 2000, 39, 44 ± 122

okadaic acid exhibits potent inhibition of certain phosphatases and is a strong tumor promotor. With its three spiroketal
moieties and seventeen stereogenic centers, the molecules
polycyclic structure presented a serious challenge to synthetic
chemistry. The first total synthesis of okadaic acid was
achieved in 1984 by the Isobe group in Japan[119] and was
followed by those of Forsyth[120] and Ley.[121] The Isobe
synthesis of okadaic acid, summarized in Scheme 21, highlights the use of sulfonyl-stabilized carbanions in synthesis, the
67


REVIEWS

K. C. Nicolaou et al.

Scheme 21. a) Strategic bond disconnections and retrosynthetic analysis of okadaic acid and b) total synthesis (Isobe et al., 1986).[119]


control of stereochemistry through chelation, and the power
of the anomeric effect to exert stereocontrol in spiroketal
formation.
Amphotericin B (1987)
The polyene macrolide family of natural products is a
subgroup of the macrolide class, which poses formidable
challenges to synthetic organic chemistry. Among them,
68

amphotericin B[122] (1 in Scheme 22), isolated from Streptomyces nodosus, occupies a high position as a consequence of
its complexity and medical importance as a widely used
antifungal agent. Its total synthesis[123] in 1987 by our group
represented the first breakthrough within this class of complex molecules. This total synthesis featured the recognition
of subtle symmetry elements within the target molecule that
allowed the utilization of the same starting material to
construct two, seemingly unrelated, intermediates and the
Angew. Chem. Int. Ed. 2000, 39, 44 ± 122


×