(Raymond Cooper, John K. Snyder) Biology - Chemist (BookFi) PDF
(Raymond Cooper, John K. Snyder) Biology - Chemist (BookFi) PDF
(Raymond Cooper, John K. Snyder) Biology - Chemist (BookFi) PDF
edited by
Raymond Cooper
Pharmanex, Inc.
San Francisco, California
John K. Snyder
Boston University
Boston, Massachusetts
MARCEL
MARCELDEKKER,
INC. NEWYORK BASEL
D E K K E R
ISBN: 0-8247-7116-8
This book is printed on acid-free paper.
Headquarters
Marcel Dekker, Inc.
270 Madison Avenue, New York, NY 10016
tel: 212-696-9000; fax: 212-685-4540
Eastern Hemisphere Distribution
Marcel Dekker AG
Hutgasse 4, Postfach 812, CH-4001 Basel, Switzerland
tel: 41-61-261-8482; fax: 41-61-261-8896
The publisher offers discounts on this book when ordered in bulk quantities. For more
information, write to Special Sales/Professional Marketing at the headquarters address
above.
TM
TM
TM
TM
Raymond Cooper
John K. Snyder
TM
Preface
Contributors
Tribute Letters
TM
TM
Rajni Govindjee Center for Biophysics and Computational Biology and De-
partment of Cell and Structural Biology, University of Illinois at Urbana-Cham-
paign, Urbana, Illinois
TM
Harold V. Meyers Chemistry and Drug Discovery Group, New Chemical Enti-
ties, Inc., Framingham, Massachusetts
Chris A. Moore Department of Chemistry, San José State University, San José,
California
Ian F. Musgrave Prince Henry’s Institute for Medical Research, Clayton, Vic-
toria, Australia
TM
Roy K. Okuda Department of Chemistry, San José State University, San José,
California
* Current affiliation: Immunobiology Unit, Departments of Medicine and Microbiology, Boston Uni-
versity Medical School, Boston, Massachusetts.
TM
Dear Koji,
On the occasion of your 70th birthday, many of your old friends and colleagues
came to Columbia for a wonderful celebration in 1995. Now we are assembling
a permanent record of our appreciation for your friendship and as a tribute to
your many elegant and important contributions to the chemistry and biology of
natural products.
I can remember our first meeting, 34 years ago, in Tokyo, at the Presympo-
sium to the IUPAC meeting in Kyoto. In the next two weeks you were our host
almost every evening and introduced us to Japanese food and customs. It was a
great awakening to the realization that Japanese chemistry was rapidly gaining
tremendous momentum and a turning point in my career. Since that meeting I
have had the pleasure and privilege of working with 26 Japanese colleagues (sev-
eral of whom came from your lab).
It was a pleasure to repay a little of your hospitality when you visited our
homes in Sussex, New Haven and College Station, and you know that you and
your wife are always welcome in Texas.
You are a true pioneer in solving different problems at the chemistry–
biology interface using every possible technique on vanishingly small amounts
of material and your work continues to be an inspiration to all of us. Most impor-
tantly your personal qualities have ensured a permanent legacy in your many
students who have done so well in our profession. It must make you feel very
TM
As always!
Yours very sincerely,
A. I. Scott, F.R.S.
Davidson Professor of Science
Director of Center for Biological NMR
Texas A&M University, College Station, Texas
March 8, 1998
Dear Koji,
‘‘They’’ never stop celebrating you! ‘‘They,’’ of course, are those who have had
the privilege to obtain from you, as post-docs or as Ph.D. students, part of their
life baggage. They have been also kind enough to associate to them some of your
long time friends, and it was indeed a great pleasure to have the opportunity to
pay tribute to you in Columbia nearly half a century after we had first met in
the basement of Converse Laboratory, at Harvard, in Louis Fieser’s group.
When I was invited to contribute to this volume with a letter, I tried to call
back the oldest memories of our meetings I could muster. For some odd reason,
even though I am neither a gourmet nor a gourmand, they were nearly all memo-
ries of food. The experience of learning from you (and from Huang Wey Yuan)
to use chopsticks (a very useful lesson), the dinners of frog1 or lamb2 legs in
1
My wife was working at Harvard Medical School in Pharmacology with Fieser’s friend Prof. O.
Krayer, on the action of the Veratrum alkaloids on frog heart. A frog: one heart, two legs. We
always had a few dozen frozen legs aside for our friends. These legs, and frozen guinea pigs (one
heart, one guinea pig), helped us survive on our starvation scholarships.
2
On affluent months, for a change from frogs and guinea pigs, I was buying lamb by the half at the
TM
Italian market. The subway passengers had to sit next to a very un-American young man carrying
half a lamb protruding from his rucksack.
3
We were living in the small rooms above the stables built by a former U.S. Ambassador to Italy
and to Japan, in the middle of the huge estate he had bequeathed to the township of Brookline,
Lars Andersen Park. The stables were in the form of the Château de Chambord, or nearly so, and
were the seat of the Veteran Motor Cars Association of America, the ‘‘vie de château,’’ which we
shared with some 100 old cars.
TM
Munehiro Nakatani
Kagoshima University, Kagoshima, Japan
I. INTRODUCTION
TM
TM
TM
TM
TM
Scheme 1
TM
B. Bioassay of Antifeedants
The antifeedant assay results reported here were obtained with the Southern army
worm Spodoptera eridania (Boisduval) [20] as the test species. Spodoptera spe-
cies are distributed throughout the world and constitute a major agricultural
threat. The feeding bioassay was carried out by the conventional leaf disk method
[21], using 2-cm-diameter leaf disks cut from the Chinese cabbage Brassica
campestris L. var. chinensis (Cruciferae) with a cork-borer. Each disk was dipped
TM
Figure 5 Isolation scheme for the limonoids from the root bark of M. azedarach.
for 2 s in an acetone solution of the sample; 5 treated disks were arranged alter-
nately with another 5 control disks (immersed for 2 s in acetone alone), all
concentrically placed near the periphery in a Petri dish, as illustrated in Fig. 6.
Subsequently, 10 third-instar larvae were placed in the center of the dish, and
the treated and untreated leaves eaten by the larvae in 2-hour to 24-hour periods
were evaluated at appropriate intervals. The bioassay was terminated after the
TM
larvae had eaten approximately 50% of the control disks, which usually took 6–
12 h. When the average eaten area of the treated disks was visually judged to
be less than 50% of that of the control disks, the test compound was judged to
be active. This bioassay was used to guide the isolations to the active limonoids,
as well as to assess their relative antifeedant activities. To determine the minimum
inhibitory concentration, this choice test was done at 50, 100, 150, 200, 300, 400,
500, and 1000 ppm, with 50 ppm corresponding to a concentration of ca. 1 µg/
leaf-cm2.
1. Irradiation of the 13- and 8-Me peaks induced 30% and 12% Nuclear
Overhauser Effect (NOE) enhancements of the 9-H and 19-H signals,
respectively (see conformational drawing, Fig. 7).
2. The assignment of the 12-OH stereochemistry as β in trichilin A and
α in trichilin B (17) was deduced from the finding that in their 12-p-
bromobenzoates, the aromatic protons of the benzoate and furan rings
were at higher field in trichilin B. Thus, the shifts of the p-bromoben-
zoate protons were (for trichilins A/B) o-H δ 7.99/7.65, m-H δ 7.51/
TM
7.59, and those of the furan were 21-H δ 7.20/7.02, 22-H δ 6.36/5.98,
and 23-H δ 7.37/7.10. The higher-field chemical shifts of these aro-
matic protons in trichilin B can be accounted for by the mutual
shielding induced by the ring currents of the two aromatic rings, which
are located on the same side of the molecule. More directly, however,
the 12-OH configurations were independently derived from a new addi-
TM
Scheme 2
TM
Figure 8 1H NMR data for meliacarpinin A (46, cin ⫽ cinnamoyl) in CDCl3, 400 MHz,
in δ (multiplicity and J values), and NOE correlations (double-headed arrows). NMR,
nuclear magnetic resonance; NOE, nuclear overhauser effect.
TM
The limonoids isolated from M. azedarach and their effects on insect feeding
are summarized in Table 1. To develop a quantitative understanding of structure–
activity relationships among limonoids, the data are summarized primarily for
the activity against the third-instar larvae of a voracious pest insect, Spodoptera
eridania (Boisduval). It is important to note that differences in the response of
test insects when compared with different test species, or even different life cycle
stages of the same species, can mask any meaningful observations of structure–
activity relations. This is apparent from a comparison of the data reported against
different insects for compounds (11, 22, 39, and 40) in Table 1. Even where the
same bioassay species has been used, differences in the larval stage tested may
make comparisons invalid.
Some quantitative trends are apparent in the data in Table 1. Aside from
the highly oxidized C-seco meliacarpinins and azadirachtins, the most active
compounds appear to be intact apo-euphol limonoids with a 14,15-epoxide and
a C-19/C-29 acyl acetal bridged system. Structure–activity relationships in this
class will be discussed next.
Within the C-seco class, the highly oxidized meliacarpinins are the most
active of the limonoids from M. azedarach. Although their activities may be
weaker than that of the azadirachtins from M. azadirachta (effective dose of 50
ppm for 46–52 against S. eridania in comparison to 14 ppm for azadirachtin
against Epilachna varivestis), they are much more potent antifeedants than the
less oxidized class of C-seco limonoids exemplified by salannin (39), nimbolin
B (32), and ohchinolide B (37). The second most active class of limonoids from
M. azedarach is the C-19/C-29 bridged acetal class with the intact apo-euphol
skeleton (10–28) (effective dose 150–500 ppm). The two members of the intact
apo-euphol limonoid class without this C-19/C29 acetal bridge that were tested
against S. eridania, azadiron (1), and nimbolin A (9) showed little or no activity
(effective dose ⱖ 1000 ppm).
During the course of our structural and chemical correlation studies, about 50
compounds belonging to the C-19/C-29 bridged acetal class of limonoids were
made available. Antifeedant assays with S. eridania (leaf disk method) showed
several interesting structure–activity correlations.
First, activity is insensitive to substituent variation in the A-ring except for
the nature of the C29 bridging position: A hemiacetal bridge (hydroxyl group at
C29) increases the activity in comparison to that of an acylated hemiacetal (com-
TM
Effective References
concentration
Limonoid Test insect (ppm) Isolation Activity a
TM
TM
Effective References
concentration
Limonoid Test insect (ppm) Isolation Activity a
TM
TM
TM
TM
REFERENCES
TM
TM
TM
Isao Kubo
University of California, Berkeley, California
I. INTRODUCTION
Many human pathogenic microorganisms can now be controlled with the antibiot-
ics that are presently available. However, with the increase in drug resistance
and prevalence of opportunistic infections, there is still a great need for new,
more effective agents. For example, systemic infections caused by filamentous
fungi, especially in patients with impaired immune defense mechanisms, have
become an increasingly serious, worldwide problem. Although various new anti-
fungal agents have been introduced, control of most fungal diseases has not yet
been achieved [1]. Hence, in our continuous search for antimicrobial agents from
tropical plants, a new emphasis has been placed on antifungal agents, particularly
against Candida albicans, considered one of the most devastating fungi responsi-
ble for human opportunistic systemic infections [2].
Tropical plants are exposed throughout the entire year to predation by vari-
ous parasites such as bacteria, fungi, and insects. In order to survive these de-
manding conditions, efficient built-in chemical defense mechanisms have
evolved, and thus tropical plants offer a rich and intriguing source of secondary
metabolites possessing attractive biological activities with potential medicinal
applications. These plants remain a good source of new antifungal agents [3].
However, there is a need to investigate them from a point of view different from
that of the traditional approach of mere isolation and structure determination of
biologically active principles. Today, understanding the mechanism, as well as
TM
The genus Warburgia, endemic to East Africa, belongs to a small family, Canella-
ceae, and consists of only two species: W. ugandensis and W. stuhlmannii. We
first became interested in the tree W. ugandensis while I was on a tenured appoint-
ment at the International Centre of Insect Physiology and Ecology (ICIPE), lo-
cated in Nairobi, Kenya, because of its strong, hot taste. The methanolic extract
of the bark of W. ugendensis collected on the ICIPE grounds was originally found
to exhibit potent insect antifeedant activity against the African armyworm,
Spodoptera exempta [7]. Fractionation guided by the antifeedant assay led to
three active principles isolated from the bark, leaves, and fruit of this tree [8,9].
These active compounds, isolated after repeated chromatography and character-
ized by their unique hot taste, were the sesquiterpene dialdehydes [10] warburga-
nal (1) [11] and muzigadial (also known as canellal) (2) [12,13], in addition to
a known congener, polygodial (3). Most of the initial chemical study was per-
formed in Professor Koji Nakanishi’s labs together with his colleagues at Colum-
bia University in New York [11,12,14]. Subsequently, these antifeedant sesquiter-
pene dialdehydes received considerable attention from synthetic organic
TM
Since the Warburgia plants are widely used in folk medicine in East Africa
[19], their extracts were also submitted for testing in various available pharmaco-
logical assays together with other plant extracts. Among the extracts tested, the
two from the Warburgia species were found by Professor Makoto Taniguchi of
Osaka City University to exhibit a broad-spectrum antimicrobial activity, includ-
ing activity against Candida utilis [5]. This finding initiated an investigation to
determine the antifungal principles, which was achieved in collaboration with
Professor Taniguchi and his coworkers [20–24].
TM
TM
MIC (µg/mL)
Microorganisms tested 1 2 3 4 6
TM
Scheme 2
TM
TM
(b)
Figure 1 (a) Section of untreated control cell of Saccharomyces cerevisiae. (b) Section
of S. cerevisiae cell treated with 50 µg/mL of polygodial (3) for 10 min. CW, cell wall;
PM, plasma membrane (cell membrane); N, nucleus; M, mitochondrion; V, vacuole.
TM
dant Mg 2⫹, such as its larger radius, its lower energy of hydration, and the pres-
ence of d-orbitals, allowing it to chelate more readily. In addition, both 3 and 1
have been reported to exhibit the membrane leakage activity in human neuro-
blastoma cells [43]. The binding site of 3 on the cell membrane and the mecha-
nism of the suppressive effects of Ca 2⫹ on the polygodial-induced leakage remain
to be established.
TM
In our continuing search for antimicrobial agents from tropical plants, several
phenylpropanoids, such as anethole (16), isolated from the seeds of Pimpinella
anisum (Umbelliferae) [44], and eugenol (17), methyleugenol (18), and safrole
(19), in addition to anethole from the seeds of Licaria puchuri-major (Lauraceae)
[45], were isolated as antimicrobial principles in rather large quantities. All exhib-
ited moderate but broad-spectrum activity. Their MICs, ranging from 100 to 800
µg/mL, were not potent enough to warrant further studies on the individual com-
pounds. However, it was still considered worthwhile to investigate the possibility
of their use as antifungal agents in combination with other antimicrobial natural
products, since they were all isolated from various food spices that have long
been consumed by people from many different cultures. Hence, they were first
examined in combination with 3 in an attempt to enhance their antifungal activity
against several fungi such as C. albicans and S. cerevisiae as well as the dermato-
mycotic fungus Pityrosporon ovale. Unexpectedly, 3 did not synergize the anti-
fungal activity of any of these phenylpropanoids, though its antifungal activity
was significantly increased when combined with one of the phenylpropanoid
compounds. Notably, a dramatic increase in the antifungal activity of 3 occurred
when it was combined with a sublethal amount (1/2MIC) of 16: the activity of
3 against C. albicans and S. cerevisiae was increased 32- and 64-fold, respec-
tively. Thus, the MIC of 3 against C. albicans was lowered from 3.13 to 0.098
µg/mL, and against S. cerevisiae, from 1.56 to 0.024 µg/mL, when 3 was com-
bined with 100 µg/mL of 16 (1/2MIC for both C. albicans and S. cerevisiae)
[44].
These combination effects, based on the MICs obtained after 48-h incubations,
do not fully characterize the antifungal activity. For example, it was not clear
whether the combination of 3 and 16 was fungicidal or fungistatic. Hence, the
viable count method [46] to analyze the growth curve of C. albicans was used
TM
Figure 3 Growth curves of Candida albicans in the presence of anethole (16), polygod-
ial (3), and their combination.
TM
MIC (µg/mL)
Candida Saccharomyces Pityrosporon
Compounds combined albicans cerevisiae ovale
TM
other compounds therefore depends on the species of fungi being tested and the
antifungal agents being combined. Although accumulation of this kind of knowl-
edge may provide new insight into the molecular basis of fungicidal activity, the
mechanism of the combination treatment remains to be established.
TM
It should also be noted that the activity of 3 is enhanced under acidic condi-
tions [22]. It is known that yeast cells are able to maintain a normal internal pH
when suspended in an acidic medium with relatively little change in the intracel-
lular pH. The acidic conditions appear to stimulate the plasma membrane H⫹-
ATPase activity and ‘‘excess’’ protons are pumped out to the external medium,
maintaining constant internal pH during growth [48]. As a result of the inhibition
of the plasma membrane H⫹-ATPase by 3, the intracellular pH may drop into
the range where phosphofructokinase is sensitive [51]. The subsequent inhibition
of glycolysis caused by this inactivation of phosphofructokinase results in a drop
in adenosine triphosphate (ATP) levels and thus restricts growth [49]. This ratio-
nale may explain why 3 is more potent in the acidic conditions.
In contrast to the potent antifungal activity of 3, its congener 4 did not
exhibit any activity up to 800 µg/mL; 1 was active, though to a lesser extent
than 3 [33]. Thus, the activity decreased for each additional hydroxyl group
‘‘added’’ to the molecular framework of 3. The fungicidal activity of 3 was ex-
plained as the result of the structural disruption of the cell membrane [21,52,53],
as illustrated in Fig. 1. Moreover, in previous work using human neuroblastoma
cells [43], the increase in membrane permeability was demonstrated to depend
TM
TM
VIII. CONCLUSIONS
The data so far obtained indicate that polygodial (3) initially acts as a nonionic
detergent. More importantly, 3 inhibits the plasma-membrane H ⫹-ATPase by dis-
rupting and disorganizing the hydrogen bonds at the lipid bilayer–protein inter-
face. It seems that 3 targets the extracytoplasmic region and thus does not need
to enter the cell, thereby avoiding most cellular pump–based resistance mecha-
nisms.
For the otherwise healthy person, fungal infections are more of a nuisance
than health-threatening and are normally kept in check by a strong immune sys-
tem and by otherwise innocuous bacteria of the throat and gut. However, when
outside forces such as cancer chemotherapy or heavy doses of antibiotics disrupt
the body’s natural defenses, fungal populations can sharply increase and cause
serious health problems. Synergistic substances that enable physicians to use
lower, safer antifungal dosages would be a useful addition to the therapeutic
TM
ACKNOWLEDGMENT
REFERENCES
TM
TM
TM
I. INTRODUCTION
TM
II. BACKGROUND
A. The Enzymes
The first haloperoxidase to be characterized was named chloroperoxidase and
was identified from the fungus Caldariomyces fumago [10]. This enzyme con-
tains heme-bound iron as the cofactor and has been extensively investigated [11].
Several bacteria are reported to contain a nonheme chloroperoxidase, which is
involved in the chlorination step of chlorinated metabolites [4]. Except for a small
number of cases of enzymes from a marine sponge [12] and a marine worm
[13], all chloroperoxidases reported to date have been isolated from nonmarine
fungi [6].
Bromoperoxidases are so named because of their requirement for hydrogen
peroxide and bromide (or iodide) for catalytic activity. If an organic substrate
containing a functional group that is susceptible to electrophilic attack (e.g. al-
kene, alkyne, aromatic, β-diketone) is also present, an electrophilic substitution
or addition occurs. In the mechanistic sense, the bromide is oxidized to a bromon-
ium ion equivalent and reacts as an electrophilic species. In some cases, the bro-
mide becomes incorporated into the substrate; in others, it may be part of a tran-
sient intermediate and may not be found in the product [6].
In addition to being classified according to the halides used by an enzyme,
haloperoxidases are also categorized by the metal cofactor present. Two general
classes are recognized—those that contain heme (the ‘‘H’’ haloperoxidases, such
as chloroperoxidase), and the nonheme type (the ‘‘NH’’ enzymes, which include
TM
TM
TM
C. Mechanism of Action
Several articles have reviewed the mechanism of action of marine bromoperoxi-
dases [5,30], and thus only a brief discussion will be given here.
Heme bromoperoxidases react via oxidation of the heme iron by hydrogen
peroxide to yield an activated species, called Compound 1, which then oxidizes
the halide to a halonium ion equivalent. It is unclear whether the halonium ion
is enzyme bound or exists as a hypohalous acid (Fig. 2) [6]. Nonheme vanadium
bromoperoxidases from marine algae are reported to yield an ‘‘enzyme-trapped’’
bromonium species; if an organic substrate is present, it is bound to the enzyme,
whereupon reaction occurs [31].
Bromoperoxidases react primarily with substrates that are susceptible to
electrophilic attack. In biosynthesis, the bromoperoxidases obviously exert their
halogenating capabilities in a regio- and stereoselective fashion, which gives rise
to many natural products containing chiral halogens. Since chemical routes to
chiral halogen moieties starting from achiral substrates are still rather circuitous,
if bromoperoxidases can be induced to perform enantioselective halogenation in
a chemoenzymatic mode, these enzymes could be extremely useful for synthetic
applications. To date, enantioselective halogenation has not been achieved by
using bromoperoxidases. However, work in this area is still in its preliminary
development, and the enzymes may require selection of proper experimental con-
ditions before progress is made.
TM
TM
1. pH
Nearly all marine bromoperoxidases have a pH optimum for reactivity that falls
in the range of 5.0 and 7.0. This is in marked contrast to fungal chloroperoxidase,
which has a pH optimum near pH 3.3 [10]. The pH optimum for most reported
marine bromoperoxidases is a fairly sharp peak, and generally activity drops to
less than 50% of the maximum if the solution is ⫾2 pH units from the optimal
value [9]. Fig. 4 shows examples of pH optima for some marine algal bromoper-
oxidases.
TM
TM
3. Bromide Concentration
Optimal bromide concentrations for bromoperoxidase activity vary rather widely
among different sources of the enzyme. In some cases, the enzyme activ-
ity diminishes significantly upon addition of additional bromide. Interestingly,
some bromoperoxidases are still active at very high levels of bromide (e.g., 1M
NaBr) [9].
TM
TM
Cinnamyl C. pilulifera 38
alcohol [α] D ⫽ 0
Cinnamic C. pilulifera 38
[α] D ⫽ 0
acid
Cyclohexene C. pilulifera 38
[α]D ⫽ 0
Styrene C. pilulifera 38
[α]D ⫽ 0
cis-Propenyl C. pilulifera 38
[α]D ⫽ 0
phosphonic
acid
Geraniol Corallina 35
vancouveriensis
X ⫽ Br, Y ⫽ OH
X ⫽ OH, Y ⫽ Br
TM
Linalool Corallina 35
X ⫽ Y ⫽ Br vancouveriensis
X ⫽ OH, Y ⫽ B
X ⫽ BR, Y ⫽ OH
α-Pinene a Corallina 35
vancouveriensis
a
See Fig. 7.
TM
TM
3-oxo-octanoic acid
R ⫽ CH 2 Br a
CHBr2
CBr3
solutions, the enol form, which has an ⑀ value of 20,000/M-cm at 290 nm, pre-
dominates. If a haloperoxidase is added to a solution of MCD, hydrogen peroxide,
and bromide, the dihalo product is formed, which has an ⑀ of 200/M-cm. One
unit (U) of bromoperoxidase activity is defined as the amount of enzyme neces-
sary to convert one micromole of MCD to dihalo product at 25°C in 1 min [10].
The role of bromoperoxidases in the biosynthesis of brominated carbonyl
compounds was investigated by Theiler et al. [7]. When reacted with 3-oxoocta-
noic acid, the enzyme from Bonnemaissonia hamifera yielded a mixture of dibro-
momethane, bromoform, and 1-bromopentane. In subsequent studies, the bro-
moperoxidase from Penicillus capitatus produced brominated heptanones and
bromoform from the same substrate, which were similar or identical to com-
pounds found naturally in B. hamifera. The authors suggest that the difference
in product formation may be due to either distinctive catalytic abilities of each
or two simultaneous reaction pathways [39].
The ready formation of bromoform by action of bromoperoxidases has led
to the suggestion that these enzymes, through their natural catalytic activity with
simple carbonyl compounds, are responsible for a significant percentage of atmo-
spheric halomethanes, which are contributing factors to the degradation of the
earth’s ozone layer. Indeed, bromoperoxidases from marine algae are suggested
to be ultimately responsible for the formation of significant quantities of volatile
TM
4. Heterocycles (Table 5)
A variety of nitrogen heterocycles and one sulfur heterocycle have yielded prod-
ucts with bromoperoxidases. Itoh et al. have compared the relative halogenating
abilities of fungal chloroperoxidase and the bromoperoxidase from the alga
Corallina pilulifera and report some significant differences between them. Using
these enzymes, they were able to produce halogenated derivatives of nucleoside
and nucleotide bases, which are known to have activity as potential anticancer
agents. In this study, the reactions of the same substrates with fungal chloroperox-
idase were compared. With the exception of the chlorinated products, most of
the brominated and iodinated nucleic acid base products were similar. However,
the report indicates several practical differences that favor the use of
bromoperoxidase over the fungal enzyme [44]. Barbituric acid and derivatives
were brominated by the enzyme from Ascophyllum nodosum to only mono- or
dibromo products [45].
TM
(91% yield)
Thymol C. vancouveriensis 33
56 : 31
1-Methoxynaphthalene C. pilulifera 38
TM
Thiophene C. pilulifera 38
Pyrazole C. pilulifera 44
R ⫽ Br, I
Indoles ‘‘V-bromoperoxidase’’ 31
R ⫽ Me, Phenyl (from marine alga)
R1 ⫽ R2 ⫽ R3 ⫽ H R1 ⫽ R2 ⫽ H
Nucleic acid bases and nucleosides
Cytosine 5-Bromocytosine C. pilulifera 44
Uracil 5-Bromouracil, 5-iodouracil
Cytidine 5-Bromocytidine
Thymine (Decomposes)
(No reaction with adenine, adenosine, guanine, and 2-deoxyuridine)
Amino acid
P. capitatus 47
R ⫽ CH 2-CN
CH2-CHO
TM
IV. CONCLUSIONS
TM
TM
ACKNOWLEDGMENTS
The authors wish to gratefully acknowledge the financial support of the National
Institutes of Health (GM-38040), the CSU Awards for Research, and the Tropical
Technology Center (Okinawa, Japan).
TM
TM
TM
I. INTRODUCTION
The use of medicinal plants for curing illnesses can be traced back over five
millennia to written documents of the early civilizations in China, India, and the
Near East, but it is doubtless an art as old as mankind. Even today, plants are a
major and significant source of drugs for the majority of the world’s population.
Although in industrialized countries medicinal plant research has faced cyclic
fortune during the last few decades [1], nonetheless, substances derived from
higher plants still constitute ca. 25% of prescribed medicines [2,3].
The plant kingdom represents an extraordinary reservoir of novel mole-
cules. Of the estimated 400,000–500,000 plant species around the globe, only a
small percentage has been investigated phytochemically and the fraction submit-
ted to biological or pharmacological screening is even lower. Since plants may
contain hundreds, or even thousands, of metabolites, there is currently a resur-
gence of interest in the vegetable kingdom as a possible source of new lead
compounds for introduction into therapeutic screening programs. Furthermore,
the rapid disappearance of tropical forests and other important areas of vegetation
has added renewed pressure for the rapid isolation and identification of bioactive
natural products.
In general, the approach adopted to obtain an exploitable pure plant constit-
TM
Figure 1 Procedure for obtaining active principles from plants and use of LC hyphen-
ated techniques as strategic analytical screening tools during the isolation process of a
plant extract. LC, liquid chromatography.
TM
TM
the same analysis the detection of all the secondary metabolites encountered in
a plant extract; each method has its own selectivity.
TM
Figure 3 Schematic representation of the experimental setup used for the postcolumn
addition of UV shift reagents. UV, ultraviolet.
TM
TM
TM
4). Four ionization modes exist with this interface. Besides chemical ionization
initiated by conventional electron bombardment using a heated filament (‘‘fila-
ment-on’’ mode) or a discharge electrode (‘‘discharge-on’’ mode), a third ioniza-
tion method, in which ammonium acetate or some other volatile buffer in the
mobile phase plays a major role (‘‘thermospray buffer’’ ionization) is available;
the fourth ionization mode is based on an ion evaporation mechanism. This latter
mechanism may be operative in TSP buffer ionization mode as well [30].
The chemical process of ionization plays an important role and should al-
ways be taken into account when analyzing samples with LC/TSP-MS. Usually,
as for CI, the values of the proton affinity of the analytes, the solvent, and the
buffer should be considered. In our experience, for the analysis of plant metabo-
lites (150–1000 µ, aglycones, mono- and diglycosides), the TSP operated mainly
in the positive-ion mode in ‘‘thermospray buffer’’ ionization generally afforded
MS spectra similar to those produced by (desorption chemical ionization (DCI)
with NH 3 as reagent gas, positive-ion mode).
In order to perform satisfactory LC/TSP-MS investigations of plant metab-
olites, or any other sample, different parameters should be optimized. Many inter-
dependent experimental parameters are important in this respect: the mobile
phase composition (amount and type of buffer and organic modifier), the solvent
flow rate, the temperature of the vaporizer and the ion source, together with the
geometry, the position, and the potential of the repeller electrode. The tuning of
TM
TM
Figure 5 Schematic representation of the experimental setup used for LC/UV/MS (1)
and LC/UV/NMR (2) analyses. LC, liquid chromatography; UV, ultraviolet; MS, mass
spectroscopy; NMR, nuclear magnetic resonance.
TM
When searching plant extracts for compounds with interesting properties, a multi-
dimensional approach to their chromatographic analysis is of great significance.
By combining HPLC with on-line UV, MS, and NMR, a large amount of prelimi-
nary information can be obtained about the constituents of an extract before isola-
tion of the compounds of interest. In the case of polyphenols, for example, UV
spectra recorded on-line give useful information (type of chromophore or pattern
of substitution) complementary to that obtained with LC/MS and provide an
initial assignment of the peak of interest at an early stage. When this information
is not sufficient for a precise structural assignment, LC/NMR provides a useful
complement for a full structural identification on-line. To illustrate this approach,
examples of on-line LC/UV/MS, LC/MS/MS, and LC/NMR analyses of xan-
thones, flavones, and secoiridoids in crude extracts of plant species belonging to
the Gentianaceae family will be discussed. Gentianaceae plants have indeed been
extensively screened in our laboratory, especially in the search for novel antide-
pressive agents.
TM
TM
TM
lecular [M ⫹ H] ⫹ ion at m/z 259. The molecular weight of this xanthone was
thus 258 µ. As the mass of a bare xanthone nucleus is 196 µ, this indicated that
the polyphenol was substituted by two hydroxyls and one methoxyl group. A
comparison of the on-line UV spectra with those of trisubstituted xanthones from
an in-house UV database confirmed that this compound was indeed gentisin, the
main xanthone of the yellow gentian [52]. This example shows that for simple
known compounds like xanthones, a precise identification can be performed on-
line with the help of LC/UV and LC/MS, provided some information about the
TM
TM
TM
TSP spectra of 9, Fig. 8, or 4 and 8, Fig. 9). After subsequent isolation of these
glycosides, the disaccharide was shown to be primeverose, a β-d-xylopyranosyl-
(1 → 6)-β-d-glucopyranoside disaccharide unit, often encountered in the Gen-
tianaceae family [54].
Since MS detection allows the selective recording of each ion trace, it was
possible to reconstruct the chromatogram of the ion m/z 657 and to obtain a
precise assignment of the peak corresponding to compound 9 in the extract. Simi-
larly, the specific ion trace at m/z 363 gave responses for the HPLC peaks of 9
TM
TM
D. Liquid Chromatography/Thermospray–Mass
Spectroscopy and Liquid Chromatography/Continuous
Flow–Fast-Atom Bombardment–Mass Spectroscopy of
Gentiana rhodantha
Among the other Gentianaceae species screened by LC/UV/MS, Gentiana rho-
dantha from China presented a completely different extract composition profile
from that found for C. krebsii. The LC/UV analysis of G. rhodantha showed
the presence of only one predominant xanthone 17 and different secoiridoids.
Compound 17 was easily identified as the widespread xanthone C-glycoside man-
giferin (see structure in Fig. 7). The TSP-MS spectrum of 17 exhibited a pseudo-
molecular ion at m/z 423, characteristic fragments for C-glycosides at [M ⫹ H-
90] ⫹ and [M ⫹ H-120] ⫹, and a weak aglycone ion at m/z 261 (xanthone with
four OH groups). The computer fit for the UV spectrum of 17 with our in-house
UV spectral library allowed the definitive characterization of 17.
With the help of the LC/TSP-MS spectra recorded on-line, chemotaxonom-
ical considerations, and comparison with pure standards, the secoiridoids with
retention times less than 10 min were also easily identified [58]; among them a
very minor secoiridoid, sweroside (3)(MW 358), was found to be present. The
slower eluting peaks 18 and 19 also exhibited the same characteristic UV spectra
of secoiridoids (one band around 240 nm, Fig. 10). These compounds, which
were less polar than common secoiridoids, were studied in more detail. The LC/
TSP-MS analysis of 18 and 19 in each case gave a spectrum identical to that
TM
TM
obtained for 3; all exhibited an intense ion at 359 u and no ion at higher mass
(Fig. 10). However, the chromatographic behavior was quite different for 18, 19,
and 3. In order to obtain complementary information about these constituents, a
second LC/MS analysis with CF-FAB was undertaken, using the same HPLC
conditions. The total ion current recorded for the whole chromatogram showed
a very important MS response for compounds 18 and 19, while the more polar
metabolites were only weakly ionized (Fig. 10). The CF-FAB spectrum of 18
recorded on-line in negative-ion mode exhibited a very intense pseudomolecular
TM
E. Liquid Chromatography/Thermospray–Mass
Spectroscopy and Liquid Chromatography/
Electrospray–Mass Spectroscopy of
Halenia corniculata
As shown in the previous example, the use of different complementary MS ion-
ization techniques is often important for unambiguous on-line determination of
molecular weights. The analysis of another of the Gentianaceae species collected
in Mongolia, Halenia corniculata, by TSP and LC/ES-MS also demonstrated
the importance of an LC/MS analysis using two independent ionization methods
for screening unknowns.
TM
TM
TM
TM
TM
TM
TM
TM
Figure 16 Summary of all the spectroscopic data obtained on-line for naphthoquinone
35 in the CH 2 Cl 2 extract of Swertia calycina.
TM
TM
TM
IV. CONCLUSIONS
TM
TM
ACKNOWLEDGMENT
REFERENCES
TM
TM
TM
Takenori Kusumi
Tokushima University, Tokushima, Japan
Ikuko I. Ohtani
University of the Ryukyus, Okinawa, Japan
I. INTRODUCTION
Most biologically active natural products have stereogenic centers and are ob-
tained from their natural sources enantiomerically pure. In pharmaceutical drug
development of naturally occurring and synthetic compounds, identification of
the absolute configuration of the drug is crucial. In many cases, when a drug is
chiral, only one enantiomer is effective, and sometimes the other enantiomer
may be seriously hazardous [1]. Because of the continuous development of new
spectroscopic techniques, unambiguous structure determination of natural and
synthetic products has become relatively routine even in cases in which the mo-
lecular weight of a compound is greater than 1000 daltons.
Few methods exist, however, to elucidate the absolute configuration of or-
ganic compounds; the most commonly employed nonchemical procedures are x-
ray crystallography and the exciton chirality method [2]. The latter circular di-
chroic methodology has been most prominently developed by Nakanishi [3] and
Harada [4] (see Chapter 5) over the past decade.
We have been developing methodology to elucidate the absolute configu-
ration of organic compounds using nuclear magnetic resonance (NMR) spectros-
TM
Figure 1 Conformational model proposed by Mosher for MTPA esters. MTPA, 1-me-
thoxy-1-phenyl-1-trifluoromethylacetic acid. (Source : Ref. 6.)
TM
this concept in a qualitative sense. As Mosher pointed out, when the MTPA ester
exists in this conformation, resonances of ligand L 3 in the (S)-MTPA ester are
upfield in the 1 H NMR spectrum relative to those of the (R)-MTPA ester as a
result of the diamagnetic shielding effect of the benzene ring (Fig. 1). The reverse
sense of nonequivalence is true for the 1 H resonances of ligand L 2, with those
of the (R)-MTPA ester being upfield relative to those of the (S)-MTPA ester.
Therefore, comparison of the chemical shifts of the proton resonances of L 2 and
L 3 of the diastereomeric esters allows assignment of the absolute configuration
of the secondary alcohol.
When Mosher first put forward his analysis in the early 1970s however,
the NMR instruments available were mostly 60–100 MHz in proton frequency,
and the resolution and assignment of protons of complex organic molecules were
practically impossible with such limited field strengths (14.1–23.6 kG). Conse-
quently, the use of 19 F NMR [6,14] or a lanthanide-shift reagent to resolve the
easily identified OMe singlets of the MTPA-esters [15] was often preferred to the
original 1 H NMR method, which required identification of the usually overlapped
protons bonded to the original alcohol (on L 2 and L 3). Since Mosher’s conforma-
tional analysis did not address the relative shielding of these resonances, unam-
biguous assignment of absolute stereochemistry was difficult (if not impossible!),
and early applications of the MTPA ester method were primarily limited to deter-
mination of diastereomeric excess (enantiomeric excess of original alcohol) of
enriched synthetic mixtures.
The intrinsic drawback in these early applications of Mosher’s method for
absolute stereochemistry assignment was that it usually depended on only two
data points—the chemical shift difference of the two CF 3’s (19 F) or OMe’s (1 H)
of the (R)- and (S)-MTPA esters—and not on the more difficulty assigned reso-
nances of the substrate alcohol. The modified Mosher’s method, on the contrary,
uses the chemical shifts of many protons as Mosher originally intended and thus
is more reliable.
TM
protons on the left side of the plane should have negative values (∆δ ⬍ 0, Fig.
2B).
Mosher’s original method can be extended as follows: (1) assign as many
proton signals as possible of the (R)- and (S)-MTPA esters; (2) obtain ∆δ values
for the protons; (3) place the protons with positive ∆δ on the right side, and those
with negative ∆δ on the left side of Model A (Fig. 2b); (4) construct a molecular
model of the compound in question and confirm that all the assigned protons
with positive and negative ∆δ values are actually found on the right and left sides
of the MTPA plane, respectively. The absolute values of ∆δ should be propor-
tional to the distance from the MTPA moiety. When these conditions are all
satisfied, Model A will indicate the correct absolute configuration of the com-
pound.
This modified Mosher’s method allows examination of chemical shift dif-
ferences of as many protons as may be assigned by means of up-to-date NMR
techniques including two-dimensional (2D) spectra such as H,H-Correlation
Spectroscopy (H,H-COSY) and Homonuclear Hartmann-Hahn spectroscopy
(HOHAHA) spectra. We feel that as this method relies on a larger number of
TM
data points (i.e., all the protons assignable by 2D NMR techniques), it is far more
reliable than the previously mentioned variations of Mosher’s method using 19 F
NMR or a lanthanide shift reagent.
1. Preparation of 1-Methoxy-1-Phenyl-1-Trifluoromethylacetic
Acid Esters
The MTPA esters of a secondary alcohol are easily prepared [5] by treatment of
the alcohol (1) with (R)- or (S)-MTPA acid and 1-ethyl-3-(3-dimethylaminopro-
pyl)-carbodiimide (EDC) and 4-dimethylaminopyridine (DMAP) in dichloro-
methane, or (2) with (R)- or (S)-MTPA chloride and DMAP/Et 3N in dichloro-
methane or pyridine. Usually the reaction is completed within 15 h at room
temperature. In a few cases, the reaction rate of (R)- and (S)-MTPA acids or
acid chlorides with the chiral alcohol differs significantly (kinetic resolution),
and either diastereomer may be slow, or even difficult to obtain.
Note the following: because of the nomenclature priority rules, the chloride
of (R)-MTPA acid is (S)-MTPA chloride and the chloride of (S)-MTPA acid is
(R)-MTPA chloride. The stereochemical correlation of MTPA acids and their
chlorides is correctly shown in the scheme below.
When the amount of sample is limited, less than 1 mg of the sample per
esterification is usually sufficient for assignment of the absolute configuration
thanks to the high sensitivity of superconductive NMR instruments. One advan-
tage of MTPA esters is that the MTPA moiety possesses ultraviolet (UV) ab-
sorbing phenyl chromophore. Combinatory use of preparative thin layer chroma-
tography (TLC) (fluorescent under UV 254 nm) and HPLC successfully achieves
purification of the MTPA esters.
TM
Figure 3 ∆δ Values (ppm) recorded for the MTPA esters of (⫺)-menthol (1). MTPA,
1-methoxy-1-phenyl-1-trifluoromethylacetic acid. (Source : Ref. 5.)
TM
Figure 4 ∆δ Values (ppm) recorded for the MTPA esters of marine terpenoids 2–6.
MTPA, 1-methoxy-1-phenyl-1-trifluoromethylacetic acid. (Source: Ref. 5.)
uration of compounds that are not secondary alcohols yet possess functionality
routinely converted into one with high stereoselectivity are also amenable to this
method. For example, reduction of the aragupetrosine A (59) ketone group to
the secondary alcohol with cyanoborohydride allowed absolute stereochemistry
assignment by the modified Mosher’s method [23].
TM
Figure 5 Examples of natural and synthetic secondary alcohols whose absolute config-
urations were determined by the modified Mosher’s method.
TM
TM
TM
TM
TM
TM
TM
TM
Figure 7 Auto oxidation product 69 of isoclavukerin A (68) and ∆δ values (ppm) re-
corded for the MTPA esters of 69. MTPA, 1-methoxy-1-phenyl-1-trifluoromethylacetic
acid. (Source : Ref. 73.)
TM
chirality method is not applicable, the modified Mosher’s method may be suc-
cessfully applied.
Figure 8 The ∆δ values (ppm) recorded for the MTPA esters of sipholenol-A (70).
MTPA, 1-methoxy-1-phenyl-1-trifluoromethylacetic acid. (Source: Ref. 78.)
TM
Figure 9 The ∆δ values (ppm) recorded for the MTPA esters of friedelan-3β-ol (72)
with the axially oriented C3 alcohol, and those of the MTPA esters of its C3 epimer 73.
MTPA, 1-methoxy-1-phenyl-1-trifluoromethylacetic acid. (Source : Ref. 5.)
TM
Figure 10 Synthetic compounds 74a′ and 74b′ with irregular ∆δ values for one of the
α-protons (shown in boldface) and other model compounds 75–80. (Source : Ref. 81.)
TM
H B). Although these methylene protons should have positive ∆δ values, in many
compounds they were negative or zero. The conformations of these esters de-
duced from their NMR properties (Fig. 11) were supported by molecular model-
ing studies of 74b′ (Fig. 12). Since the α-protons are located near the β-aromatic
ring, these were anisotropically affected by both this β-aromatic ring as well as
the benzene ring of the MTPA moiety. Even though a difference of the dihedral
angles (H c-C-C-C1) in the predicted most stable conformations for 74b′ was not
observed between the (S)- and (R)-MTPA esters (352° and 351°, respectively)
by molecular mechanics calculation, subtle differences in the conformation
around the β-carbon between (S)- and (R)-MTPA esters might cause these irregu-
larities of the ∆δ values.
A similar anomaly was observed in tanabalin [81], a natural product iso-
lated from the dried flowers of the Brazilian medicinal plant Tanacetum bal-
samita, which has a secondary alcohol adjacent to a furan ring [82]. One of the
C-11 methylene protons showed an anomalous positive ∆δ value thought to be
caused by the furan ring anisotropy. The anomaly was ‘‘verified’’ by using model
TM
Figure 13 The ∆δ values (ppm) obtained for the MTPA esters of tanabalin 81a and
model compound 82a. MTPA, 1-methoxy-1-phenyl-1-trifluoromethylacetic acid. (Source:
Ref. 82.)
compound 82, which also showed the irregular ∆δ value of the α-proton denoted
by a thick arrow (Fig. 13).
Minale et al. [40] encountered similar anomalies in the structure determina-
tion of superstolide A that can be interpreted in the same manner: an irregular
∆δ value (denoted by the arrow) in both 28a and 28b (Fig. 14). This anomaly
was apparently caused by the N-acetyl group anisotropy, the conformation of
which may be slightly different between the (R)- and (S)-MTPA diastereomers,
affecting the chemical shifts of the protons close to the carbonyl group. Since
other ∆δ values were systematically arranged, their absolute configurations were
safely assigned.
In conclusion, the modified Mosher’s method can be applied to secondary
alcohols possessing other π-systems, such as the aromatic rings and carbonyl
groups that can fully or partially rotate along a single bond, in the vicinity of
the hydroxy group. In such cases, however, one must be aware of the possible
impact of this second anisotropy in generating irregular ∆δ values.
TM
TM
Figure 15 Strategy for the determination of the absolute configuration of synthetic prod-
ucts obtained from enantioselective synthesis using the modified Mosher’s method.
TM
TM
Figure 17 The ∆δ values (ppm) recorded for amino compounds 87–97. Proton assign-
ments that could be interchanged lead to two possible values, as indicated by the second
value in parentheses. Values for the methoxyl group of the MTPA subunit (all positive)
are also given. MTPA, 1-methoxy-1-phenyl-1-trifluoroacetic acid. (Source: Ref. 84.)
TM
examples demonstrate, a large number of research groups have now found this
procedure to be tremendously useful for the assignment of absolute stereochemis-
try for primary amines as well as secondary alcohols.
TM
assignment, new chiral anisotropic reagents may be developed for the compounds
possessing functional groups other than a secondary alcohol. In this context, we
have elaborated new chiral anisotropic reagents, which allow determination of
the absolute configuration of carboxylic acids, in which the carboxylic group is
attached to a methine carbon [88].
The general molecular design for such chiral anisotropic reagents is shown
in Fig. 19. A chiral carboxylic acid of type I is converted to acyl derivative II,
where X can be any heteroatom. In order for the phenyl ring to cast its diamag-
netic field effectively, and selectively, onto R 1 or R 2, coplanarity of C1 to Y5 in
the dominant conformation is necessary. When X is NH, coplanarity from C1 to
C4 may be achieved because of the planar structure as well as the preferred s-
trans conformation of the amide group. If Y is also a polar group that would
interact repulsively with the polar carbonyl group at C2 in a syn relationship,
the C2 through Y5 orientation about the X3-C4 bond may also adopt a trans
conformation. We chose the dimethylaminocarbonyl and methoxycarbonyl
groups for Y, the former being preferable because of the enhanced dipole caused
by the greater participation of the lone-pair electrons of the nitrogen atom with
carbonyl resonance. Furthermore, hydrogen bonding between N 3H and C 5 ⫽ O,
TM
TM
were nearly coplanar, and, because the phenyl plane faces the methylcyclohexane
residue, the anisotropy of the phenyl group was effectively cast over the methyl-
cyclohexyl protons. The ∆δ values for this amide were calculated (Fig. 20): they
are positive and negative on the right and left sides of the PGDA plane, respec-
tively, as shown in 101a. Analogous results were also obtained with the PGDA
derivative of (S)-2-methylbutanoic acid 102a. The absolute configuration de-
Figure 21 The ∆δ values (ppm) and NOEs supporting the proposed conformation of
the PGME derivative of 4-methylcyclohexanecarboxylic acid 101c and the ∆δ values of
the PGME derivative of (S)-2-methylbutanoic acid 102c. NOE [nuclear Overhauser ef-
fect], PGME, phenylglycine methyl ester. (Source : Ref. 88.)
TM
Figure 22 The ∆δ values (ppm) recorded for PGME derivatives 103–106. PGME, phe-
nylglycine methyl ester. (Source: Ref. 88.)
TM
amide 102c are smaller than those for the PGDA amide 102a as expected, they
are still of practical magnitude for absolute stereochemistry assignment. Consid-
ering the ease of the preparation, the authors prefer PGME to PGDA, and thus
for further derivatizations, PGME was subsequently applied to several carboxylic
acids with known absolute configurations [88]. The results are summarized in
103–106 (Fig. 22).
In all cases, the protons of ∆δ ⬎ 0 and ∆δ ⬍ 0 are on the right and left
sides of the PGME plane, respectively, without exception, confirming the validity
of the method. The absolute configurations predicted by the present method are
identical with the known ones. It should be noted that PGME was applicable to
rather complex compounds such as gibberellic acid (106). It was also found that
the magnitude of the absolute values of ∆δ was proportional to the distance from
the PGME moiety.
The present method should be appropriate for absolute stereochemical as-
signment of natural products possessing a carboxyl group of type (I) as well as for
primary alcohols 107 that can be routinely oxidized into carboxylic acid 108 (Eq. 1).
IV. CONCLUSIONS
REFERENCES
1. (a) D. Parker, Chem. Rev., 91: 1441 (1991). (b) S. Ahuja, in Chiral Separations,
Applications and Technology (S. Ahuja, ed.), American Chemical Society, Washing-
ton, D. C., chapters 1 and 6 (1997).
TM
TM
25. H. Itokawa, E. Kishi, H. Morita, K. Takeya, and Y. Iitaka, Tetrahedron Lett., 32:
1803 (1991).
26. M. Natsume, K. Yasui, S. Kondo, and S. Marumo, Tetrahedron Lett., 32: 3087
(1991).
27. M. Nishizawa, M. Emura, Y. Kan, H. Yamada, K. Ogawa, and N. Hamanaka, Tetra-
hedron Lett., 21: 2983 (1992).
28. M. Mori, F. Saitoh, N. Uesaka, and M. Shibasaki, Chem. Lett., 213 (1993).
29. I. Kubo, and T. Kusumi, unpublished results.
30. J. Tanaka, T. Higa, K. Suwanborirux, V. Kokpol, G. Bernardinelli, and C. W. Jef-
ford, J. Org. Chem., 58: 2999 (1993).
31. K. Shin-ya, M. Tanaka, K. Furihata, Y. Hayakawa, and H. Seto, Tetrahedron Lett.,
34: 4943 (1993).
32. K. Iguchi, M. Fujita, H. Nagaoka, H. Mitome, and Y. Yamada, Tetrahedron Lett.,
34: 6277 (1993).
33. M. Ochi, S. Ariki, A. Tatsukawa, H. Kotsuki, Y. Fukuyama, and K. Shibata, Chem.
Lett., 89 (1994).
34. Y. Inouye, Y. Sugo, T. Kusumi, and N. Fusetani, Chem. Lett., 419 (1994).
35. M. Kobayashi, S. Aoki, and I. Kitagawa, Tetrahedron Lett., 35: 1243 (1994).
36. M. Sugano, A. Sato, Y. Iijima, K. Furuya, H. Haruyama, K. Yoda, and T. Hata, J.
Org. Chem., 59: 564 (1994).
37. G. Trimurtulu, I. Ohtani, G. M. L. Patterson, R. E. Moore, T. H. Corbett, F. A.
Valeriote, and L. Demchik, J. Am. Chem. Soc., 116: 4729 (1994).
38. H. Shimada, S. Nishida, S. Singh, M. Sahai, and Y. Fujimoto, Tetrahedron Lett.,
35: 3961 (1994).
39. H. Yada, H. Sato, S. Kaneko, and A. Ichihara, Tetrahedron Lett., 35: 4393 (1994).
40. M. V. D’Auria, C. Debitus, L. G. Paloma, L. Minale, and A. Zampella, J. Am. Chem.
Soc., 116: 6658 (1994).
41. Z. Gu, X. Fang, L. Zeng, and J. L. McLaughlin, Tetrahedron Lett., 35: 5367 (1994).
42. T. Hashimoto, M. Morie, M. Toyota, Z. Taira, R. Takeda, M. Tori, and Y. Asakawa,
Tetrahedron Lett., 35: 5457 (1994).
43. S. Paik, S. Carmeli, J. Cullingham, R. E. Moore, G. M. L. Patterson, and M. A.
Tius, J. Am. Chem. Soc., 116: 8116 (1994).
44. T. R. Hoye, J. I. Jiménez, and W. T. Shier, J. Am. Chem. Soc., 116: 9409 (1994).
45. M. L. Ciavatta, E. Trivellone, G. Cimino, and M. J. Uriz, Tetrahedron Lett., 35:
7871 (1994).
46. K. Stratmann, D. L. Burgoyne, R. E. Moore, G. M. L. Patterson, and C. D. Smith,
J. Org. Chem., 59: 7219 (1994).
47. G. J. Hopper and M. T. Davies-Coleman, Tetrahedron Lett., 36: 3265 (1995).
48. F. Kong and R. J. Andersen, Tetrahedron, 51: 2895 (1995).
49. S. Fukuzawa, S. Matsunaga, and N. Fusetani, J. Org. Chem., 60: 608 (1995).
50. W. Y. Yoshida, P. J. Bryan, B. J. Baker, and J. B. McClintock, J. Org. Chem., 60:
780 (1995).
51. Y. Matsuo, M. Suzuki, and M. Masuda, Chem. Lett., 1043 (1995).
52. J. Tanaka, T. Higa, G. Bernardinelli, and C. W. Jefford, Chem. Lett., 255 (1996).
53. G. Shi, D. Alfonso, M. O. Fatope, L. Zeng, Z. Gu, G. Zhao, K. He, J. M. MacDougal,
and J. L. McLaughlin, J. Am. Chem. Soc., 117: 10409 (1995).
TM
TM
TM
Nobuyuki Harada
Tohoku University, Sendai, Japan
I. INTRODUCTION
TM
TM
(ϕa | ∇| ϕb ) z ⫽ 冱 (C C
bonds
ra sb ⫺ Csa Crb ) 〈∇rs 〉 cos Zrs (3)
(ϕa | r ⫻ ∇ | ϕb ) z ⫽ 冱 (C
bonds
ra Csb ⫺ Csa Crb )
(4)
〈∇rs〉 (Xrs cos Yrs ⫺ Yrs cos Xrs )
cos Zrs ⫽ (Zr ⫺ Zs )/Rrs (5)
Xrs ⫽ (Xr ⫹ Xs )/2 (6)
where Cra is the coefficient of atomic orbital r in the wave function ϕa ; 〈∇rs 〉 is
the expectation value of a dipole velocity vector ∇rs that is directed along the
bond rs in the direction r → s; Xr, Yr, and Zr are the x, y, and z coordinates of
an atom r, respectively; and Rrs is the interatomic distance between atoms r and
s. In a similar way, the x and y components of the electric and magnetic transition
moments were calculated.
In the π-electron SCF-CI-DV MO calculation, the configuration interac-
tions (CIs) among all singly excited states were included, and the following
standard values of atomic orbital parameters were used: for aromatic carbon,
W(C) ⫽ ⫺11.16 eV, (rr | rr) (C) ⫽ 11.13 eV, β(C ⫺ C, 1.388 Å) ⫽ ⫺2.32 eV,
〈∇〉 (C ⫺ C, 1.388 Å) ⫽ 4.70 ⫻ 107 cm⫺1; for carbonyl oxygen, W(O)
TM
TM
TM
A. Molecular Structure
Model compound (8aR)-9 has the following stereochemical features: The 1,8a-
dihydroazulene ring skeleton is conformationally very rigid. The triene subunit
TM
TM
TM
Scheme 1 Synthesis of model compounds (⫹)-7 and (⫹)-8. (Source: Ref. 5.)
TM
TM
TM
a weak, long wavelength transition centered around 315–325 nm, with a stronger,
oppositely signed band around 225 nm for these models, and the natural dihy-
droazulenes, along with the agreement of this pattern with the theoretical CD
spectrum of (8aR)-9, indicated that the absolute stereoconfiguration of (⫹)-1 and
related dihydroazulenes may be assigned on the basis of the theoretical calcula-
tions.
Scheme 2 Synthesis of model compounds (⫹)-5 and (⫹)-6. (Source: Ref. 6.)
TM
In recent years, there have been many reports concerning isolation, structure de-
termination, and biological activity studies of marine natural products. Many
novel, biologically active compounds have been isolated from marine sponges.
For example, Scheuer and coworkers isolated halenaquinone [(⫹)-18], an antibi-
otic with a novel pentacyclic skeleton, from the tropical marine sponge Xestos-
pongia exigua collected off the Western Caroline Islands (Chart 2) [40]. The
structure of (⫹)-18 was determined by x-ray crystallographic analysis. The abso-
lute configuration, however, remained undetermined since the x-ray Bijvoet
TM
method for determination of the absolute configuration on the basis of the anoma-
lous dispersion effect of heavy atoms could not be applied.
Subsequent to Scheuer’s work, Kitagawa and coworkers isolated halenaqui-
nol [(⫹)-19], the hydroquinone form of (⫹)-18, from the Okinawan sponge Xes-
tospongia sapra, together with halenaquinol sulfate [(⫹)-20] [41]. Halenaquinol
[(⫹)-19] was easily oxidized either at UV irradiation or through heating in air
to give (⫹)-18. Furthermore, Nakamura and coworkers isolated xestoquinone
[(⫹)-21], a powerful cardiotonic, from the same Okinawan sponge, X. sapra [42].
More recently, Schmitz and Bloor isolated a series of similar natural products,
tetrahydroxestoquinol (23), the related dihydrofuran compound 24, adociaqui-
none A [(⫹)-25], adociaquinone B [(⫹)-26], and 3-ketoadociaquinone A (27),
from a marine sponge, Adocia sp. collected in Truk Lagoon, in addition to (⫹)-
18 and (⫹)-21 (Chart 3) [43]. They also revealed that halenaquinone [(⫹)-18]
and adociaquinone B [(⫹)-26] showed mild cytotoxicity. Harada and coworkers
also reported (⫺)-prehalenaquinone [(⫺)-22], a putative biosynthetic precursor
to (⫹)-18 and (⫹)-21 (Chart 2) [11]. Considering the interest in the halenaquinol
family of marine natural products due to the physiological activity of these novel
TM
TM
TM
TM
of (⫹)-28, and the results suggested the 12bS absolute configuration for (⫹)-28
(as later shown to be correct), we don’t consider these results a convincing and
unambiguous determination of the absolute configuration because of the small
∆ε values of CD Cotton effects.
TM
Figure 8 NOE and coupling constant data and conformation of naphthalene–diene de-
rivative (⫹)-34. (Source: Ref. 8.)
TM
showed two intense π → π* bands (Fig. 9), a broad band at 324 nm (ε 27,000)
with complex vibrational structure, and a sharp band at 218 nm (ε 42,000). In
these corresponding regions, the CD spectrum of (⫺)-33 exhibited three major,
intense Cotton effects: λext 338 nm (∆ε ⫹6.4), 301 nm (∆ε ⫺23.3), and 229 nm
(∆ε ⫹40.9). The other naphthalene–diene compound (⫹)-34 also exhibited three
major CD Cotton effects of similar intensity and of the same sign as those of
(⫺)-33. These results clearly indicated that the CD Cotton effects originated
mainly from the π-electron chromophore composed of the naphthalene–diene
moiety, which is twisted by the angular methyl group at the 12b-position. Further-
more, the additional chiralities due to the silyloxy group at the 3-position and
the methoxy group at the 4-position are only minor contributors to the CD Cotton
TM
Figure 10 Model compound (12bS)-35 and its stable conformation calculated by the
MOPAC 93, AM1 programs.
TM
Figure 11 Circular dichroism and ultraviolet curves of the model compound (12bS)-
35 calculated by the π-electron SCF-CI-DV MO, self-consistent field/configuration
interaction/dipole velocity molecular orbital method. (Source: Refs. 7 and 8.)
TM
TM
TM
TM
TM
TM
TM
afford naphthalene derivative (⫺)-37 in 89% yield, which was then subjected to
air oxidation in the presence of base [48]. Thus, oxygen was bubbled through a
solution of (⫺)-37 and potassium t-butoxide in t-butyl alcohol for 5 h, and the
mixture was worked up with aqueous ammonium chloride to give diosphenol 56
in 90% yield. The structure of 56 was secured by the 1H NMR (a sharp singlet
at δ 7.60 ppm disappeared upon addition of D2O), UV spectra (a red shift and
a hyperchromic effect of the longer-wavelength UV absorption band upon adding
aqueous NaOH), along with high-resolution MS data. Deprotection of the aceto-
nide group of 56 upon treatment with 60% aqueous acetic acid yielded triol 36,
which was subjected to the next reaction without purification (Scheme 10). The
oxidation of the primary and secondary hydroxyl groups of 36 and subsequent
cyclization to form the furan ring were accomplished by treatment with dimethyl
sulfoxide (DMSO) and 1,3-dicyclohexylcarbodiimide (DCC) in benzene in the
presence of trifluoroacetic acid and pyridine (PTFA), giving the desired halena-
quinol dimethyl ether [(12bS)-(⫹)-28] with the furan–diketone system, in 44%
overall yield from (⫺)-37. All of the spectroscopic data of the synthetic sample
of (⫹)-28 were identical with those of an authentic sample of (⫹)-28 prepared
from natural halenaquinol.
The dimethyl ether groups of (12bS)-(⫹)-28 were next oxidatively cleaved
with cerium (IV) ammonium nitrate (CAN) in aqueous methanol, affording halen-
aquinone [(12bS)-(⫹)-18] in 45% yield (Scheme 10). The 1H NMR and UV spec-
tra of the synthetic sample agreed with those of natural (⫹)-18. Finally, (12bS)-
TM
(⫹)-18 was reduced with aqueous sodium hydrosulfite in acetone to give halena-
quinol [(12bS)-(⫹)-19] in near-quantitative yield. The 1H NMR spectrum of
(12bS)-(⫹)-19 in DMSO-d6 exhibited two broad singlets at δ 9.6 and 9.8 that
were due to the phenolic hydroxyl groups, which disappeared when D2O was
added. The remaining 1H NMR peaks and UV spectrum curve were in good
agreement with those of the natural sample. The first total synthesis of halenaqui-
nol [(12bS)-(⫹)-19] and halenaquinone [(12bS)-(⫹)-18] with a novel polyketide
skeleton had been thus accomplished.
TM
halenaquinol. This was verified as shown in Figs. 7 and 14: the CD spectra of
the synthetic and natural samples are identical. These results lead to the experi-
mentally confirmed, unambiguous determination that the absolute stereochemis-
try of (⫹)-halenaquinol and (⫹)-halenaquinone is 12bS. In addition, these syn-
theses also proved that the theoretically predicted absolute configurations of
halenaquinol compounds were correct.
TM
TM
[11]. This synthetic route is more efficient than our previous total synthesis of
xestoquinone and xestoquinol. We have thus achieved the total syntheses of xes-
toquinol, halenaquinol, and related compounds from the common synthetic inter-
mediate (⫺)-57.
TM
TM
The title biflavone was isolated from Garcinia mangostana L. [13] and was iden-
tified as optically active cupressuflavone tetramethyl ether [(⫺)-61] by a direct
comparison of spectral data with those of the authentic sample isolated earlier
from Araucaria cunninghamii and A. cookii (Chart 4) [14]. Compound 61 is a
dimeric flavone as indicated by the MS spectral data ([M]⫹, m/z 594) and has a
chiral C2-symmetric structure that exhibits an optical rotation of [α]D27 ⫺25.3°,
with only seven signals in the 1H NMR spectrum, and 15 peaks in the 13C NMR
spectrum. We established the dimeric structure with an 8-8″-linkage of 61 on the
basis of detailed 1H and 13C NMR studies, including 1H NOESY, heteronuclear
gated decoupled 13C NMR, 1H-13C COSY, and long-range 1H-13C COSY
(COLOC) methods, and have achieved full assignment of all proton and carbon
TM
TM
TM
tive Cotton effect that appeared as a shoulder around 290 nm and a positive one
at λext 263.2 nm (∆ε ⫹21.7). These calculated CD bands were in excellent agree-
ment with those observed in the CD spectrum of (⫺)-61, including sign, intensity,
and position of the Cotton effects (compare Figs. 15 and 17): λext 362.0 nm (∆ε
⫹25.6), 326.2 (∆ε ⫺54.4), a negative shoulder around 300 nm, and 267.5 (∆ε
⫹21.3). Since the calculation was performed for the enantiomer with the aR con-
figuration, the absolute stereochemistry of natural biflavone, (⫺)-4′,4,7,7″-tetra-
O-methylcupressuflavone [(⫺)-61], was thus determined to be aR (or M helicity)
in a nonempirical manner.
TM
TM
TM
TM
TM
c: concentration.
a
From Ref. 18. The sign of optical rotation was taken from the [α]D value of the natural sample in
MeOH.
b
From Ref. 13.
c
From Ref. 15.
tion, we adopt both methods here; the natural product (⫺)-61 is thus defined
as [CD(⫹)362.0]-(aR)-(⫺)-4′,4,7,7″-tetra-O-methylcupressuflavone. Since it is
also often useful to include P and M designations for these enantiomers, the (aR)
enantiomer has M helicity and the (aR) isomer has P helicity.
TM
TM
TM
TM
TM
Figure 22 Circular dichroism and ultraviolet spectra of the synthetic sample of biflav-
one,4′,4,7,7″-tetra-O-methylcupressuflavone {[CD(⫹)362.0]-(aR)-(⫺)-61]} in EtOH.
(Source: Ref. 18.)
TM
VI. CONCLUSIONS
REFERENCES
TM
TM
TM
TM
This chapter is intended to show that circular dichroism (CD), although a well-
recognized spectroscopic technique, is still being developed, and that new appli-
cations are yet to be discovered. In spite of the existence of many useful empirical
rules and of the nonempirical CD exciton chirality method for determination of
absolute configuration and conformational analysis, this technique remains to be
developed fully.
Two recent CD applications are described: (1) First is the conformational
analysis of carbohydrates, exemplified by the CD and 1 H-nuclear magnetic reso-
nance (1 H-NMR) study of the rotational population dependence of the hydroxy-
methyl group in alkyl glucopyranosides on the aglycon and its absolute con-
figuration, a clear correlation between the rotamer distributions and the
stereoelectronic exo-anomeric effect being observed. As a consequence of these
results the absolute configuration of secondary alcohols can be determined by
either CD or 1 H NMR, as a single enantiomer is necessary for this purpose;
(2) The direct determination of drug levels in pharmaceutical formulations (oral
suspensions, injections, and capsules), as well as in human biological fluids
(urine and serum), with specific reference to the β-lactam antibiotics, is de-
scribed.
TM
Circular dichroism (CD) is a powerful technique that has been mainly used for
the determination of absolute configuration of a great number of compounds of
both natural and synthetic origin. This technique has its origins in the 1960s
through important contributions by scientists such as C. Djerassi, P. Crabbé, and
G. Snatzke, who introduced many effective empirical rules for absolute configu-
ration determinations of organic compounds [1]. Further developments came with
the exciton chirality method [2], mainly through the seminal scientific contribu-
tions of N. Harada and K. Nakanishi. This method, which allows the determina-
tion of absolute configurations in a nonempirical way, offers great analytical po-
tential and, in my opinion, establishes a second era for CD. Applications of
circular dichroism have also been widely used in the conformational studies in
solution of small to medium size molecules, as well as macromolecules, espe-
cially proteins [3].
The development of the CD exciton chirality method [2] has extended the
utility of this technique for determining the absolute configuration of organic
compounds. The high sensitivity and straightforward spectral interpretation of
this method, as a consequence of the general validity of the pairwise additivity
in exciton-coupled systems [4], have not been fully exploited. CD is still the
technique that is least known and used by most organic chemists.
The CD exciton chirality method is based on the interaction through space
of the electric transition moments of two chromophores, which gives rise to an
excited state split into two energy levels. Excitations to these levels lead to a CD
spectrum with two Cotton effects of opposite signs, namely, to a ‘‘split’’ CD
curve [2]. The chiral environment of the two chromophores determines the sign
of the Cotton effects; the sign of the exciton chirality is that of the first Cotton
effect, the one at longer wavelength. Furthermore, the existence of an additivity
relation in multiple-chromophoric systems [4] allows an easy spectral interpreta-
tion of this type of compound, since the observed CD spectrum is the sum of
CDs arising from all pairwise interactions present in the system: namely, the CD
spectrum of a system composed of three interacting chromophores XYZ is the
sum of the CDs arising from X/Y, X/Z, and Y/Z interactions. The general valid-
ity of the pairwise additivity in exciton-coupled systems is valid independently
of whether the interacting chromophores are the same or not [4].
In the belief that CD is a spectroscopic technique that offers a series of
analytical advantages over other techniques, our group led the research to apply
the CD exciton chirality method to the conformational analysis of carbohydrates,
namely, to study the rotational population dependence of the hydroxymethyl
group in alkyl glycopyranosides on the structural nature of the aglycon and its
absolute configuration.
TM
Figure 1 Torsion angles about the glycosidic linkage and around the C5–C6 bond (top);
Newman projections of the gg, gt, and tg rotamers around the C5–C6 bond (bottom).
TM
TM
Moreover, these spectral differences are only consistent with a gradual de-
crease and an increase in the population of the gg and gt rotamers, respectively,
as can be deduced by analyzing the pairwise interactions involved in these com-
pounds. Fig. 3 shows for the glucopyranosyl system those pairwise interactions
involving the chromophore at the 6 position, the 2/6, the 3/6, and the 4/6 interac-
tion, for the three rotamers. As can be observed, those interactions involving the
gg rotamer have a net positive contribution; those involving the gt rotamer have
a net negative contribution; and those interactions involving the tg rotamer have
a nil contribution. Therefore, the general decreases in the CD Cotton effects can
be explained by a decrease in the population of the gg rotamer (net positive
contribution) and an increase in the population of the gt rotamer (net negative
contribution). The other three existing pairwise interactions, the positive 2/3, the
negative 3/4, and the nil 2/4, having constant intensities and signs, do not affect
this CD interpretation. Furthermore, the striking increase in the first positive and
second negative Cotton effects observed on these spectra by lowering the temper-
ature (MeOH, ⫺80°C) can only be explained by an increase in the rotational
population of the energetically favored gg rotamer.
Rotamer distribution is normally determined by 1 H NMR analysis of the
coupling constants of the prochiral protons at C6 [11], which are assigned on
TM
the basis of their chemical shifts. That is, the coupling constant values of the
usually well-differentiated doublet of doublets for H6R and H6S are substituted
in empirical equations to calculate the rotamer distribution [12].
With the exception of the cyclohexyl derivative 1d, a doublet, rather than
the usual clearly differentiated doublet of doublets for each proton at C6, was
obtained for the bichromophoric compounds. Therefore, their rotamer distribu-
tions could not be calculated by means of 1 H NMR coupling constants. However,
the coupling constant J H5,H6, which gradually increased from methyl (3.9 Hz), to
ethyl (4.1 Hz), to iso-propyl (4.4 Hz), and to tert-butyl (4.6 Hz) glucopyranoside
derivatives (1a–c,e), corroborated the increase in the gt population.
Since the results show that the gt population increases as the pK a of the
aglycon increases, and it is accepted that the stereoelectronic exo-anomeric effect
[13] increases with increasing ease for charge delocalization from the aglycon
to the anomeric carbon [14], this stereoelectronic effect must be responsible for
the rotational population dependence of the hydroxymethyl group on the aglycon,
because for these low-molecular-size alcohols the existence of nonbonded inter-
actions with the chromophore at C6 cannot be expected.
This correlation was confirmed by the CD and NMR data comparison of
TM
TM
Figure 4 Circular dichroism spectrum (CH 3CN) of methyl, (⫹)- and (⫺)-2-octyl (left),
and (⫹)- and (⫺)-menthyl (right) 2,3,4,6-tetrakis-O-( p-bromobenzoyl)-β-d-glucopyrano-
sides. (Source : Ref. 9.)
TM
TM
Table 2 Circular Dichroism Data (CH 3CN) and 1 H Nuclear Magnetic Resonance
Coupling Constants JH5,H6 (CDCl 3, 400 MHz) for 2,3,4,6-Tetrakis-O-Benzoyl-β-
Glucopyranosides of Secondary Alcohols
Glucose ∆ε
Entry Alcohol series Cl′ 233/220 nm A Value J H5,H6R J H5,H6S
Source: Ref. 9.
TM
TM
NMR shifts (Fig. 7), constituted the basis of a new method to determine the
absolute configuration of secondary alcohols [18]. The differences between the
proton chemical shifts of the d-glucosylated derivative and the free alcohol (∆δ ⫽
δ D ⫺ δ ROH ) or, more significantly, between their chemical shifts in the d- and l-
glucosylated derivatives (∆δ ⫽ δD ⫺ δL) are characteristic of the absolute config-
uration of the secondary chiral alcohol. Moreover, this method involves the use
of one enantiomer and generally a single derivatization is sufficient.
Fig. 8 shows for (⫺)-menthol (in hertz, at 500 MHz, 25°C, CDCl 3) the
chemical shift differences (δD ⫺ δL) obtained by applying our method, on the
basis of the tetra-O-benzoylglucosylation, as well as those (δ S ⫺ δ R) acquired
by the advanced Mosher’s method [19]. The larger values obtained by the former
can make it a suitable alternative method to determine the absolute configuration
of secondary alcohols.
Figure 8 Comparison of the δ S ⫺ δ R and the δ D ⫺ δ L values of the MTPA [19d] (left)
and the tetrakis-O-benzoyl-glucopyranoside (right) derivatives of (⫺)-menthol, respec-
tively. MTPA, 1-methoxy-1-phenyl-1-trifluoromethylacetic acid. (Source: Ref. 18.)
TM
TM
TM
TM
TM
1. Human Urine
At the time the study that follows [27] was performed, only two reports had been
published for the determination of drugs in human biological fluids by CD, one
describing the determination of tetracycline in urine [28] and the other the deter-
mination of tri- and tetra-(hydroxyethyl)-rutosides in urine and serum [29].
In collaboration with the urological department of the Hospital Universita-
TM
TM
TM
Figure 12 Circular dichroism spectra of the urine of three patients under cefoxitin ther-
apy (250 times dilutions): without proteins, solid line (23.25 µg/mL of cefoxitin); with
proteins, dashed line (15.29 µg/mL of cefoxitin, and 11.19 µg/mL of proteins); dotted
line, (7.48 µg/mL of cefoxitin, and 17.16 µg/mL of proteins). (Source: Ref. 27.)
TM
TM
urinary protein was identical with that of the main human urinary protein, human
albumin. Similarly to the β-lactam antibiotics, the analysis of the variance, (AN-
OVA) of the linear regression of the calibration line carried out with human
albumin at 220 and 209 nm confirmed the linearity of the method (Table 7); the
statistical analyses showed precise and accurate values for both wavelengths,
although slightly better for the one measured at 220 nm.
The direct quantitative determination of a CD-active drug in protein-free
urine samples from type-1 CD spectra or the determination of the total urinary
proteins in urine samples lacking CD-active drugs from type-2 CD spectra is
achieved in a straightforward manner by measuring in each case the ellipticity
angle (mdeg) at the selected wavelength and using the corresponding regression
line equation (Table 7).
The procedure to determine a CD-active drug from a ‘‘complex’’ type-3
CD spectrum depends on the wavelength of the first Cotton effect of the drug
to be analyzed: (1) If the drug exhibits a Cotton effect above 250 nm, its determi-
nation can be performed by direct measurement of the ellipticity at the wave-
length of this Cotton effect, regardless of the presence of proteins (Fig. 12). This
can also be determined by CD spectral subtraction of the CD-active drug contri-
bution to the ‘‘complex’’ CD spectrum; (2) the simultaneous determination of a
TM
Figure 14 Complex circular dichroism spectra, from top to bottom, of in-house mix-
tures of albumin (48 µg/mL) and cefoxitin (80, 60, 40, 20, 5, and 1 µg/mL), respectively
[30]. (Source: Ref. 27.)
TM
TM
of optically active absorbing drugs. As can be observed in Fig. 16, a clear gradual
decrease in the intensity of the Cotton effects of the urine samples along a 6-h
period was obtained from a patient receiving intravenously (i.v.) the first dose of
1 g of cefoxitin. The corresponding CD hourly determinations were in excellent
agreement with those obtained from HPLC, showing the applicability of this
method to monitor drugs. The total amounts of cefoxitin recovered from the first
up to the sixth hour were 349.4, 227.1, 113.4, 74.1, 59.7, and 39.3 mg, respec-
tively.
2. Human Serum
The most complex and important of the biological fluids is blood, and thus once
the study in urine was completed, the possibility of applying CD to establish a
method to determine CD-active drugs in this biological fluid was analyzed [33].
Since the concentration of proteins in human serum or plasma is very high,
it was necessary to dilute the serum sample to 1 : 1000, in order to have a suitable
voltage in the CD photomultiplier (below 460 V). Consequently, the direct deter-
mination of a CD-active drug is not possible from these very dilute solutions;
deproteinization of the serum sample is necessary before CD analysis.
The precipitation of the serum/plasma proteins by means of acetonitrile
proved to be an excellent way to perform CD analysis of drugs. The method
TM
TM
tion of drugs in biological fluids. Its high degree of selectivity, together with the
fact that it is quick and economical, could make CD a common technique in
clinical laboratories in the near future.
ACKNOWLEDGMENTS
I wish to thank all the persons who worked out the results described: Drs. J. I.
Padrón, E. Q. Morales, P. Gortázar, M. Ravina, and Mrs. M. Trujillo. Support
of this work by the Dirección General de Investigación Científica y Técnica
(DGICYT), Ministerio de Educación y Ciencia (Spain), through grant PB93-
0559, and by the Gobierno de Canarias is gratefully acknowledged.
1. (a) F. Ciardelli and P. Salvadori, eds., Fundamental Aspects and Recent Develop-
ments in Optical Rotatory Dispersion and Circular Dichroism, Heyden & Son, Lon-
don (1973). (b) K. Nakanishi, N. Berova, and R. W. Woody, eds., Circular Dicho-
ism, Principles and Applications, VCH Publishers, New York (1994).
2. (a) N. Harada and K. Nakanishi, Circular Dichroic Spectroscopy-Exciton Coupling
in Organic Stereochemistry, University Science Books, Mill Valley, Calif. (1983).
(b) Ref. 1b, p. 361–398.
3. C. D. Fasman, ed., Circular Dichroism and the Conformational Analysis of Biomo-
lecules, Plenum Press, New York, p. 25–157 (1996).
4. (a) W. T. Wiesler, J. T. Vázquez, and K. Nakanishi, J. Am. Chem. Soc., 109: 5586
TM
TM
TM
Steven P. Tanis
Pharmacia & Upjohn, Inc., Kalamazoo, Michigan
I. INTRODUCTION
Approximately 23 years ago the sesquiterpene warburganal (1) was isolated from
the East African medicinal plant Warburgia ugandensis [1], and we had the op-
portunity to embark on a synthesis prior to the publication of the structure of 1.
Our initial approach constructed the sesquiterpene framework through a Diels-
Alder cycloaddition of 1-vinyl-2,6,6-trimethyl-1-cyclohexene with dimethyl acet-
ylenedicarboxylate, as shown in Scheme 1 [2]. The Diels-Alder adduct 2 (83%)
resisted the development of the desired trans-ring fusion via catalytic hydrogena-
tion, affording instead only cis-fused products. While we were wrestling with
this problem, which was eventually solved [2], we considered alternatives that
might directly construct the desired trans-fused sesquiterpene skeleton.
The cationic π-cyclization, which has been widely utilized in the construc-
tion of polycyclic ring systems, has been the object of intense study since the
early 1950s [3]. Initial forays into this arena demonstrated the syntheses of fused
ring terpenoid-type systems, and later efforts demonstrated the construction of
spiro and bridged ring carbocyclic systems [4]. These studies have demonstrated
that a wide variety of initiating functions (e.g., epoxides, allylic alcohols, enones,
olefins, carbinolamides) and terminating moieties (e.g., aromatic rings, acety-
TM
lenes, allylsilanes, allenes, olefins) can be incorporated into the cyclization sub-
strate to lead to terpenoids and alkaloids. In the synthesis of simple fused ring
systems, such as the perhydronaphthalene moiety of warburganal (1), this method
has produced the target 4,4,8a-trimethyl-trans-fused skeleton as shown in Eq. (1)
[5]. The issue to be considered prior to examining a cationic π-cyclization for
the preparation of 1 was the identity of the 1,4-dialdehyde equivalent as the termi-
nator function for the cyclization.
(1)
TM
moiety and the reaction conditions would be required if the projected chemistry
were to be successful.
Our interest in developing substituted furans as nucleophilic synthons in
annulative processes stemmed from the variety of useful functional groupings
that might be realized from the relatively unreactive furyl nucleus. As mentioned
(Fig. 1), a furan can serve as the operational equivalent of a 1,4-dialdehyde; Fig.
2 indicates additional acyclic, heterocyclic, and carbocyclic subunits that can be
derived from a furan after standard chemical manipulations.
TM
TM
(2)
TM
(3)
(4)
1. Cyclization Studies
Relatively potent Lewis acids such as boron trifluoride etherate are often selected
to catalyze epoxy olefin cyclizations [8]. Given the acid lability of the starting
furans and the increased acid sensitivity of the products, the choice of Lewis acid
should have a profound effect on the partitioning of the reaction between a fruitful
cyclization and acid-mediated decomposition. Six Lewis acids were utilized
TM
the initial study [7a], BF3⋅OEt2 [8], EtAlCl2 [14], Et2AlCl [14], Al2O3 [13], Ti(Oi-
Pr)3Cl [15], and ZnI2 [16]. The choice of Lewis acid was dictated by (1) the
ability to modify the potency of a group of Lewis acids with a common metal
center readily and (2) the possibility of moderating the Brønsted acidity of the
medium through the choice of the Lewis acid. Adventitious protic acid might be
scavenged by a Lewis acid possessing a metal–carbon bond releasing an alkane;
alternatively with the proper choice of metal, the product metal–alcohol complex
should be a much weaker protic acid compared to a BF3 –alcohol complex.
The substrate epoxy furans were exposed to boron trifluoride etherate (0.3
eq.) in methylene chloride at ⫺25°C as the standard cyclization conditions (Table
2). Five-membered ring precursors 22, 23, and 30, when treated with BF3⋅OEt2,
afforded only allylic alcohols 32 (62%), 34 (53%), and 36 (49%), respectively.
Only epoxy furans 24 and 25 gave appreciable quantities of cyclized products,
37 (47%) and 39 (30%), respectively. The majority of the material isolated from
the BF3⋅OEt2-mediated cyclizations consisted of the depicted allylic alcohols, and
the mass balances were poor (ca. 60% or less). The poor yields of cyclized materi-
als and low mass balances are in agreement with the surmises regarding the
Lewis/Brønsted acidity requirements for these reactions.
The aluminum-based Lewis acids, EtAlCl2, Et2AlCl, and Al2O3, provided
better mass balance (ca. 70% or more), and variable amounts of cyclized prod-
ucts. The treatment of 22, 23, and 30, five-membered ring precursors, with either
EtAlCl2 or Et2AlCl led to good yields of allylic alcohols. Only furans 24 (6-
endo), 25 (6-exo), and 26 (7-endo) gave any cyclized material (10%–22%) with
EtAlCl2 or Et2AlCl. Alumina (Al2O3) provided only allylic alcohol for all epoxy
furans except 24, which furnished a 32% yield of 37 and a 51% yield of allylic
alcohol 38. The modification of the Lewis acid to provide a protonolyzable M-C
bond, thus reducing Brønsted acidity of the medium, did lead to improved mass
balance; however, elimination was the dominant path with the alkylaluminum
halides. Further modification of the aluminum-centered Lewis acid to alumina
also provided improved mass balance but little cyclization.
The next most Lewis acidic compound in the series Ti(Oi-Pr)3Cl [15]
proved to be an efficient and useful promoter of epoxy furan cyclization. As
before, five-membered-ring precursor oxiranes 22, 23, and 30 did not lead to
desired cyclized products, with elimination products 32 (80%) and 34 (72%)
coming from 22 and 23, respectively. The monosubstituted epoxide 30 could not
be induced to react, even after exposure to 3 eq. of Ti(Oi-Pr)3Cl for 24 h at room
temperature. Similar treatment of epoxy furan 24 led to the formation of the
desired cyclized adduct 37 in 78% yield, with no elimination product observed.
6-exo-Epoxide 25 and 7-endo-precursor 26 afforded cyclized adducts 39 (89%)
and 41 (87%), respectively, with the latter accompanied by 8% of allylic alcohol
42. Even 7-exo-precursor epoxide 29 gave a respectable yield of cyclized product
43 (36%) when treated with Ti(Oi-Pr)3Cl.
Epoxides 22, 23, and 30, five-membered ring precursors, were next exposed
TM
Furyl-epoxide Product(s)
22 31 32
BF3-OEt2 (0.3 eq.) 0% 62%
Et2AlCl (2 eq.) 0% 85%
Al2O3 0% 83%
Ti(OiPr)3Cl (3 eq.) 0% 80%
ZnI2 (3 eq.) 0% 76%
23 33 34
BF3-OEt2 (0.3 eq.) 0% 53%
Et2AlCl (2 eq.) 0% 85%
Ti(OiPr)3Cl (3 eq.) 0% 72%
ZnI2 (3 eq.) 0% 70%
30 35 36
BF3-OEt2 (0.3 eq.) 0% 49%
Et2AlCl (2 eq.) 0% 78%
Ti(OiPr)3Cl (3 eq.) no reaction
ZnI2 (3 eq.) 25% 44%
24 37 38
BF3-OEt2 (0.3 eq.) 47% 0%
EtAlCl2 (2 eq.) 16% 57%
Et2AlCl (2 eq.) 22% 49%
Al2O3 32% 51%
Ti(OiPr)3Cl (3 eq.) 78% 0%
ZnI2 (3 eq.) 71% 0%
TM
25 39 40
BF3-OEt2 (0.3 eq.) 30% 10%
EtAlCl2 (2 eq.) 0% 73%
Et2AlCl (2 eq.) 10% 70%
Al2O3 0% 81%
Ti(OiPr)3Cl (3 eq.) 89% 0%
ZnI2 (3 eq.) 70% 0%
26 41 42
BF3-OEt2 (0.3 eq.) 0% 41%
EtAlCl2 (2 eq.) 0% 76%
Et2AlCl (2 eq.) 10% 69%
Al2O3 0% 83%
Ti(OiPr)3Cl (3 eq.) 87% 8%
ZnI2 (3 eq.) 88% 9%
29 43 44
BF3-OEt2 (0.3 eq.) 10% 12%
EtAlCl2 (2 eq.) 0% 64%
Et2AlCl (2 eq.) 0% 73%
Al2O3 0% 79%
Ti(OiPr)3Cl (3 eq.) 36% 47%
ZnI2 (3 eq.) 23% 52%
TM
TM
bonucleic acid (DNA) polymerase-α [19g]. McMurry [23] has reported the syn-
thesis of (⫾)-46 from diketone 49. We envisioned accomplishing a synthesis of
(⫾)-49 and (⫺)-49 [24] via furan-terminated cationic π-cyclization, as shown in
Scheme 4.
The chemistry outlined in Scheme 4 would exploit a furan-terminated cat-
ionic-π-cyclization to establish the carbon framework of 49 rapidly with the cor-
rect relative and absolute configuration created at carbons 4, 5, and 10; a furan
to dione conversion would provide 49. The potential availability of the modified
epoxygeranyl chloride precursor to furan 51 in optically pure form from a
Sharpless asymmetric epoxidation [25] is an additional advantage to the route
depicted.
In the event, geraniol was converted to the related benzoate (Scheme 5),
which readily underwent allylic hydroxylation (SeO2, TBHP) [26] to give alcohol
52 (70%). Sharpless asymmetric epoxidation of hydroxybenzoate 52 gave (⫺)-
53 in 71% yield [25a]. The optical purity of (⫺)-53 was judged to be ⱖ95% ee
after an examination of (⫺)-53 by nuclear magnetic resonance (NMR) in the
presence of Eu(hfpc)3 [25a] and high-performance liquid chromatography
(HPLC) analysis of the related Mosher ester [27]. Alternatively, (⫺)-53 could
be prepared in 83% yield and ⱖ95% ee via the Sharpless catalytic asymmetric
epoxidation protocol [25b]. We considered a number of blocking groups with
which to protect the free hydroxyl group of (⫺)-53 during the halogenation/
coupling/cyclization steps. Initially we selected a t-butyldimethylsilyl (TBDMS)
moiety for this function. This proved to be somewhat difficult at the cyclization
TM
TM
Figure 3 Reduction selectivity of (⫾)-58 with L-Selectride in the presence and absence
of added metal salts.
TM
(5)
(6)
TM
TM
and 75 (90%), respectively. The inclusion of only six- and seven-membered pre-
cursor chain lengths (m ⫽ 2, 3) in the compounds of Schemes 6 and 7 was based
upon the results presented in Table 2.
1. Cyclization Studies
Allylic alcohols have been extensively employed as initiators in cationic π-cycli-
zations [3], and the reaction conditions that have been employed generally in-
volve a protic acid of reasonable strength in a solvent in which it is soluble. Of
the many conditions reported in the literature, the two-phase mixture of anhy-
drous formic acid and cyclohexane [38] was selected for this study as the mildest
method for initiating the cyclization of allylic alcohols 70, 71, 74, and 75 (Scheme
8). Alcohols 70 and 71, designed to form six- and seven-membered rings, respec-
tively, from an allylic alcohol with the hydroxyl center exocyclic to the existing
ring, were separately exposed to formic acid in cyclohexane for 20 min at room
temperature. Alcohols 70 and 71 smoothly cyclized to give the six-membered
target 76 and the seven-membered target 77 in 68% and 61% yields, respectively,
as mixtures of exo-ethylidene double-bond isomers. The endocyclic allylic alco-
hols 74 and 75 were next treated with formic acid in cyclohexane to give tricyclic
furans 77 (73%) and 78 (56%), respectively. The studies of Scheme 8 firmly
established allylic alcohols as suitable initiators in furan-terminated cationic π-
cyclizations. Next we examined the cyclization of enals 79 and 80, which were
prepared as intermediates from the oxidation of 68 and 69, as previously illus-
trated in Scheme 6.
Enals 79 and 80 [Eq. (8)] were treated with BF3⋅OEt2, SnCl4, TiCl4, EtAlCl2,
Et2AlCl, MgBr2⋅OEt2, ZnI2, and Ti(Oi-Pr)3Cl, in a variety of solvents at various
temperatures, to no avail. Extensive decomposition was observed when BF3⋅OEt2,
SnCl4, TiCl4, EtAlCl2, and Et2AlCl were employed; the use of MgBr2⋅OEt2, ZnI2,
and Ti(Oi-Pr)3Cl led to recovered starting materials. Acylative-type enal cycliza-
tions similar to those reported by Andersen [39], Marshall [40], and Harding [41]
TM
were also examined. The treatment of 79 and 80 with either Ac2O/HClO4 /EtOAc
or (CF3CO)2 O/CF3CO2H resulted in a facile and high-yield acylation of the furyl
nucleus at the 2-position. Protic acid treatment (HCOOH, c-C6H12) of 79 and 80
led to recovered starting material or, after extended treatment, extensive decom-
position.
(8)
TM
(9)
TM
TM
The relative orientation of the methoxyl group of the cyclopentanone, to the exo-
cyclic stereocenters, could not be ascertained spectroscopically; therefore we
elected to continue the synthesis and determine the relative orientation when a
suitable crystalline derivative was obtained. Initial attempts to cyclize (⫾)-88
with Hg(OTFA)2 gave only a 12% yield of (⫾)-89, with the bulk of the recov-
ered material corresponding to the carboxylic acid equivalent of (⫾)-88. Ex-
posing (⫾)-88 to mercuric triflate/N,N-dimethylaniline complex [51], with a less
nucleophilic triflate counteranion, afforded an excellent 96% yield of the target
bicyclo[5.3.0]decane (⫾)-89. The reduction of (⫾)-89 was modified to employ
Luche conditions (NaBH4, MeOH, CeCl3 ⋅ 7H2O, ⫺78°C [52], after sodium boro-
hydride in ethanol afforded mixtures. This modification led to the isolation of
(⫾)-90 in 93% yield as a single stereoisomer, establishing the fifth stereocenter
about the seven-membered B-ring. The C-9-OH was protected as the correspond-
ing TBDMS-ether, giving (⫾)-91 (99%).
Numerous attempts to block the 4-ketone of (⫾)-91 as the related
TM
TM
TM
(10)
TM
(12)
TM
(Scheme 13), alcohols 105a and 105b were exposed to the two-phase mixture
of HCOOH/cyclohexane to give spiro[4,5]decane 107a (58%) and spiro[4,6]un-
decane 107b (53%), respectively. Similarly, alcohols 105c–e were treated with
HCOOH/cyclohexane to furnish the formate of 105c (84%), spiro[5,5]undecane
107d (72%), and spiro[5,6]dodecane 107e (58%), respectively.
The enones 106a–e, produced as illustrated in Scheme 12, were dissolved
in cyclohexane and treated with formic acid. Of the five substrates shown in
Scheme 14, only enones 106a and 106d, leading to spiro[4,5]decane 108a (72%)
and spiro[5,5]dodecane 108d (66%), respectively, provided any cyclized prod-
ucts. Enones 106b,c and 106e were recovered unchanged. Compounds 106b,c
TM
and 106e also proved resistant to cyclization under a wide variety of other reac-
tion conditions including Lewis acids and acylating agents.
In order to examine whether the formation of a six-membered ring is the
necessary factor in a successful enone-initiated/furan-terminated cyclization, the
furan 2-to-3 closure related to the chemistry of Scheme 13 was examined. The
Grignard reagent derived from 3-(2-furyl)-1-bromopropane was added to the iso-
butyl enol ether of 2-methyl-1,3-cyclohexanedione (Scheme 15), [60] to give
enone 109 (85%) after acidic workup. Enone 109 was exposed to a variety of
Brønsted and Lewis acids, as well as acylating agents, to no avail. The target
spiro[5,5]undecane furan 110, regioisomer of 108d, was not observed. This rela-
tively less favored (electronically) furan 2-to-3 cyclization [61] was observed
when allylic alcohol 111, prepared from 109 in 93% yield (NaBH4, CeCl3), was
treated with formic acid/cyclohexane, affording olefin 111 (68%).
TM
TM
TM
on the N-acyliminium ion precursor as illustrated in Fig. 5. The first foray into
alkaloid synthesis via N-acyliminium ion furan-terminated cyclization [68] was
directed to the preparation of the spirolactam intermediate 119 in Kishi’s synthe-
sis of perhydrohistrionicotoxin (120) [69].
The Kishi spiropiperidine 119 [69] was approached by the spirocyclic discon-
nection presented in Fig. 5. For the present application it was necessary to alter
the substrate from the illustrated carbinolamide-3-substituted furan to a 3-(5-
ethyl-2-furyl)propyl carbinolamide in order to provide a proper ketone–side chain
location. In the event (Scheme 18), the Grignard reagent prepared from 1-bromo-
3-(5-ethyl-2-furyl)propane was added to the iodomagnesium salt of glutarimide
[67b] to furnish the related carbinolamide 121 in excellent crude yield. Without
TM
VI. CONCLUSIONS
TM
ACKNOWLEDGMENT
I wish to thank those people who made all this possible; the grad students, post-
docs, and coworkers at Michigan State University, the Upjohn Co., and Phar-
macia and Upjohn Inc., who did the overwhelming majority of the work men-
tioned. Sincere thanks go to Yu-Hwey Chuang, Mark Collins, Melissa Deaton,
Lisa Dixon, Dave Head, Paul Herrinton, Mark McMills, Tim Parker, and Ed
Robinson. Thanks also to the Camille and Henry Dreyfus Foundation, the Na-
tional Institutes of Health, and the Upjohn Postdoctoral Research Scholar Pro-
gram for financial support.
REFERENCES
TM
TM
TM
TM
TM
Timothy A. Blizzard
Merck Research Laboratories, Rahway, New Jersey
I. INTRODUCTION
TM
been widely reviewed [1–4,6] and will not be comprehensively reviewed again
here. This article will describe only work in the late stages of the medicinal
chemistry program with which I was involved, with reference to other work only
as necessary to place our work in context.
II. CHEMISTRY
A. Avermectin Epoxides
1. Avermectin B1 8,9-Epoxide
Avermectin B1 8,9-epoxide (3) (Fig. 2) is a highly active anthelmintic that was
synthesized at Merck as a potential agricultural pesticide [7]. Although 3 is more
stable than 1 because of the elimination of the photosensitive diene, the presence
of a potentially reactive epoxide functionality still raised concerns about the
chemical stability of 3. In order to address these concerns we explored the reactiv-
ity of 3 with various nucleophiles. Not surprisingly, we found that the epoxide
reacted readily with strong nucleophiles such as thiophenoxide [8]. This result
was not of great importance by itself; however, the fact that the thiophenol adduct
4 crystallized from methanol turned out to be highly significant. When we began
our studies on 3, the epoxide stereochemistry had not been unambiguously estab-
lished. X-ray diffraction analysis of the crystals of 4 allowed the definitive assign-
ment of the stereochemistry of 4 as 8S,9R. Since 4 is derived from 3 with inver-
sion at C-9, then 3 must be the α-epoxide. Although 4 proved to be useful for
structural determination, epoxide-opened analogues of 3, including 4, were uni-
formly inactive (see Section III.B for a discussion of bioactivity).
TM
2. Avermectin B1 3,4-Epoxide
Although disappointed by the inactivity of epoxide-opened derivatives of 3, we
were encouraged by the activity of intact 3 and decided to explore the possibility
of preparing additional avermectin epoxides. Since 3 had been prepared by util-
izing a hydroxyl-directed epoxidation, we felt that it might be possible to ap-
ply similar chemistry to epoxidize the 3,4-double bond of 1. In fact, once the
C-7 hydroxyl group was blocked as the trimethylsilyl (TMS) ether 6 (Fig. 3) by
persilylation of 1 with bis(trimethylsily)trifluoroacetamide (BSTFA) followed by
selective hydrolysis of the secondary TMS ethers, the C-5 hydroxyl group effi-
ciently directed epoxidation of the 3,4 double bond. The β-3,4-epoxide 7 was
thus readily prepared [9].
As observed with the 8,9-epoxide, the 3,4-epoxide 7 also reacted readily with
strong nucleophiles to afford inactive ring-opened adducts (e.g., 8). In another par-
TM
TM
TM
nor, in this case a glycosyl fluoride, was not involved in the reaction. However,
both SnCl2 and AgClO4 were required, for reasons that we do not currently under-
stand. Furthermore, the reaction was independent of C-13 stereochemistry, but
was greatly accelerated by the presence of a TMS ether at C-7, possibly as a
result of relief of steric strain. It was subsequently shown that 13 also produced
15 under the glycosidation conditions. Recall that we had previously observed
a similar accelerating effect of a 7-OTMS ether in the rearrangement reaction of
avermectin B1-8,9-oxide derivative 9 described.
2. Disaccharide Excision
Although the fragmentation and rearrangement reactions discussed previously
are potentially useful as sources of intermediates for avermectin synthesis, we
did not utilize them for that purpose. We did, however, apply the concept of using
a natural avermectin as a source of synthetic intermediates when we required a
source of the avermectin disaccharide. By modifying the multistep procedure
developed by Hanessian et al. [17], we were able to prepare the desired disaccha-
ride 17 successfully from bis-silylated avermectin B1 (16) on a multigram scale
in 57% yield via a one-pot procedure (Fig. 7) [18]. Reaction of disaccharide 17
with diethylaminosulfur trifluoride (DAST) provided the glycosyl fluoride 18.
This convenient preparation of large quantities of the disaccharide proved to be
essential in our later synthetic studies, as described later.
TM
we were able to study the effect of modifying the linkage between the avermectin
aglycone and the disaccharide. Starting with ivermectin aglycone 19, we intro-
duced a variety of spaces (e.g., the ethylene glycol spacer shown in Fig. 8), then
attached the disaccharide to the new spacer to afford, after deprotection, ‘‘spacer-
mectin’’ analogues such as 23 and 24 [19]. Interestingly, some of the ‘‘spacer-
mectins’’ had activities approaching those of the natural compounds.
TM
active as the natural isomers, but also had a substantially better therapeutic index,
which was due to their reduced mammalian toxicity [12].
2. 19-epi-Avermectin B1 (36)
Encouraged by the outstanding activity of the 13-epi-avermectins,we decided to
undertake the synthesis of 19-epi-avermectin B1 (36) (Fig. 10). The synthesis of
36 was complicated somewhat by the propensity of the avermectin C-3,4 double
bond to move into conjugation with the lactone carbonyl under basic conditions.
Thus, although formation of the seco-acid 32 and inversion of the C-19 stereo-
chemistry via Mitsunobu lactonization to give 33 are relatively straightforward,
reestablishment of the C-2 stereocenter is more difficult. After considerable ex-
perimentation, we were able to accomplish the deconjugation successfully by
TM
III. BIOLOGY
Although much interesting chemistry was discovered during this avermectin proj-
ect, our primary interest was in the biological activity of the novel avermectin
analogues that we prepared. The next section describes some of our efforts to
measure biological activity and discusses the interesting structure–activity rela-
tionships that were uncovered.
TM
TM
TM
IV. CONCLUSIONS
The most enjoyable aspects of the avermectin project were the rich diversity of
the chemistry and the multidisciplinary nature of the project, which incorporated
elements of both chemistry and biology. In this respect, the avermectin project
at Merck resembled many of the projects in Professor Nakanishi’s lab at Colum-
bia. In both cases, effective collaboration with scientists in other fields was not
only critical to success, but also a key element in making the project a lot more
fun for all concerned.
ACKNOWLEDGMENTS
I would like to thank Professor Koji Nakanishi for a highly educational and very
enjoyable postdoctoral experience. My years in his laboratory were the best possi-
ble preparation for the interdisciplinary nature of research in the pharmaceutical
industry. I am also indebted to the many excellent scientists at Merck who were
involved in the avermectin project, especially Dr. Helmut Mrozik and Ms. Gaye
Margiatto.
REFERENCES
TM
TM
Harold V. Meyers
New Chemical Entities, Inc., Framingham, Massachusetts
I. INTRODUCTION
TM
This section discusses some of the key biological and chemical methodologies
available for the creation of molecular diversity. In particular, the molecular bio-
logical generation of peptide and protein libraries by phage display techniques
is discussed; then an overview of chemical strategies and methods developed for
the combinatorial synthesis of peptide and small molecule libraries used in drug
discovery and lead optimization processes is presented. Several good review arti-
cles describing molecular biological and chemical methodologies are available
for a more comprehensive discussion of these topics [2–6].
A. Biological Approaches
Several techniques for creating large libraries of filamentous phage clones dis-
playing unique peptides or proteins on their protein coat surface have been devel-
oped [7–11]. By inserting fully random cassettes of synthetic oligonucleotides
into targeted loci, tens of millions of mutant strains can be isolated [12–15].
Each unique peptide is structurally encoded by its respective single-stranded viral
deoxyribonucleic acid (DNA) sequence and can be expressed as copies as a fusion
TM
TM
B. Chemical Approaches
Early chemical combinatorial efforts were focused on synthesizing libraries of
peptides [27], followed by the progression to peptidelike oligomeric libraries
[28], and most recently to small-molecule, nonpeptidyl libraries [29,30]. The use
of solid-phase chemistry, originally developed for peptide synthesis [31] and later
amply demonstrated for organic synthesis [32–34], has experienced a renaissance
of late spurred by combinatorial chemistry. Solid-phase synthesis precludes often
TM
1. Synthesis of Mixtures
Combinatorial libraries can be synthesized and screened as mixtures. The synthe-
sis of large mixtures of compounds (105 –108) was made possible by employing
the split synthesis approach on solid supports.
a. Split Synthesis (Split-Pool, Split-Mix, or Portion Mixing). Furka first dem-
onstrated the split-mix strategy on solid supports to synthesize large mixtures of
peptides in equimolar quantities rapidly [35]. This one bead/one unique peptide
approach is schematically represented (Fig. 1). A solid support material is split
into equal-sized portions and placed in separate reaction vessels. Each portion is
coupled with excess building blocks (e.g., unique amino acid derivatives), which
provide uniform coupling since competition between reactants is eliminated. The
coupled resins are then pooled together into one reaction vessel for the removal
of common protecting groups. The resin is partitioned again into separate pools,
each containing mixtures of unique peptides, for coupling new reactants. This
iterative split-couple-mix strategy theoretically enables libraries of oligomers of
Xn members to be assembled, where X is the number of monomers, amino acids
in this example, in the basis set and n is the length of the oligomer. A pure
statistical distribution of sequences results.
When screening split-pool libraries on solid supports, bioactive compounds
can be physically isolated through incubation with a tagged ligate. Polymeric
beads possessing peptides bound to a ligate have been identified by visual inspec-
tion, physical removal, and microsequencing of the active peptides [36]. Alterna-
tively, the supported peptides are cleaved into solution for screening. When
screening mixtures of compounds in solution from split-synthesis libraries, sev-
TM
TM
TM
TM
TM
TM
TM
Figure 4 Synthetic strategy for the ‘‘double combinatorial’’ approach to the biphenyl
library.
TM
III. CONCLUSIONS
TM
REFERENCES
1. P. van Eikeren, in Chiral Separations, Applications and Technology (S. Ahuja, ed.)
American Chemical Society, Washington, D.C., chapter 2 (1997).
2. W. H. Moos, G. D. Green, and M. R. Pavia, Annu. Rep. Med. Chem., 28: 315 (1993).
3. M. R. Pavia, T. K. Sawyer, and W. H. Moos, Bioorg. Med. Chem. Lett., 3: 387
(1993).
TM
TM
TM
TM
Colin J. Barrow
The University of Melbourne, Parkville, Victoria, Australia
Ian F. Musgrave
Prince Henry’s Institute for Medical Research, Clayton, Victoria,
Australia
I. INTRODUCTION
TM
A. I1-Imidazoline-Binding Sites
The I1-imidazoline-binding sites are characterized by having nanomolar affinity
for clonidine (3), naphazoline (2), cirazoline (4), and idazoxan (10), with cloni-
dine having a higher affinity than idazoxan [7,15–18] (see Table 2). Moxonidine
(7) and rilmenidine (8), recently developed antihypertensives, are relatively selec-
tive for the I1-imidazoline-binding sites over I2-imidazoline-binding sites and α2-
adrenoceptors, although the exact degree of selectivity varies in different reports
[2,15,17]. There may be more than one form of I1-imidazoline-binding site; for
example, human I1-imidazoline-binding sites differ from bovine sites in having
lower affinity for guanabenz (13) and moxonidine [2,15–18]. The binding of
ligands to the I1-imidazoline-binding sites is strongly inhibited by cations such
as K⫹ and tetraethyl ammonium (TEA) [7,15] (see Table 1).
I1-Imidazoline-binding sites are mostly distributed in the brain (corpus stri-
atum, hippocampus) and the brain stem (medulla oblongata, rostral ventrolateral
medulla) [2,11]. I1-Sites have also been found in the prostate [19] and the kidney
[16], in both cases localized to the epithelium.
Virtually nothing is known about the structure of I1-imidazoline-binding
sites. They are present in the plasma membrane, and although some investigators
have reported that [3H]-clonidine binding to these sites is modulated by guanosine
triphosphate (GTP) suggesting a G-protein-linked receptor [19,20], to date there
is no consensus [7,15,21]. Furthermore, binding of ligands to I1-sites does not
TM
Non-I1, non-I2-imidazoline-
I1-Imidazoline-binding sites I2-Imidazoline-binding sites binding sites References
Selectivity Cirazoline ⫽ clonidine ⫽ napha- Cirazoline ⫽ idazoxan ⬎⬎ cloni- Various, often idazoxan ⱖ 6,7,14,74,77,
zoline ⬎ idazoxan ⬎ phentol- dine ⬎ naphazoline ⫽ phentol- clonidine 80,81
amine ⬎⬎⬎ adrenaline amine ⬎⬎⬎ adrenaline
Cation 4-AP ⬎ CsCl ⫽ TEA ⬎ RbCl 4-AP ⬎ CsCl ⬎ TEA ⬎ RbCl Unknown 7,15,77,81
sensitivity ⬎ KCl ⬎ LiCl ⬎ NaCl ⬎ KCl ⬎ NaCl ⬎ LiCl
Location Plasma membrane Outer mitochondrial membrane Unknown 7,40,80,82
G-protein Unknown No Unknown 77,78,79,81,
linked 82
Subtypes Probably A,B Probably a number of sites 7,77,80,81,
82
Structure Unknown Monoamine oxidase, other? Unknown 41
Tissue location Brain stem, basal ganglia, hippo- Adipose tissue, liver, placenta Brain, stomach 2,11,12,16,19,
campus, prostate, kidney 14,37,78,81
Action Modulation of blood pressure Growth factor–like effects? Unknown, modulation of 29,30,31,33,
Modulation of ocular pressure? gastric secretion? 34,44,71
Modulation of gastric secretion?
Modulation of renal secretion?
a
TEA, tetraethyl ammonium.
TM
TM
a
Ki Values are nanomolar.
b
Amiloride distinguishes between two forms of the I2-imidazoline, binding site, one with a high
affinity for amiloride (I2A) and one with a low affinity (⬎1000 nM: I2B); n.d., Not determined.
TM
TM
TM
TM
TM
TM
CDS was not agmatine, again indicating the presence of multiple CDS com-
pounds in multiple tissue types [50].
Thus, agmatine and other CDS compounds appear to be widely distributed
in mammalian tissue. Agmatine present in the stomach appears to be contained
in endothelium and vascular smooth muscle and so may play a role in modulating
gastric secretion [66]. However, there is substantial variation in the quantity of
agmatine in individual animals [66], greater than 10-fold in some cases, raising
concern about a functional role for agmatine. Furthermore, the imidazoline-bind-
ing site in the stomach appears to be a non-I1, non-I2-site [14].
TM
TM
TM
TM
TM
TM
TM
TM
REFERENCES
TM
TM
TM
TM
Valeria Balogh-Nair
The City College of the City University,
New York, New York
I. INTRODUCTION
Over the last decade, the incidence of fungal infections has increased dramati-
cally. This can be attributed to drug resistance, the emergence of new pathogens
and resurgence of old ones, and the lack of effective therapeutics. As patients
become severely immunocompromised because of underlying diseases such as
leukemia or acquired immunodeficiency syndrome (AIDS), and as immune sup-
pression is becoming routine in cancer and organ transplantation patients, oppor-
tunistic pathogens have begun preying on a growing population, causing life-
threatening systemic fungal infections. Candida species, the ordinarily harmless
denizens of the gastrointestinal and genitourinary tracts, are now becoming one
of the largest threats in hospitals. Infections by Candida albicans and other oppor-
tunistic fungi are often the most devastating in AIDS patients [1]. Pneumocystis
carinii pneumonia [2] affects ⬎70% of AIDS patients and is one of the most
common lethal infections among them. Cryptococcus neoformans causes poten-
tially fatal meningitis [3] in 7%–10% of AIDS patients and Candida albicans
causes oral and esophageal candidiasis [4] in ⬎70% of them. In this chapter,
focus will be on the development of novel oxidoredox pharmacophores and their
effectiveness against the three selected pathogens, C. albicans, C. neoformans,
and P. carinii.
TM
TM
TM
the target proteins that evoke the functional response. It is also understood that
coordination to the enzyme-bound ferrous heme by •NO triggers guanylate cy-
clase-catalyzed generation of cyclic guanosine monophosphate (GMP) from gua-
nosine triphosphate (GTP). The increases in intracellular second messenger cyclic
GMP concentrations are then responsible for the cellular responses, such as vas-
cular smooth muscle contraction. However, there is no consensus concerning the
role of •NO in the immunological response leading to the microbicidal effects
and accompanying tissue injury. Along with the mechanisms of action associated
with •NO itself, it has been postulated that redox-related forms of •NO play a
major role. Beckman et al. [17] were the first to point out that under pathophysio-
logical conditions, both •NO and superoxide anion (O •⫺ 2 ) are produced by macro-
phages at high rates. These can react to form the more potent oxidant peroxynitrite
(ONOO⫺), the conjugate base of peroxynitrous acid.
Peroxynitrite is known to react with a variety of biologically important
molecules. Nitric oxide or O •⫺ 2 reacts only slowly with the iron–sulfur cluster in
mitochondrial aconitase, a major target of oxidant-mediated toxicity in cells, but
peroxynitrite rapidly inactivates it [18]. Peroxynitrite oxidizes sulfhydryls, ascor-
bate, α-tocopherol, and lipids. Nitration of phenolic compounds, such as tyrosine,
in a metal-catalyzed process also can occur. Therefore, peroxynitrite is considered
sufficiently reactive to mediate nitric oxide–dependent microbial killing. During
macrophage activation, its precursors •NO and O •⫺ 2 are formed simultaneously,
in high concentrations, and peroxynitrite is formed at almost diffusion-controlled
rates (6.7 ⫻ 10 9 M⫺1s⫺1). This indicates that significant concentrations of peroxy-
nitrite should be attainable in vivo [19,20]. The observation that superoxide dis-
mutase (SOD), a known quencher of superoxide, enhances •NO levels in vivo,
and that nitrotyrosine is detected in biological fluids that produce high levels of
•
NO, casts peroxynitrite in a key role. However, these lines of evidence attributing
to peroxynitrite a key role in vivo are considered by some as circumstantial.
TM
TM
TM
TM
TM
A. Sulfonyloxaziridines
Sulfonyloxaziridines, known for two decades, are among the most versatile in
the repertory of oxidizing agents [41]. They are best known as the reagents of
choice for chiral oxidations, but it is other aspects of their chemical characteristics
that work in their favor in the context of pharmacophore development. Because
the oxygen in sulfonyloxaziridines is more electrophilic than is the oxygen in
oxaziridines, sulfonyloxaziridines, with few exceptions, are more reactive to
nucleophiles than are oxaziridines. Oxygen transfer by sulfonyloxaziridines
(Scheme 2) involves S N 2-type displacement of the sulfimine, facilitated by a rela-
tively weak oxygen–nitrogen bond and by the favorable enthalpy of the C N
π bond formed, with a slight bias in favor of a planar rather than a spiro transition
state. Their reactivity is modulated by the substituents attached to the carbon and
Scheme 2
TM
TM
TM
B. Oxaziridines
Despite an idea advanced in the early seventies that the critical oxygenating spe-
cies of flavin oxygenases [50] is an oxaziridine, before our studies [51] there were
no efforts to explore the pharmacophore potential of the oxaziridine functionality.
These oxygenases are unique in that they carry out the only known non-metal-
ion-requiring oxygen activation reactions in biological systems. They bind and
activate molecular oxygen, ultimately transferring one oxygen atom to substrate
and releasing the second as water. On the basis of mechanistic studies of these
enzymes, in 1974 Orf and Dolphin [50] proposed an oxaziridine intermediate as
the monooxygenating species, derived by rearrangement of an initial intermedi-
ate, 4α-hydroperoxyflavin (Scheme 6). Working with nonenzymatic models, Ras-
tetter et al. [52] proposed that a nitroxyl radical derived from Dolphin’s oxaziri-
dine (Scheme 6) is a viable candidate for the oxygenating species. However,
many other mechanisms were proposed later, not involving the oxaziridine, and
these received more attention [43].
The lack of interest in possible biological roles for oxaziridines can be
attributed to the fact that at the time of the mechanistic proposals discussed not
much was known about the chemical properties of of oxaziridines. Despite the
substantial number of papers published since then, many details in the mechanism
TM
of oxygen transfer from structures where the active site oxygen is part of a three-
membered ring, e.g., in metal peroxides, dioxiranes, oxaziridines, and sulfonylox-
aziridines, remain obscure. The similarity in the active site structures, however,
suggests that they may have a common mechanism of oxygen transfer, an S N2-
type displacement by the nucleophilic substrate (Z) on the electrophilic oxygen
atom (Scheme 7). Kinetic studies of deoxygenation of oxaziridines and sulfony-
loxaziridines [53], epoxidation of alkenes by sulfonyloxaziridines [54], and theo-
retical calculations on the transfer of oxygen from oxaziridines to ethylene [55]
support this mechanism.
Oxaziridines are active oxygen compounds. They transfer their ring oxygen
atom to tertiary phosphines and oxidize HI, but they are considered to be poor
reagents to oxidize sulfides to sulfoxides, or to epoxidize alkenes. However, their
reactivity is strongly dependent on the substitution pattern of the oxaziridine.
Thus, a bis-oxaziridine oxidized thiacycloalkanes to sulfoxides in only 5%–7%
yields, but an N-tert-butyloxaziridine converted dimethylsulfide to the sulfoxide
quantitatively [41]. An oxaziridine substituted with electron withdrawing groups,
2-(trifluoromethyl)-3,3-difluorooxaziridine, epoxidized alkenes under extremely
mild conditions (⫺50°C, ⬃1 h), and perfluorodialkyloxaziridines performed hy-
droxylation of unactivated tertiary aliphatic C-H bonds at room temperature, in
high yields [56,57]. Moreover, aza-aromatic N-oxides, when irradiated with ultra-
violet (UV) light, transfer their oxygen atom. It was proposed that oxaziridines,
formed by photoisomerization of the N-oxides, are the oxygen transfer agents,
TM
C. Nitrones
Nitrones could directly modulate the levels of oxidoredox species important in
the biocidal action of macrophages and neutrophils by preferential trapping of
one or more of the biocidal species. However, it is more likely that a complex
series of events takes place, in which nitrone-derived radicals and/or nitrone-
derived hydroxylamines are major contributors to the biocidal effects:
1. Nitrones, by undergoing ring closure to the corresponding oxaziridines,
could act as precursors of these efficient oxygen transfer agents dis-
cussed in Section IV.B.
2. Nitrones are well known spin traps [59]. Depending on their structure,
they can react with C-, O-, N-, and S-centered radicals, including those
produced by macrophages and neutrophils, to modulate their in vivo
concentrations (Scheme 8). Recently, efficient spin trapping of O-, C-,
and S-centered radicals and peroxynitrite, but not superoxide, was re-
ported when using 2H-imidazole-1-oxides as spin traps [60]. Because
the radical trapping reactions of nitrones yield stable nitroxyl radicals,
these radicals might participate in controlling the levels of oxidoredox
species present in vivo by serving as traps for short-lived radicals. On
the other hand, stable nitroxyl radicals are SOD and catalase mimics,
Scheme 8
TM
Scheme 10
TM
TM
D. Macrocyclic Amines
The rationale for selecting macrocyclic amines as potential microbicidal pharma-
cophores is based on consideration of the reactions of amines with the potent
oxidants produced by macrophages and neutrophils. The macrocyclic structure,
as carrier of the amine functionalities, was chosen because such structures can
TM
From the limited set of macrocyclic amines tested to date, it is not likely that
N-oxidation of tertary amines confers antifungal activity. Another possibility is
that biologically abundant metal nitrosyl complexes acting as electrophilic ni-
trosating agents transfer NO⫹, thus converting the amines to N-nitrosamines.
However, the biological relevance of NO⫹ under physiological conditions has
been disputed [67]. An alternate, and most likely explanation for the mode of
action of macrocyclic amines is that they function as prodrugs; that is, they are
converted by the HOCl of neutrophils into N-Cl derivatives, potent and long-
lasting oxidizing agents. It is interesting to note that two marine alkaloids, pa-
puamine and haliclondiamine, containing unusual 13-membered macrocyclic
rings encompassing 1,3-diaminopropane moieties, display significant antifungal
activity against C. albicans. However, the recent total synthesis of papuamine
by Weinreb’s group [68] required 16 steps, whereas the macrocyclic imines re-
ported here, highly active against C. albicans, can be obtained from commercially
available starting materials, in 2 steps, in ⬎70% overall yields.
E. Metallomacrocycles
The rationale for employing metallomacrocycles as antifungals was based on
consideration of the reactivities of their biological counterparts, as follows:
TM
Scheme 12
TM
Scheme 13
TM
Of the five pharmacophores selected for synthesis, at least one compound in each
class showed high activity against P. carinii, except the macrocyclic polyamines,
which showed activity only against Cryptococcus neoformans and Candida albi-
cans. Further, we recently established, in the case of the metallomacrocycle phar-
macophore VBN-10, active against P. carinii, that the in vitro activity closely
correlates with in vivo efficacy. Here we summarize the syntheses of the lead
compounds and their respective activities:
TM
A. Synthesis of Sulfonyloxaziridines
Finding novel structural classes, hitherto not employed as therapeutic agents, is
usually difficult and seldom attempted. We overcame this hurdle by synthesizing
the lead compounds, a sulfonyloxaziridine podand, VBN-1, and the macrobicy-
clic oxaziridine, VBN-2 (see Section V.B). However, since these are truly new
structural leads and thus have not been employed in any drug development, it is
necessary to determine which types of molecular structures are best as vehicles
for delivery of the active pharmacophore.
Podand VBN-1 contains two sulfonyloxaziridine moieties; it was obtained
in 92% yield by m-CPBA (meta-chloroperbenzoic acid) oxidation of the corre-
sponding sulfonimine [76] in the presence of a phase transfer catalyst [77]
(Scheme 14). It is a stable white solid that can easily be obtained on ⬎10 g scale.
The two oxaziridine moieties in VBN-1 have identical geometry, as ascertained
by spectral data. In vitro, VBN-1 was active against P. carinii, at a concentration
of 2.1 µM, comparable to that of pentamidine. In vivo testing in mice indicates
it has no gross toxicity at 50 mg/kg/day intraperitoneally the highest concentra-
tion tested to date.
B. Synthesis of Oxaziridines
Since oxaziridine as a pharmacophore was unknown prior to our studies, in the
first attempt to establish its usefulness against P. carinii we prepared a compound
containing several oxaziridine units, so as to have as many active oxygens as
possible delivered per mole. We were aware of the formidable difficulties, since
one can theoretically obtain a large number of isomers. To gain some control
over the stereochemistry we chose a rigid macrocyclic framework as carrier for
the oxaziridine groups. The hexaimino macrobicyclic precursor of VBN-2, which
the spectral and x-ray crystallographic analysis showed to contain six trans-im-
ines [78], was deemed to be a good candidate. Inspection of its structure generated
with MacroModel and those of two related hexaimino macrobicycles allowed us
to visualize the likely mode of attack of the peracid on these imines. Even though
the initial trans-geometry of the imines is essential, this alone was not expected
to provide stereocontrol over the production of the oxaziridines, since in nonmac-
TM
TM
TM
C. Synthesis of Nitrones
During our studies on retinoids, we synthesized highly conjugated open chain
nitrones [85] and found that this conversion reduced the teratogenicity of reti-
noids [86,87]. However, we decided to try a less conjugated molecule containing
two nitrone groups and an aromatic ring to employ as a pharmacophore against
P. carinii. Bis-(amidine oxides) were very attractive as target structures for test-
ing the nitrone pharmacophore. Nitrones are known to react with the biologically
important radicals, which we believe to be particularly relevant to the inhibition
of P. carinii (e.g., •OH, •OOH, O ⫺• •
2 , and NO), and are thereby themselves con-
verted to more stable aminoxyl (nitroxyl) radicals. These aminoxyls elicit height-
ened interest as potential therapeutic agents, spin traps, and synthetic intermedi-
ates [88]. The aminoxyl radicals generated from the reaction of nitrones are in
themselves of considerable interest because their drug potential is enhanced, as
evidenced by the observation that they are typically quite nontoxic in vivo [59].
Nitronyl nitroxides were also of interest because they react specifically with nitric
oxide [89], an important product of iNOS in macrophages, to yield imino nitroxyl
radicals.
We prepared the target nitrones, VBN-3 and VBN-4, and nitronyl ni-
troxides, VBN-3-ox and VBN-4-ox, by condensation of 2,6-pyridinedicarboxal-
dehydes and 1,4-benzenedicarboxaldehydes with 2,3-bis(hydroxylamino)-2,3-
dimethylbutane sulfate followed by oxidation of the condensation products
according to the procedure of Ullman [90,91]. In the case of 2,6-pyridinedicar-
boxaldehyde, yellow crystals of two products, VBN-3 and VBN-3′, were ob-
tained in 85% and 5% yields, respectively. The structures of both were ascer-
tained by UV, infrared (IR), 1 H, and 13 C NMR spectra and by MS. Interestingly,
these conformers do not interconvert in solution at room temperature and can be
readily distinguished by the low-field region 1 H NMR spectra. In VBN-3′ only
one of the pyridine protons (C-5H) shows the typical deshielding (δ 9.5 ppm)
[92] by the nitrone oxygen, whereas both 3 and 5 protons are deshielded in VBN-
3 (δ 9.44 ppm). In VBN-3′ the pyridine-3-proton appears at δ 8.21 ppm.
TM
TM
hydrogen bond formation favored ring closure. Lehn was among the first in 1987
[96] to report efficient syntheses of macrocyclic Schiff bases via template-free
[3 ⫹ 2] condensations. Since that time, diverse needs for metal-free macrocyclic
ligands led to the development of methodologies to obtain them by simple [2 ⫹
2] and [3 ⫹ 2] macrocyclizations. We synthesized the Schiff base macrocycles,
such as the macrobicyclic imine precursor of VBN-2, and a series of iminomacro-
cycles, such as VBN-5 to VBN-9, and the macrocyclic amide precursor of VBN-
10 via these nontemplate syntheses [51,77]. These, together with the syntheses
of aminomacrocycles screened for antifungal activity against Candida albicans
and Cryptococcus neoformans, result to date in over 50 macrocycles made in
template-free manner. The success of these syntheses (yields ⬎ 50%), in which
we obtained the metal-free macrocycles by direct [3 ⫹ 2] and [2 ⫹ 2] condensa-
tions (step a, Scheme 16; Scheme 17), depends on the careful selection of inher-
TM
TM
TM
TM
taining unlike functionalities that render them chemically very different, each of
these compounds is able to cause oxidative damage to P. carinii; i.e., the only
common feature of these compounds, which are structurally so widely dissimilar,
e.g., VBN-10, molecular weight of 1412, a metallomacrocycle, and VBN-3, a
relatively small, molecular weight (MW) ⬍ 400 aromatic entity, is that all are
oxidizing agents. This gives credence to our hypothesis, the basis of our design
of these compounds, namely, that P. carinii is susceptible to oxidative damage.
The in vivo studies discussed in Section VI.C confirmed this.
TM
Analogue Formula 1 2 1 2
1 C 28 H 38N6O4 5/⬎5 5/5 ⬎5/⫺ ⬎5/⫺
2 C 32 H 46N6O4 6.25/⬎6.25 6.25/6.25 6.25/⬎6.25 6.25/⫺
3 C 28 H 46N6O4 5/⬎5 5/5 ⬎5/⫺ ⬎5/⫺
4 C 32 H 54N6O4 ⱕ0.3/ⱕ0.3 ⱕ0.3/0.6 ⱕ0.3/0.6 0.6/2.5
5 C 34 H 50N6O4 ⬎5/⫺ 5/5 ⫺/⫺ ⬎5/⫺
6 C 28 H 38N6 ⬎5/⫺ ⬎5/⫺ ⫺/⫺ ⫺/⫺
7 C 40 H 46N6 5/⬎5 2.5/2.5 5/⬎5 ⬎5/⫺
8 C 30 H 42N6 ⬎5/⫺ 5/⬎5 ⫺/⫺ ⫺/⫺
9 C 44 H 54N6 5/5 2.5/2.5 ⫺/⫺ 5/5
10 C 32 H44N4O6 ⬎5/⫺ 5/⬎5 ⬎5/⫺ ⬎5/⫺
11 C 28 H 46N6 0.6/1.25 1.25/1.25 1.25/2.5 1.25/1.25
12 C 46 H 50N6 5/⬎5 1.25/2.5 ⬎5/⫺ ⬎5/⫺
13 C 40 H 38N6 5/⬎5 ⱕ0.3/2.5 ⬎5/⫺ ⬎5/⫺
14 C 44 H 46N6 ⬎5/⫺ 5/5 ⫺/⫺ 5/⬎5
15 C 44 H 44N4O 2 ⬎5/⫺ 2.5/2.5 ⫺/⫺ 2.5/⬎5
16 C 44 H 48N8 ⬎5/⫺ 1.25/2.5 ⫺/⫺ 5/⬎5
17 C 26 H 32N4O6 ⬎5/⫺ ⬎5/⫺ ⫺/⫺ ⫺/⫺
18 C 16 H 26N4O 2 ⬎5/⫺ ⬎5/⫺ ⫺/⫺ ⫺/⫺
19 C 18 H 30N4O4 ⬎5/⫺ ⬎5/⫺ ⫺/⫺ ⫺/⫺
(data not shown), one of the macrocyclic compounds showed encouraging activ-
ity in in vitro tests. Since this active compound could be synthesized in two easy
steps, and in high yields, using the methodology we developed, we prepared 19
analogues and tested them against these fungal species. The results of the assays
of these analogues, compounds 1–19, are shown in Tables 2 and 3.
Compound 4 (Table 2) showed activity against C. neoformans strains 1
and 2 at a minimum inhibitory concentration (MIC) ⱕ0.5 µM (ⱕ0.3 µg/ml).
This level of activity compares favorably with that of fluconazole, the positive
control, especially, since an end point was not established; i.e., the lowest concen-
tration tested was 0.5 µM. Compound 11 (Table 3) was found to be fungistatic
against C. albicans strains A–E at a MIC ⱕ 0.5 µM (ⱕ0.3 µg/mL) and also
showed both fungistatic and fungicidal activity, though more modest, against C.
neoformans strains 1 and 2 at a MIC of 1.3–2.6 µM (0.6–1.25 µg/mL) (Table 2).
Because of the high levels of in vitro activity, the syntheses of both compounds 4
and 11 were scaled-up to gram amounts for in vivo assays on mice. Initial toxicity
studies are encouraging, since all mice survived for at least 1 month when daily
doses of 15 mg/kg of compound 11 and 10 mg/kg of compound 4 were given
TM
No. Formula A B C D E A B C D E
1 C 28 H 38N6O4 ⬎5/⫺ ⬎5/⫺ ⬎5/⫺ ⬎5/⫺ ⬎5/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺
2 C 32 H 46N6O4 6.25/⬎6.25 ⬎6.5/⫺ 6.25/⬎6.25 6.25/⬎6.25 6.25/⬎6.25 ⬎6.5/⫺ ⫺/⫺ ⬎6.5/⫺ ⬎6.5/⫺ ⬎6.5/⫺
3 C 28 H 46N6O4 ⬎5/⫺ ⬎5/⫺ ⬎5/⫺ ⬎5/⫺ ⬎5/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺
4 C 32 H 54N6O4 ⬎5/⫺ ⬎5/⫺ 5/⬎5 ⬎5/⫺ ⬎5/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺
5 C 34 H 50N6O4 ⬎5/⫺ ⬎5/⫺ ⬎5/⫺ ⬎5/⫺ ⬎5/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺
6 C 28 H 38N6 ⬎5/⫺ ⬎5/⫺ 5/⬎5 ⬎5/⫺ ⬎5/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺
7 C 40 H 46N6 ⬎5/⫺ ⬎5/⫺ 5/⬎5 ⬎5/⫺ ⬎5/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺
8 C 30 H 42N6 ⬎5/⫺ ⬎5/⫺ 5/⬎5 ⬎5/⫺ ⬎5/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺
9 C 44 H 54N6 ⬎6.5/⫺ ⬎6.5/⫺ ⬎6.5/⬎6.5 ⬎6.5/⫺ ⬎6.5/⫺ ⫺/⫺ ⫺/⫺ ⬎6.5/⫺ ⫺/⫺ ⫺/⫺
10 C 32 H 44N6O6 ⬎5/⫺ ⬎5/⫺ ⬎5/⫺ ⬎5/⫺ ⬎5/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺
11 C 28 H 46N6 ⱕ0.3/5 ⱕ0.3/5 ⱕ0.3/5 ⱕ0.3/5 ⱕ0.3/5 ⬎5/⫺ 5/5 5/⬎5 5/⬎5 ⬎5/⫺
12 C 40 H 50N6 ⬎5/⫺ ⬎5/⫺ 5/⬎5 ⬎5/⫺ ⬎5/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺
13 C 40 H 38N6 5/⬎5 ⬎5/⫺ 5/⬎5 ⬎5/⫺ ⬎5/⫺ ⬎5/⫺ ⫺/⫺ ⬎5/⫺ ⫺/⫺ ⫺/⫺
14 C 44 H 46N6 5/⬎5 ⬎5/⫺ 5/⬎5 ⬎5/⫺ 5/⬎5 ⫺/⫺ ⫺/⫺ ⬎5/⫺ ⫺/⫺ ⬎5/⫺
15 C 44 H 44N4O 2 ⬎5/⫺ ⬎5/⫺ 5/⬎5 ⬎5/⫺ ⬎5/⫺ ⫺/⫺ ⫺/⫺ ⬎5/⫺ ⫺/⫺ ⫺/⫺
16 C 44 H 48N8 5/⬎5 5/⬎5 5/⬎5 5/⬎5 ⱕ0.3/⬎5 ⬎5/⫺ ⬎5/⫺ ⬎5/⫺ ⬎5/⫺ ⬎5/⫺
17 C 26 H 32N4O6 ⬎5/⫺ ⬎5/⫺ 5/⬎5 ⬎5/⫺ ⬎5/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺
18 C 16 H 26N4O 2 ⬎5/⫺ ⬎5/⫺ 5/⬎5 ⬎5/⫺ ⬎5/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺
19 C 18 H 30N4O4 ⬎5/⫺ ⬎5/⫺ 5/⬎5 ⬎5/⫺ ⬎5/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺ ⫺/⫺
TM
VII. CONCLUSIONS
TM
TM
The author thanks Drs. Nancy A. Roth and Ronald J. Roth for more than the
critical reading of the manuscript. She would also like to express her appreciation
for the contributions of her coworkers. Financial support from NIH, National
Institute of Allergy and Infectious Diseases grant ROI AI39418, is gratefully
acknowledged.
REFERENCES
TM
TM
TM
TM
I. INTRODUCTION
TM
TM
Figure 1 (a) Concerted model for the β-elimination catalyzed by enoyl-CoA hydratase.
Evidence supports the assignment of a glutamate residue as the active-site amino acid
responsible for both protonation and deprotonation. (b) Stepwise model for the reaction.
In this model, an acidic residue protonates the carbonyl and lowers the pK a of the α-
proton so a single basic residue can both abstract the proton and protonate the leaving
hydroxyl.
CEH bond to occur [8]. In the hydration direction, the driving force of this
catalysis has been attributed to the capability of crotonase to polarize the π-elec-
trons of the α,β-unsaturated double bond, thus enhancing the electrophilicity of
the β-carbon [9]. Consequently, the corresponding active-site acid/base pair plays
a ‘‘push–pull’’ role in the reverse (dehydration) reaction and promotes the α-H
abstraction and the β-OH cleavage.
Sequence alignments with other members of this family implicated a highly
conserved glutamate (Glu) residue (Glu164 in rat liver enzyme) as the catalytic
residue. In fact, substitution of this residue with glutamine (Gln) led to a dramatic
decrease of k cat by more than 100,000-fold [10]. Since the K m remained essentially
unaffected by the mutation, it was concluded that Glu164 assumes the role as
the active-site base to abstract the α-proton. A similar conclusion was reached
in studying the enoyl-CoA hydratase activity associated with the large α-subunit
of the multienzyme complex (79 kDa) of fatty acid oxidation of Escherichia coli
[11–14]. Again, a glutamate residue (Glu139) was identified as the catalytic resi-
due by site-directed mutagenesis [15]. Although additional work is necessary to
elucidate the mechanistic details, sufficient data have been collected to establish
that the hydratase activity in this multifunctional protein proceeds through a gen-
eral acid–general base mechanism with no required cofactors.
TM
TM
3. 3-Dehydroquinase
Enzymes with no cofactor requirements often recruit amino acid residues other
than general acids or bases to participate in the catalytic mechanism, and 3-dehy-
droquinase (DHQase, EC 4.2.1.10) is a good example. This enzyme catalyzes
the dehydration of 3-dehydroquinic acid (4) to 3-dehydroshikimic acid (5) in
the third step of the shikimate pathway, leading to the biosynthesis of aromatic
metabolites in microorganisms, fungi, and plants [27,28]. There are two distinct
classes of DHQase, designated type I and type II, and they exhibit different struc-
tural properties and mechanisms [29–31]. Chemical modification and differential
peptide mapping studies revealed that the type I enzyme, exemplified by the E.
coli enzyme, catalyzes a syn elimination of a water molecule [32] via an imine
intermediate linked to a conserved lysine residue (Lys170) [33,34], and it uses
a conserved histidine residue (His143) as a general base [35]. Subsequent experi-
ments with site-directed metagenesis of the monofunctional, homodimeric E. coli
enzyme [36] confirmed these conclusions, and the deduced mechanism is illus-
trated in Fig. 3. The covalent imine serves as an electron sink to stabilize the
carbanion intermediate after the pro-R hydrogen is abstracted from the C-2 posi-
tion. As a result of a proposed distortion of the carbocyclic ring of dehydroquinate
upon formation of the Schiff base [32], the critical CEH bond is apparently
reactive enough to be abstracted by His143. Interestingly, this histidine residue
Figure 3 Mechanism for 3-dehydroquinase (DHQase) from the third step of the shiki-
mate pathway in Escherichia coli.
TM
TM
on the specific isozyme and tissue type [40], though both the forward and reverse
reactions proceed through the highly reactive Zn-OH ⫺.
In spinach CA, and perhaps in other plant CAs, extended x-ray absorption
fine structure (EXAFS) studies have shown that the coordination sphere for the
zinc ion includes at least one sulfur ligand [47]. Though the exact ligand structure
is not yet known, plant CAs apparently have a quite different active-site organiza-
tion than mammalian CAs. Regardless of the structural differences, the kinetics
of spinach CA are consistent with the mammalian zinc-hydroxide mechanism,
including a rate-determining proton transfer step that requires the participation
of buffer for maximum efficiency [47]. Despite the lack of conclusive evidence
for a proton-shuttle system in the spinach CA [47], it appears that a similar mech-
anism is shared by all known CAs. Further structural studies of the convergently
evolved mammalian and plant enzymes will be instrumental in developing a
deeper understanding of the catalysis of CA. Sly and Hu [40] summarize the
information on this enzyme more comprehensively.
TM
TM
in 3-OH activation [56]. Notably, the relative positions of Lys345 and Glu211
in the Reed and Rayment proposal are consistent with the antistereochemistry
of the elimination reaction.
The x-ray crystal structure obtained by Larsen et al. [66] also revealed
that both essential divalent cations interact with the carboxylate of the substrate/
product. Together with Lys396, they appear to promote the redistribution of the
negative charge developed during the formation of the aci-carboxylate intermedi-
ate. Thus, a relatively weak active-site base, most likely assisted by significant
substrate orientation by the enzyme [68], can abstract a nonacidic (pK a ⬃ 28–32)
proton and expel a poor leaving group. Another example of this class of hydrolase
is galactonate dehydratase, which depends on Mg 2⫹ [69] and appears to utilize
His285 as the active-site base [70]. The Mn 2⫹-dependent imidazole glycerol phos-
phate dehydratase [71] may also be related to this mechanistic class of hydrolase.
The properties of the latter enzymes remain to be further characterized.
TM
constituted with 57 Fe revealed a single labile iron site, which is called Fe a [75,78].
Protein ligands of the nonlabile iron atoms are cysteine residues [79–82], and
the labile iron is ligated to a hydroxyl ion, which becomes protonated to water
when substrate is bound [81–84].
Results gathered from x-ray analysis [82], mutation experiments [85], and
inhibitor studies [86] support a general mechanism (amino acid residue number-
ing for porcine heart aconitase), that starts with substrate’s binding to the enzyme
and causes the active-site cleft to close. Then Arg452 and Arg580 act in concert
with His101 and His147 to bind and orient the α- and γ-carboxyl groups of citrate
(Fig. 8). At the same time, both the β-carboxylate and β-hydroxyl of citrate coor-
dinate to Fe a, thus expanding the coordination sphere to octahedral, and the iron-
bound hydroxyl group becomes protonated to form water. The proton to be ab-
stracted is aligned with serine (Ser642), which is in an oxyanion hole (Arg644)
and consequently deprotonated as a result of a significant lowering of pK a. An-
ionic Ser642 accepts the proton from the substrate and eliminates the hydroxide
via a carbanion intermediate [86] to form cis-aconitate. The positive charge on
Figure 8 Binding of citrate to the [4Fe-4S] cluster of aconitase and partial active-site
structure. (Source: Adapted from Ref. 74.)
TM
Fe a also assists in the dissociation of the hydroxyl group from the substrate. As
shown in Fig. 8, cis-aconitate, with its γ-carboxyl bound to Arg580, then flips
180° as a result of a conformational change [87]. Such a flip switches the positions
of the α- and β-carbons with respect to Fe a and allows addition of water onto
the α-carbon. The overall reaction is readily reversible.
2. Fumarate Hydratase
Fumarate hydratase (fumarase, EC 4.2.1.2) catalyzes the reversible hydration of
fumarate (14) to (S)-malate (15) in the citric acid cycle (Fig. 9). The mammalian
enzyme is extremely efficient and catalyzes the reaction with an apparent second-
order rate constant (k cat /K m ⫽ 2.4 ⫻ 10 8) approaching the diffusion limit [88,89].
Unlike fumarases of mammals, yeast, and Bacillis subtilis (class II), fumarase A
from E. coli (class I) is an iron–sulfur-containing hydrolase [90]. The active
fumarase A is a homodimer (60 kDa per monomer), and each subunit contains
a [4Fe-4S] cluster that can be oxidized to the inactive [3Fe-4S] state in the same
fashion as shown in Fig. 7 for aconitase [90]. The reduced [4Fe-4S] form of
fumarase A is much less active (10- to 50-fold) than the oxidized [4Fe-4S] 2⫹
form, and substrate binding dramatically affects the [4Fe-4S] ⫹ EPR signal. These
results are used as evidence to support a catalytic mechanism for fumarase that
parallels the one for aconitase [90]. Other class I fumarases may include E. coli
fumarase B [91], Euglena gracilis fumarase [92], and Rhodobacter capsulatus
fumarase [93]. It should be noted that class II fumarases bear little homology
to fumarase A and fumarase B from E. coli [94,95], and they have no cofactor
requirements, so these other enzymes do not function by the same mechanism.
TM
TM
ies support the development of a partial positive charge at β-C [101]. The sub-
strate must have the R-configuration at α-C, but DHAD surprisingly accommo-
dates either the 2R,3R or 2R,3S isomer with nearly equal efficiency. Solvent
hydrogen incorporation at the β-position proceeds with retention of configuration
[102]. Because DHAD from all sources examined thus far also requires a divalent
metal ion for activity [101], it would be interesting to find out whether the mecha-
nism of DHAD bears any similarity to those discussed in Section II.B as well
as to that of aconitase.
4. L-Serine Dehydratase
Recently, an l-serine dehydratase (EC 4.2.1.13) involved in both the biosynthesis
and metabolism of serine was purified from Peptostreptococcus asaccharolyticus
and shown to be a heterodimer of 55 kDa (α, 30 kDa; β, 25 kDa) containing a
[3Fe-4S] center [103,104]. The enzyme could be inactivated by exposure to air
and reactivated by incubation with Fe 2⫹ under anaerobic conditions, presumably
by regenerating an oxidized [4Fe-4S] 2⫹ center. Though EPR spectroscopic exper-
iments confirmed the presence of an oxidized [3Fe-4S] ⫹ center, direct evidence
for a [4Fe-4S] center is still lacking [104]. Instead, the purported [4Fe-4S] 2⫹
center was inferred by titration of the anaerobically purified enzyme with
K 3Fe(CN) 6 and observation of a successively larger [3Fe-4S] ⫹ EPR signal, which
correlated with a decrease in activity. Interaction of the substrate with the Fe-S
center was also inferred from the observation that l-serine could both shield the
[3Fe-4S] ⫹ center from buffer-dependent changes in the EPR signal and prevent
oxidative loss of activity when the active enzyme is exposed to air. This evidence
has been used to propose a mechanism in which the labile iron of the [4Fe-4S] 2⫹
center binds the hydroxyl and carboxyl groups of serine [104]. This iron atom
would thus serve as a Lewis acid and facilitate loss of the hydroxyl group, an
otherwise poor leaving group (Fig. 11). Such a mechanism is vastly different
from that for most l-serine dehydratases from nonbacterial sources, which utilize
PLP as the cofactor in the irreversible deamination of serine [105]. In the latter
case, the α-H and not the β-hydroxyl group is activated during catalysis.
TM
2. Nicotinamide-Adenine-Dinucleotide-Dependent
Dehydratases: TDP-Glucose 4,6-Dehydratase
Temporarily converting a hydroxyl group to a keto function via intramolecular
oxidation–reduction is another strategy adopted by biological systems for activat-
ing an adjacent CEH bond in CEO bond cleavage reactions. For example, this
TM
TM
TM
dehydratases, this unique group of catalysts includes a few other enzymes, such
as 3-dehydroquinate synthase (EC 4.6.1.3) [120,121] and 2-deoxy-scyllo-inosose
synthase [122].
TM
TM
40 before regenerating the FAD cofactor and completing the catalysis. Support
for the latter hypothesis derives from two primary sources: (1) There is a close
relationship with the (R)-2-hydroxyglutaryl-CoA dehydratase, as discussed in the
previous section, and (2) the FAD of 4-hydroxyglutaryl-CoA dehydratase is
readily reduced to the neutral semiquinone whereas reduction to the hydroqui-
none appears to be kinetically hindered [133]. Though the details of this mecha-
TM
nism still need to be elaborated and verified, it seems clear that the flavin coen-
zyme directly participates in the deoxygenation reaction, and it needs to be
oxidized to do so.
c. Chorismate Synthase. Another deoxygenation in which the cleavage of a
C-O bond may occur prior to the rupture of a C-H bond is catalyzed by chorismate
synthase (EC 4.6.1.4) in the seventh step of the shikimate pathway (also see
Section II.A.3). The enzyme from all known sources depends on a reduced flavin
coenzyme in the conversion of 5-enolpyruvylshikimate 3-phosphate (EPSP) (41)
to chorismate (42) [135–139], as shown in Fig. 16. Although this reduced flavin
is often generated by an alternative pathway, the enzyme from some sources is
bifunctional and can provide the reduced flavin via an additional diaphoraselike
activity using NADPH as the cosubstrate [28]. After the catalytic cycle, the flavin
cofactor remains reduced, so there is no net change in the oxidation state of the
enzyme.
Though the precise role of the flavin is unknown, studies using flavin and
Figure 16 Possible nonconcerted pathways for the chorismate synthase reaction. Cur-
rent data cannot distinguish between a cationic (top) and a radical (bottom) mechanism.
(Source: Adapted from Ref. 158.)
TM
E. Protein-Radical-Dependent Dehydratases
1. Diol Dehydrases
Though the aforementioned flavin-dependent enzymes may include a free-radical
intermediate in the reaction pathway, none of those reactions requires the forma-
tion of a protein radical as part of the catalysis. However, there are a few systems
in which the generation of a protein radical appears to be a prerequisite to abstract
a hydrogen atom from the substrate before deoxygenation can occur. The reac-
tions mediated by diol dehydrases and glycerol dehydrases are two likely exam-
ples [160]. These multimeric enzymes (⬃230 kDa) catalyze the irreversible dehy-
dration of vicinal diols into corresponding aldehydes or ketones [161] in the
anaerobic fermentation of glycerol and other glycols in several bacterial organ-
isms [Table I of Ref. 162]. Adenosylcobalamin (AdoCbl) is a required cofactor
of these enzymes [163], and the adenosyl radical [164] that is generated by the
homolytic cleavage of the reactive Co-C bond of AdoCbl [165] is the initiator
of the reaction. A monovalent cation, such as K ⫹, NH 4⫹, Tl ⫹, or Rb ⫹, is also
required to assist in the binding of AdoCbl to the enzyme [166,167]. In the cases
with methylcobalamin-dependent methionine synthase [168] and methylmalonyl-
TM
TM
fully elucidated and the protein-based radical has not been directly detected
in diol dehydrase, the participation of the adenosyl radical in the reaction is con-
clusive. However, a reactive protein radical may actually be responsible for
abstracting a relatively inert C-H bond of the substrate to initiate the deoxygen-
ation.
TM
2. Ribonucleotide Reductase
Of the many deoxygenation reactions that have been characterized, the conver-
sion of ribonucleotides to 2′-deoxyribonucleotides is perhaps the most prominent
because of its role in deoxyribonucleic acid (DNA) biosynthesis. Because of the
sheer importance of the reaction it catalyzes, ribonucleotide reductase (RNR)
has been studied extensively over the years, and the mechanism is now well
characterized. Ribonucleotide reductase utilizes a stable protein radical to cata-
lyze the irreversible, reductive replacement of the 2′-hydroxyl group with hydro-
gen in ribonucleotides. Interestingly, at least three classes (I–III) of RNR have
been discovered, and each employs different cofactors and protein radical–gener-
ating mechanisms, though the subsequent catalytic steps are the same. In all
known forms of RNR, radical-mediated abstraction of a nonacidic hydrogen atom
precedes C-O bond cleavage, similar to most anionic deoxygenation mechanisms.
Several reviews summarizing the work that has been done on this enzyme are
available [189–195], and information pertaining to cofactor requirements and the
mechanism is briefly presented.
Class I ribonucleotide reductases are characterized by a binuclear high-spin
Fe(III) complex and a tyrosyl radical, and these enzymes aerobically deoxygenate
ribonucleotide diphosphates. The 258-kDa (α 2β 2: α 2 usually known as R1 and
β 2 as R2) protein from Escherichia coli is the best studied member of this class
[196–202]. Crystal data of R2 confirmed the presence of a binuclear µ-oxo
bridged iron cluster that is close enough to stabilize the tyrosine–122 (Tyr122)
radical (Fig. 18) [203]. Scavenging of this Tyr122 radical by hydroxyurea [204]
or by the substrate analogue 2′-deoxy-2′-mercaptouridine 5′-diphosphate [205]
leads to irreversible inactivation of the enzyme. Extensive study of the spontane-
ous assembly of the iron cluster and the Tyr122 radical in the presence of Fe(II)
and O 2 revealed the involvement of a diferric Fe(III)/Fe(IV) species and a neutral
tryptophan–48 (Trp48) radical [206–211]. The Trp48 occupies a site in a hy-
drophobic region of R2 that is believed to interact with R1 [212]. It may also
be part of an uncharacterized electron relay, involving His118, Asp237, and
Trp48, that transfers the radical from Tyr122 in R2 to a critical protein residue
in R1 (a distance of 35 Å) where the substrate binds and reacts. This protein
residue is proposed to be Cys439 on the basis of mutagenesis studies [213] and
of sequence comparisons with AdoCbl-dependent RNR from Lactobacillus leich-
mannii, in which a kinetically competent thiyl radical (Cys408) was characterized
by EPR spectroscopy [188].
As shown in Fig. 19, deoxygenation is initiated by the abstraction of a
hydrogen atom from the 3′-position of the ribose ring by the thiyl radical. The
2′-hydroxyl group is then protonated by one cysteine of a redox-active thiol pair
(Cys225 and Cys462) and is lost as water, resulting in a resonance-stabilized
cation radical intermediate. Though such a substrate-based radical has not yet
TM
TM
TM
TM
TM
1. CDP-6-Deoxy-L-Threo-D-Glycero-4-hexulose-3-
Dehydrase (E1)
Enzyme E 1 is a red–brown protein consisting of two identical subunits (49 kDa
per monomer). Both radiometric [242] and fluorometric [243] quantification re-
vealed one equivalent of PMP per monomer for full activity. Analytical assays
detected stoichiometric amounts of iron and sulfur, and EPR analysis of the re-
duced enzyme confirmed the presence of an adrenodoxin/putidaredoxin-type
[2Fe-2S] center (g ⫽ 2.007, 1.950, and 1.930) [244]. It should be noted that
purified E 1 is a mixture of apo- and holo-enzyme, so exogenous PMP, iron, and
sulfur are required to reconstitute its activity fully. Incubation of the enzyme with
substrate revealed that the PMP of E 1 forms a Schiff base with the C-4 keto group
of the substrate [245]. Then, His220 abstracts the pro-S hydrogen from the C-4′-
position of the PMP in the Schiff base complex and triggers the expulsion of the
C-3 hydroxyl group of the substrate to give PMP-∆ 3,4-glucoseen complex 52 (Fig.
20) [243]. Although intermediate 52 has never been isolated, it is reversibly hy-
drated on the si face of the conjugated imine, as demonstrated by isotopic incorpo-
ration with [18 O]H 2O [245]. Also, replacement of the C-3 hydroxyl group by
a solvent hydrogen in the E 1 –E 3 coupled reaction proceeds with retention of
configuration [246]. Together, these observations imply that the overall catalysis
in the active site of E 1 is most likely a suprafacial process occurring on the si
face of the PMP–substrate complex, as is the case with most other coenzyme
B 6 –dependent enzymes [247]. In fact, sequence alignments demonstrated a clear
relationship between E 1 and other PLP/PMP enzymes [243,248]. One significant
difference highlighted by the comparisons is the replacement of a highly con-
served lysine by a histidine at position 220 in E 1. This single replacement may
be responsible for converting a PLP-dependent transaminase to a PMP-dependent
dehydrase.
The preceding results clearly showed that E 1 behaves as a normal coenzyme
B 6-dependent enzyme that catalyzes a dehydration reaction. However, the C-3
deoxygenation product is formed only after NADH and E 3 are added to the reac-
tion mixture. Other reductases, such as diaphorase and methane monooxygenase
(MMO) reductase, are also able to generate small amounts of product [245].
Removal of the [2Fe-2S] center of E 1 impaired its ability to promote the final
product formation, though E 1(apoFeS) could still abstract the C-4′ hydrogen of
PMP [244]. This requirement of the [2Fe-2S] center, which is an obligatory one-
electron carrier, for product formation strongly implicated the intermediacy of
free radicals in the reaction. An EPR analysis of the chemically reduced E 1 sub-
TM
2. CDP-6-Deoxy-L-Threo-D-Glycero-4-hexulose-3-Dehydrase
Reductase (E 3)
CDP-6-Deoxy-l-threo-d-glycero-4-hexulose-3-dehydrase reductase (E 3) is a mo-
nomeric (36-kDa) protein of a red–brown color, and it contains 1 mol of FAD
as revealed by high-performance liquid chromatography (HPLC) and UV-Vis
analysis [250]. Iron and sulfur quantification combined with UV-Vis and EPR
spectroscopy indicated the presence of a plant-type [2Fe-2S] center. Sequence
alignments established a close relationship between E 3 and other iron–sulfur fla-
voproteins in the ferredoxin-NADP ⫹ reductase (FNR) family [251]. Like other
members in its family, E 3 will transfer reducing equivalents from NADH to a
variety of one-electron acceptors, including O 2, with varying degrees of efficiency
[252]. Removal of the [2Fe-2S] center impairs the ability of E 3 to catalyze final
product formation in the presence of E 1 and substrate, but the FAD of E 3(apoFeS)
remains functional as a two-electron/one-electron switch in the reduction of O 2
to H 2O 2 [250]. Therefore, the FAD can operate independently of the iron–sulfur
center when E 3 acts as a NADH oxidase, but the electron-transfer relay from
NADH to substrate during sugar reduction clearly includes both FAD and the
[2Fe-2S] center.
Stopped-flow spectroscopic experiments confirmed the independent nature
of each cofactor by showing that FAD is reduced by NADH at a similar rate in
both E 3(apoFeS) and holo-E 3 [253]. After the FAD is reduced to the hydroqui-
none form, the subsequent electron transfer to the [2Fe-2S] center was found to
be pH-dependent. At pH 7, the equilibrium favors the hydroquinone and oxidized
[2Fe-2S] 2⫹ state of the two-electron reduced enzyme, whereas at pH 10 the flavin
semiquinone radical and reduced [2Fe-2S] ⫹ state are favored. Spectroelectro-
chemical studies substantiated this pH dependence by showing that the redox
potentials of both the FAD and the [2Fe-2S] center change with respect to pH
[254]. Importantly, the midpoint potential of the FAD was dramatically altered
by pH, changing from ⫺212 mV at pH 7.5 to ⫺273 mV at pH 8.4 versus a shift
of ⫺257 to ⫺279 mV for the [2Fe-2S] center under the same conditions. The
ionizable group responsible for the pH dependence is estimated to have a pK a
of 7.3 by the stopped-flow studies [253], and N-1 of the flavin is likely the protic
position in terms of work with other flavoenzymes [255]. Overall, the proposed
electron transport sequence is consistent with the role of E 3 as a 2e ⫺ /1e ⫺ switch,
and the evidence provides compelling support for a radical mechanism in the
C-3 deoxygenation of 51 (Fig. 20). In light of the fact that PMP–glucoseen adduct
TM
TM
TM
TM
TM
IV. CONCLUSIONS
In this review, we have illustrated the unique nature of the C-O bond cleavage
event catalyzed by the E 1 –E 3 coupled enzyme system. These enzymes recruit
several cofactors to achieve a common biological transformation, albeit via a
novel mechanism. The PMP cofactor is directly responsible for the reversible
dehydration catalyzed by E 1. Electrons are then transferred from the biological
reducing agent NADH through an FAD cofactor and two [2Fe-2S] centers of
E 1 and E 3 to drive the deoxygenation to completion and regenerate PMP. This
TM
REFERENCES
TM
TM
TM
TM
TM
TM
TM
TM
TM
TM
Michael G. Zagorski
Case Western Reserve University, Cleveland, Ohio
I. INTRODUCTION
TM
Figure 1 Overview of the formation of β-peptide from the amyloid precursor protein
(APP) that contains 695 residues. The amino acid sequences of the β-(1–28) and β-
(1–42) peptides are shown. Depending upon conditions, in solution the β-peptide exists
in distinct conformations, whereas in the amyloid deposit only the oligomeric β-pleated
sheet structure is present.
TM
TM
TM
TM
TM
TM
TM
TM
line broadening and prevents the detection of well-resolved NMR signals. Addi-
tional serious problems are the subsequent aggregation and precipitation, particu-
larly with the longer β-(1–42) peptide, that readily occur at the higher peptide
concentrations required for NMR spectroscopy. It is possible, however, to per-
form NMR studies of fragments that are less likely to aggregate such as the β-
(1–28), β-(12–28), β-(10–35), and β-(25–35) in water solution, or with the com-
plete β-(1–42) peptide in water solution containing TFE or micelles [61,70,74,
81–83].
One of the most useful NMR parameters is the nuclear Overhauser effect
Figure 3 Graphs depicting the variation of random coil and α-helical secondary struc-
tures of the β-(1–28) peptide with pH in SDS, DPC, and DTAC micelle solutions from
CD data. The chemical structures of the micelles are shown in Fig. 4. The charged SDS
and DTAC micelles stabilize the α-helix at physiological pH and β-sheet structure was
not observed in any micelle solution. SDS, sodium dodecyl sulfate; DPC, dodecyl phos-
phocholinc; DTAC, dodecyl trimethylammonium chloride; DPC, dodecyl phosphocholine;
CD, circular dichroism.
TM
TM
d3 and DPC-d38 solutions, the smaller α-helix (residues 2–11) rapidly unfolds
above 25°C with the other helix (residues 13–27) unfolding above 30°C. Both
α-helices unfold to random coil structures and refold back to α-helices with cool-
ing. The NH temperature coefficients were determined by recording NOESY data
at 10, 25, 35, 40, and 50°C. The random pattern of coefficients suggests that the
peptide is not buried within the hydrophobic interior of the SDS-d25 micelle,
but instead is located at the surface near the lipid/water interface. These studies
demonstrate that the negatively charged SDS micelle provides an environment
most conducive to α-helix formation and that this stability is not due to the loca-
tion of the peptide within the hydrophobic interior of the SDS micelle.
To evaluate further the effect of pH on the loss of α-helical and the forma-
tion of β-sheet structure, we are currently determining the dissociation constants
(pKa) of the aspartic acid (Asp), glutamic acid (Glu), histidine (His), and tyrosine
(Tyr) residues under different solution conditions. We have used the β-(1–28)
TM
TM
TM
rather than the hydrophobic interior, is consistent with the amide-NH temperature
coefficients and the rapid deuterium exchange rates. The irregular pattern of the
NH-temperature coefficients and the rapid NH-exchange rates are consistent with
the peptide being located on the surface of the SDS micelle. If residues were
buried within the hydrophobic interior of the micelle, then significantly lower
temperature coefficients and NH-exchange rates should be observed. Overall, the
secondary structure is defined by an extended structure (residues 1–9), α-helix
(residues 10–24), loop (residues 25–27), and α-helix (residues 28–42). The pres-
ence of a loop within the Gly25–Ser26–Asn27 segment is supported by the ab-
sence of any major backbone NOEs and the downfield chemical shift of the
Asn27 αH. Interestingly, the location of this loop is nearly identical to a reverse
turn suggested for the β-sheet structure of a β-(10–43) peptide, in which the
turn was proposed to be within the Ser26–Asn27–Lys28–Gly29 sequence [44].
Recent NMR studies of a β-(10–35)-CONH2 fragment suggests a turn–strand–
turn motif between His13 and Val24 in aqueous solution [51], whereas the β-
(1–40) peptide forms two α-helical segments between Gln15–Asp23 and lle31–
Met35 in 30% TFE-d3 solution [81].
TM
TM
TM
TM
TM
TM
(b)
TM
TM
Despite the importance to AD, the underlying mechanisms for the accumulation
of the β-peptide into insoluble amyloid remain largely unknown. A popular con-
cept involves an ‘‘amyloid-initiated-cascade’’ phenomenon, in which altered pro-
duction, removal, and aggregation of the amyloid β-peptide initiate a sequence
of events that leads to neuronal death [11,38,115]. It should be kept in mind that
other lesions such as the neurofibrillary tangles are also abundant in AD brains,
and it may happen that the tangles are more important in the pathogenesis of the
disease [116]. The tangles are intracellular deposits consisting of twisted fila-
ments of the cytoskeletal tau protein. However, since the majority of genetic and
biological data support a critical role for amyloid formation in AD, the ‘‘amyloid-
initiated-cascade’’ hypothesis appears to be the most promising model for drug
discovery [11,117].
A hypothetical model for β-amyloidosis is outlined in Scheme 1. This
mechanism attempts to take into account the results of our work, as well as work
from other laboratories. The various pathways shown in the scheme connect the
in vitro biophysical studies to a natural situation that may exist in the brain. The
model emphasizes a conformationally driven mechanism, in which the three ma-
jor solution structures of the β-peptide coexist in equilibrium: random coil (mono-
meric), α-helix (monomeric), and β-sheet (oligomeric). The aggregated β-sheet
TM
is the predominant structural motif in the amyloid plaques. When the aggregation
reaches a critical mass, the β-sheet structure precipitates as amyloid, and recon-
version back to soluble random coil or α-helical structures is no longer possible
[48,62,118]. Recent data suggest that soluble β-sheet dimers are toxic [47], with
the larger amyloid plaques that form later further disrupting neuronal circuitry.
During β-sheet formation, the β-peptide-induced neurotoxicity increases [31,32],
suggesting that the processes involved in amyloid plaque formation are critical
in the pathogenesis of AD. The mechanisms whereby the β-sheet structure exerts
its neurotoxicity remain unknown [119].
Once released by proteolysis of APP, the β-peptide would be expected
initially to adopt a monomeric, random coil conformation. This interpretation is
based on the known physiological concentrations of the β-peptide in cerebrospi-
nal fluid, which are in the nano- to picomolar range [91,120]. At these low con-
centrations, the production of aggregated β-peptide as amyloid is very unlikely
[28,29,48,66] since the threshold concentrations required for generation of amy-
loidlike oligomeric β-sheet structure are much higher 50–350 µM. Presumably,
the random coil structure is biologically inactive and would not be involved in
specific binding processes to other macromolecules. This suggests that, in order
TM
TM
V. CONCLUSIONS
There are currently no effective treatments for AD, and the current drug develop-
ment process is often cumbersome, taking an average 9 years for approval [153].
The methods discussed in this review, notably the solution NMR technique, can
TM
ACKNOWLEDGMENTS
I would like to thank the many excellent colleagues who completed the studies
shown in this review. These include Colin Barrow, Takashi Iwashita, and Vassil
Vassilev from the Suntory Institute for Bioorganic Research, Osaka, Japan, and
likewise Keith Marcinowski, Joseph Talafous, Haiyan Shao, Kan Ma, Arthur
Salomon, Shu-chuan Jao, Charlene Keane, Erin Clancy, Jing Yang and Rob
Friedland from Case Western Reserve University, Cleveland, Ohio. The work
was supported in part by grants from the American Health Assistance Foundation,
the Suntory Institute for Bioorganic Research, the National Institutes of Aging
(AG-08992-06 and AG-14363-01), American Federation of Aging Research,
Philip Morris, Inc., Gliatech, Inc., Pfizer, Inc., Smokeless Tobacco Agency, and
a Faculty Scholars Award from the Alzheimer’s Association (FSA-94-040). The
600-MHz NMR spectrometer was purchased with funds provided by the National
Science Foundation, the National Institutes of Health, and the state of Ohio. I
would also like to thank Anita Hong (Anaspec, Inc.), Larry Sayre, and Witold
Surewicz for helpful discussions. Finally, I would like especially to thank Koji
Nakanishi for his sound advice, particularly his support and encouragement
throughout my stay in his groups in both New York and Japan.
REFERENCES
TM
TM
TM
TM
TM
TM
TM
I. INTRODUCTION
Bacteriorhodopsin (bR), the only protein in the purple membrane of the archae-
bacterium Halobacterium salinarium, is a retinal-based, light-transducing pig-
ment [1]. The dark adapted form contains roughly equivalent amounts of the all-
trans and 13-cis isomers of retinal (1, Fig. 1), the aldehyde of vitamin A, as its
chromophore [2]. The protein functions as a light driven transmembrane proton
pump and is of quite some interest for its potential to harvest light energy. The
discovery that bR with its seven transmembrane α-helices has the same spatial
arrangement as the large class of biologically important G-protein receptors has
further increased interest in this protein. In addition, the protein is easily grown
and purified, making it readily available for a wide range of spectral and physio-
logical studies.
The photocycle of all-trans bR [WT (wild type)/1a] initiates with the isom-
erization of all-trans retinal to the 13-cis isomer and consists of a sequence of
at least five steps (Fig. 2). The 13-cis pigment [WT/1] is also photochemically
active and undergoes a distinct photocycle [3] in which proton translocation is
not observed [4]. The photocycle of the all-trans isomer has been studied in the
TM
TM
TM
TM
TM
C. Site-Directed Mutagenesis of bR
Oligonucleotide-directed, site-specific mutagenesis was performed essentially as
described by Menick [27]. The 2.7-kb BamHI-HindIII bop gene insert was re-
stricted from the H. salinarium shuttle vector pMC-1 (a generous gift of R. Nee-
dleman) and ligated into M13mp19. The Arg82 was replaced with Ala by using
the mutagenic primer 5′CTGG GCG GCG TAC GCT GAC 3′, which contained
the indicated two-nucleotide mismatch. Mutants were initially screened by colony
blot hybridization and then plaque purified, and the mutation was verified by
dideoxy sequencing [28]. The entire coding region of the bop gene was sequenced
to ensure that no additional mutations had occurred. The BamHI-HindIII bop
gene insert was restricted from the M13mp19RF and religated into pMC-1.
TM
G. Pigment Regeneration
The native and transformed pigments were bleached according to the method of
Tokunaga et al. [6]. Pigments were regenerated by additions of the all-trans reti-
nals in ethanol (⬍1% vol) to bleached membranes (50 mM Tris-HCl, pH 8.8) at
25°C. Pigment formation was measured after 24 h by absorption changes with
bleached membrane as reference.
H. Spectroscopic Methods
Absorption spectra were recorded in 5-mm quartz cuvettes thermostatted at 20°C
using a Cary-Aviv 14 spectrophotometer (Aviv Associates, Lakewood, NJ) by
digital measurement with 1-nm steps and a spectral bandwidth of 1 nm. Kinetic
measurements of dark adaptation were done at 580 nm using the kinetic mode
of the spectrophotometer. A homemade cryostat was used for low-temperature
measurements [33].
Flash-induced absorption changes were measured on a kinetic spectropho-
tometer [34,35]. The actinic pulse was from a Quanta Ray DCR11 Nd:YAG laser
(532 nm, 7 ns, 5–10 mJ/pulse; Spectra Physics, Mountain View, CA). Flash-
induced pH changes were measured at pH 7–7.2 using the pH-sensitive dye py-
ranine.
The retinal analogues were synthesized as outlined in Fig. 5. This scheme is quite
similar to those used in other laboratories for the syntheses of various retinal
analogues and relies primarily on Horner-Emmons olefination for chain elonga-
tion. Yields of the various steps were between 40% and 65%, but the products
were seldom isolated other than for an initial characterization, because of the
lability of these polyene compounds.
Stable pigments were formed between bacteriorhodopsin and the trans iso-
mer of all three retinal analogues 2–4. The absorption maxima of these WT pig-
ments were similar to those reported by Gärtner et al. [10] (Table 1). Pigments
TM
TM
Retinal (1) 2 3 4
Native WT 558 540 569 539
Transformed WT 558 540 567 538
R82A 555 537 566 539
a
All pigments were regenerated with the all-trans isomer of the respective retinal. Pigment solutions
were dark adapted for 1 h before measurement. Spectra were recorded in 50-mM-Tris-HCl, pH 8.8
at 24°C. Values ⫾2 nm. WT, wild type.
were formed with both bR apoprotein generated from the normal purple mem-
brane purified from cultures of the halobacteria and from protein of H. salinarium
IV-8 transformed with the WT bop gene. The transformed and native pigments
had identical absorption properties. The data presented here are from pigments
regenerated with the apoprotein isolated from the native pigment. The pigments
formed with the apoprotein of R82A and the three demethyl retinals showed
absorption maxima at 537 nm (2), 566 nm (3), and 539 nm (4), respectively
(Table 1). The pigments were reasonably stable in the presence of 10-mM hy-
droxylamine, pH 7.5, with only modest degradation noted for the pigments
formed with 4 (⬍25% over 2 h). The deletion of the 9-methyl group shifted the
absorption maxima in both of the pigments, whereas the removal of the 13-methyl
and the mutation at the 82-position had little effect on the absorption properties.
The R82A pigment has a pK a for the purple to blue transition of 8.7 in
25% glycerol/water (Fig. 6). For the R82A/3, the pK a was shifted to 8.1. The
origin of this shift can be attributed to the finding that for the 3 pigment most
of the chromophore is probably in the 13-cis conformation and the pK a of the
13-cis form of the pigment is lower than that for the all-trans form [36].
The photocycle intermediates were detected by monitoring the flash-in-
duced absorption changes at 410 nm, 540 nm, and 600 nm. The WT changes
(Fig. 7a) represent predominantly M (410 nm) and O (640 nm) intermediates as
well as the recovery of the pigment (540 nm). The WT/2 bR converted to a 410-
nm M intermediate and a 600-nm species, as a result of the O intermediate, after
20 ms (Fig. 7c). The formation of the M intermediate is in agreement with the
results of Yamazaki et al. [12].
Excitation of WT/3 produced only a small amount of the 410-nm M inter-
mediate species (Fig. 7e). The half-time of decay observed for this species was
about 3 s, whereas the decay of M in the WT bR is 10 ms. The predominant
conversion of WT/3 was to a 600-nm species that is much less in WT/2 (Fig.
7c). It has been proposed that this 600-nm species, which is different from the
TM
O intermediate of the trans cycle, arises from the 13-cis cycle. [7,12] This identi-
fication of the photoproduct and a small amount of M is consistent with the fact
that the 13-cis isomer accounts for at least 80% of the retinal in the binding
site of 3 pigment. Furthermore, upon the illumination of WT/4 pigment, the M
intermediate species from the trans cycle was quite small and only an unusual,
long-lived 600-nm species was found (Fig. 7g). These results of flash-induced
spectroscopic experiments are consistent with the conclusion that the 13-position
of retinal is important for the stability of the all-trans configuration of retinal in
the binding site [7].
In R82A, the light-induced absorption change at 410 nm decayed much
faster than the absorption at 580 nm recovered, and an O intermediate was not
observed (Fig. 7b). Interestingly, a rapidly forming 600-nm species is observed
in all three R82A/2–4 pigments (Fig. 7d, f, and h), indicating an increase in the
proportion of the 13-cis isomer in the pigment. The R82A/3 and R82A/4 pig-
ments have particularly long-lived 600-nm species.
The flash-induced proton changes were measured by the transient absorp-
tion changes of the pH indicator pyranine at pH 7.2. The fast (⬍1 ms) decrease
in dye absorbance observed in WT corresponds to release of protons, and
the subsequent restoration of absorbance (rise time 12 ms) corresponds to the
uptake of protons (Fig. 8a) [11]. The signals from all three WT/2–4 pigments
were ⬍20% that of WT (Fig. 8b, c, and d). The proton signal from R82A was
about 40% that of WT, and the order of proton release and uptake is reversed
TM
TM
compared to that of the WT, as observed previously (Fig. 8e) [11]. No absorption
change of the dye in R82A/3 was observed (Fig. 8f), demonstrating that the
combination of the deletion of the 13-methyl and the removal of the charged
amino acid at position-82 prevents the light-induced movement of protons for
this pigment.
IV. CONCLUSIONS
TM
ACKNOWLEDGMENTS
This work was supported by NIH (EY04939 and GM52023), DOE (95ER20171),
and Research to Prevent Blindness, Inc. We thank Dr. R. Needleman for provid-
ing the shuttle vector containing the bop gene, pMC-1, and bR-defective H. sali-
narium strain, IV-8. We particularly thank Dr. Koji Nakanishi, without whose
initial guidance and enthusiasm this collaboration would not have existed, and
therefore these studies would not have been accomplished.
REFERENCES
TM
TM
David G. Lynn
The University of Chicago, Chicago, Illinois
I. INTRODUCTION
In 1995, the first complete genomic sequence was published [1], and since then
several others have become available. Of particular importance to the bioorganic
chemist, this list includes Escherichia coli [2] and Bacillus subtilis [3], the organ-
isms whose proteins and molecular genetics have been used so extensively to
define and characterize the chemical reactions of biology. These primary se-
quence data will certainly extend and enrich our understanding of the reactions
that drive these organisms, but it also opens an entirely new set of questions. It
is likely that no scientific discipline will benefit more directly from this knowl-
edge than bioorganic chemistry and chemical biology. Just as the structure and
the synthesis of the steroid nucleus placed human reproduction and hormone
physiological processes at the atomic level, and the structure of proteins revealed
the exquisite sophistication of the mechanisms of biological catalysts, the struc-
ture of the genome, the molecule that codes for the entire organism, places the
most central questions of biology on an atomic scale.
More than just understanding the molecular machines of biology, it is now
possible to consider how chromosomes function: to wonder about the logic for
the organization and positions of genes that are functionally interdependent, the
function of intervening sequences that tie the coding regions together, the reason
for the positioning of the origin and terminus of replication, the functional sig-
nificance of G-C-rich domains; the positions, frequencies, and function of re-
peating sequences; and the organization and utility of coding redundancy. With
data on multiple genomes, strategies for organization and stability can be com-
pared and the rules governing success may be understood. Questions as to how
TM
Maniloff [6] argued that small autonomous genomes could have arisen along two
independent paths, which he referred to as ‘‘top-down’’ and ‘‘bottom-up.’’ The
bottom-up path is essentially an origin of life argument, to which I will return.
The top-down path requires the evolutionary development of symbiotic, parasitic,
or other interdependent associations such that genomic information can be re-
duced. The biosphere today is a tightly woven network of interdependent organ-
isms. Plants can utilize CO 2, O 2, and NO 2, inorganic materials, and photons for
energy, seemingly functioning in the absence of other organisms. There are many
TM
TM
Figure 1 Light micrograph of a 2-day-old Striga asiatica seedling that has been stained
with pyrogallol. The red coloration [shown here as a darkening at the root meristem (see
arrow)], of the root meristem highlights the H 2O 2 production. (Source: Ref. 32.)
TM
ent there is no evidence for the exudation of these more easily oxidized substrates
to the parasite cell surface. Evidence for a membrane-bound NADPH oxidase
complex, homologous to the human neutrophil oxidative burst machinery, has
also been found in plants [41–44]. In that regard, SXSh quinones were recently
shown to be detected via a benzoquinone-dependent oxidoreductase [45,46]. In
terms of the observed redox range (Fig. 3), the terminal acceptor could be O 2,
via a cytochrome-mediated reduction. It may be, therefore, that signal detection
is coupled with oxidative signal release through the production of H 2O 2 in an
autocatalytic host detection mechanism. It has been suggested that a mutation in
the oxidative burst machinery controlling H 2O 2 production in a dicotyledonous
plant has made it possible for the parasite to exploit monocot cell surfaces via
a dicot defense pathway. A good defense is converted into an offensive host
TM
detection strategy [32]. Such a gain of function mutation could have easily
evolved and played a critical role in the development of parasitism. Further analy-
sis of the oxidoreductase genes responsible for H 2O 2 production in Striga should
allow the model to be evaluated.
The study of Striga spp. makes possible a structure/activity analysis of the
parasitic genome. By defining the elements that lead to parasitism, a functional
understanding for the process of genome reduction can be obtained. Through
comparisons with other related parasites, both the rate of gene loss and the re-
sulting changes in genomic organization may be understood. Analogous ‘‘com-
pare and contrast’’ approaches have taught us much about the forces that drive
biology, and it is clear that these analyses have now moved to an atomic level.
In our case, the genomic reorganization that occurs between parasitic and nonpar-
asitic relatives provides glimpses into the structure/function analysis of their ge-
nome.
TM
Figure 4 The timeline for the infection process of Agrobacterium tumefaciens. Across
the top, the oncogenetically transformed plant cell displays tumorigenic growth. In the
left-hand panel, the gene transfer event is induced by specific signaling molecules (repre-
sented by phenol) that originate from the plant cell. The T-DNA is transferred after activa-
tion of the vir regulon of the Ti plasmid (pTiA6). In the middle panel, octopine is produced
by the transformed plant cell and activates both occ and traR expression. In the right-
hand panel, CF, a factor whose biosynthetic genes are constitutively expressed from Ti,
together with TraR, activates conjugal transfer of the Ti plasmid.
TM
tent plant cell and recognized via the activation of the virulence regulon of the
Ti plasmid. A histidine two-component sensor kinase and response regulator are
constitutively expressed from the regulon and integrate information from mono-
saccharide, phenol, and acidic pH to control expression of the remaining genes
(Fig. 5). The virulence regulon encodes the machinery that catalyzes the excision
of the T-DNA, its transfer into the plant cell, and targeting aspects of its incorpo-
ration into the genome of the plant cell. The center panel shows the result of
incorporation of the T-DNA. Driving the prokaryotic genes for critical steps in
the biosynthesis of auxin and cytokinin behind eukaryotic promoters leads to the
uncontrolled growth of the plant cell. In addition to the oncogenes, an expressed
dehydrogenase catalyzes the reductive amination of pyruvate with arginine to
generate octopine, a substance not metabolized by the plant, but that can serve
as the sole carbon and nitrogen source for A. tumefaciens that harbor the catabolic
genes of the Ti plasmid. Octopine accumulation activates both the expression of
the octopine catabolism regulon (occ) and the regulatory protein, TraR. TraR and
a homoserine lactone adduct autoinducer, AI, constitutively biosynthesized from
TM
TM
TM
of template turnover was reduced [79] such that the template efficiently catalyzed
the amplification of the encoded information. This opportunity to have multiple
turnovers on the template allows for the information to be transferred accurately
and amplified many times with catalytic amounts of template.
This ligation reaction gives a different product backbone from the template
and, in that way, has characteristics of a translation process. To extend this pro-
cess to replication would require the amine product to catalyze the construction
of the template as shown in Fig. 7. The ethyl amine linkage is more hydrophobic
than the phosphate diester, and its reduced hydration and increased flexibility
TM
should enable polymers containing such linkages to fold and pack in a way that
is more like that of proteins. The basic amine in the linkage should expand the
catalytic competence of the polymer much as the 2′-OH does in RNA [68,74].
Therefore it may be possible to reduce the central dogma of biology, DNA to
RNA to protein, to a two-step process as shown in Fig. 7.
The ligation efficiency of the reaction outlined in Fig. 6 has allowed us to
consider new approaches to template-directed synthesis, even a synthetic ap-
proach that extends the architectural limits of the bottom-up construction of a
genome. Now the term autonomy becomes more relevant because this genome
would be synthetic. The extension of the ligation reaction to template-directed
polymerizations would be necessary, and eventually, for the template to be auton-
omous, it must encode the capability of catalyzing a ‘‘metabolism’’ sufficient
for monomer generation [80–82]. Therefore, the central question becomes how
the required diversity of catalysts might be produced from the template. The
principles inherent in the imine coupling reactions can be extended to other cou-
pling reactions, giving different functionality in the backbone of the translation
product. For example, DNA has recently been shown to direct the condensation of
peptide nucleic acid amides under dehydrating conditions [83,84]. These products
have new backbones and represent new translation products that could be selected
for new catalytic function. In other words, one template could be translated into
many different products in which the base sequence is the same, but the different
TM
V. CONCLUSIONS
ACKNOWLEDGMENTS
The ideas presented here have grown from discussions and experiments with very
talented coworkers, most of whom are referenced here. I am grateful for a long-
standing and valuable collaboration with Andy Binns, University of Pennsylva-
nia, and support from the NIH, NSF, DOE, Novartis, and the Rockefeller Founda-
tion. Most specifically, I acknowledge Koji Nakanishi for establishing a breadth
of inquiry and an enthusiasm for learning that have so broadly impacted science,
and for the personal support that has made these investigations possible.
REFERENCES
TM
TM
TM
TM
I. INTRODUCTION
TM
TM
To test the hypothesis that the organization of the µA and µB sites was
critical for enhancer activity in B cells, we constructed mutated µ enhancers with
either 4 or 10 base pairs inserted between the µE3 and µB sites (Fig. 2) and
tested these mutations in DNA binding and functional assays. Insertion of 4 or
10 bp rotates the µB site approximately 145° or 350°, respectively, relative to
the µA or µE3 sites, assuming a typical DNA helical periodicity of 10.5 bp.
Protein binding to the three sites was not affected in the rotation mutations. How-
ever, transcription activation in B cells was significantly impaired in both the ⫹4
TM
TM
TM
Figure 4 Model of how a DNA binding protein can alter DNA structure. A DNA bind-
ing protein that increases DNA flexure can result in numerous conformations indicated
by thin and dotted arrows. A DNA binding protein that induces a directed DNA bend
locks the DNA into a particular structure, represented by the bold arrow.
TM
TM
tive site in the noncoding strand between the C and A residues opposite the
T-C junction of the coding strand, suggesting protein induced distortion of the
phosphodiester backbone [18]. We approximated that this hypersensitive site is
the center of the PU.1-induced DNA bend and calculated the lengths between
µB and the poly A tract bend centers with reference to the position of the hyper-
sensitive site. The ETS(PU.1)-induced DNA bend reinforced the poly A tract
bend, resulting in the slowest mobility nucleoprotein complex when the µB site
was positioned 36 bp (3.4 helical turns) from the poly A tract. Because ETS(PU.1)
reinforced the minor groove-directed poly A tract bend when the µB element
was on the opposite side of the DNA helix, we propose ETS(PU.1) bends µ
enhancer DNA toward the major groove (Fig. 6).
TM
Figure 5C The best-fit cosine curve of ETS (PU.1)–µ enhancer nucleoprotein com-
plexes from Figure 5B. Analysis indicated ETS (PU.1) induces a 24° directed bend in the
µ enhancer DNA. ETS (PU.1) ⫽ the ETS domain of PU.1.
TM
III. DISCUSSION
TM
CONCLUSIONS
REFERENCES
TM
TM