Download as pdf or txt
Download as pdf or txt
You are on page 1of 29

Current Medicinal Chemistry, 2002, 9, 2005-2032 2005

Fostriecin: Chemistry and Biology

D.S. Lewy†, C.-M. Gauss, D.R. Soenen and D.L. Boger*

Department of Chemistry and The Skaggs Institute for Chemical Biology, The Scripps Research Institute, 10550
North Torrey Pines Road, La Jolla, California 92037, USA

Abstract: A review of the current status of the chemistry and biology of fostriecin (CI-920) is provided.
Fostriecin is a structurally unique, naturally-occurring phosphate monoester that exhibits potent and
efficacious antitumor activity. Initially it was suggested that its activity could be attributed to a direct, albeit
weak, inhibition of the enzyme topoisomerase II. However, recent studies have shown that fostriecin inhibits
the mitotic entry checkpoint through the much more potent and selective inhibition of protein phosphatase 2A
(PP2A) and protein phosphatase 4 (PP4). In fact, it is the most selective small molecule inhibitor of a protein
phosphatase disclosed to date. The contribution, if any, that topoisomerase II versus PP2A/PP4 inhibition
makes to fostriecin's antitumor activity has not yet been fully defined. Initial phase I clinical trials with
fostriecin never reached dose-limiting toxicity or therapeutic dose levels and were halted due to its storage
instability and unpredictable chemical purity. Hence, the total synthesis of fostriecin has been pursued in order
to confirm its structure and stereochemistry, to provide access to quantities of the pure natural product, and to
access key partial structures or simplified/stable analogs. Several additional natural products have been
isolated which contain similar structural features (phospholine, phoslactomycins, phosphazomycin,
leustroducsins, sultriecin, and cytostatin), and some exhibit comparable biological properties.

INTRODUCTION novel, clinically unexplored mechanisms worthy of pursuit


for introduction of a new class of antitumor agents. Herein,
Fostriecin (1, CI-920), Fig. (1), is a structurally novel we review the present status of the chemistry and biology of
phosphate ester produced by Streptomyces pulveraceus that fostriecin.
is active in vitro against leukemia (L1210, IC50 = 0.46
µM), lung, breast, and ovarian cancer, and which exhibits 3 15 17
OH
2 4 14
efficacious in vivo antitumor activity. It has been H2O3PO OH
16 18
investigated in a Phase I clinical trial at NCI that was halted 5
8 10 13
even before dose-limiting toxicities or therapeutic plasma O O
6
7 9 11
12
levels were reached when concerns regarding drug purity and Me OH
storage stability proved problematic with the naturally- 1, Fostriecin
occurring material. Although fostriecin inhibits DNA
topoisomerase II (IC50 = 40 µM) through a novel, non- R1 R2
DNA-strand cleavage mechanism, the activity is weak and it H2 O3PO OH
does not induce G2 arrest like other topoisomerase II
inhibitors. Thus, it is unlikely this initially suggested target O O
is responsible for the antitumor activity of 1. Instead, Me OH
fostriecin has been shown to inhibit the mitotic entry
checkpoint through more potent and selective inhibition of 2, R1 = R2 = H
protein phosphatases 1 (PP1), 2A (PP2A), and 4 (PP4) 3, R1 = R2 = OH
(IC50 = 45 µM, 1.5 nM, and 3 nM, respectively). Notable
in this regard is the fact that fostriecin is the most selective Fig. (1).
protein phosphatase inhibitor known to date, exhibiting a
10 4 -fold selectivity for PP2A/PP4 versus PP1. Also
contributing to its selective antitumor properties, fostriecin ISOLATION AND STRUCTURE ELUCIDATION
was shown to be transported into tumor cells via the reduced
folate carrier system. Inhibition of the mitotic entry Fostriecin (1), along with two structurally related
checkpoint and selective protein phosphatase inhibition are compounds PD 113,270 (2) and PD 113,271 (3), were
isolated from the fermentation broth of an unidentified
actinomycete found in a Brazilian soil sample [1,2]. The
*Address correspondence to this author at the Department of Chemistry organism was characterized as Streptomyces pulveraceus
and The Skaggs Institute for Chemical Biology, The Scripps Research subspecies fostreus ATCC 31906. It was found that
Institute, 10550 North Torrey Pines Road, La Jolla, California 92037, USA;
FAX: 1-858-784-7550; E-mail: boger@scripps.edu fostriecin and its congeners were unstable above pH 8 and
†This article is dedicated in memory of David and all that he meant to us. very labile in dilute acid. Therefore, they could only be

0929-8673/02 $35.00+.00 © 2002 Bentham Science Publishers Ltd.


2006 Current Medicinal Chemistry, 2002, Vol. 9, No. 22 Boger et al.

obtained in salt form, and isolation steps had to be 11 Hz) appeared as apparent triplets, whereas the one
performed within the pH range 5.5–7.5. The fermentation observable proton with a vicinal single bond and trans
broth was filtered at pH 6.5 and passed through an ion- couplings (J vic = 11 Hz, J trans = 15 Hz) appeared as a
exchange column. Dissolution of the filtrate residue in doublet of doublets. 1H NMR decoupling data indicated that
acetone and purification on a Diaion HP-20 resin column the conjugated triene chain terminated with an allylic
(water eluent) yielded two fractions which were combined hydroxymethyl group and that an isolated trans carbon–
and further purified. The resulting mixture was carbon double bond connected the 5,6-dihydropyran-2-one
chromatographed on a C18-reverse phase column utilizing 5– ring to the quaternary carbon C8. In addition, a 1 H NMR
7% CH3CN–0.05 M phosphate buffer (pH 6.8). The eluate signal at δ 4.17 corresponded to a proton adjacent to a
was filtered and desalted, and the resulting solution was phosphate. Simple chemical transformations further
lyophilized at pH 6.5 to afford 40 g of fostriecin as a nearly confirmed the connectivity of fostriecin. Acetylation of the
white solid. The purity of this product was about 90%, and natural product provided 4, Fig. (2). As mentioned above,
a sample of more than 96% purity was prepared by C8- or dephosphorylation of 1 with alkaline phosphatase afforded 5.
C18-reverse phase chromatography using 2 or 10% CH3CN Hydrolysis of the 5,6-dihydropyran-2-one ring yielded the
in 0.1 M phosphate buffer (pH 6.8). carboxylic acid 6 and treatment of 5 with sodium periodate
provided the ketone 7 and aldehyde 8 as cleavage products
Preliminary characterization of fostriecin and its [3].
congeners 2 and 3 revealed that all three compounds were
monosodium salts of phosphate monoesters containing a The relative and absolute stereochemistry of 1 was
conjugated triene and an α,β-unsaturated δ-lactone [1,3]. determined later by Boger, et al. [5]. The relative
Elemental analysis showed that 1 was composed of 6.4% stereochemistry between the quaternary C8 carbon and C9,
phosphorus, and dephosphorylation by alkaline phosphatase which contains the phosphate ester, was elucidated upon
demonstrated that it was a monosubstituted phosphate ester. formation of the five-membered cyclic phosphodiester 9,
The IR spectra of the three isolates were nearly identical, Fig. (3), and observation of NOE’s (2D 1 H– 1 H ROESY
displaying a prominent peak at 1710 cm–1 , which was NMR, D2O, 400 MHz, δ) between H6,7 (5.93–6.00) and H9
attributed to the presence of an α,β-unsaturated δ-lactone. (4.34), as well as between the C8-CH3 (1.40) and H1 0
FABMS analysis of fostriecin revealed a molecular weight (1.75–1.90). This established the cis relationships of the
of 452 for the monosodium salt which, together with C6–C7 vinyl side chain with H9 and the C8-CH3 with the
elemental analysis and 1 3 C and 1 H NMR spectral C9 side chain on the five-membered cyclic phosphodiester
information, led to the assignment of a molecular formula of and a syn-1,2-diol relationship between C8 and C9.
C19H26O9PNa. Derivatization of fostriecin to provide the C9/C11 acetonide
10 led to the elucidation of the relative stereochemistry of
The 13 C NMR spectrum of 1 displayed characteristic the 1,3-diol. The acetonide methyl groups displayed similar
couplings indicative of two- and three-bond couplings of 13 C NMR chemical shifts of δ 24.49 and 24.89 (CDCl ,
3
three assignable carbon atoms to phosphorus. Inspection of 100 MHz) within the range characteristic of an anti-1,3-diol
the 1H NMR spectrum established the Z,Z,E configuration acetonide (23.96–25.22) and different from the distinct
of the triene system based on first-order analysis of coupling chemical shifts observed with syn-1,3-diol acetonides
constants. Consistent with model trienes [4], the four triene (18.67–19.98 and 29.74–30.16) [6,7]. In addition,
protons with a vicinal single bond and cis coupling (J = 10– diagnostic NOE cross-peaks in the 2D 1H–1H NOESY NMR
spectrum of 10 were seen between one acetonide methyl
R3 group (1.35) and H9 (3.72) and between the other acetonide
OR1 OR2 methyl group (1.39) and H11 (4.73). This data further
supported the assignment of an anti-1,3-diol acetonide
O O
adopting a twist boat conformation. Consistent with this
Me OH assignment, H9 (J(H 10 ) = 6.3, 9.6 Hz) and H11 (J(H 10 ) =
6.3, 9.5 Hz) exhibited 1H NMR coupling constants expected
4, R 1 = PO3 H2 , R2 = Ac, R3 = OAc of this twist boat conformation and the observed NOEs.
5, R 1 = R 2 = H, R3 = OH
In order to establish the absolute stereochemistry, (5R)-7
OH was synthesized through an unambiguous method by
OH H2 O3 PO OH
Hokanson, et al. [3]. Comparison (TLC, [α]25D, IR, NMR)
with the analogous degradation product from fostriecin led
to the assignment of the C5 R configuration. During the
NaO2 C Me OH
final assignment of the relative and absolute stereochemistry
6 of fostriecin by Boger, et al. [5], this designation of the 5R
stereocenter was confirmed by an alternative enantioselective
synthesis of (5R)-7. Additionally, the assignment of the
OH
OH absolute configuration at C11 was established by correlation
(NMR, IR, MS, Chiralcel OD-H HPLC) of the protected
O O OHC triol 11, derived from degradation of the natural product,
O 8 with the (R)-(+)-tribenzoate 11 prepared from (R)-(+)-1,2,4-
7
butanetriol. The absolute stereochemical determinations at
Fig. (2). C5 and C11, combined with the relative configuration
Fostriecin: Chemistry and Biology Current Medicinal Chemistry, 2002, Vol. 9, No. 22 2007

13C NMR
OH δ = 24.49 and 24.89
OH
nOe's
nOe's
H H OH
a b O O H
1 H
O O O
Me O O O
P ONa Me OH
C8/C9 syn diol c d C9/C11 anti diol
O 9 10
PMBO2C
OBz OBz 11 7
OBz O O COCH3 OH
HO
C11 R configuration C5 R configuration

(a) 4-BrC 6H4COCl, pyr, 70%. (b) alkaline phosphatase, 100%; Me 2C(OMe) 2, TsOH, 40%. (c) alkaline phosphatase; TBDPSCl; NaIO4; NaBH4 ; BzCl;
O3; NaBH4 ; BzCl. (d) alkaline phosphatase, 100%; NaIO4.

Fig. (3). Assignment of relative and absolute stereochemistry for fostriecin.

assignments, provided the complete stereochemical coupling of 13 and 19 proceeded in 50% yield, affording the
assignments of fostriecin as 5R,8R,9R,11R. C6–C7 trans olefin. Subsequent treatment of the α,β-
unsaturated ketone with trimethylaluminum provided the
tertiary alcohol 20 (65%) containing an 8R stereocenter
SYNTHETIC STUDIES derived from a chelation-controlled nucleophilic addition.
Removal of the protecting groups provided 21, and its 13C
In an initial attempt to establish the stereochemistry of NMR spectrum (D2O, 400 MHz) was found to be different
the natural product, Just and O'Connor prepared a from that of the fostriecin degradation product 5.
diastereomer of dephosphorylated fostriecin containing the
5R,8R,9S,11R configuration [8]. The α,β-unsaturated δ- In our own efforts, a total synthesis of fostriecin was
lactone 13, containing the previously established 5R viewed as an opportunity to not only confirm our
stereocenter, was prepared in three steps from the known stereochemical assignments, but also to address the issues of
methyl ester 12, Fig. (4a). The known thioacetal 14, chemical purity and storage stability currently limiting
obtained in four steps from 1,2-O -isopropylidene-D- clinical trials of the natural product. In addition, it could
glucofuranose and protected as the cyclopentanone ketal, was provide access to more stable structural analogs not
tranformed to the cis bromoalkene 1 5 by thioacetal accessible by degradation or functionalization of fostriecin
deprotection and a Wittig reaction (61% for two steps) Fig. itself and was designed to be sufficiently general as to be
(4b). This provided the C9 and C11 stereocenters from adaptable to the structurally-related leustroducsins,
naturally-occurring D-glucofuranose. Carbonate phoslactomycins/phospholine, sultriecin, and cytostatin, for
saponification (100%), oxidative diol cleavage, and which no synthetic efforts have yet been described. The
subsequent oxidation with PCC/CH3OH (60%) yielded the strategy for our total synthesis of fostriecin is highlighted in
methyl ester 16. In order to complete the installation of the Fig. (5) [9]. The key elements include the late stage
Z,Z,E-triene, 16 was coupled to the TBS derivative of deprotection and sequential introduction of the C-terminus
commercially available (2E)-penten-4-yn-1-ol utilizing a lactone and C9 phosphate, a Wadsworth–Horner–Emmons
modified Sonogashira reaction (88%), and the resulting installation of the C6–C7 trans double bond, and stepwise
alkyne 17 was reduced by catalytic hydrogenation utilizing assembly of the sensitive Z,Z,E-triene. The absolute
Brown's catalyst (NaBH4, Ni(OAc)2). Conversion of 18 to stereochemistry of the C9 center was set by utilizing an
the β-ketophosphonate 19 and Wadsworth–Horner–Emmons optically active starting material, the C5 and C11

OH
SEt
p-TsOH
MeO2C SEt SEt
< 35% O O
OH OH
SEt
12
MsCl,
95%
Et3N

Br 2, Et2O/H2O
SEt
O O CHO O O

13 SEt
2008 Current Medicinal Chemistry, 2002, Vol. 9, No. 22 Boger et al.

(Fig. 4). contd.....

OR1 OR1 O O Br
Ph3P=CHBr
O 61%, 2 steps O
R2
O O
O O 15
R 1 = H, R2 = CH(SEt) 2 (14)
R1/R 1 = C 5H8, R2 = CH(SEt) 2 NaOH 100%
R1 /R1 = C5H8, R 2 = CHO

OTBS

OTBS O O Br

(Ph3P) 4Pd, CuI


R
O O 88%
R = CH(OH)CH2OH
M eO2C NaIO4
MeOH, PCC R = CHO
17 DMF, 60% R = CO2Me (16)

NaBH4,
OTBS
Ni(OAc) 2 O O
R

O
(MeO)2POCH2Li R = OM e (18)
R = CH2PO(OMe) 2 (19)

50% 13, NaH

OTBDPS
O O

O O
O
65% AlMe3

OTBDPS
O O

O O
Me OH 20

OH
OH OH

O O
Me OH 21
Fig. (4).

stereocenters were installed by Sharpless AD reactions, and (75%) via 23, conversion to the α-phenylselenyl lactone
the labile C8 center was installed enlisting a Felkin–Anh (50%), and subsequent oxidation to the corresponding
versus chelation-controlled nucleophilic addition. selenoxide provided the unsaturated lactone 24 following in
situ elimination (85%) [5]. Reduction of the lactone with
The 5R stereogenic center found in the lactone was Dibal-H and formation of the isopropyl acetal yielded 25 in
installed by a Sharpless asymmetric dihydroxylation (AD) a 90% overall yield. Removal of the primary alcohol
reaction [10] on p-methoxybenzyl 5-hexenoate (22, 79%, protecting group (91%) and subsequent oxidation with
88–92% ee), Fig. (5). Acid-catalyzed lactone formation TPAP–NMO provided the key aldehyde 26 (90%).
Fostriecin: Chemistry and Biology Current Medicinal Chemistry, 2002, Vol. 9, No. 22 2009

6 W adsw orth-Horner-E mmons 4 Wit tig Synthesi s of C1-C6


S harpless A D 17 S harpless A D TFA , 75%
OR
OR OR 14 PM BO 2C PM BO 2C
18 79%, 90% ee
HO CH 2OR
8 10 13
RO O 5 Stille coupling R =H
22 TB DP SCl, 89%
- O 3 Still -G enna ri R = T BD PS (23)
7 Me
R
Fel kin Addition 2 Sha rple ss AD H 2O2
1 from D -Glu
OT BD PS 85% O T BD PS
OR O O O O
OR OR
24
RO R=H
P PhSe Br, Dibal-H ;
RO O CH O 90%
RO L D A, 50% R = S ePh PPT S, i-PrO H
O O
C 1-C6 C7-C18

T PA P-N M O
OR OR
90% OR
O O C HO O O
CHO Bu 3S n OR
R = T BDPS (25)
O OR 26 Bu 4N F, 91%
OH R=H
C 1-C6 C8-C12 C 16-C18
OH O TBS
D ibal -H Sha rple ss AD T BSO T f, E t3N
RO PM BO PMBO PM BO
O M sCl, E t3N (D H QD ) 2AQ N
O OH 90%
O OH 88%
O O
73% >83%, > 10:1
PM BO C(=N H)CCl3
28 30
R = H (27) 29
95% Br
R = P MB

OR OT BS CO 2M e T ES O OT BS R TE SO O T BS Br
90% 94% , 2 ste ps
P M BO P MB O RO
Diba l-H CBr 4, PPh 3
B u 3SnH , (Ph 3P )4P d 84%
D DQ , 95% R = PM B
E t3S iOT f De ss-M a rti n
R = H (31) R = CH 2O H (32) R= H
95% 100% Ac 2O , 93%
R = TE S R = CH O (33) R = Ac (34)

OT BD PS O T BDP S
T E SO O T BS T ES O OT BS 82% >98:2 TE S O O TBS Br
Et O A cO
P (E tO )2PO CH2 Li R Bu 3Sn O T BDP S
Et O De ss-M a rt in 90% , 2 steps
O O 40 R = CH2O A c (37) 36 35
Diba l-H , 98%
R = CH2O H (38)
26, t-BuO K, tolue ne, 91% De ss-Ma rtin, 91%
R = CHO (39)
O T BD PS OT BD PS
T E SO O TBS TE SO O TB S
M e Li-Ce Cl3
O O 41 O O 42
96%
O Me OH
3:1
93% P PT S-E tOH
O T BD PS OT BD PS
OR O TB S A g 2CO 3, 81% OH O T BS

O O RO O
1
M e OT BS Me OR
R = E t, R 1 = H (43)
PCl3 ; PM BO H; R = H (48) T BSO T f, 70% 1
H2O2, 91% R = P O(OP MB) 2 (49) R = E t, R = T BS (46)
a q H Cl, 50% R = H , R 1 = T BS (47)
H F;
43
H F-Pyr

OH OR
H 2O 3PO OH OR OR

O O O O
M e OH Me OH
R = H (44)
1, fostrie cin A c2O , 70%
R = Ac (45)
Fig. (5). First total synthesis of fostriecin.
2010 Current Medicinal Chemistry, 2002, Vol. 9, No. 22 Boger et al.

The C7–C18 subunit of 1 was prepared by first allow for the introduction of the phosphate ester, yielded 43
assembling C8–C12 with introduction of the C9 and C11 (93%). However, prior to completion of the synthesis,
stereocenters, followed by a stepwise introduction of the correlation of synthetic and naturally-derived 44 and 45 was
sensitive triene, and finally conversion to the β - carried out to ensure that the intermediates bore the correct
ketophosphonate 39 for linkage to C1–C6. The stepwise relative and absolute configurations. Reprotection of the
order of the triene introduction permitted the assessment of secondary alcohol 43 as its OTBS ether (TBSOTf, Et3 N,
the relative stability of the C12–C13 Z-alkene, the C12–C15 CH 2 Cl 2 , –78 °C, 4 h, 77%), followed by aqueous acid
Z,Z-diene, and the C12–C17 Z,Z,E-triene, and offered the hydrolysis of the acetal (0.5 N aq HCl–acetone 1:4, 50%),
potential of postponing the introduction of any of its provided the lactol which was cleanly oxidized to the lactone
sensitive segments to a late stage of the synthesis. The (Ag2CO 3–Celite, 80%). Deprotection (HF–pyr, THF–pyr,
absolute stereochemistry of the C9 center was set by 25 °C, 4 d, 32%) furnished a 3:1 mixture of C8
utilizing the optically active hydroxylactone 27, available diastereomers of 44, of which the major C8 diastereomer
from D-glutamic acid in two steps [11,12]. The absolute correlated with 5 derived from the natural product [5]. In
stereochemistry at C11 was set utilizing a Sharpless AD addition, the 9,11,18-triacetate 45 was prepared from the
reaction on the advanced intermediate 28 (>83% yield, naturally-derived 5 by acetylation (Ac2 O, 70%) and
>10:1 ee), with clean delivery of the C3 alcohol cis to the correlated with the major isomer of a sample of 45 prepared
C5 hydroxymethyl substituent. Selective protection of the from synthetic 43 (i. Bu4NF; ii. Ac2O; iii. 0.5 N aq. HCl–
C3 secondary alcohol over the anomeric alcohol was acetone 1:4; iv. Ag2CO3–Celite).
achieved by direct, low temperature (–78 °C) treatment of 29
with TBSOTf (1.0 equiv, CH2Cl2) in the presence of Et3N In the course of handling 42 and its related C9 OTBS
(3.0 equiv), providing lactol 30 in 90% yield. Subsequent derivative, a base-catalyzed migration of the C9 silyl group
condensation with the Still–Gennari phosphonate [13] to the adjacent C8 tertiary alcohol was observed. We took
((CF 3 C H 2 O) 2 POCH 2 C O 2 Me, KHMDS, 18-crown-6) advantage of this observation and developed conditions
afforded 31 (88%) which contains the first cis olefin of the whereby treatment of 43 with TBSOTf (2.0 equiv, 5 equiv
sensitive Z,Z,E-triene. TES protection of the secondary of 2,6-lutidine, –20 °C, 45 min) provided predominantly
alcohol (95%), followed by stepwise reduction–oxidation, 46. Hydrolysis to the lactol 47, and oxidation to the lactone,
provided aldehyde 33. The cis olefin at C14–C15 was cleanly provided 48, at which point the separation of the C8
introduced utilizing a two-step protocol which enlisted a diastereomers was straightforward. Thus, only four
Corey–Fuchs reaction [14] (94% from 32), an intervening intermediates in the total synthesis (42, 43, 46, and 47) were
protecting group exchange, followed by a selective reduction prepared and characterized as the 3:1 mixture of C8
of 34 with Bu3 SnH–Pd(PPh 3 )4 [15–17] to provide the Z- diastereomers. Global silyl ether deprotection of 48 was
bromoalkene 35 (84%). In the initial efforts to elaborate the accomplished by treatment with HF–pyr in THF–pyr (25
cis bromide to the requisite cis,cis,trans-triene, typical Stille °C, 4 d) with a solid NaHCO3 workup, providing the tetraol
coupling conditions [18,19] provided low yields for the 4 4 which also proved to be spectroscopically and
homologation of 3 5 with vinylstannane 36 [20,21]. chromatographically indistinguishable from naturally-derived
However, the Stille coupling conducted in i-Pr 2 N E t 5. Conversion of 48 to the bis-PMB protected C9 phosphate
provided the key triene intermediate 37 in excellent yield occurred upon treatment with PCl3 (5 equiv, pyridine, 25
(82%) with near perfect stereochemical integrity (>98:2). °C), followed by PMBOH (15 equiv) and subsequent
Deprotection of the C8 alcohol and oxidation to the phosphite oxidation (50 equiv H2O2–H2O), to provide 49 in
aldehyde, followed by addition of α -lithio diethyl a superb conversion (91%). Subsequent global deprotection
methylphosphonate to the aldehyde 39 and oxidation of the (HF–pyr) provided fostriecin identical in all aspects with
secondary alcohol with Dess–Martin periodinane [22], authentic material.
afforded the key β-ketophosphonate 40 (90% from 39).

This set the stage for a Wadsworth–Horner–Emmons BIOLOGICAL PROFILE OF FOSTRIECIN


reaction to unite the C1–C6 and C7–C18 subunits, which
was carried out smoothly to furnish the desired trans α,β- Mode of Action
unsaturated ketone 41 in 91% yield. Direct addition of MeLi
to 41 via Felkin–Anh attack was anticipated to provide 42. Inhibition of Topoisomerase II
However, treatment of 4 1 with MeLi in THF gave
predominantly the undesired 1,4-adduct [23] (90%, >10:1). The biological activity and preclinical studies of
Alternative conditions including MeLi/18-crown-6 [24], fostriecin have been reviewed [27]. Fostriecin was initially
MeLi/LiCl, or MeTi(OiPr)3 [25], either did not substantially reported to inhibit macromolecular synthesis in the L1210
improve the selectivity or led to no reaction. Ultimately, it cell line [28]. Experimental data suggested that the mode of
was found that when ketone 41 was introduced as a toluene action was through the inactivation of an essential
solution into a MeLi/CeCl3 [26] slurry in THF, the 1,2- component of the cell cycle responsible for mediating DNA
adduct 42 was formed (>20:1) in 96% yield as a mixture of and RNA synthesis. Subsequently, it was proposed that
two diastereomers (3:1, 1H NMR). Correlation of the major fostriecin's antitumor activity and macromolecular synthesis
isomer with natural product derivatives established its inhibition could be attributed to the direct inhibition of
stereochemistry as 8R derived from a Felkin–Anh addition. topoisomerase II (topo II) [29]. The role that topo II plays in
the cell cycle, as well as its mechanism of action, have been
Exchange of isopropyl acetal to an ethyl acetal, with reviewed [30].
concomitant deprotection of the C9 secondary alcohol to
Fostriecin: Chemistry and Biology Current Medicinal Chemistry, 2002, Vol. 9, No. 22 2011

Topo II functions to unwind double-stranded DNA carcinoma. The minimum inhibitory concentration (MIC)
during two distinct phases of the cell cycle. In the S phase, was determined to be 50 µM with a modest K i,app of 110
topo II relaxes supercoiled DNA generated for transcription µM. Unlike the classical topo II inhibitors, fostriecin did
and DNA synthesis, but this function is thought to be not induce DNA strand breaks [29]. Several subsequent
nonessential and can be replaced by topoisomerase I. In the reports have described studies which have tried to establish
M phase, topo II is required for the unwinding of sister fostriecin’s similarities to known topo II catalytic inhibitors.
chromatids and subsequent chromatid segregation [31]. The Withoff, et al. reported a significant correlation between
enzyme functions by passing an intact DNA helix through a decreased topo IIα mRNA levels and increased fostriecin
transient double-stranded break which it generates by sensitivity in human small-cell lung carcinoma cells [53],
cleaving a phosphodiester bond in each strand of a second supporting fostriecin's classification as a topo II catalytic
DNA helix [30]. inhibitor. In addition, an etoposide-resistant GLC4/ADR cell
line (etoposide IC 50 = 10.2 µ M), characterized by a
Classical inhibitors of topo II (topo II poisons), decreased topo II level, showed a modest 3-fold increase in
including the epipodophyllotoxins (etoposide and sensitivity to fostriecin (IC50 = 4.1 µM) relative to the wild-
teniposide) [32–34] and 4'-(9-acridinylamino)methanesulfon- type GLC4 cell line (etoposide IC50 = 0.16 µM, fostriecin
m-anisidide (m-AMSA) [32,33,35], induce irreversible DNA IC 50 = 11.2 µM) [54]. However, a related study reported
strand cleavage during replication in the S phase of the cell only a 1.3-fold increased sensitivity to fostriecin in a
cycle [36–38]. These agents stabilize the covalent interaction teniposide-resistant CCRF-CEM cell line also characterized
between topo II and double-stranded DNA, serving to trap by a decreased topo II level [55].
the enzyme–DNA complex. It is the trap of this complex
and subsequent DNA cleavage, rather than the inhibition of The effects of various catalytic inhibitors on the cell
the catalytic activity of topo II, that is important, and a cell's cycle distribution have been evaluated and could potentially
sensitivity to such drugs can be expected to be directly be used to determine whether fostriecin’s profile merits its
proportional to the level of topo II activity. The classical classification as an inhibitor of topo II. Merbarone,
inhibitor’s activity as an anticancer drug is the result of RP60475F, SN 22995, and ICRF-187 induced G2 arrest
induced G2 arrest and/or apoptosis in replicating cells. Other without accumulation of cells in another stage, while the
topo II inhibitors do not stabilize topo II–DNA complexes. bis(2,6-dioxopiperazine) derivative ICRF-193 caused CHO
Rather, they prevent the binding of the enzyme to the DNA, cells to pass through the S phase and G2 checkpoint and
or block additional steps in the topo II catalytic cycle. A enter mitosis, resulting in polyploidy and tangled chromatin
cell's sensitivity to such catalytic inhibitors, therefore, [55,40,56]. Clearly there is some variability in the point of
should be inversely proportional to the level of the enzyme the cell cycle at which these drugs act. However, both
activity [37,39,40]. The mechanism of such inhibitors, observations are consistent with inhibited topo II activity
which include the bis(2,6-dioxopiperazines) [41,42], [40,57]. Flow cytometry studies with fostriecin have
merbarone [43], aclarubicin [43], amsacrine [44,45], suramin produced results that, in some cases, vary significantly from
[46], quinobenoxazine [47], chloroquine [48], and those of the topo II catalytic inhibitors. Initial studies with
novobiocin [49] have not been as well elucidated as those of L1210 cells showed an increase in the G2/M fraction and a
the classical inhibitors. It has been proposed that fostriecin's constant S fraction at 1.5 µM (growth inhibitory but not
properties could be attributed to a direct inhibition of the highly toxic) and 5 µM, but a 15 µM concentration of drug
topo II enzyme prior to or without an interaction with DNA. produced a lower proportion of cells in the G2/M phase and
However, recent studies suggest that the drug's site of action a greater number in the S phase [29,51]. Hotz, et al. reported
may be upstream of topo II within the cell cycle control that a concentration of 1 µM fostriecin resulted in significant
pathways. accumulation in the S phase for both promyelotic (HL-60)
and lymphocytic (MOLT-4) leukemia cells. At
In 1984, a series of studies that attempted to elucidate concentrations between 5 and 500 µM, the drug triggered
fostriecin’s mode of action were reported by groups within endonucleotic DNA degradation (apoptosis) in HL-60 cells,
Warner–Lambert/Parke–Davis. Fry, et al. showed that resulting in cell death with little change in phase proportions
L1210 leukemia cells exposed to 10 µ M fostriecin versus control, whereas no apoptosis was observed in
demonstrated a decreased rate of DNA, RNA, and protein MOLT-4 cells at any concentration [58].
synthesis. The reduction in nucleic acid synthesis was
shown not to be the result of nucleic acid precursor The drug also displays additional characteristics that
depletion, and 50 µM of fostriecin had no effect on DNA or differ from those expected for a topo II catalytic inhibitor. In
RNA polymerase activity. In addition, adding DNA to drug- a study similar to that of the etoposide-resistant GLC4/ADR
treated cells restored DNA polymerase activity, indicating cell line (vida supra), no cross-resistance or increased
that the limiting factor was the quantity of functional DNA sensitivity to fostriecin was observed in a K562 cell line
[50,51]. Gedik, et al. also found a strong and delayed with a 26-fold resistance to etoposide, whereas a 3.4-fold
inhibition of DNA synthesis by fostriecin in human HeLa sensitivity was exhibited in the presence of the topo II
cells [27,52]. It was also shown that transport of fostriecin catalytic inhibitor ICRF-187. Alone, this lack of cell
into L1210 leukemia cells occurs through the reduced folate sensitivity to fostriecin casts doubt on its classification as a
carrier system which contributes to its properties [50]. catalytic inhibitor. However, the catalytic inhibitor
merbarone also showed results (4-fold resistance) that would
The inhibition of topo II catalytic activity by fostriecin not be predicted [59]. Nonetheless, additional studies also
was first observed in a supercoiled DNA decatenation assay cast doubt on topo II as fostriecin’s primary target. It has
using partially purified topo II from Ehrlich ascites been reported that the catalytic inhibitors merbarone,
2012 Current Medicinal Chemistry, 2002, Vol. 9, No. 22 Boger et al.

aclarubicin, and the bis(2,6-dioxopiperazine) derivatives can up to 800 µM fostriecin in CHO cell extracts [61], and,
inhibit etoposide- and m-AMSA-mediated DNA–topo II unlike fostriecin’s effects on L1210 cells (vida supra),
complex formation in vitro [42,43,60]. These findings were merbarone and SN 22995 caused no inhibition of DNA
mirrored by Chen, et al. for teniposide-mediated DNA–topo synthesis in CCRF-CEM cells [55].
II complex formation with merbarone (IC50 = 66 µM),
aclarubicin (IC50 = 2.5 µM), RP60475F (IC50 = 7.2 µM), Inhibition of Ser/Thr Protein Phosphatases
and SN 22995 (IC 5 0 = 19.8 µ M) (two amsacrine
derivatives), whereas no inhibition was observed with More recent studies suggest a different mechanism for
fostriecin (up to 200 µM) under identical conditions. In fostriecin’s antitumor effects, implicating the Ser/Thr protein
addition, Frosina, et al. reported that no inhibitory effect on phosphatase family as the key cellular target [62]. This
topo II’s decatenating activity was detected in the presence of family is a highly homologous (35–65%) class of

Acidic group interacting with dinuclear metal center mimics Inhibitor IC 50 /PP1 IC 50 /PP2A selectivity: PP1/PP2A
substrate phosphate
Fostriec in 45 µM 1.5 nM 10 4-10 5
Elect rophile like microcystin-LA H-bond to Arg221
dehydroalanine?
Microcystin-L A 0.1 nM 0.1-1 nM 1-0.1
OH
H2 O3 PO OH
Tautomyc in 0.2 nM 1 nM 0.2
O O
Me OH
T hr substrate M e Hydrophobic tail tha t binds in (–)-Calyculin 0.5-2 nM 0.1-1 nM ca. 1
mimic substrate binding site

Oka daic acid 20 nM 0.2 nM 100


Proximal OH that mimics substrateThr -OH or displaces PP2A
active site bound water nucleophile? selectivity?
Ca ntharidin 470 µ M 40 µ M 10

CO2 H Me O Forms covalent Michael adduct with thiol of Cys273 of PP1 (x-ray)
N
HN NH
O
OM e O
O

NH HN
H H
N N
O
M icrocystin-LA O
O
CO2 H

O OH O

O OH O OH H
O O
O
O H O
OH O O Me
O
Me O N T automycin
H
(Me )2 N OH N

(–)-Calyculin

N O
H2 O3 PO
C
OH
O

OH OH O Me

O OH
H
O O
HO
O O O
O HO
O OH O O
O H
OH
Common M otifs: O kadaic acid
O Me 1. acidic group: carboxylate (masked) or phosphate
Me 2. methyl substituent
Cantharidin 3. hydrophobic segment

Fig. (6). Structural features of fostriecin fitting the pharmacophore model.


Fostriecin: Chemistry and Biology Current Medicinal Chemistry, 2002, Vol. 9, No. 22 2013

bimetalloenzymes of which several subtypes have been lower concentrations (1–50 nM), similar to those at which
identified. These include PP1, PP2A, PP2B (calcineurin), fostriecin was tested, indicated that okadaic acid arrests cell
PP2C, PP4, PP5, PP6, and PP7. PP1 and PP2A have been growth, and no changes in cell morphology were detected
the most studied [63–69]. Because of the high degree of (72 h at 2 nM okadaic acid) [82]. Furthermore, 8 nM
homology that exists in the family, differentiating which okadaic acid prevented transformed NIH 3T3 cells from
enzyme is responsible for regulation of a particular forming colonies, indicating reversion to a normal
biochemical pathway has proved challenging. The use of phenotype [83].
natural product inhibitors that target individual members of
the family has become an indispensible tool in determining Implication of PP2A/PP4 in G2/M Checkpoint Control
specific function [70], Fig. (6). As discussed in detail
below, fostriecin is a potent and selective inhibitor of two of Regardless of the origin of the selectivity demonstrated
the phosphatases, PP2A and PP4, both of which are by fostriecin for PP2A and PP4, links have been established
implicated in regulating the progression of the cell cycle between its inhibition of these enzymes and interference with
[65,71,72]. The fact that fostriecin and other known PP2A- normal regulation of mitosis [82,84–86]. Although PP2A
selective inhibitors, including okadaic acid and calyculin, and PP4 share 65% sequence homology, their functions are
induce premature entry into mitosis is further evidence independent [65,72,86]. There are two proposals on the
against the topo II inhibition model [73–76], as topo II is possible role of fostriecin-mediated PP2A inhibition in
not required for entry into mitosis [77]. The PP2A substrate regulating the tumor cell's entry into mitosis. One suggests
p34cdc2 kinase, together with cyclin B, however, comprises that fostriecin induces premature mitosis, whereas the second
the maturation promoting factor (MPF), which is the key suggests that it prevents mitosis from beginning. Less is
enzyme involved in initiating the M phase of the cell cycle known about the function of PP4, although it has been
[78–80]. shown to be involved in centrosome-mediated mitotic
spindle formation, and this will be described below in the
Binding Affinity of Fostriecin for Ser/Thr Protein discussion of the PP2A inhibition mechanisms.
Phosphatases
Premature Mitosis Hypothesis
A number of studies have established that fostriecin is a
potent inhibitor of the Ser/Thr protein phosphatases PP1, A basic outline of the cell cycle is represented in Fig.
PP2A, and PP4 (Table 1). These studies also demonstrated (7). Cells proceed through a resting G1 phase (first gap),
that the molecule is on the order of 100–10,000 times more followed by the S phase during which the synthesis of DNA
selective for PP2A and its close homolog PP4, than PP1, and histones occur. These first two stages are followed by
making it the most selective inhibitor known for any the "quality control" phase, G2 (second gap), in which
member of this class of phosphatases. To date, the only transcriptional errors are corrected [87]. Entry into the M
other PP2A-selective inhibitors known are okadaic acid (mitotic) phase, during which cell division occurs, is
(~100:1), cantharidin (~10:1) [81], and calyculin (~5:1). regulated by the maturation promoting factor (MPF)
complex which consists of the Ser/Thr kinase p34cdc2 and
Table 1. Fostriecin Inhibition of Ser/Thr Protein cyclin B [79,88]. When activated, the complex is thought to
Phosphatases stimulate normal chromosome condensation [84]. Many
antitumor agents interfere with the normal cell cycle control
Substrate IC50 PP1 IC50 PP2A IC50 PP4 mechanism or mitotic entry checkpoint that occurs at the
interface between the G2 and M phases. Usually the agents
1a phosphorylase a 45 nM 1.5 nM 3 nM do so by initiating G2 arrest, a normal control mechanism
2b phosphorylase a 131 µM 3.2 nM - for preventing a cell from undergoing division with
imperfect DNA [87].
3c phsophorylase 4 µM 40 nM -
chromosome condensation
4d phosphohistone 58 µM 5.6 nM -
aref 86. bref 127. cref 62. dref 82. M
MPF MPF
inactive active
Okadaic acid inhibits both PP2A and PP4 phosphatases Mitotic Entry
with the same IC50 (0.2 nM) [70], but microcystin-LR is Checkpoint
PP4-selective (8 pM vs 0.1 nM for PP2A). This selectivity G2 G1
is interesting considering that PP2A and PP4 are ~65% DNA repair
homologous. Unlike fostriecin, okadaic acid and cantharidin S
do not inhibit topoisomerase. Fostriecin does not inhibit
PP2B (calcineurin) or protein tyrosine phosphatases, as is DNA and histone
typical of the other PP1/PP2A-specific natural product synthesis
inhibitors. Notably, fostriecin and cantharidin are established Fig. (7). Schematic of the cell cycle.
antitumor agents while okadaic acid is categorized as a
tumor promoter. However, the tumor promoting activity of An early study of fostriecin's effects on the cell cycle
okadaic acid was determined in a mouse skin model at high showed that at a 375 µM concentration it induced baby
concentrations (123 µM). Experiments in CHO cells at hamster ovary (BHK) cells to enter mitosis prematurely
2014 Current Medicinal Chemistry, 2002, Vol. 9, No. 22 Boger et al.

rather than suffer G2 arrest, as evidenced by chromosome (0.5–15 nM) in CHO cells [27,82,100,101]. The antitumor
condensation, separation of spindle poles, and formation of activity occurs at a level that completely inhibits PP2A but
asters [62]. Premature mitosis with damaged, incomplete, or not PP1. If some PP1 is available, the cells will not proceed
incorrect DNA sequences ultimately results in cell death into mitosis because they are damaged, resulting in G2
[89,90]. Okadaic acid also instigated BHK21 cells to arrest. The mechanism by which fostriecin-induced G2 arrest
proceed through the mitotic entry checkpoint prematurely at occurs is likely through disruption of the normal changes
0.5 µM concentration [75]. Subsequently, several studies that take place in the cytoskeleton during mitosis. The two
were undertaken to define how this process occurs with major abnormalities noted by Honkanen and coworkers were
fostriecin and its implications. aberrant mitotic spindle formation and atypical centrosome
replication [82]. Normally, the microtubules which form the
Based on substantial precedent, one question deserving mitotic spindle radiate from phosphorylated centrosomes.
investigation was the nature of fostriecin's effect on the Fostriecin-, okadaic acid-, and cantharidin-treated cells
regulation of p34cdc2 kinase, histone H1, and histone H3, all displayed numerous spindle poles extending from multiple
of which are implicated in stimulating chromosome centrosomes. All three of these agents are PP2A-selective,
condensation [84]. Roberge and coworkers discovered that and mutant Drosophila which do not express PP2A also
FT210 cells in G 2 arrest underwent chromosome exhibited multiple centrosomes and associated spindles
condensation without active p34cdc2 or histone H1 when [102].
treated with fostriecin (≥25 µM) or okadaic acid (1 µM).
However, an increase in phosphorylated histones 2A and 3 It is unclear whether exclusive inhibition of PP2A or
was observed. In the presence of the kinase inhibitor PP4 results in fostriecin’s activity, or if inhibition of both
staurosporine, fostriecin was not able to induce chromosome enzymes is necessary for optimal activity. In Drosophila and
condensation, implicating activation of H2A and H3 in mammalian cells, centrosomes are stained by PP4-sensitive,
stimulating this process. Pulse-chase experiments with 32P but not PP2-sensitive, antibodies, and PP4 is also thought
also showed that, in mitotic cells treated with fostriecin, to be necessary for the initiation of microtubule growth from
there was an increase in the amount of radioactive label centrosomes [65,72]. Therefore, PP4 may also play a role in
incorporated into H2A and H3 compared with controls. This fostriecin’s antitumor effects. In conclusion, considerable
suggested that either H2A and H3 phosphatases were evidence supports the inhibition of protein phosphatases as
inhibited or H2A and H3 kinases, as yet unidentified, were fostriecin’s mode of action.
activated. An additional interesting finding in this study was
that p34cdc2 kinase activity was decreased in FT210 cells in Is There a Pathway Connecting PP2A and Topo II?
the presence of fostriecin. By contrast, BHK cells showed an
increase in p34cdc2 kinase activity when treated with Some of the known substrates of PP2A are enzymes
fostriecin (375 µM) [62]. Okadaic acid also stimulated an which have been shown to regulate topo II. It is possible
increase in p34cdc2 kinase activity in BHK21 cells [75]. that by inhibiting PP2A, fostriecin may indirectly affect the
activation of topo II, Fig. (8). PP2A is a negative regulator
A second investigation led to the discovery that the of MPF both directly, via dephosphorylation of a threonine
filamentous protein vimentin, which is commonly found in residue of active MPF, and indirectly, through deactivation
developmental and tumor cell lines, was
hyperphosphorylated in cells administered with 50–200 µM
Inactive MPF
fostriecin [85]. The same effect has been observed with
okadaic acid in rat brain tumor cells and with calyculin A in Cyclin B
NIH3T3 fibroblasts [91–93]. During mitosis, TP
TP
hyperphosphorylation of vimentin destabilizes the normal YP
p34cdc2
cytoplasmic filament network by causing the formation of
aggregates. Although vimentin is phosphorylated by p34cdc2
kinase in mitosis [94–96], the results of this study indicate wee1
that hyperphosphorylation of vimentin occurs even in cells cdc25
where p34cdc2 activity is suppressed. Ultimately, one of the Inactive MPF
Active MPF
isoforms of protein kinase C (PKC) was implicated as the
enzyme responsible for vimentin hyperphosphorylation. Cyclin B Cyclin B
TP PP2A T
Protein kinase C is known to be deactivated by PP2A and is T T
Y Y
implicated in affecting cell morphology. Indeed, the p34cdc2
MO15
p34cdc2
overactivation of PKC effected by the phorbol esters and
related compounds results in tumor formation [97,98]. It has
been suggested that the tumor promoting activity of okadaic
acid is also the result of PKC activation through the Activates
inhibition of PP2A [85,99].

Activates
Cas ein Kinas e 2 Topoisomerase II
G2 Arrest Hypothesis

The G2 arrest hypothesis is based on the observation that Fig. (8). Activation of MPF and a biochemical pathway linking
fostriecin exerts its antitumor effects at low concentrations topoisomerase II and MPF.
Fostriecin: Chemistry and Biology Current Medicinal Chemistry, 2002, Vol. 9, No. 22 2015

of the phosphatase cdc25 and dephosphorylation of the carbon bearing the important methyl substituent. The role of
inhibitory kinase wee1, which must be dephosphorylated in this hydroxy group is presently not well understood, but
order to be active. PP2A's main function is to phosphorylate could serve as a mimic of the substrate Thr-OH or, more
tyrosine and threonine residues on MPF, keeping MPF in a provocatively, serve to displace the enzyme's metal bound
hyperphosphorylated, inactive state [103–107]. There may be active site water nucleophile.
a link between activated MPF and topoisomerase II through
MPF-mediated activation of the intermediary enzyme casein Additional studies indicate some analogy between an
kinase 2 (CK2), which is the major kinase known to important recognition element not directly implicated as part
phosphorylate topo II [108,109]. The other major enzyme of the pharmacophore and a similar structural feature present
that phosphorylates topo II is protein kinase C, which is in the other natural product inhibitors. Acetylation of the
also known to regulate MPF [110–113]. C11 alcohol of fostriecin, located α to the hydrophobic
triene, destroys activity. Similarly, a calyculin analog
Fostriecin May Fit the Okadaic Acid Pharmacophore displayed a dramatic increase in IC50’s with PP1 and PP2A
Model for PP1/PP2A Inhibition when its C11 and C13 alcohols, adjacent to its hydrophobic
tetraene, were protected as the acetonide [117]; and
Several groups have proposed a common pharmacophore phosphatase binding was abolished in tautomycin analogs in
based on extensive SAR of the natural product inhibitors of which the C22 hydroxyl, also located near its purportedly
PP1 and PP2A, which include okadaic acid, the calyculins, hydrophobic spiroketal segment, was removed by
the microcystins and nodularins, cantharidin, and deoxygenation or β-elimination [121]. Based on studies
tautomycin, Fig. (6). Further support for this model comes illustrating that this residue is important in recognition,
from two X-ray crystal structures: the first, of the phosphate molecular modeling has indicated that the calyculin and
mimic tungstate coordinated with the metal ions in the tautomycin alcohols may form key hydrogen bonds with
catalytic core of PP1, and the second, of the toxin Arg221 of PP1 [115,120,122,123].
microcystin-LR complexed in the PP1 active site [114,115].
Conserved features of this pharmacophore are (1) an acidic Specificity for PP2A/PP4– Does it Reside in the β 12–β 13
moiety – either a phosphate or a carboxylate – proposed to Loop?
interact with the phosphatase metal ions (as does tungstate
in the X-ray), (2) a methyl group proximal to the acidic The remarkable selectivity for PP2A and PP4
moiety, proposed to mimic the methyl group in demonstrated by fostriecin is noteworthy. The elements of
phosphothreonine, and (3) an extended hydrophobic segment the pharmacophore common to the natural product inhibitors
(except for cantharidin which is a µM inhibitor), thought to are only thought to affect binding affinity and not
mimic the hydrophobic amino acid residues adjacent to selectivity. One hypothesis that has emerged is that a
phosphothreonine in endogenous proteins which bind segment of the phosphatase catalytic pocket, known as the
PP1/PP2A [116–120]. Notably, the elements of the β12–β13 loop, is responsible for the selectivity imparted by
pharmacophore are also present in fostriecin and are the inhibitors. The catalytic pockets of PP1 and PP2A show
highlighted in Fig. (6). >90% sequence homology except in this region, and assays
of the inhibitors with β12–β13 deletion mutants of PP1
Unfortunately, systematic studies establishing whether indicate that the loop is critical for binding [124]. In
the pharmacophore features are required for binding of addition, most of the inhibitors are believed to form specific
fostriecin with PP1 or PP2A have not yet been conducted. contacts with this area based on SAR data [70].
However, some indirect information is available which Interestingly, the only study regarding fostriecin which has
supports this tentative pharmacophore correlation. For attempted to address this issue suggests that the β12–β13
example, dephosphorylated fostriecin shows considerably loop is not implicated in fostriecin’s selectivity.
diminished cytotoxic and antitumor activity. A naturally-
occurring congener of fostriecin, PD 113,270 (2), which Shenolikar and Honkanen constructed a chimera of
lacks only the C18 primary alcohol, is equipotent in human PP1α (CRHM2) in which the entire amino acid
cytotoxic assays, nearly as efficacious in antitumor assays in sequence of the β12–β13 loop was replaced with that of
vivo, and is transported into cells by the reduced folate PP2A (bovine) [124]. Fostriecin, okadaic acid, tautomycin,
carrier at least as effectively, and perhaps better, than 1. The microcystin-LR, calyculin A, and endogenous protein
studies also indicate there may be additional recognition or inhibitors of PP1 (I–1, I–2, NIPP1), which are substrates of
selectivity elements not yet implicated as part of the PP2A, were assayed against phosphorylase a with
pharmacophore which are more easily identified within the recombinant human PP1α , and CRHM2. The results
structure of fostriecin than other natural product inhibitors. showed that all of the natural product inhibitors bound
The PP1/PP2A-mediated phosphate cleavage is catalyzed in CRHM2 and PP1α with comparable affinity suggesting that
a single step by a dinuclear metal site activating a bridging the β 1 2 – β 13 loop is not involved in the selective
water molecule for hydroxide nucleophilic attack on the recognition. In contrast, the endogenous proteins inhibited
bound substrate, and active site residues involved in the PP1 at nM concentrations while CRHM2 was effectively not
substrate activation (PP1, His125, and Asp95) have been inhibited, which was similar to reported assays with PP2A.
clearly identified. The comparison of the inhibitors that Therefore, the loop was implicated in preferential PP1
display significant and potent PP2A versus PP1 selectivity recognition of the endogenous inhibitors. This disparity is
(fostriecin, okadaic acid) with the remainder which do not especially interesting when considered with data reported for
reveals that these two uniquely bear a hydroxy group another PP1 β12–β13 loop chimera. Previously, Zhang and
immediately adjacent to the anionic group on the same coworkers substituted a short segment of the loop in PP1
2016 Current Medicinal Chemistry, 2002, Vol. 9, No. 22 Boger et al.

(GEFD, residues 274–277) with the corresponding residues Fostriecin was reported to inhibit the decatenation of
(YRCG) of PP2A. Okadaic acid showed a lower IC50 than kinetoplast DNA by partially purified topoisomerase II in an
for PP1 with this mutant suggesting that the origin of its in vitro assay with an MIC of 50 µM (Table 2). Several
PP2A selectivity could arise from interactions with this natural or semisynthetic analogs of the natural product were
segment. In addition, the IC50's of microcystin and calyculin examined in the same assay. Addition of a hydroxyl group
were also affected, showing an increase with the mutant at the C4 position on the lactone ring increased potency 4-
enzyme [125]. Shenolikar and Honkanen concede that there fold. A 2-fold increase in potency was observed upon
may be a conformational difference in this region between removing the phosphate ester or opening the lactone ring to
the chimera and PP2A arising from incorporating the whole the free carboxylic acid. Acetylation of the primary and
loop sequence. Perhaps a more conclusive study would be to secondary hydroxyl groups decreased the inhibition to an
test the inhibitors against a PP2A chimera constructed with IC50 of 200 µM (4-fold) [29]. Notably, these trends do not
the loop of PP1 and determine whether fostriecin’s affinity follow those observed in functional cytotoxic assays.
is considerably diminished.
Table 2. Topoisomerase II Inhibition
Another finding by Honkanen and coworkers indicated
that fostriecin did not competitively inhibit PP1 in cells R1 R4
treated with both fostriecin and microcystin-LR. They OR2 OR 3
concluded that fostriecin must be binding a different site
than microcystin [126]. However, since fostriecin only binds O O
PP1 with µM affinity, whereas microcystin-LR has low nM Me OH
affinity and binds with a covalent linkage to the enzyme, it
may not be surprising that competitive inhibition is not Topoisomerase II
observed. Fostriecin was tested for PP1 and PP2A Compound R1 R2 R3 R4
µ M))
Inhibition (MIC (µ
inhibition against the known enzyme substrate
phosphorylase a, and it is difficult to imagine that its
phosphate group is not coordinating to the active site Fostriecin H PO3H 2 H OH 50
phosphate-binding metal ions. Site-directed mutagenesis PD 113,270 H PO3H 2 H H
experiments with key active site residues as have been
performed with the other inhibitors should confirm whether PD 113,271 OH PO3H 2 H OH 12.5
fostriecin binds in the active site.
PD 114,027a H PO3H 2 H OH
Undoubtedly, it will prove useful to probe the origin of PD 114,631 H H H OH 25.0
fostriecin’s selectivity with the preparation of structural
analogs. Fostriecin is structurally simple compared to the PD 116,243 H H H OH
other natural product inhibitors and these issues of binding
PD 116,245a OH PO3H 2 H OH 25.0
affinity and selectivity can be most easily and systematically
examined with its analogs. Finally, fostriecin's most PD 116,250 H PO3H 2 Ac OAc 200
significant feature that does not correspond to the
pharmacophore model is the unsaturated lactone. It is PD 116,251 OAc PO3H 2 Ac OAc
tempting to speculate that this group is necessary for the aLactone opened to carboxylic acid
observed selectivity, and to suggest a possible interaction
with the β12–β13 loop. For example, the enone may act as a
Michael acceptor analogous to the dehydroalanine moiety in Several lines of evidence indicate that fostriecin
the microcystins, or the latent acid may form ionic penetrates cells via the reduced folate carrier system. The
associations with active site residues. cultivation of a fostriecin-resistant L1210 cell line
(L1210/CI-920 r) and its cross-resistance to methotrexate,
which was shown to be derived from an impaired transport,
Biological Properties led to the initial discovery that 1 is transported into cells by
the reduced folate carrier system. It was demonstrated that
Fostriecin exhibits potent cytotoxic activity against the drug is a potent inhibitor of methothrexate influx
mouse leukemia (L1210, IC50 = 0.46 µΜ) and is about one exhibiting a mixture of competitive and noncompetitive
order of magnitude less active against human colon inhibition. When 50 µM of fostriecin was administered to
adenocarcinoma cells (HCT-8) [27]. Initial evidence of the in L1210 cells over 3-min intervals, the methothrexate influx
vivo anticancer activity of 1 was obtained against P388 rate fell to 37% of the control cells (Table 3). Interestingly,
leukemia-bearing mice (% T/C = 246). The drug was also this inhibition by fostriecin is transiently irreversible, but
active against B16 melanoma, but its most efficacious in reversed upon addition of thiols (dithiothreitol). The authors
vivo activity was observed with L1210-bearing mice, where suggest this may represent reversible covalent attachment of
an i.p. dose of 6.25 mg/kg/day on days 1–9 was curative fostriecin to the reduced folate carrier by a Cys thiol Michael
[51,100]. Fostriecin was devoid of antibacterial activity addition to the α,β-unsaturated δ-lactone. Removal of the
when tested against 11 organisms at 500 µg/mL [2], but it primary hydroxyl group showed little effect on the activity,
did exhibit significant antimycotic effects versus certain while addition of a hydroxyl group at the C4 position on the
yeasts with MIC values ranging from 30 to 300 µg/mL lactone ring decreased the inhibition of methothrexate entry
[127]. over 50% versus fostriecin. Opening the lactone ring lowered
Fostriecin: Chemistry and Biology Current Medicinal Chemistry, 2002, Vol. 9, No. 22 2017

Table 3. Fostriecin SAR Studies of Reduced-Folate Carrier, Cytotoxicity, and in vivo Antitumor Activity

R1 R4
OR2 OR 3

O O
Me OH

Methotrexate influx µ M))


Cytotoxicity (IC50 (µ Antitumor Activity (% T/C)a
Rate (% of control)
Compound R1 R2 R3 R4 L1210 HCT-8 L1210 P388

Fostriecin H PO3H 2 H OH 36.6 ± 4.4 0.46 5.1 207 246

PD 113,270 H PO3H 2 H H 31.5 ± 13.4 0.50` 3.4 180 178

PD 113,271 OH PO3H 2 H OH 57.2 ± 6.7 1.8 9.0 121 175

PD 114,027b H PO3H 2 H OH 71.2 ± 7.8 >2.1 120

PD 114,631 H H H OH 89.2 ±5.9 5.5 (6.7)c 9.9

PD 116,243 H H H OH

PD 116,245b OH PO3H 2 H OH 8.1

PD 116,250 H PO3H 2 Ac OAc 84.7 ± 1.4 >100

PD 116,251 OAc PO3H 2 Ac OAc


a(% T/C) represents increase in life-span of treated mice versus control. Sixty-day survivors not included.
bLactone opened to carboxylic acid.
cRepresents data from two separate experiments.

the activity 2-fold. Either C8 dephosphorylation or primary alcohol and the addition of a hydroxy group at the
acetylation of the C11 and C18 alcohols abolished most of C4 position had the same effect of lowering efficacy [2]. The
the inhibitory activity [50,51]. results of a more extensive optimization of the in vivo
activity are summarized in Table 4, which lists the efficacy
The cytotoxicity of fostriecin and several analogs were of fostriecin and the different analogs (i.p. administration,
investigated in both L1210 and HCT-8 cell lines (Table 3). L1210). The parent compound showed strong activity and
The parent compound proved to be an order of magnitude was "curative" on a daily dosing regimen of 5 or 9 days
more potent in the L1210 assay, but the structure–activity duration. Removal of the primary hydroxyl group increased
trends were similar in both cell lines. Removal of the potency 3-fold, but the efficacy was unchanged or lower
primary hydroxyl group at C18 had little effect on the drug compared to fostriecin. Hydroxylation of the C4 position
potency and addition of a hydroxyl group at the C4 position reduced the efficacy whereas dephosphorylation or hydrolysis
on the lactone ring decreased the cytotoxicity approximately of the lactone to the carboxylic acid virtually abolished
3- to 4-fold. Opening the lactone ring virtually abolished activity. Acetylation of the primary and C11 secondary
activity. Removal of the phosphate led to an order of alcohols of fostriecin or PD 113,271 also abolished activity
magnitude decrease in activity in the L1210 cell line, but [51,100].
only a 2-fold decrease for the HCT-8 cells. Concomitant
hydroxylation of the C4 position and lactone hydrolysis The antimycotic properties of fostriecin and four analogs
resulted in greater than a 16-fold loss of activity. As noted have also been examined (Table 5). The drug was effective
in previous studies, acetylation of the C11 and C18 alcohols against multiple yeast strains, with MIC values in the range
abolished the activity [29,51,100]. of 30–300 µg/µL. PD 113,270, which is devoid of the
primary alcohol, shows similar activity. Hydroxylation at
Tunac, et al. reported that fostriecin and its natural the C4 position of the lactone increased the antimycotic
product analogs PD 113,270 (2) and PD 113,271 (3) showed activity 10-fold. Removal of the phosphate group from
excellent antitumor activity in murine tumor models (Table fostriecin or PD 113,271 resulted in no activity [127].
3). The L1210 tumor-bearing mice treated with fostriecin
(i.p., 6.25 mg/kg/injection, daily × 9) exhibited increased
life spans of 207% relative to control animals (% T/C). CLINICAL STUDIES
Removal of the primary hydroxyl group lowered activity
slightly, whereas adding a secondary alcohol at the C4 A phase I clinical trial and pharmacokinetic study of
position lowered the activity more significantly versus the fostriecin was conducted in which the drug was administered
parent compound. Against P388, a similar but slightly intravenously over 60 min on days 1–5 at 4-week intervals.
different trend was observed, where both the removal of the The doses used ranged from 2 to 20 mg/m2 /day. The
2018 Current Medicinal Chemistry, 2002, Vol. 9, No. 22 Boger et al.

Table 4. Antitumor Activity of Fostriecin and Related Analogs

R1 R4
OR2 OR 3

O O
Me OH

Compound R1 R2 R3 R4 Dose (mg/kg/inj.) Schedule % T/Ca 60-day survivors

Fostriecin H PO3H 2 H OH 40 Days 3-7 267 0/6

25 Days 3-7 250 1/6

12.5 Days 3-7 289 1/6

12.5 Days 3-11 307 0/6

6.25 Days 1-9 Curative 6/6

PD 113,270 H PO3H 2 H H 8.12 Days 3-7 250 0/6

6.25 Days 3-7 271 0/6

6.00 Days 1-9 226 0/6

1.5 Days 1-9 180 0/6

PD 113,271 OH PO3H 2 H OH 25 Days 3-7 180 0/6

10 Days 3-7 140 0/6

17.5 Days 3-11 190 0/6

6.0 Days 1-9 128 0/6

6.0 Days 1-9 123 0/6

3.12 Days 1-9 128 0/6

PD 114,027b H PO3H 2 H OH 25c Days 3-7 100 0/6

PD 114,631 H H H OH 50c Days 3-7 109 0/6

PD 116,250 H PO3H 2 Ac OAc 50c Days 3-7 101 0/6

PD 116,251 OAc PO3H 2 Ac OAc 50c Days 3-7 102 0/6


a(% T/C) represents increase in life-span of treated mice versus control. Sixty-day survivors not included.
bLactone opened to carboxylic acid.
cHighest dose tested.

predominant toxicities included elevated liver transaminases inherent drug instability and unpredictable purity in the
and serum creatinine, but these effects showed only a limited clinical supply of the natural product.
increase with increasing doses, often recovered during drug
administration, and were fully reversible. Other frequent side Table 5. Antimycotic Activity of Fostriecin and Related
effects were grade 1–2 nausea/vomiting, fever, and mild Analogs
fatigue. Mean fostriecin plasma half-life was 0.36 h (initial;
95% Cl, 0–0.76 h) and 1.51 h (terminal; 95% Cl, 0.41–2.61 Compound Antimycotic Activity (MIC, µ g/mL)
h). A metabolite was detected in plasma and urine, and the
authors proposed that this may be dephosphorylated
Fostriecin 30-300
fostriecin. Antitumor activity was not observed; however,
the plasma concentrations reached in the patients were below PD 113,270 30-300
even those needed to cause in vitro growth inhibition, and PD 113,271 3-30
the dose-limiting toxicity was not reached with 20
mg/m2/day [128,129]. Although fostriecin was formulated PD 114,631 no activity
as the sodium ascorbate salt to increase its stability to PD 116,243 no activity
oxidation [101], clinical trials were suspended due to
Fostriecin: Chemistry and Biology Current Medicinal Chemistry, 2002, Vol. 9, No. 22 2019

3 15 17
2 4 14 16
18
H2O3PO OH
5 8 10
13 21 19
O O 7 9 11 20
6 12
OH R

H2N
O

46, R = H 53, R = O

O O

47, R = O 54, R = O

O
O

48, R = 55, R = O
O
O
O
56, R = O
49, R = O
O
O
57, R = O
50, R = O
O
O
58, R = O
51, R = O

O 59, R = OH

52, R = O

Fig. (9).

RELATED NATURAL PRODUCTS Phospholine was tested for in vitro cytotoxicity against
L1210, P388, and EL-4 tumor cell lines and was found to
Phospholine be 100-fold less potent (IC50 = 2–3.4 µg/mL) than that of
mitomycin C (IC50 = 0.02–0.07 µg/mL) [130,131].
Phospholine (46), Fig. (9), an amphoteric phosphate
monoester containing an α,β-unsaturated δ-lactone and a
conjugated diene system, was isolated from the culture broth Phoslactomycins A–F
of Streptomyces hygroscopius subsp. luteolus subsp. nov.
by Ozasa, et al. It was obtained by gel filtration, butanol Phoslactomycins A (47), B (46), and C–F (48–51), Fig.
extraction, and reverse phase chromatography. The molecular (9), phosphate monoesters containing an α,β-unsaturated δ-
formula and molecular weight were determined to be lactone and a conjugated diene system, were isolated from
C25H40NO8P (MW 513) from high resolution FABMS data the culture broth of Streptomyces nigrescens by Fushimi, et
and elemental analysis. A positive test for ninhydrin and al. They were obtained by butanol extraction, gel filtration,
ammonium molybdate-perchloric acid indicated the presence and reverse phase chromatography. The molecular formula
of an amino and a phosphate group, respectively. The IR and molecular weight for each compound were determined
absorbance at 1720 cm–1 suggested an α,β-unsaturated δ- from high resolution FABMS, 1H, 13C, and 31P NMR data
lactone and at 1100 cm–1 indicated a phosphate moiety. The (phoslactomycins A, D, and E), from high resolution
characteristic UV absorbance at 234 nm was indicative of the FABMS, 1H, 13C NMR data (phoslactomycin B), or from
α,β-unsaturated δ-lactone and/or a conjugated diene, and FABMS and 1H NMR (phoslactomycins D and F) (Table
enzymatic hydrolysis with alkaline phosphatase implied the 6). The functional groups present in the phoslactomycins A–
presence of a phosphate monoester. Phospholine's F were also initially elucidated utilizing a combination of
connectivity was determined by 1H NMR, 13C NMR, 1H– color reactions (ninhydrin: primary amino group,
1 H COSY, 1 3 C – 1 H COSY, and HMBC spectra. ammonium molybdate–perchloric acid: phosphate group).
2020 Current Medicinal Chemistry, 2002, Vol. 9, No. 22 Boger et al.

Since the UV and 1H NMR spectra of the phoslactomycins Phosphazomycin C


resembled each other so closely, they were assumed to be
structural analogs. The major component of the isolated Phosphazomycin C, an equimolar complex of two
material was phoslactomycin E, therefore its structure was phosphate monoesters, each containing an α,β-unsaturated δ-
elucidated first. lactone and a conjugated diene system, was isolated from the
culture broth of Streptomyces sp. HK-803 by Tomiya, et al.
Table 6. Molecular Formulas and Molecular Weights of
It was obtained by gel filtration, butanol extraction, and
Phoslactomycins A–F
reverse phase chromatography. The molecular formula of
each component was determined to be C30H48NO10P (MW
613) from FABMS data and elemental analysis. It gave
Phoslactomycin Molecular Formula Molecular Weight positive reactions to ninhydrin, Lemieux, and anisaldehyde–
sulfuric acid reagents. The connectivity of the two
A (47) C 29H 46O 10NP 599 components of phosphazomycin C (phosphazomycin C1 (48)
and phosphazomycin C2 (52)), Fig. (9), were determined
B (46) C 25H 40O 8NP 513 based on interpretation of 1 H NMR, 13 C NMR, 1 H– 1 H
C (48) C 30H 48O 10NP 613 COSY, 13 C– 1 H COSY, 1 H– 31 P COSY and long-range
proton decoupling (LSPD) spectra. Phosphazomycin C1 was
D (49) C 31H 50O 10NP 627
shown to be identical in structure to phoslactomycin C
E (50) C 32H 50O 10NP 639 [133]. Accordingly, the phosphazomycin C complex showed
potent antifungal activity [134].
F (51) C 32H 52O 10NP 641

Leustroducsins A–C
The IR spectrum of phoslactomycin E (50) showed major
absorptions at 1724, 1076, and 1250 cm–1 denoting the Leustroducsins A–C (53–55), Fig. (9), phosphate
presence of an α,β-unsaturated δ-lactone, P–O, and P=O, monoesters containing an α,β-unsaturated δ-lactone and a
respectively. The UV spectra (absorption at 233 nm) conjugated diene system, were isolated from the culture
suggested the presence of an α,β-unsaturated δ-lactone. Its broth of Streptomyces platensis SANK 60191 by Kohama,
connectivity was determined based on interpretation of 1H et al. They were obtained by ethyl acetate extraction and
NMR, 13 C NMR, 1 H– 1 H COSY, 13 C– 1 H COSY, DEPT, preparative reverse phase chromatography. The molecular
decoupling difference (DDS), and 3 1 P spectra. The formula and molecular weight for leustroducsins A, B, and
configuration of the conjugated diene system was established C was determined to be C3 2 H 5 2 N O 1 0 P (MW 641),
as Z,Z on the basis of the coupling constants (J 12,13 = 8.6 C 34H 56NO 10P (MW 669), and C34H 56NO 10P (MW 669),
Hz, J 14,15 = 9.5 Hz) and the C6–C7 stereochemistry was respectively, from high resolution liquid secondary ion mass
determined to be E (J6,7 = 15.1 Hz). The chemical shift of spectrometry (LSI–MS). The functional groups present in
C5-H (400 MHz, δ 5.10) indicated the formation of a the leustroducsins A–C reacted positively on TLC to iodine,
lactone ring including C1 to C5 which, the authors pointed sulfuric acid, ninhydrin, and ammonium molybdate–
out, is almost identical to the chemical shift value of the perchloric acid, indicating the presence of amino and
corresponding proton in fostriecin. The 1H and 13C NMR phosphorus groups. The initial UV, IR, 1H and 13C NMR
spectra suggested the linkage of a cyclohexane ring to a spectra of the leustroducsins resembled those of the
carbonyl carbon and its presence was confirmed by detecting phoslactomycins so closely that compounds 53–55 were
cyclohexane carboxylic acid in the alkaline hydrolysate of assumed to be structural analogs. The structural assignment
phoslactomycin E by GC–MS. Analysis of the 13C and 2D of the acyl chains, the only novel portion of the molecules
COSY NMR data (phoslactomycin A, C, D, F) or the 1 H throughout the phospholine, phoslactomycin,
and 13 C NMR data (phoslactomycin B) revealed that phosphazomycin C, and leustroducsin series, was made
phoslactomycin E and all five congeners contained the same based on 1H and 13C NMR, MS/MS, and LSI–MS/MS. In
carbon skeleton from C1 to C25. The only structural addition to the three novel leustroducsins isolated from the
difference between the compounds was the acyl moiety bacterial broth, the previously reported phoslactomycins F
bound to C18. By interpretation of the 2D COSY NMR (51) [132,133], I-h (56), I-i (57), and I-j (58) [135] were also
spectra, the different acyl subunits for phoslactomycins A–F identified [136,137].
were assigned. Phoslactomycin B was shown to be identical
in structure to phospholine [131]. A single congener of the leustroducsins was obtained
from treating a mixture of leustroducsins A, B, and C and
The phoslactomycins A–F showed strong activity against phoslactomycin F with porcine liver esterase (leustroducsin
various fungi, especially phytopathogenic fungi such as H, 59), Fig. (9). The absolute configuration of the common
Botrytis cinerea and Alternaria kikuchiana. The antibacterial core structure 59 was elucidated by Shibata, et al. utilizing
activity, however, was weak. Comparison of the the Mosher ester method. Both protected (+)-(R)- and (–)-(S)-
antimicrobial activity between the six phoslactomycin MTPA esters of the 11-hydroxy group (60) were prepared,
compounds showed almost an identical profile. Since the Fig. (10). The proton signals of each derivative were
difference in the chemical structures was ascribed to a assigned by COSY, and the systematic arrangement of
substituent bound to a cyclohexane ring, the substituent is positive and negative ∆δ's revealed the R configuration at
considered not to be important for antimicrobial activity C11. An analogous assignment of the configuration at C18
[132,133]. utilizing the (+)-(R)- and (–)-(S)-MTPA esters of the 18-
Fostriecin: Chemistry and Biology Current Medicinal Chemistry, 2002, Vol. 9, No. 22 2021

OMTPA
determined [138]. The stereochemical assignments also
O H correlated with the absolute configuration determined for
O fostriecin at C5, C8, C9, and C11 [5].
OHC
N 11

OH The leustroducsins A–C were isolated while screening for


O compounds which induce the production of colony-
stimulating factors (CSF) in KM-102 cells [136]. It was
Br O demonstrated that the growth of CSF-dependent cell lines
60
were significantly stimulated by media of KM-102 treated
O with leustroducsin A, B, or C, whereas the activity of the
18 OM TPA
phoslactomycins F, I-h, I-i, and I-j was slightly lower. The
O H OBz E D 50 of these compounds correlated positively with
16
O hydrophobicity, hence leustroducsins B and C were the most
OHC
N potent inducers. The cytotoxicity of leustroducsin B against
H various cell lines was approximately two-fold more potent
than the CSF-inducing ability, suggesting that these
compounds may have limited potential as therapeutic agents.
O Like the phoslactomycins, the leustroducsins showed no
61 antibacterial activity but did exhibit significant antifungal
O activity.
MTPAO OBz
H
HO
OHC 9 Sultriecin
N 8
H OBz Sultriecin (64), Fig. (11), a sulfate monoester containing
an α,β-unsaturated δ-lactone and a conjugated triene system,
O was isolated from the culture broth of Streptomyces
roseeiscleroticus by Ohkuma, et al. as the sodium salt. It
O 62 was obtained by butanol extraction and reverse phase
chromatography. The molecular formula and molecular
H weight was determined to be C23H33O8SNa (MW 492) from
N O FABHRMS and FABMS data. Both the sultriecin
OHC
monosodium salt, as well as its desulfated derivative,
showed positive responses to iodine and sulfuric acid, while
demonstrating negative results for ninhydrin and anthrone
MTPAO 5 tests. The UV absorbance maxima (260, 269, 279 nm)
indicated the presence of a triene while the IR spectrum
MeOOC
63
(1720 and 1260 cm–1) was indicative of an α,β-unsaturated
δ-lactone and a sulfate, respectively. The structure clearly
Fig. (10).
resembled that of fostriecin, but differed in the MS, NMR,
hydroxy group (61) established the S configuration. The and in the absence of a phosphate group as shown by
relative stereochemistry on the cyclohexane ring, and hence negative Hanes reaction. The connectivity of 64 was
the absolute configuration of C16 (R), was determined by determined based on interpretation of 1H NMR, 13C NMR,
1 H– 1 H COSY, and 13 C– 1 H COSY spectra. Sultriecin
examination of the coupling constants of 61. The remaining
relative stereochemistry of the (+)-MTPA of leustroducsin H demonstrated moderate and broad spectrum in vitro
derivative 61 was determined from the NOESY spectrum. It antifungal activity. In addition, 64 and the desulfated
was deduced that both the C7 and H9 of the acetonide ring derivate 65 exhibited weak in vitro cytotoxicity against
and the H4 and H5 of the lactone were in a cis relationship. various cancer cell lines. However, both analogs inhibited
RNA and protein, but not DNA, synthesis in cultured
In order to assign the absolute configuration of the C9
OH
stereocenter, the protected (+)-(R)- and (–)-(S)-MTPA esters OR OH
of the 9-hydroxy group (62) were prepared which led to the
assignment of the R configuration. According to the relative O O
stereochemistry mentioned above, the absolute configuration
Me 64, R = SO3 H
of C8 was assigned as R. Finally, the absolute configuration
of C5 was assigned as S based on proton assignments of 65, R = H
both the (+)-(R)- and (–)-(S)-MTPA esters of the 5-hydroxy Me
group from the lactone derivative 6 3 . The absolute H2O3P OH
configuration of leustroducsin H was hence determined to be
4S,5S,8R,9R,11R,16R,18S. Since all the leustroducsins and O O
phoslactomycin F possess the same core structure, the Me Me
absolute configurations of the four isolates could be 66
Fig. (11).
2022 Current Medicinal Chemistry, 2002, Vol. 9, No. 22 Boger et al.

L1210 murine leukemia cells. The natural product 6 4 selective protein phosphatase inhibition is a novel, clinically
demonstrated in vivo antitumor activity comparable to unexplored mechanism worthy of pursuit for introduction of
mitomycin C in mice innoculated with P388 and B16 a new class of antitumor agents. Unfortunately, phase I
melanoma, but 65 was devoid of activity against P388 clinical trials on fostriecin failed to reach therapeutic dose
leukemia. The antitumor activity, as well as the levels when the trials were suspended due to concerns over
macromolecular biosynthesis inhibition, is strikingly similar its inherent storage stability and an unpredictable purity in
to the activity profile of fostriecin [139]. the clinical supplies of the natural product. These issues,
along with the opportunity to probe the mechanism of
action, may now be addressed with the development of a
Cytostatin total synthesis of the natural product that is amenable to the
preparation of key substructure analogs that may be used to
Cytostatin (66), Fig. (11), is a phosphate monoester define its pharmacophore.
containing an α,β-unsaturated δ-lactone and a conjugated
triene system. The compound was discovered in the cultured
broth of Streptomyces sp. MJ654-NF4 by Amemiya, et al. ADDENDUM
as the sodium salt. It was obtained by ethyl acetate
extraction, size-exclusion chromatography, and reverse phase Several important papers have appeared while this review
chromatography. The molecular formula and molecular was at the publishers. Chavez and Jacobsen also recently
weight was determined to be C21H33O7PNa (MW 427) from completed a total synthesis of fostriecin [146]. Their
high resolution FABMS. Cytostatin showed positive approach integrates highly effective asymmetric catalytic
responses to ferric chloride and anisaldehyde-H2SO4 reagents reactions to generate key chiral building blocks, and efficient
on TLC. In addition, a positive color reaction to ammonium coupling reactions to enable their convergent assembly. The
molybdate–perchloric acid indicated the presence of a synthetic plan, shown in Fig. (12), involves the assembly of
phosphate ester. The UV absorbance maxima (260, 267, 277 four fragments of similar complexity. Epoxyketone 68 plays
nm) indicated the presence of a triene while the IR spectrum a central role in the synthetic strategy serving not only as the
(1715 cm–1) was indicative of an α,β-unsaturated ester or source of the C9 stereocenter but also as a lynchpin for
lactone. The connectivity of 66 was determined based on joining the left-hand vinyl lactol 67 and the right hand
interpretation of 1 H NMR, 13 C NMR, 31 P NMR, DEPT, triene-diol fragment. Key features of the synthesis include
1 H– 1 H COSY, HOHAHA, LSPD, and HMBC spectra.
the use of the hydrolytic kinetic resolution (HKR) to provide
Cytostatin was intially identified as a novel inhibitor of B16 enantiomerically pure epoxyketone 68, an asymmetric
melanoma cell adhesion to laminin and collagen type IV in catalytic hetero Diels–Alder reaction to afford
the extracellular matrix. An in vitro cytotoxicity assay enantiomerically pure lactol 67, a highly diastereoselective
showed that cytostatin is 5- to 10-fold less toxic to the B16 1,2-addition used to introduce the C8 stereocenter, and a
cell line than to other tumor cells [140,141]. Yamazaki, et catalytic asymmetric transfer hydrogenation was used to set
al. demonstrated that the natural product increased the rate of the C11 stereocenter. A late stage Stille coupling followed
DNA fragmentation in FS3 fibrosarcoma cell, thereby by phosphorylation of the C9 alcohol provided the Z,Z,E-
causing apoptosis [142]. Another study reported that triene system and the C9 phosphate, respectively.
cytostatin was cytostatic to both B16 and EL-4 cell lines
with an IC50 of 1–2 µg/µL and cytotoxic at a concentration Highly enantioenriched epoxyketone 68 was prepared by
of 10 µg/µL [143]. In in vivo studies conducted by Ishizuka epoxidation of methyl vinyl ketone and subsequent
and coworkers, cytostatin inhibited lung metastasis of B16- hydrolytic kinetic resolution [147] of this epoxide using
F10 and B16-BL6 melanoma cells by 76.5 and 72.9%, (salen)CoIIIOAc 73 under an oxygen atmosphere (>99% ee,
respectively, at an i.p. dose of 1.25 mg/kg in BDF1 mice 80%). Chiral building block 67 was accessed through a Cr-
[140,144]. It has recently been demonstrated that cytostatin catalyzed asymmetric hetero Diels–Alder reaction recently
inhibits cell adhesion through the selective noncompetitive developed in the Jacobsen group [148]. 1-Benzyloxy-1,3-
inhibition of serine/threonine phosphatase PP2A (IC50 = butadiene underwent reaction with triisopropylpropar-
0.09 µ g/mL) [145] which, considering its structural galdehyde in the presence of the tridentate Cr Schiff base
similarities to fostriecin, is not unexpected. catalyst 74 (3 mol%) to provide the cycloadduct 75 in good
yield, diastereoselectivity and enantioselectivity.
Desilylation and acidic workup gave the more stable trans-
CONCLUSIONS acetal, which was recrystallized to provide the highly
enantioenriched alkyne 76 (99% ee).
Fostriecin is a naturally-occurring potent and efficacious
antitumor agent that most likely derives its activity through The acetal 76 was converted to the more robust isopropyl
the selective inhibition of protein phosphatases 2A and/or 4 derivative 7 7 , and was subjected to a one-pot
(PP2A and/or PP4), resulting in inhibition of the mitotic hydrozirconation-Zn transmetallation sequence developed by
entry checkpoint. Contributing to its activity is an active Wipf [149] followed by coupling with epoxyketone 68, and
transport into tumor cells by the reduced folate carrier. resulted in the isolation of the desired 1,2-addition product
Fostriecin also inhibits topoisomerase II, albeit much more 78 in >30:1 diastereoselectivity favoring the desired
weakly, by a mechanism which is not yet clear, and whether chelation control product in modest yield (45%). The epoxy
this also contributes to its antitumor properties remains to alcohol was protected as the triethylsilyl ether 79 (TESCl,
be established. Inhibition of the mitotic entry checkpoint via Im., DMF, 99%) and was subsequently coupled with
Fostriecin: Chemistry and Biology Current Medicinal Chemistry, 2002, Vol. 9, No. 22 2023

Me Chiral Epoxide
Asym. H DA 17
14 18 OH
H H NaH O3 PO OH
N N 10 13
Adama ntyl 8 Stille Coupling
Co O O
t
Bu O t N O OH Asym. Hydrogenation
O Bu 1,2-Add' n
Cr Chelation Control 1
Epoxide Ring Opening
t
Bu t
Bu Cl
O
73 74 OR

Bu3 Sn 70
RO O M
67 O
(S,S)-73 (2 mol%) O OH
O S S
HOAc (4 mol%) + OH
OR O
O 68 H
O 2 Balloon O O
O 69
5-25 o C, 48 h
(R)-(+)-68 H TMS
40% yield 71 72
TIPS
>99% ee

TIPS 1) TBAF, T HF Cp 2Z r(H)Cl;


(R ,S)-74 (3 mol%) 2) TsOH M e2Z n;
H
+
36 h, RT BnO O 3) recryst., 65%, RO O 68
O 45%, >30:1 dr
75 TIPS 99%ee
OBn H
72
90% yield R = Bn (76)
71 90%
89% ee, >95:5 dr R = i Pr (77)

n-BuLi, 69, HO S S
T HF OH S S
O PPTS
i 89% A cetone/H2 O X O
PrO O i
PrO O
Me OR Me OT ES Me O TES T MS
TM S
80
E t3SiCl R = H (78)
MnO2 X = H, OH (81)
99% R = T ES (79)
X = O (82)
65%, 2 ste ps
Ts
Ph
N
Ru
Ph N
H
85
PM BO OR
PhI(O2 CCF3 )2 RO O A gNO3 , NIS

93%, >95:5 dr O O 95%


70% O O Me OT ES
Me OT ES TM S
T MS
T BSOT f R = H (86)
PM BO (C=NH)CCl3 R = H (83)
Trityl-BF4 , 83% 80% R = TBS (87)
R = PMB (84)

RO OT BS HO OTBS
"H N=NH" 70, PdCl2
O O O O
Me OT ES I Me OTES
85% 85%
90 I
D DQ R = PMB (88)
100% R = OH (89)

OH
OR NaHO 3PO OH
R' O OTBS HF;
HF, Pyr, 45% O O
O O
M e OH
Me OTE S

PCl3 ; PM BOH; R' = H (91) 1, fostriecin


t
BuOOH, 65% R' =PO (OPMB)2 (92)
Fig. (12). Jacobsen total synthesis of fostriecin.
2024 Current Medicinal Chemistry, 2002, Vol. 9, No. 22 Boger et al.

(+)-Ipc2 BOM e
4 M etathesis MgBr
1 Asym. Allylation 17 OR
OH –100 ˚C; OsO4, NMO;
H2 O3 PO OH 14 18 OH C TMS
H2 O2 , NaOH NaIO4, SiO2
8 10 13
O O 5 Suzuki Coupling 93
Me OH 2 Sharpless AD T BDPSCl, Im. R = H (94)
3 Asym. Allylation 48%, 2 steps R = TBDPS (95)
1
O
PPh 3
RO OR X OT BDPS EtO OT BDPS
OHC + Me EtO2 C AD-mix β;
BR2 OR
O
Me OR 56%, 2 steps , PPTS
C5–C13 C14–C18 96 97
Me OMe
73%, 2 steps, 3:1 dr
OR
O O O O
RO2 C Me Me
OTBDPS OTBDPS
E tO2C
OR NBS, AgNO3 ; EtO2 C Br LiAlH4
98 HN NH 99
C7–C13 C14–C18 58%, 3 ste ps

O
O (+)-Ipc2BOMe
O O O O MgBr O O Cl
Me PPh3 Me
OTBDPS H OT BDPS Me OT BDPS i
–100 ˚C; Pr2 NEt
R Br Br Br
8 8% H2 O2, NaOH 71%, 2 steps
OHC 10 2 1 03
98% de
Swern R = CH2 OH (100) OH
86% R = CHO (101)

T BDPSO Br
O O OH OTBDPS
Me OTBDPS Montmorillo nit e K-10
Br O O O O O
Grubb's ca t.
Me 65%, 2 steps Me OH
10 4 O Br
O 10 5 106 Pd(PPh 3) 4
Ag2 O
O O
74%
BH
O
TBD PSCI, Im O B
OH OT BDPS O OTBDPS
[Rh(cod)Cl] 2
10 7 108 109
PCy3 , Et3 N
63%, 2 ste ps
O
TMSEO P
T MSEO OP
OP O OP
OH OP PCl3 , pyr;
T MSCH2 CH2 OH; HF•pyr
O O 1, fostriecin
O O H2 O2 CH3 CN/H2O
Me OTBS
Me OR 83% 52%
P = TBDPS
TBSOT f R = H (110)
2,6-lutidine 112
86% R = TBS (111)

Fig. (13). Falck total synthesis of fostriecin.

fragment 69 through epoxide ring opening (69, n-BuLi, butyldimethylsilyl ether 87 (TBSOTf, CH2Cl 2, 80%) and
THF, 89%). Acetal hydrolysis (PPTS, acetone/H2 O) was converted to iodoacetylene 88 (AgNO3, NIS, 95%). The C9
followed by selective lactol oxidation (MnO2, CH2Cl2) to alcohol was deprotected at this stage under oxidative
afford lactone 82 (65%, 2 steps). Dethioketalization using conditions (DDQ, quantitative yield) and the iodoacetylene
Stork’s protocol [150] (bis(trifluoroacetoxy)iodosobenzene, reduced to the vinyl iodide 90 using a very selective diimide
70%) was followed by protection of the ß-hydroxy ketone as reduction [152] (o-nitrobenzenesulfonylhydrazide, THF/i-
the p-methoxybenzyl ether 84 (PMBO(C=NH)CCl3, trityl PrOH, 85%).
BF 4 , 83%). The C11 stereocenter was installed using
Noyori’s asymmetric catalytic transfer hydrogenation catalyst Stille coupling of the vinyl iodide with stannane
85 [151] and provided the propargyl alcohol 86 in excellent fragment 70 [153] installed the Z,Z,E-triene system with
yield and diastereoselectivity (92% yield, >95:5 dr). The complete retention of olefin stereochemistry in the presence
resulting propargyl alcohol was protected as the tert- of the free C9 alcohol under ligand free conditions
Fostriecin: Chemistry and Biology Current Medicinal Chemistry, 2002, Vol. 9, No. 22 2025

(PdCl2(CH3CN)2, DMF, 85%). Phosphorylation of 91 was detailed by Boger [9] was found necessary to prevent cyclic
completed by treatment of the free alcohol with PCl3 in phosphate formation during the phosphorylation procedure
pyridine followed by addition of PMBOH and subsequent [9]. As such, 111 was subjected to Evans’ three-stage
oxidation using t-BuOOH [154] to provide 92 (65%). phosphorylation protocol [154] (PCl 3 , pyr;
Global deprotection (HF, CH3 CN; then Pyr) provided TMSCH2CH2OH; H2O2) to provide protected fostriecin 112
fostriecin (45%) identical in all aspects to authentic material. in 83% yield. Global desilylation was achieved by treatment
with excess HF·pyr at room temperature followed by
Soon after the reported synthesis by Chavez and NaHCO3 quench to afford natural (+)-fostriecin (52%).
Jacobsen, Reddy and Falck also disclosed a total synthesis
of fostriecin [155]. The strategy, shown in Fig. (13), is The most recent total synthesis of fostriecin was
based on late stage introduction of the sensitive unsaturated completed by Imanishi and coworkers [167]. The key bond
lactone, phosphate and triene functional groups. The construction steps, shown in Fig. (14), include a Horner–
unsaturated lactone was formed by ring-closing metathesis Wadsworth– Emmons reaction of fragment C1–C6 with C7–
and the Z,Z,E-triene was established by Suzuki coupling of C13 to create the C6–C7 double bond, acid-catalyzed
the corresponding Z-vinyl bromide and a Z,E-dienylboronic cyclization of the C5 alcohol to form the δ-lactone and Stille
acid derivative. The four stereocenters found in fostriecin coupling to form the Z,Z,E-triene. The use of (R)-malic acid
were constructed by two asymmetric allylations (C5, C11) as starting material established the absolute stereochemistry
and a Sharpless asymmetric dihydroxylation of a C8–C9 of C11 and a Sharpless asymmetric dihydroxylation
trisubstituted double bond. established the remaining two chiral centers of fragment C7–
C13. The stereochemistry at C5 was set by asymmetric
The synthesis commenced with a low temperature reduction of the corresponding ketone with (R)-BinalH.
asymmetric allylation [156] of 93 [157] with (+)-B -
methoxydiisopinocampheylborane and allylmagnesium Alcohol 113 was synthesized from (R)-malic acid
bromide to provide alcohol 94 [158] in 98% ee. TBDPS according to literature procedure [168]. Subsequent Swern
protection of the alcohol (48%, 2 steps) and dihydroxylation oxidation and Wittig reaction provided unsaturated ester 114
followed by oxidative cleavage of the resultant diol (OsO4, (86%). Reduction to the allylic alcohol and benzyl
NMO; NaIO4, SiO2) afforded aldehyde 96. Homologation to protection yielded the tri-substituted olefin 115 for Sharpless
the α,β-unsaturated ester (E)-97 was achieved by Wittig asymmetric dihydroxylation. The C8 and C9 chiral centers
reaction with (carboethoxyethylidene)triphenylphosphorane were formed with high diastereoselectivity (95:5 anti:syn,
(56%, 2 steps). Next, the C8 and C9 stereocenters were 92%) with the (DHQD)2PHAL ligand, thereby establishing
established by Sharpless asymmetric dihydroxylation [159] three of the four chiral centers of fostriecin. Acetonide
(AD-mix β) with 3:1 diastereoselectivity. Diol protection as formation followed by selective removal of the terminal
the corresponding 1,2-isopropylidene allowed acetonide with zinc nitrate [169] afforded 117 (81%, 2
chromatographic separation of the diastereomers (73%, 2 steps). Full orthogonal protection of fragment C7–C13 was
steps). Bromination of alkyne 98 [160] followed by diimide completed by selective 4-methoxybenzyl protection of the
cis-hydrogenation [161] provided vinyl bromide 99. primary alcohol (Bu2SnO, PMBCl, Bu4NI; 84%) and TBS
Reduction of the ester in 99 (LiAlH 4 , 58%, 3 steps) protection of the remaining secondary alcohol (TBSOTf,
followed by Swern oxidation afforded aldehyde 101 in good 2,6-lutidine, 100%).
yield (86%). Subsequent Wittig homologation
(OHCCHPPh3, 88%) established the C6–C7 double bond of Fragment C7–C13 was coupled with fragment C1–C6
fostriecin. before performing the Stille coupling to avoid complications
with the acid-labile conjugated triene in the course of the
Construction of the unsaturated lactone was initiated by synthesis. As such, the benzoyl group of 119 was removed
an asymmetric allylation of α,β-unsaturated aldehyde 102 by basic hydrolysis (3 M KOH–MeOH–THF (1:2:2), 87%)
((+)-Ipc2BOMe, allylmagnesium bromide; H2O2, 98% de). and the resultant alcohol 120 was oxidized to the aldehyde
Acylation of the newly-formed alcohol (acryloyl chloride, i- (TPAP, NMO). Subsequent Horner–Wadsworth–Emmons
Pr2NEt, 71%, 2 steps) provided the ring-closing metathesis coupling [170] (DBU, LiCl) provided vinyl ketone 121 with
substrate 104 in good yield and treatment with Grubb’s complete stereocontrol (79%, 2 steps). The last stereocenter
catalyst [162] ([1,3-bis-(2,4,6-trimethylphenyl)-4,5- of fostriecin was introduced by (R)-BinalH reduction [171]
dihydroimidazol-2-ylidene] [benyzlidene] ruthenium (IV) of the C5 ketone, affording 1 2 2 with high
dichloride) formed the α,β-unsaturated lactone of fostriecin. diastereoselectivity (>20:1, 85%). Acid treatment of 122
Subsequent acetonide deprotection of 105 (Montmorillonite effected a facile lactonization (92%), and a two-step
K-10 [163], 65%, 2 steps) provided 106, set for introduction procedure through selenide 124 (NaHMDS, TMSCl,
of the triene system. PhSeBr, 94%; H 2 O 2 , NaHCO 3 , 94%) introduced the
unsaturation in the δ-lactone.
The C14–C18 subunit was prepared from 107 [164] via
silyl protection (TBDPSCl, Im.) and [Rh(cod)Cl]2-mediated Oxidative removal of the PMB protecting group (96%)
trans-addition [165] of pinacolborane to the terminal alkyne and a buffered Dess-Martin oxidation of the resulting
of 108 (63%, 2 steps). Suzuki–Miyaura coupling [166] of primary alcohol set the stage for introduction of the Z,Z,E-
subunits 106 and 109 (Pd(PPh3 ) 4 , Ag2 O) afforded the triene system. Wittig reaction (Ph3 PCH 2 I 2 , NaHMDS,
fostriecin carbon backbone with complete olefin HMPA) of 126 afforded iodo-olefin 127 in modest yield
stereoretention in 74% yield. Selective protection of the C8 (61%, 2 steps) and selectivity (4:1 Z:E). It was determined
alcohol (TBSOTf, 2,6-lutidine, 86%) by the method first that acetonide removal was essential before introduction of
2026 Current Medicinal Chemistry, 2002, Vol. 9, No. 22 Boger et al.

4 Asym. Hyd rogenation 1 from R-M alic ac i


d 17 OH
OH O
H2 O3 PO OH 14 HO2 C BH3•Me2 S, 87%
O Swern;
18 CO2 H
8 10 13 HO
O O 5 Stille Coupling Me2 C(OMe) 2, 95%
PPh3
Me OH 2 Sharpless AD 113
3 Horner– CO2 Me
1 86%
Wadsworth–
E mmons
(DHQD)2 PHAL
OH OH I OR K2OsO2 (OH) 4
+ O K3Fe(CN) 6 OBz OH O Me2 C(OMe)2
O O O O pTsOH;
Bu3 Sn R
Me OH Zn(NO3 )2 •6H2 O
C1–C13 C14–C18 K2CO3 Me OH 81%, 2 steps
Me
MeSO2N H2 116
t
OH OH OH BuOH–H2 O

+ OHC HO2 C 92%, 95:5 anti:syn


EtO2 C CHO CO2 H
O
Me OH LiAlH4 ; R = CO2 Me (114)
(MeO)2 OP R-malic acid BzCl, Et3 N
R = CH2 OBz (115)
C1–C6 C7–C13 84%, 2 steps

Bu2 SnO ; TBSOTf


O O PMBCl, Bu4 NI O O 2,6-lutidine O O TPAP, NM O;
Me OH Me Me
OH O TBS
BzO 100 % DBU, L iCl
84% BzO RO
OH O O
117 118 OPM B OPMB
(Me O)2 P
3 M KOH R = Bz (119) (CH2 )3 CO2 Et
87% R = H (120) 79%, 2 steps

OTBS
O O O O OPMB
Me OTBS R-BinalH Me pTsOH 94%
OTBS
85%, >20:1 92% O O O NaHMD S
E tO 2C 3 EtO2 C Me O
OPMB 3 OPMB TM SCl
O 121 122 123
OH PhSeBr

OTBS OTBS OTBS


PhSe OPMB OPMB O
H2O2 Ph 3P CHI
NaHCO3 DDQ, 96% HM PA
O O O O O O O O O
Me O 94% Me O Dess–Martin Me O 61%, 2 steps
124 125 126 4:1 Z:E

T BSO I
TBSOTf,
OH OH I 2,6-lutidin e; OR 1 OR 2 I
1M H C l–M eOH TE SOTf PdCl2 (Me CN) 2
O O O O O O O
89% separation OTBDPS
Me O Me OH Me O TES
127 128 56% Z-i some r
SnBu3 80%
1M HCl R1 = TES, R2 = TBS (129)
70% R1 = H, R 2 = T BS (130)
O
PMBO P
O TBDPS OT BDPS
OH OTBS PCl3 , pyr PMBO O OTBS
PMBOH; HF–CH3 CN;
1, fostriecin
O O t O O
BuOOH, 56% Pyr, 38%
Me OTES Me O TES
131 132

Fig. (14). Imanishi total synthesis of fostriecin.

the acid-labile triene system. As such, treatment of 127 with separation of the Z-isomer (56%). Selective deprotection of
1 M HCl–MeOH provided simultaneous removal of the the C9 TES group (1 M HCl–THF–MeCN (1:3:6), –10ºC)
acetonide and TBS protecting groups. Phosphorylation of affords 130, intersecting the late stage intermediate first
the C9 alcohol required selective protection of the tertiary disclosed in the Jacobsen synthesis.
and allylic secondary alcohols. All alcohols were protected
by sequential treatment with TBSOTf (2,6-lutidine, –78 ºC) Several groups have also recently published synthetic
and TESOTf (–78 to 25 ºC) in one pot, followed by work toward fostriecin. Esumi and coworkers [172]
Fostriecin: Chemistry and Biology Current Medicinal Chemistry, 2002, Vol. 9, No. 22 2027

completed the synthesis of 8,9-acetonide dephospho- Waldmann have published the synthesis of the
fostriecin and Cossy and coworkers have detailed the (4S,5S,6S,9S,10S,11S)-isomer of cytostatin [175]. The
synthesis of a protected C1-C12 subunit of fostriecin [173]. synthesis was designed to have maximal flexibility since the
Kobayashi et al. [174] have also disclosed a study on the relative and absolute configuration of cytostatin is unknown.
dihydroxylation of a series of dienes as synthetic Installation of the chiral centers with asymmetric Evans
intermediates for fostriecin. alkylation (C6), asymmetric Evans aldol addition [176] (C4,
While this review was at the publishers, Bialy and C5, C9, and C10), and enantioselective reduction of an

O O OH M OMCl, i Pr 2N Et
OH
2 Evans' Aldol Swern; 97%
Still–Gennari 3 Evans' Aldol
O N
5 4 134, Bu2 BOTf, LiBH3OH, 78%
H2 O3 PO OH 17 Me i Me Me
1 Me Pr2 NEt; H2 O2
6 Ph Me
O O 9 11 6 Stille Coupling 133 68%, 2 steps 135
H
Asym. Me Me
Alkylation 1 4 Asym. Hydrogenation
OR OMOM 9-BBN; T BSO O MOM OH
(4S,5S,9S,10S,11S)-isomer of 66
H2 O2, NaOH Swern;

83% 134, Bu2 BOTf,


Me Me Me Me i
Pr 2NE t; H2 O2
Me
OH I 138
H2O3 PO TBSCl R = H (136)
Me
+ Im., 97%
O O Bu 3Sn C14–C18 R = T BS (137)
H
Me Me
OR OMOM OH O O Me3 Al
C1–C13
(H2NM eOMe)+Cl-
N O
O O 84%
O OR OR OR O Me Me Me
Me Ph
+ O N
TBSCl R = H (139) H Ph
Me Me Me
Ph Me Et3N, DMAP
R = TBS (140) Ph
C3–C13 1 34 86%, 3 steps
N O
B
(R )-145
n Me
TBSO OMOM OR O BuLi T BSO OMOM OMOM O TBSO OMOM OMOM O
SiMe3 borax BH3 •Me2 S
NMe 96%
95%, 2 steps
Me Me Me OM e Me Me Me TM S Me Me Me

MOM Cl R = H (141) 143 144


O
i
Pr2 NEt R = MOM (142)
92% (CF3CH2 O)2 P CO2M e
TBSO OMOM OMOM OH HO OMO M OM OM OTBDPS O OMOM OM OM OTBDPS 18-c-6,
TBDPSCl, Im.; Dess–Ma rtin KHMDS

HF•pyr, THF 93% 92%


Me Me Me RT, 40 min Me Me Me Me Me Me
146 88%, 2 steps 147 148
OFm
FmO
MO MO MOMO OT BDPS Me P O
Me i
CBr4 OH OTBDPS (FmO) 2PN Pr2 O OTBDPS
HF•pyr
tetrazole;
O O I 2, pyr, H2O O O 82%
Me Me Me OH , 82 ˚C H 95% H
Me O O Me Me Me Me
149 83% 150 151

OFm OFm Me
FmO FmO
Me P O Me P O Bu3 Sn 154
NIS, AgNO3
O OH O OH I
100% PdCl2 (CH3CN)2
O O HN NH O O
H H 62%
Me Me 63% Me Me
152 153

OFm ONa
FmO HO
P O Me P O
Me Et3N , CH3CN (1:5)
O OH Me O OH Me
+
Na Dowex resin
O O O O
H 85% H
Me Me Me Me
155 (4S,5S,9S,10S,11S)-isomer of 66

Fig. (15). Waldmann synthesis of a cytotastin isomer.


2028 Current Medicinal Chemistry, 2002, Vol. 9, No. 22 Boger et al.

acetylenic ketone [177] (C11) allows access to every possible opening. After much experimentation, it was found that
isomer. The C2–C3 (Z)-double bond was formed by a Still- conversion to the alkynyl iodide with AgNO3 followed by
Gennari coupling and the C13–C14 bond was established by reaction with in situ generated diimide [161] ((KO2CN) 2,
Stille coupling. HOAc, 2-propanol/dioxane) afforded vinyl iodide (Z)-153
(63%). The sensitive Z,Z,E-triene was then formed by Stille
Chiral alcohol 133 was prepared by Evans asymmetric coupling with Z,E-dienyl stannane 154 without additional
alkylation as previously described [178] (134, NaHMDS, phosphine [182] ([PdCl2 (CH 3 CN) 2 ], DMF/THF, 62%).
allyl iodide; LiAlH4) and subjected to Swern oxidation. The Treatment of the phosphate triester 155 with excess
corresponding aldehyde was used in an asymmetric aldol triethylamine induced β-elimination of the fluorenylmethyl
condensation with the dibutylboron enolate of phosphate ester [183] and the (4S ,5S ,9S ,10S ,11S ) -
acyloxazolidinone 134 [179] to provide 135 in good yield diastereomer of cytostatin was obtained in 85% yield. It was
(68%, 2 steps) and with high diastereoselectivity. Trace not reported whether this isomer correlates with the natural
amounts of the undesired diastereomers were readily product. However, this isomer of 66 was established to be
separated by recrystallization. Protection of the C5 alcohol an effective inhibitor of PP2A (IC50 = 33 nM) comparable
as the MOM ether and reductive removal of the chiral or more potent than the natural product itself (IC50 = 200
auxiliary with in situ generated LiBH3 OH [180] (LiBH4 , nM).
H2O) afforded alcohol 136 in 78% yield. TBS protection of
the primary alcohol (TBSCl, Im., 97%), hydroboration of Since this review was originally submitted, a number of
the terminal double bond (9-BBN; H2O2, NaOH; 83%) and studies continue to strengthen the expectation that selective
Swern oxidation provided the aldehyde for the second inhibition of PP2A among the phosphatases is the target to
asymmetric syn aldol condensation. As such, the C9 and address for the treatment of cancer. PP2A is the protein
C10 stereocenters were installed with high phosphatase that fostriecin is uniquely selective against.
diastereoselectivity under the same conditions used for the PP2A has one of the most highly expressed catalytic
C5 and C6 chiral centers (134, Bu2BOTf, i-Pr2NEt; H2O2). subunits and it combines to form a large repertoire of
The primary TBS group was cleaved under the aldol reaction heteromultimeric holoenzymes composed of the catalytic
conditions but was readily resilylated (TBSCl, Et3 N , subunit (PP2Ac), a structural A subunit, and at least three
DMAP; 86%, 3 steps). Conversion of acyloxazolidinone distinct B subunits encoded by at least 13 genes and
1 4 0 to fully protected Weinreb amide 1 4 2 (Me3 A l , multiple splice variants. Recent studies make it clear that a
(H 2 N M e O M e ) + Cl – , 84%; MOMCl, i - Pr 2 NEt, 92%) major function of these regulatory subunits is to target the
provided the starting material for preparation of the C11 PP2A holoenzyme to distinct intracellular locations,
acetylenic ketone. The desired ketone was prepared by signaling complexes, and substrates, including those
treatment of Weinreb amide 142 with lithium involved in regulation of the cell cycle [184, 185]. The
trimethylsilylacetylide followed by borax-mediated TMS breadth of PP2A function reported since this review was
deprotection (95%, 2 steps). After screening several reagents submitted extends its role from regulating the initiation of
for the asymmetric reduction of the C11 ketone, the best DNA replication to regulating cell death by apoptosis and it
conversion was obtained with two equivalents of (R)-145 is this latter role that may be most closely linked to a
[Parker, Corey 177]. The last stereocenter was thus potential treatment of cancer. Several recent studies implicate
established as a single diastereomer in 96% yield. PP2A as the phosphatase that associates with several
proteins involved in apoptosis including Bcl-2, caspase-3,
The next stage of the synthesis required formation of the and the adenovirus E4orf4 protein [186]. In particular, PP2A
unsaturated lactone. The propargylic alcohol was protected as is emerging as a target with powerful antiapoptotic activity
the TBDPS ether, a blocking group which would survive the that is mediated by a regulating subunit (R5/B56). The
Still-Gennari olefination and acid-mediated cyclization indirect evidence suggests a role for PP2A in promoting cell
conditions as well as allow selective deprotection of the survival through inhibition of apoptosis [187, 188].
primary TBS ether. Thus, treatment of the TBDPS
propargylic ether with HF·pyr (THF, 25 ºC, 40 min; 88%, 2
steps) selectively cleaved the primary TBS ether and Dess– ACKNOWLEDGEMENTS
Martin oxidation [22] provided aldehyde 148 in 93% yield.
Condensation with the Still–Gennari phosphonate [13] We gratefully acknowledge the financial support of the
((CF3CH2O)2P(O)CH2CO2Me, 18-c-6, KHMDS) provided National Institute of Health (CA93456), the Skaggs Institute
Z-olefin 149 in 92% yield. Treatment of 149 with CBr4 (2- for Chemical Biology, the award of a NIH postdoctoral
propanol, 82 ºC) caused deprotection of the MOM ethers and fellowship (C.-M.G.), and a Bristol-Myers Squibb
simultaneous cyclization to unsaturated lactone 150, likely fellowship (D.R.S.). D.R.S. is a Skaggs Fellow.
due to in situ formation of HBr [181].

Next, the C9 phosphate was installed by phosphitylation REFERENCES


((FmO)2PNiPr2, tetrazole) and subsequent oxidation (I2, pyr,
H 2 O; 95%). Selective deprotection of the propargylic [1] Stampwala, S.S.; Bunge, R.H.; Hurley, T.R.; Willmer,
TBDPS ether with HF·pyr proceeded smoothly to provide N.E.; Brankiewicz, A.J.; Steinman, C.E.; Smitka, T.A.;
alkyne 152 in 82% yield. Reduction of the triple bond to the French, J.C. J. Antibiot. 1983, 36, 1601.
(Z)-vinyl iodide proved problematic, as a variety of reagents [2] Tunac, J.B.; Graham, B.D.; Dobson, W.E. J. Antibiot.
caused reduction of the α,β-unsaturated lactone and ring 1983, 36, 1595.
Fostriecin: Chemistry and Biology Current Medicinal Chemistry, 2002, Vol. 9, No. 22 2029

[3] Hokanson, G.C.; French, J.C. J. Org. Chem. 1985, 50, [29] Boritzki, T.J.; Wolfard, T.S.; Besserer, J.A.; Jackson,
462. R.C.; Fry, D.W. Biochem. Pharmacol. 1988, 37, 4063.

[4] Baker, S.R.; Jamieson, W.B.; McKay, S.W.; Morgan, S.E.; [30] Osheroff, N.; Zechiedrich, E.L.; Gale, K.C. BioEssays
Rackham, D.M.; Ross, W.J.; Shrubsall, P.R. Tetrahedron 1991, 13, 269.
Lett. 1980, 21, 4123.
[31] Holm, C.; Stearns, T.; Botstein, D. Mol. Cell. Biol. 1989,
[5] Boger, D.L.; Hikota, M.; Lewis, B.M. J. Org. Chem. 1997, 9, 159.
62, 1748.
[32] Pommier, Y.; Kohn, K.W. In Topoisomerase II inhibition
[6] Rychnovsky, S.D.; Skalitzky, D.J. Tetrahedron Lett. by antitumor intercalators and demethylepipodophyl-
1990, 31, 945. lotoxins; Gazer, R. I., Ed.; CRC Press: Boca Raton, FL,
1989, p. 175.
[7] Evans, D.A.; Rieger, D.L.; Gage, J.R. Tetrahedron Lett.
1990, 31, 7099. [33] Liu, L.F. Annu. Rev. Biochem. 1989, 58, 351.

[8] Just, G.; O'Connor, B. Tetrahedron Lett. 1988, 29, 753. [34] Chen, G.L.; Yang, L.; Rowe, T.C.; Halligan, B.D.; Tewey,
K.M.; Liu, L.F. J. Biol. Chem. 1984, 259, 13560.
[9] Boger, D.L.; Ichikawa, S.; Zhong, W. J. Am. Chem. Soc.
2001, 123, 4161. [35] Nelson, E.M.; Tewey, K.M.; Liu, L.F. Proc. Natl. Acad.
Sci. USA 1984, 81, 1361.
[10] Kolb, H.C.; VanNieuwenhze, M.S.; Sharpless, K.B. Chem.
Rev. 1994, 94, 2483. [36] Pommier, Y.; Leteurtre, F.; Fesen, M.R.; Fujimori, A.;
Bertrand, R.; Solary, E.; Kohlhagen, G.; Kohn, K.W.
[11] Herdeis, C. Synthesis 1986, 232. Cancer Invest. 1994, 12, 530.
[12] Lee, Y.S.; Kang, S.S.; Choi, J.H.; Park, H. Tetrahedron [37] Nitiss, J.L.; Beck, W.T. Eur. J. Cancer, Part A 1996, 32A,
1997, 53, 3045. 958.
[13] Still, W.C.; Gennari, C. Tetrahedron Lett. 1983, 24, [38] Cummings, J.; Smyth, J.F. Ann. Oncol. 1993, 4, 533.
4405.
[39] Gorczyca, W.; Gong, J.; Ardelt, B.; Traganos, F.;
[14] Corey, E.J.; Fuchs, P.L. Tetrahedron Lett. 1972, 3769. Darzynkiewicz, Z. Cancer Res. 1993, 53, 3186.
[15] Uenishi, J.; Kawahama, R.; Yonemitsu, O.; Tsuji, J. J. [40] Ishida, R.; Sato, M.; Narita, T.; Utsumi, K.R.; Nishimoto,
Org. Chem. 1998, 63, 8965. T.; Morita, T.; Nagata, H.; Andoh, T. J. Cell Biol. 1994,
126, 1341.
[16] Uenishi, J.; Kawahama, R.; Shiga, Y.; Yonemitsu, O.;
Tsuji, J. Tetrahedron Lett. 1996, 37, 6759. [41] Ishida, R.; Miki, T.; Narita, T.; Yui, R.; Sato, M.; Utsumi,
K.R.; Tanabe, K.; Andoh, T. Cancer Res. 1991, 51, 4909.
[17] Uenishi, J.i.; Kawahama, R.; Yonemitsu, O.; Tsuji, J. J.
Org. Chem. 1996, 61, 5716. [42] Tanabe, K.; Ikegami, Y.; Ishida, R.; Andoh, T. Cancer
Res. 1991, 51, 4903.
[18] Evans, D.A.; Black, W.C. J. Am. Chem. Soc. 1993, 115,
4497. [43] Drake, F.H.; Hofmann, G.A.; Mong, S.-M.; Bartus, J.O.;
Hertzberg, R.P.; Johnson, R.K.; Mattern, M.R.; Mirabelli,
[19] Gibbs, R.A.; Krishnan, U. Tetrahedron Lett. 1994, 35, C.K. Cancer Res. 1989, 49, 2578.
2509.
[44] Lavelle, F.; Nguyen, C.H.; Bissery, M.C.; Riou, J.F.;
[20] Jung, M.E.; Light, L.A. Tetrahedron Lett. 1982, 23, Bisagni, E. Proc. Am. Assoc. Cancer Res. 1991, 32, 405.
3851.
[45] Woynarowski, J.; McCarthy, K.; Reynolds, B.; Beerman,
[21] Bansal, R.; Cooper, G.F.; Corey, E.J. J. Org. Chem. 1991, T. Proc. Am. Assoc. Cancer Res. 1990, 31, 438.
56, 1329.
[46] Bojanowski, K.; Lelievre, S.; Markovits, J.; Couprie, J.;
[22] Dess, D.B.; Martin, J.C. J. Am. Chem. Soc. 1991, 113, Jacquemin-Sablon, A.; Larsen, A.K. Proc. Natl. Acad. Sci.
7277. USA 1992, 89, 3025.
[23] Leonard, J.; Mohialdin, S.; Reed, D.; Ryan, G.; Swain, [47] Permana, P.A.; Snapka, R.M.; Shen, L.L.; Chu, D.T.W.;
P.A. Tetrahedron 1995, 51, 12843. Clement, J.J.; Plattner, J.J. Biochemistry 1994, 33,
[24] Yamamoto, Y.; Maruyama, K. J. Am. Chem. Soc. 1985, 11333.
107, 6411. [48] Jensen, P.B.; Soerensen, B.S.; Sehested, M.; Grue, P.;
[25] Weidmann, B.; Seebach, D. Helv. Chim. Acta 1980, 63, Demant, E.J.F.; Hansen, H.H. Cancer Res. 1994, 54, 2959.
2451. [49] Utsumi, H.; Shibuya, M.L.; Kosaka, T.; Buddenbaum,
[26] Imamoto, T. Pure Appl. Chem. 1990, 62, 747. W.E.; Elkind, M.M. Cancer Res. 1990, 50, 2577.

[27] de Jong, R.S.; de Vries, E.G.E.; Mulder, N.H. Anti-Cancer [50] Fry, D.W.; Besserer, J.A.; Boritzki, T.J. Cancer Res. 1984,
Drugs 1997, 8, 413. 44, 3366.

[28] Fry, D.W.; Boritzki, T.J.; Jackson, R.C. C a n c e r


Chemother. Pharmacol. 1984, 13, 171.
2030 Current Medicinal Chemistry, 2002, Vol. 9, No. 22 Boger et al.

[51] Jackson, R.C.; Fry, D.W.; Boritzki, T.J.; Roberts, B.J.; [75] Yamashita, K.; Yasuda, H.; Pines, J.; Yasumoto, K.;
Hook, K.E.; Leopold, W.R. Adv. Enzyme Regul. 1985, 23, Nishitani, H.; Ohtsubo, M.; Hunter, T.; Sugimura, T.;
193. Nishimoto, T. EMBO J. 1990, 9, 4331.

[52] Gedik, C.M.; Collins, A.R. Nucleic Acids Res. 1990, 18, [76] Tosuji, H.; Mabuchi, I.; Fusetani, N.; Nakazawa, T. Proc.
1007. Natl. Acad. Sci. USA 1992, 89, 10613.

[53] Withoff, S.; de Vries, E.G.E.; Keith, W.N.; Nienhuis, E.F.; [77] Uemura, T.; Ohkura, H.; Adachi, Y.; Morino, K.; Shiozaki,
van der Graaf, W.T.A.; Uges, D.R.A.; Mulder, N.H. Br. J. K.; Yanagida, M. Cell 1987, 50, 917.
Cancer 1996, 74, 1869.
[78] Nurse, P.; Bissett, Y. Nature (London) 1981, 292, 558.
[54] de Jong, S.; Zijlstra, J.G.; Mulder, N.H.; de Vries, E.G.E.
Cancer Chemother. Pharmacol. 1991, 28, 461. [79] Nurse, P. Nature (London) 1990, 344, 503.

[55] Chen, M.; Beck, W.T. Cancer Res. 1993, 53, 5946. [80] Moreno, S.; Nurse, P. Cell 1990, 61, 549.

[56] Anderson, H.; Roberge, M. Cell Growth Differ. 1996, 7, [81] Honkanen, R.E. FEBS Lett. 1993, 330, 283.
83.
[82] Cheng, A.; Balczon, R.; Zuo, Z.; Koons, J.S.; Walsh, A.H.;
[57] Downes, C.S.; Clarke, D.J.; Mullinger, A.M.; Gimenez- Honkanen, R.E. Cancer Res. 1998, 58, 3611.
Abian, J.F.; Creighton, A.M.; Johnson, R.T. Nature
(London) 1994, 372, 467. [83] Sakai, R.; Ikeda, I.; Kitani, H.; Fujiki, H.; Takaku, F.;
Rapp, U.; Sugimura, T.; Nagao, M. Proc. Natl. Acad. Sci.
[58] Hotz, M.A.; Del Bino, G.; Lassota, P.; Traganos, F.; USA 1989, 86, 9946.
Darzynkiewicz, Z. Cancer Res. 1992, 52, 1530.
[84] Guo, X.W.; Th'ng, J.P.H.; Swank, R.A.; Anderson, H.J.;
[59] Fattman, C.L.; Allan, W.P.; Hasinoff, B.B.; Yalowich, J.C. Tudan, C.; Bradbury, E.M.; Roberge, M. EMBO J. 1995,
Biochem. Pharmacol. 1996, 52, 635. 14, 976.

[60] Jensen, P.B.; Sorensen, B.S.; Demant, E.J.F.; Sehested, [85] Ho, D.T.; Roberge, M. Carcinogenesis 1996, 17, 967.
M.; Jensen, P.S.; Vindelow, L.; Hansen, H.H. Cancer Res.
1990, 50, 3311. [86] Hastie, C.J.; Cohen, P.T.W. FEBS Lett. 1998, 431, 357.

[61] Frosina, G.; Rossi, O. Carcinogenesis 1992, 13, 1371. [87] Hartwell, L.H.; Weinert, T.A. Science (Washington, D. C.)
1989, 246, 629.
[62] Roberge, M.; Tudan, C.; Hung, S.M.F.; Harder, K.W.;
Jirik, F.R.; Anderson, H. Cancer Res. 1994, 54, 6115. [88] Lewin, B. Cell 1990, 61, 743.

[63] da Cruz e Silva, O.B.; da Cruz e Silva, E.F.; Cohen, P.T.W. [89] Murray, A.W. Nature (London) 1992, 359, 599.
FEBS Lett. 1988, 242, 106. [90] Steinmann, K.E.; Belinsky, G.S.; Lee, D.; Schlegel, R.
[64] Cohen, P. Annu. Rev. Biochem. 1989, 58, 453. Proc. Natl. Acad. Sci. USA 1991, 88, 6843.

[65] Brewis, N.D.; Street, A.J.; Prescott, A.R.; Cohen, P.T.W. [91] Eriksson, J.E.; Brautigan, D.L.; Vallee, R.; Olmsted, J.;
EMBO J. 1993, 12, 987. Fujiki, H.; Goldman, R.D. Proc. Natl. Acad. Sci. USA
1992, 89, 11093.
[66] Huang, X.; Cheng, A.; Honkanen, R.E. Genomics 1997,
44, 336. [92] Lee, W.-C.; Yu, J.-S.; Yang, S.-D.; Lai, Y.-K. J. Cell.
Biochem. 1992, 49, 378.
[67] Chen, M.X.; McPartlin, A.E.; Brown, L.; Chen, Y.H.;
Barker, H.M.; Cohen, P.T.W. EMBO J. 1994, 13, 4278. [93] Chartier, L.; Rankin, L.L.; Allen, R.E.; Kato, Y.; Fusetani,
N.; Karaki, H.; Watabe, S.; Hartshorne, D.J. Cell Motil.
[68] Bastians, H.; Ponstingl, H. J. Cell Sci. 1996, 109, 2865. Cytoskel. 1991, 18, 26.

[69] Zhuang, Z.; Dean, N.M.; Honkanen, R.E. J. Biol. Chem. [94] Chou, Y.-H.; Rosevear, E.; Goldman, R.D. Proc. Natl.
1998, 273, 12250. Acad. Sci. USA 1989, 86, 1885.

[70] Sheppeck, II, J.E.; Gauss, C.-M.; Chamberlin, A.R. [95] Chou, Y.-H.; Bischoff, J.R.; Beach, D.; Goldman, R.D.
Bioorg. Med. Chem. 1997, 5, 1739. Cell 1990, 62, 1063.

[71] Usui, T.; Marriott, G.; Inagaki, M.; Swarup, G.; Osada, H. [96] Chou, Y.-H.; Ngai, K.-L.; Goldman, R.D. J. Biol. Chem.
J. Biochem. 1999, 125, 960. 1991, 266, 7325.

[72] Helps, N.R.; Brewis, N.D.; Lineruth, K.; Davis, T.; Kaiser, [97] Nishizuka, Y. Nature (London) 1984, 308, 693.
K.; Cohen, P.T.W. J. Cell Sci. 1998, 111, 1331.
[98] Mechanisms of Tumor Promotion, Slaga, T.J., Ed., CRC
[73] Felix, M.-A.; Cohen, P.; Karsenti, E. EMBO J. 1990, 9, Press: Boca Raton, FL, 1984.
675.
[99] Gatti, A.; Robinson, P.J. Eur. J. Biochem. 1997, 249, 92.
[74] Picard, A.; Labbé, J.-C.; Barakat, H.; Cavadore, J.-C.;
Dorée, M. J. Cell. Biol. 1991, 115, 337. [100] Leopold, W.R.; Shillis, J.L.; Mertus, A.E.; Nelson, J.M.;
Roberts, B.J.; Jackson, R.C. Cancer Res. 1984, 44, 1928.
Fostriecin: Chemistry and Biology Current Medicinal Chemistry, 2002, Vol. 9, No. 22 2031

[101] Susick, R.L. Jr.; Hawkins, K.L.; Pegg, D.G. Fundam. Appl. [126] Walsh, A.H.; Cheng, A.; Honkanen, R.E. FEBS Lett. 1997,
Toxicol. 1990, 15, 258. 416, 230.
[102] Snaith, H.A.; Armstrong, C.G.; Guo, Y.; Kaiser, K.; [127] Mamber, S.W.; Okasinski, W.G.; Pinter, C.D.; Tunac, J.B.
Cohen, P.T.W. J. Cell Sci. 1996, 109, 3001. J. Antibiot. 1986, 39, 1467.
[103] Solomon, M.J.; Glotzer, M.; Lee, T.H.; Philippe, M.; [128] de Jong, R.S.; Mulder, N.H.; Uges, D.R.A.; Sleijfer, D.T.;
Kirschner, M.W. Cell 1990, 63, 1013. Hoppener, F.J.P.; Groen, H.J.M.; Willemse, P.H.B.; Van
der Graaf, W.T.A.; de Vries, E.G.E. Br. J. Cancer 1999, 79,
[104] Lee, T.H.; Solomon, M.J.; Mumby, M.C.; Kirschner, M.W.
882.
Cell 1991, 64, 415.
[105] Clarke, P.R.; Karsenti, E. J. Cell Sci. 1991, 100, 409. [129] de Jong, R.S.; de Vries, E.G.E.; Meijer, S.; de Jong, P.E.;
Mulder, N.H. Cancer Chemother. Pharmacol. 1998, 42,
[106] Krek, W.; Nigg, E.A. EMBO J. 1991, 10, 305. 160.

[107] Norbury, C.; Blow, J.; Nurse, P. EMBO J. 1991, 10, 3321. [130] Ozasa, T.; Suzuki, K.; Sasamata, M.; Tanaka, K.; Kobori,
M.; Kadota, S.; Nagai, K.; Saito, T.; Watanabe, S.;
[108] Bosc, D.G.; Lüscher, B.; Litchfield, D.W. Mol. Cell. Iwanami, M. J. Antibiot. 1989, 42, 1331.
Biochem. 1999, 191, 213.
[131] Ozasa, T.; Tanaka, K.; Sasamata, M.; Kaniwa, H.; Shimizu,
[109] Ackerman, P.; Glover, C.V.; Osherhoff, N. Proc. Natl. M.; Matsumoto, H.; Iwanami, M. J. Antibiot. 1989, 42,
Acad. Sci. USA 1985, 82, 3164. 1339.
[110] Sahyoun, N.; Wolf, M.; Besterman, J.; Hsieh, T.-S.;
[132] Fushimi, S.; Nishikawa, S.; Shimazu, A.; Seto, H. J .
Sander, M.; LeVine III, H.; Chang, K.-J.; Cuatrecasas, P.
Antibiot. 1989, 42, 1019.
Proc. Natl. Acad. Sci. USA 1986, 83, 1603.
[133] Fushimi, S.; Furihata, K.; Seto, H. J. Antibiot. 1989, 42,
[111] Rottmann, M.; Schröder, H.C.; Gramzow, M.; Renneisen,
1026.
K.; Kurelec, B.; Dorn, A.; Friese, U.; Müller, W.E.G. EMBO
J. 1987, 6, 3939. [134] Tomiya, T.; Uramoto, M.; Isono, K. J. Antibiot. 1990, 43,
118.
[112] Eckberg, W.R.; Johnson, M.R.; Palazzo, R.E. Invertebr.
Reprod. Dev. 1996, 30, 71. [135] Maeda, M.; Kodama, T.; Asami, S. 2 - P y r a n o n e
derivatives: Jpn. patent, 304893, Dec. 8, 1989.
[113] Takuwa, N.; Zhou, W.; Kumada, Z.; Takuwa, Y. Biochem.
Biophys. Res. Commun. 1992, 188, 1084. [136] Kohama, T.; Enokita, R.; Okazaki, T.; Miyaoka, H.;
[114] Egloff, M.-P.; Cohen, P.T.W.; Reinemer, P.; Barford, D. J. Torikata, A.; Inukai, M.; Kaneko, I.; Kagasaki, T.;
Mol. Biol. 1995, 254, 942. Sakaida, Y.; Satoh, A.; Shiraishi, A. J. Antibiot. 1993, 46,
1503.
[115] Goldberg, J.; Huang, H.; Kwon, Y.; Greengard, P.; Nairn,
A.C.; Kuriyan, J. Nature (London) 1995, 376, 745. [137] Kohama, T.; Nakamura, T.; Kinoshita, T.; Kaneko, I.;
Shiraishi, A. J. Antibiot. 1993, 46, 1512.
[116] Quinn, R.J.; Taylor, C.; Suganuma, M.; Fujiki, H. Bioorg.
Med. Chem. Lett. 1993, 3, 1029. [138] Shibata, T.; Kurihara, S.; Yoda, K.; Haruyama, H.
Tetrahedron 1995, 51, 11999.
[117] Fujiki, H.; Suganuma, M. Adv. Cancer Res. 1993, 61,
143. [139] Ohkuma, H.; Naruse, N.; Nishiyama, Y.; Tsuno, T.;
Hoshino, Y.; Sawada, Y.; Konishi, M.; Oki, T. J. Antibiot.
[118] Taylor, C.; Quinn, R.J.; McCulloch, R.; Nishiwaki- 1992, 45, 1239.
Matsushima, R.; Fujiki, H. Bioorg. Med. Chem. Lett.
1992, 2, 299. [140] Amemiya, M.; Ueno, M.; Osono, M.; Masuda, T.;
Kinoshita, N.; Nishida, C.; Hamada, M.; Ishizuka, M.;
[119] Ubukata, M.; Cheng, X.-C.; Isobe, M.; Isono, K. J. Chem. Takeuchi, T. J. Antibiot. 1994, 47, 536.
Soc. Perkin Trans. I 1993, 617.
[141] Amemiya, M.; Someno, T.; Sawa, R.; Naganawa, H.;
[120] Gauss, C.-M.; Sheppeck, II, J.E.; Nairn, A.C.; Chamberlin, Ishizuka, M.; Takeuchi, T. J. Antibiot. 1994, 47, 541.
R. Bioorg. Med. Chem. 1997, 5, 1751.
[142] Yamazaki, K.; Amemiya, M.; Ishizuka, M.; Takeuchi, T. J.
[121] Sugiyama, Y.; Ohtani, I.I.; Isobe, M.; Takai, A.; Ubukata, Antibiot. 1995, 48, 1138.
M.; Isono, K. Bioorg. Med. Chem. Lett. 1996, 6, 3.
[143] Kawada, M.; Amemiya, M.; Ishizuka, M.; Takeuchi, T.
[122] Bagu, J.R.; Sykes, B.D.; Craig, M.M.; Holmes, C.F.B. J. Jpn. J. Cancer Res. 1999, 90, 219.
Biol. Chem. 1997, 272, 508.
[144] Masuda, T.; Watanabe, S.; Amemiya, M.; Ishizuka, M.;
[123] Lindvall, M.K.; Pihko, P.M.; Koskinen, A.M.P. J. Biol. Takeuchi, T. J. Antibiot. 1995, 48, 528.
Chem. 1997, 272, 23312.
[145] Kawada, M.; Amemiya, M.; Ishizuka, M.; Takeuchi, T.
[124] Connor, J.H.; Kleeman, T.; Barik, S.; Honkanen, R.E.; Biochim. Biophys. Acta 1999, 1452, 209.
Shenolikar, S. J. Biol. Chem. 1999, 274, 22366.
[146] Chavez, D. E.; Jacobsen, E. N. Angew. Chem. Int. Ed.
[125] Zhang, Z.; Zhao, S.; Long, F.; Zhang, L.; Bai, G.; Shima, 2001, 40, 3667.
H.; Nagao, M.; Lee, E.Y.C. J. Biol. Chem. 1994, 269,
16997. [147] Tokunaga, M.; Larrow, J. F.; Jacobsen, E. N. Science
1997, 277, 936.
2032 Current Medicinal Chemistry, 2002, Vol. 9, No. 22 Boger et al.

[148] Dossetter, A. G.; Jamison, T. F.; Jacobsen, E. N. Angew. [171] Noyori, R.; Tomino, I.; Tanimoto, Y.; Nishizawa, M. J.
Chem. Int. Ed. 1999, 38, 2398. Am. Chem. Soc. 1984, 106, 6709. Noyori, R.; Tomono, I.;
Yamada, M.; Nishizawa, M. J. Am. Chem. Soc. 1984, 106,
[149] Wipf, P.; Xu, W. Tetrahedron Lett. 1994, 35, 5197. 6717.
[150] Stork, G.; Zhao, K. Tetrahedron Lett. 1989, 30, 287. [172] Esumi, T.; Okamoto, N.; Iwabuchi, Y.; Hatakeyama, S.
Tennen Yuki Kagobutsu Toronkai Koen Yoshishu 2000,
[151] Matsumura, K.; Hashiguchi, H.; Ikariya, T.; Noyori R. J. 42, 643.
Am. Chem. Soc. 1997, 119, 8738.
[173] Cossy, J.; Pradaux, F.; BouzBouz, S. Org. Lett. 2001, 3,
[152] Myers, A. G.; Zheng, B.; Movassaghi, M. J. Org. Chem. 2233.
1997, 62, 7507. Hunig, S.; Muller, H. R.; Their, W.
Angew. Chem. 1965, 77, 368; Angew. Chem. Int. Ed. Engl. [174] Kiyotsuka, Y.; Igarashi, J.; Kobayashi, Y. Tetrahedron
1965, 4, 241. Lett. 2002, 43, 2725.
[153] Mapp, A. K.; Heathcock, C. H. J. Org. Chem. 1999, 64, 23. [175] Bialy, L.; Waldmann, H. Angew. Chem. Int. Ed. 2002, 41,
1748.
[154] Evans, D. A.; Gage, J. R.; Leighton, J. L. J. Am. Chem. Soc.
1992, 114, 9434. [176] Braun, M. Methoden Org. Chem., 4th Ed; 1952–;
Houben-Weyl: 1995; Vol. E21b, 1612.
[155] Reddy, Y. K.; Falck, J. R. Organic Lett. 2002, 4, 969.
[177] Midland, M. M.; McDowell, D. C.; Hatch, R. L.;
[156] Brown, H. C.; Jadhav, P. K. J. Am. Chem. Soc. 1983, 105, Tramontano, A. J. Am. Chem. Soc. 1980, 102, 867.
2092. Ramachandran, P. V.; Teodorovic, A. V.; Rangaishenvi,
[157] Bertus, P.; Pale, P. Tetrahedron Lett. 1997, 38, 8193. M. V.; Brown, H. C. J. Org. Chem. 1992, 57, 2379. Parker,
K. A.; Ledeboer, M. W. J. Org. Chem. 1996, 61, 3214.
[158] Smith, A. B., III.; Ott, G. R. J. Am. Chem. Soc. 1998, 120, Corey, E. J.; Bakshi, R. K.; Shibata, S. J. Am. Chem. Soc.
3935. Glanzer, B. I.; Faber, K.; Griengl, H. Tetrahedron 1987, 109, 5551. Bach, J.; Berenguer, R.; Garcia, J.;
1987, 43, 5791. Loscertales, T.; Vilarrasa, J. J. Org. Chem. 1996, 61,
9021. Matsumura, K.; Hashiguchi, S.; Ikariya, T.; Noyori,
[159] Kolb, H. C.; VanNieuwenhze, M. S.; Sharpless, K. B. R. J. Am. Chem. Soc. 1997, 119, 8738.
Chem. Rev. 1994, 94, 2483.
[178] Evans, D. A.; Bender, S. L.; Morris, J. J. Am. Chem. Soc.
[160] Hofmeister, H.; Annen, K.; Laurent, H. Wiechert, R. 1988, 110, 2506.
Angew. Chem., Int. Ed. Engl. 1984, 23, 727.
[179] Evans, D. A.; Bartroli, J. A.; Shih, T. L. J. Am. Chem. Soc.
[161] Duffault, J.-M.; Einhorn, J.; Alexakis, A. Tetrahedron 1981, 103, 2127.
Lett. 1991, 32, 3701.
[180] Penning, T. D.; Djuric, S. D.; Haack, R. A.; Kalish, V. J.;
[162] Grubbs, R. H.; Chang, S. Tetrahedron 1998, 54, 4413. Myashiro, J. M.; Rowell, B. W.; Yu, S. S. Synth. Commun.
1990, 20, 307.
[163] Gautier, E. C. L.; Graham, A. E.; McKillop, A.; Standen, S.
P.; Taylor, R. J. K. Tetrahedron Lett. 1997, 38, 1881. [181] Lee, A. S.-Y.; Hu, Y.-J.; Chu, S.-F. Tetrahedron 2001, 57,
2121.
[164] Rama Rao, A. V.; Reddy, E. R.; Sharma, G. V. M.;
Yadagiri, P.; Yadav, J. S. Tetrahedron Lett. 1985, 26, [182] Stille, J. K.; Groh, B. L. J. Am. Chem. Soc. 1987, 109, 813.
465.
[183] Watanabe, Y.; Nakamura, T.; Mitsumoto, H. Tetrahedron
[165] Ohmura, T.; Yamamoto, Y.; Miyaura, N. J. Am. Chem. Soc. Lett. 1997, 38, 7407.
2000, 122, 4990.
[184] Virshup, D. M. Curr. Opin. Cell Biol. 2000, 12, 180.
[166] Miyaura, N.; Suzuki, A. Chem. Rev. 1995, 95, 2457.
[185] Janssens, V.; Goris, J. Biochem. J. 2001, 353, 417.
[167] Miyashita, K.; Ikejiri, M.; Kawasaki, H.; Maemura, S.;
Imanishi, T. Chem. Comm. 2002, 742. [186] Gjertsen, B. T.; Doskeland, S. D. Biochim. Biophys. Acta
1995, 1269, 187.
[168] Mori, K.; Takigawa, T.; Matsuo, T. Tetrahedron 1979,
35, 933. Smith, A. B. III.; Chen, S. S.-Y.; Nelson, F. C.; [187] Silverstein, A. M.; Barrow, C. A.; Davis, A. J.; Mumby, M.
Reichert, J. M.; Salvatore, B. A. J. Am. Chem. Soc. 1997, C. Proc. Natl. Acad. Sci. USA 2002, 99, 4221.
119, 10935.
[188] Fladmark, K. E.; Brustugun, O. T.; Hovland, R.; Bøe, R.;
[169] Vijayasaradhi, S.; Singh, J.; Aidhen, I. S. Synlett 2000, Gjertsen, B. T.; Zhivotovsky, B.; Døskeland, S. O. Cell
110. Death Different. 1999, 6, 1099.

[170] Wenkert, E.; Guo, M.; Lavillla, R.; Porter, B.;


Ramachandran, K.; Sheu, J.-H. J. Org. Chem. 1990, 55,
6203.

You might also like