Gillian M.
Nicholas
a and
Andrew J.
Phillips
*b
aRoche Colorado Corporation, 2075 North 55th St., Boulder, Colorado 80301
bDepartment of Chemistry and Biochemistry, University of Colorado, Boulder, Colorado 80309, USA. E-mail: Andrew.Phillips@colorado.edu; Fax: 1 303 492 0439; Tel: 1 303 735 2049
First published on 5th January 2006
Covering: January to December 2004. Previous review: Nat. Prod. Rep., 2005, 22, 144
An overview of marine natural products synthesis during 2004 is provided. As with the previous installment of this series, the emphasis is on total syntheses of molecules of contemporary interest, new total syntheses, and syntheses that have resulted in structure confirmation or stereochemical assignments.
![]() Gillian M. Nicholas | Gillian Nicholas received her B.Sc., M.Sc., and Ph.D. degrees from the University of Canterbury, where she worked on the isolation and structure elucidation of biologically active compounds from fungal sources with John Blunt and Murray Munro. Postdoctoral research with Ted Molinski at UC Davis and Carole Bewley at the National Institutes of Health was followed by a 2 year stay at Hauser Inc. in Denver, CO. She is currently a Principal Analytical Research Chemist at Roche Colorado Corporation in Boulder, CO. |
![]() Andrew J. Phillips | Andy Phillips obtained his B.Sc. (Hons) and Ph.D. degrees from the University of Canterbury in Christchurch, New Zealand. After a postdoctoral appointment with Peter Wipf at the University of Pittsburgh, he joined the faculty at the University of Colorado. His research interests focus around the development of new methods and strategies for the synthesis of complex natural products. |
![]() | ||
Scheme 1 Chemical degradation and derivatization of azaspiracid-1 (originally proposed structure) to C1–C20 alcohol 6 and C26–C40 carboxylic acid 7. Reagents and conditions: (1) TMSCHN2, MeOH, 25 °C; (2) NaIO4, MeOH–H2O (4 : 1), 25 °C, ∼100% over 2 steps; (3) NaBH4, MeOH, 25 °C, ∼90%; (4) 5% Pd/C, H2, MeOH, 25 °C, ∼90%; (5) O3, Me2S, MeOH, −78 °C; (6) NaOH (0.1 N), MeOH–H2O (4 : 1), 25 °C, 2 h; (7) TMSCHN2, MeOH, 25 °C, ∼90% over 3 steps; (8) NaIO4, MeOH–H2O (4 : 1), 25 °C, ∼90%. |
Key steps of the synthesis of the EFGHI ring containing system 8 are shown in Scheme 2. The primary alcohol of dithiane 9 was acetylated, and subsequent dithiane deprotection with PhI(TFA)2 followed by oxidation of the lactol to the lactone with N-iodosuccinimide produced 10 in 51% over the three steps. Stille coupling with stannane 11 in the presence of Pd(0) resulted in the formation of dihydropyran 12. Removal of the TES ether with HF·pyridine, followed by treatment with N-iodosuccinimide to induce iodoetherification, produced iodoether 13 in 38% yield (for the three steps). Two further steps (removal of the iodide and cleavage of the Teoc carbamate) gave the structure corresponding to degradation product 3 and showed the structure of the compound obtained by degradation is in fact epimeric, in terms of the stereochemistry around the E ring, to that which was originally proposed. A synthesis of the compound in which the FGHI rings were enantiomeric was also completed by this route, but it did not match the data for 3. These, and related studies that involved the synthesis of acid 7, also established the absolute stereochemistry for this domain.
![]() | ||
Scheme 2 Synthesis of the EFGHI ring containing lactone. Reagents and conditions: (1) Ac2O, 2,4,6-collidine, CH2Cl2, 85%; (2) PhI(TFA)2, MeCN–pH 7 buffer, 81%; (3) NIS, TBAI, CH2Cl2, 74%; (4) 3.0 equiv. lactone 10, 90 mol% [Pd2dba3], LiCl, AsPh3, DIPEA, syringe pump additon of stannane; (5) HF·py, THF–py; (6) NIS, NaHCO3, THF, 38% (3 steps). |
At this juncture, attention turned to questions regarding the connectivity and stereochemistry of the ABCD ring containing domain. Desilylation of previously synthesized compound 14 to give 15 provided material that could be compared with degradation product 4 (Schemes 1 and 3). The spectroscopic data for these two samples differed substantially, particularly in the A ring. Progress towards a corrected structure was assisted by comparison of NMR data with a related compound, lissoketal (16).20 Based on this comparison, a new structure in which the A ring double bond has been relocated was proposed; however, final resolution of the problem did not come until the synthesis of several closely related structures had been completed. Based on this work the stereochemistry and connectivity shown in compound 17 was secured as corresponding to degradation compound 4.
![]() | ||
Scheme 3 Comparison of synthetic and proposed ABCD ring containing alcohols. Reagents and conditions: (1) TBAF, THF, 88%. |
Armed with the information gleaned from these studies, and the earlier synthesis, plans could be laid to complete the total synthesis by employing the key couplings shown in Scheme 4.
![]() | ||
Scheme 4 Key carbon–carbon bond-forming reactions in the Nicolaou azaspiracid-1 synthesis. |
Key steps of the route to the ABCDE domain are shown in Scheme 5. Malic acid derived tetrahydrofuran 18 was treated with TMSOTf in CH2Cl2 at low temperature to induce the desired deprotection–spirocyclization sequence to give 19 in 89% as a single stereoisomer. A sequence of 6 steps led to allylic carbonate 20, and deoxygenation of this compound to give 21 was achieved by employing a modification of an earlier-described Pd-catalyzed method21 which produced the desired compound in 82% yield (with 7 : 1 selectivity for the Δ7,8 olefin). After advancement to pentafluorophenyl ester 22, introduction of the C21–C27 domain involved acylation of the dithiane anion derived from 23 to give 24 in 50% yield. Six further steps provided compound 25 (see Scheme 6) which served as the key precursor to the A→E domain for the final steps of the synthesis.
![]() | ||
Scheme 5 Synthesis of the ABCDE domain. Reagents and conditions: (1) TMSOTf (4.0 equiv.), CH2Cl2, −90 °C, 30 min, 89%; (2) [Pd2dba3]·CHCl3 (0.125 equiv.), nBu3P (0.48 equiv.), LiBH4 (10.0 equiv.), DME, 0 °C, 1 h, 82%, Δ7,8/Δ8,9 = 7 : 1; (3) 23 (7.0 equiv.), nBuLi–nBu2Mg (4.7 equiv.), THF, 0 → 25 °C, 1.5 h, then −90 °C, then 22, 30 min, 50%. PFP = pentafluorophenyl. |
![]() | ||
Scheme 6 The closing stages of the Nicolaou synthesis of azaspiracid-1. Reagents and conditions: (1) 25, [Pd2dba3] (0.3 equiv.), AsPh3 (0.3 equiv.), LiCl (6.0 equiv.), DIPEA, then 26 (3.0 equiv. in THF by syringe pump addition), NMP, 40 °C, 1 h, 55%; (2) TBAF (1.2 equiv.), THF, 0 °C, 80%; (3) NIS (2.0 equiv.), NaHCO3, THF, 0 °C, 12 h, 62%; (4) Et3B (0.2 equiv.), nBu3SnH–PhMe (1 : 2), 0 °C, 5 min, 86%. |
The closing stages of the synthesis are illustrated in Scheme 6. The crucial coupling of the ABCD and FHI subunits occurred by Pd-mediated Stille-type reaction between allylic acetate 25 and dihydropyranyl stannane 26 to give 27 in 55% yield. Removal of the C34 TES ether (TBAF, 80%) and iodoetherification with N-iodosuccinimide produced 28 in an impressive 62% yield given the complexity of the substrate. Subsequent deiodination with Bu3SnH and Et3B give the fully elaborated A→I ring system (29) of the natural product, and the synthesis was completed by a short sequence of 5 steps that consisted of redox chemistry and protecting group manipulations.
In other studies on azaspiracid-1, Forsyth has descibed a concise approach to the trioxadispiroketal ABCD ring system 31 (the stereochemistry at C6 is epimeric to the final stereochemistry determined by Nicolaou). The key step consists of treatment of the ynone 30 with p-TsOH in toluene at room temperature for 1–2 days. Under these conditions the C6 and C17 TES ethers are removed and the ensuing sequence of hemiacetal formations and conjugate additions leads to 31 in 55% over two steps.22
Carter and Zhou have described a concise 23-step synthesis of the ABCDE ring system 32 from N-propionyl-(S)-benzyloxazolidinone.23 Ishikaway and Nishiyama have described a synthesis of compound 33, which contains the ABCD ring system, as well as related studies.24,25
![]() | ||
Scheme 7 Key steps of Nakata's total synthesis of brevetoxin B. Reagents and conditions: (1) ethyl propiolate, N-methylmorpholine, CH2Cl2, rt, 95%; (2) SmI2, MeOH, THF, rt; (3) p-TsOH, EtOH, 80 °C, 79% (2 steps); (4) TMSOTf, 2,6-lutidine, CH2Cl2, −78 °C, 100%; (5) DIBAL-H, PhMe, −78 °C; (6) Ph3P![]() ![]() |
Fujiwara and co-workers have reported a formal total synthesis27 of hemibrevetoxin B28 by a route that reaches 53, and bisects the Yamamoto synthesis (Scheme 8). The synthesis commences with the synthesis of 50 from butyrolactone by an 18-step sequence that employs a ring-closing metathesis for the synthesis of the 7-membered ring (46 → 48). Aldehyde 50 is coupled with the anion derived from 49 (synthesized from tri-O-acetyl-D-glucal 47 by a 14-step sequence) to give 51 in 86% yield. Acidic hydrolysis of the (methylthio)(methylsulfinyl)acetal and removal of the TBS ether gave hydroxyl ketone 52, from which compound 53 was prepared by a 7-step sequence.
![]() | ||
Scheme 8 Fujiwara's formal synthesis of hemibrevetoxin B. Reagents and conditions: (1) TBSOTf, 2,6-lutidine, CH2Cl2, 99%; (2) DDQ, CH2Cl2–H2O (10 : 1), 93%; (3) acryloyl chloride, DIPEA, CH2Cl2, 94%; (4) Grubbs’ 2nd-generation catalyst, CH2Cl2, reflux, 89%; (5) 49, LDA then 50 (0.37 equiv.), −78 °C, 86%; (6) TFA–THF–H2O (1 : 5 : 5), 0 °C → rt, 71%. |
Inoue and Hirama have described the end-game of a 2nd-generation total synthesis of ciguatoxin CTX3C29 that involves milder conditions for the preparation of O,S-acetals (Scheme 9).30 Aldehyde 55 was converted to the α-chlorosulfide 58 by a sequence of three straightforward reactions (reduction, sulfide formation and chlorination). The authors note that ‘the reproducibility of the chlorination reaction … required the addition of 1 equiv. of NCS in CH2Cl2 to a solution of 57 in CCl4 at room temperature. The obtained solution of 58 in CH2Cl2/CCl4 (1:6) was directly used in the subsequent reaction because of the instability of 58 to any standard workup.’ Coupling of this α-chlorosulfide with alcohol 54 was effected by treating a CH2Cl2 solution of the alcohol, DTBMP, and AgOTf (2.0 equiv.) with the α-chlorosulfide. Removal of the C21 TIPS ether (TBAF, 85%) and conversion to the β-alkoxyacrylate (ethyl propiolate, NMM, 100%) set the stage for the radical cyclization to form the G ring. To this end, treatment of 60 with Bu3SnH and AIBN in PhMe led to 61 in 54% yield (along with 27% of the compound obtained from cyclization onto the C21 olefin). Two further steps (DIBAL-H reduction of the ester and olefination with Ph3PCH2, 92% over two steps) provided the substrate for a RCM analogous to that of the 1st-generation synthesis. Subjecting 62 to Grubbs’ I catalyst in CH2Cl2 at reflux produced 90% of protected CTX3C, and subsequent removal of the naphthylmethyl (NAP) ethers with DDQ gave 63% of CTX3C. This sequence is 4 steps shorter than the 1st-generation synthesis. Inoue and Hirama have also provided a detailed account of their 1st-generation synthesis31 and also a review of their recently developed methods for the synthesis of CTX3C,32 and Inoue has reviewed the use of acetal-linked intermediates for the synthesis of complex polyethers.33
![]() | ||
Scheme 9 Key steps from the Inoue–Hirama 2nd-generation synthesis of ciguatoxin CTX3C. Reagents and conditions: (1) NaBH4, MeOH–CH2Cl2, 81%; (2) PhSSPh, Bu3P, py, 93%; (3) NCS, CH2Cl2–CCl4 (1 : 6); (4) 54 (1.2 equiv.), AgOTf (2.0 equiv.), DTBMP, 4 Å MS, CH2Cl2–CCl4, 70% from 58; (5) (a) TBAF, THF, 85%; (b) methyl propiolate, NMM, CH2Cl2, 100%; (6) nBu3SnH, AIBN, PhMe, 54%; (7) (a) DIBAL-H, CH2Cl2; (b) Ph3PCH3![]() ![]() |
In other studies on polyethers, Clark and co-workers34 have described a 12-step synthesis of the A ring fragment of the gambieric acids.35 The key step is a [2,3]-sigmatropic rearrangement of an allylic oxonium ylide generated from the intramolecular reaction of a crotyl ether with a copper carbenoid. Related studies have shown that sequential ring-closing enyne and cross metathesis36 of carbohydrate systems is a useful method for the synthesis of polyether building blocks. Isobe and co-workers have developed an approach to the HIJ ring system of ciguatoxin that features an intromolecular Nicholas reaction for the synthesis of the 8-membered I ring (Scheme 10).37 Full details of the Sasaki total synthesis of gambierol have been reported, along with initial SAR studies that delineate some of the features required for potent toxicity.38
![]() | ||
Scheme 10 Isobe's intramolecular Nicholas approach to the I ring of ciguatoxin. Reagents and conditions: (1) Co2(CO)8, CH2Cl2, 87%; (2) BF3·OEt2, CH2Cl2, 60%. |
An improved synthesis of key intermediate 66 in the Hirama ciguatoxin CTX3C has been reported. The sequence to this compound is now 38 steps, and displays improved stereochemical control over the earlier route.39
The synthesis of models of the CDE–FG ring systems of the prymnesins has suggested that a reassignment of relative stereochemistry should be considered in this region.40 A series of gambierol analogs have been synthesized from advanced intermediates in the Sasaki total synthesis and evaluated for toxicity against mice. These SAR studies indicate that the C28–C29 double bond (in the H ring) and the unsaturated side chain are crucial and that the C1 and C6 hydroxy groups, the C30 Me group, and the C37–C38 olefin have minimal influence on toxicity.41 Trost has described the Ru-catalyzed cycloisomerization and oxidative cyclization of bis-homopropargylic alcohols as a rapid iterative approach to the pyranopyran rings found in ladder toxins such as prymnesin and yessotoxin.42
![]() | ||
Scheme 11 Overman's synthesis of (−)-dehydrobatzelladine C. Reagents and conditions: (1) HO(CH2)4NHBoc, DMAP, PhMe, 100 °C, 78%; (2) morpholinium acetate, Na2SO4, CF3CH2OH, 65 → 70 °C; (3) CAN, MeCN, rt (45%, 2 steps); (4) 1 : 1 TFA–CH2Cl2, rt, ∼100%; (5) S![]() |
Nagasawa has described a synthesis of batzelladine A45 that is based around the use of diastereoselective 1,3-dipolar cycloadditions of nitrones with olefins to provide precursors to the bicyclic and tricyclic guanidinium units (Scheme 12).
![]() | ||
Scheme 12 An overview of the retrosynthetic analysis of batzelladine A by Nagasawa and co-workers. |
The bicyclic guanidinium domain synthesis commenced with the 1,3-dipolar cycloaddition reaction between nitrone 74 and ester 80 to give isoxazolidine 81 as a single diastereomer (Scheme 13).46 Reduction of the ester with LiAlH4 and removal of the TIPS group with CsF gave 82 in 59% overall yield from 81. After selective protection of the primary alcohol with TBSCl and pyridine, removal of the secondary alcohol at C9 was achieved by a Barton–McCombie reaction to produce 83 (44% over 3 steps). A sequence of 9 steps then led to alcohol 84.
![]() | ||
Scheme 13 Nagasawa's synthesis of batzelladine A. Reagents and conditions: (1) 1-undecene, PhMe, 90 °C, 75%; (2) methyl crotonate, PhMe, 100 °C, >60%; (3) (a) TBAF, THF, 97%; (b) Jones’ reagent, acetone; (4) PhMe, 90 °C; (5) LiAlH4, Et2O, 0 °C, then CsF, EtOH, 90 °C (59%, 3 steps); (6) (a) TBSCl, py, 81%; (b) ClC(S)OPh, py, DMAP, 58%; (c) nBu3SnH, AIBN, 94%; (7) EDCI, DMAP, CH2Cl2 (60%, 2 steps from acid); (8) HF·py, THF, 80%; (9) (a) Pd/C, H2; (b) PPh3, DEAD, PhMe; (c) TFA–CH2Cl2 (24%, 3 steps). |
The synthesis of acid 79, which ultimately serves as a progenitor of the tricyclic guanidine, commences with the diastereoselective 1,3-dipolar cycloaddition of 1-undecene and nitrone 74 to give 75 in 75% yield (see also Scheme 13). After a 4-step sequence of deprotection, acylation with ClC(S)OPh, Barton–McCombie deoxygenation, and oxidation with MCPBA, nitrone 76 was then subjected to cycloaddition with methyl crotonate to give 77. This material was advanced to compound 78, which was deprotected (TBAF, 97%) and oxidized with Jones' reagent to give acid 79. Subunit coupling of this acid with bicyclic guanidine 84 was achieved by esterification with EDCI in the presence of DMAP under carefully controlled conditions (epimerization at C24 occurred at room temperature, but could be supressed at 0 °C) to furnish 85 in 60% yield. Deprotection of the secondary alcohol with HF·pyridine yielded 86 in 80% yield. After removal of the Cbz protecting groups under standard hydrogenolysis conditions the resulting bicyclic guanidine was converted to the tricyclic guanidine by an intramolecular Mitsunobu reaction. The remaining four Boc groups were cleaved with TFA, and after purification by reverse-phase HPLC, batzelladine A was obtained as the trifluoroacetate salt in 24% yield for the 3 steps from 86.
In other studies on the batzelladines, Elliott has described the application of diastereoselective three-component coupling of an alkenylpyrrolidine, an aldehyde and Si(NCS)4 as a method to prepare the bicyclic core of the batzelladine alkaloids.47 In a collaboration between the Bewley and Overman groups, a synthetic library of 28 batzelladine analogs was tested for the ability to inhibit HIV-1 envelope-mediated cell–cell fusion.48 SAR relationships indicated that the best inhibitors of fusion were most similar in structure to natural batzelladine F, with IC50 values ranging from 0.8 to 3.0 µM.
The synthesis highlights Trost's new methodology for the construction of 1,4-dienes by the Ru-catalyzed coupling of alkenes and alkynes (Scheme 14).54 The first subunit coupling was achieved by reaction of 89 with 90 in the presence of Cp*Ru(MeCN)3PF6 as catalyst (Scheme 14). This catalyst provided the branched product 91 in 23% yield (39% yield based on recovered starting material). A straightforward sequence of 3 steps provided acid 92 which was coupled to the potentially sensitive epoxy alcohol 93 under Kita's conditions55 to give ester 94 in 51% yield. After removal of the triethylsilyl ethers with TBAF–AcOH, [Cp*Ru(MeCN)3]PF6-catalyzed macrocyclization of 95 provided amphidinolide A1, 96. Although the yield may seem modest (33% or 38% based on recovered starting material) this an impressive example of the remarkable selectivity of the Ru-catalyzed alkene–alkyne addition. The spectral data for synthetic material matched very well to the natural product, with only two protons deviating by more than 0.01 ppm from the reported values (the two deviations were by 0.03 ppm and by 0.02 ppm). The 13C NMR spectrum deviated by 0.1 ppm or less in CDCl3, J values in three solvents were also in agreement, and the optical rotation was also consistent with reported data (synthetic [α]24D +56 (c 0.05, CHCl3) cf. reported [α]24D +46 (c 1.0, CHCl3)). Even with these excellent comparisons, in the absence of authentic material for comparison, Trost and Harrington conclude their paper with guarded comments: “In conclusion, we have employed a combination of synthesis and NMR spectroscopy as tools to determine the correct structure of amphidinolide A1. Although the lack of a sample of the natural product prevents a definitive comparison, the excellent correlation [of our synthetic compound] strongly suggests it is (+)-amphidinolide A1.”
![]() | ||
Scheme 14 Trost's synthesis of amphidinolide A1. Reagents and conditions: (1) 89 (5 equiv.), [Cp*Ru(MeCN)3]PF6, 23% (39% brsm); (2) (i) [RuCl2(p-cymene)]2, ethoxyacetylene; (ii) 93, CSA, 51%; (3) TBAF, AcOH, 79%; (4) [Cp*Ru(MeCN)3]PF6, 33% (38% brsm). |
Trost has also completed a synthesis of amphidinolide P56 that employs the Ru-catalyzed alkene–alkyne coupling for subunit assembly (Scheme 15). Coupling of β-lactone 98 (3.5 equivalents) and enyne 97 in the presence of 10 mol% [CpRu(CH3CN)3]PF6 in acetone at room temperature gave 99 in 75% yield, and with no trace of linear product detectable. Four further steps led to amphidinolide P.
![]() | ||
Scheme 15 The key Ru-catalyzed alkene–alkyne coupling en route to amphidinolide P. Reagents and conditions: (1) 10 mol% [CpRu(CH3CN)3]PF6, acetone, 75%. |
Ghosh has synthesized amphidinolide W57 and corrected the stereochemistry at C6.58 The synthesis of the originally proposed structure employs a Ru-catalyzed cross-metathesis for subunit coupling (Scheme 16). Among various protecting groups and catalysts surveyed, an acetate protecting group on the secondary allylic alcohol of 102 and second generation Grubbs’ catalyst gave the best result, affording the desired compound 103 along with the Z-isomer (E/Z ratio 11 : 1) in 85% yield. A short sequence of protecting group and oxidation state manipulations provided seco-acid 104. Macrolactonization under Yamaguchi conditions provided the desired macrolactone 105 along with an epimer as a 1 : 1 mixture in 50% combined yield. Removal of dioxolane and MOM protecting groups produced the target compound, which was not identical to the natural material. After determining the stereochemistry at C6 to be the likely culprit, this sequence was repeated (details not discussed in the paper) to provide synthetic amphidinolide W whose spectral data (1H and 13C NMR) were identical to those of natural amphidinolide W.2
![]() | ||
Scheme 16 Ghosh's synthesis of the proposed structure for (+)-amphidinolide W. Reagents and conditions: (1) (H2IMes)(PCy3)Cl2Ru![]() |
Jamison has described a concise, modular synthesis of amphidinolide T159 that employs a Ni-catalyzed reductive macrocyclization of an alkyne with an aldehyde (Scheme 17).60 Esterification of alcohol 106 and carboxylic acid 107 under standard DCC-based conditions in the presence of 4-pyrrolidinopyridine gave ester 108. Removal of the TBS ether and oxidation with Dess–Martin periodinane gave the precursor to the macrocyclization, 109, in 59% yield for the three steps. Exposure of 109 as a 0.05 M solution to catalytic Ni(cod)2, Bu3P, and Et3B in toluene at 60 °C produced the desired allylic alcohol 110 with excellent stereocontrol (>10 : 1 dr) and in 44% yield. Protection of the secondary alcohol as the TBS ether, ozonolysis, selective methylenation using a modification of Takai's method,61 and finally desilylation with HF·pyridine led to amphidinolide T1, 111.
![]() | ||
Scheme 17 Key steps of Jamison's synthesis of amphidinolide T1. Reagents and conditions: (1) DCC, 4-pyrrolidinopyridine; (2) TBAF, THF; (3) Dess–Martin periodinane, 59% (3 steps); (4) 20 mol% Ni(cod)2, 40 mol% Bu3P, Et3B, PhMe, 60 °C, 44% (>10 : 1 dr); (5) (a) TBSOTf, 2,6-lutidine; (b) O3, then Me2S; (c) CH2I2, Zn, ZrCl4, PbCl2; (d) HF·py, 25% (4 steps). |
Amphidinolide X62 has been synthesized by Fürstner.63 The synthesis features subunit coupling by esterification and B-alkyl Suzuki reactions (Scheme 18). Acid 112 and alcohol 113 were coupled using Yamaguchi's conditions to give ester 114 in 96% yield. Reaction of this compound with boronate 119 (which was derived from the corresponding iodide that was in turn synthesized by a 12-step sequence involving the application of a newly developed iron-catalyzed ring-opening reaction of a propargyl epoxide with a Grignard reagent as a method for the synthesis of an allenol en route to a dihydrofuran) in the presence of Pd(0) produced 115 in an excellent 74% yield. Removal of the methyl ester with LiI followed by unmasking of the ketone with aqueous AcOH led to 116 in 53% yield over the two steps, and subsequent removal of the PMB group with DDQ in the presence of pH 7 buffer produced seco-acid 117 in 84% yield. The synthesis was then completed by Yamaguchi macrolactonization to yield amphidinolide X, 118, in 62% yield.64
![]() | ||
Scheme 18 Subunit couplings and the closing stages of Furstner's synthesis of amphidinolide X. Regents and conditions: (1) 112, 2,4,6-trichlorobenzoyl chloride, Et3N, PhMe, then 113, DMAP, 96%; (2) (dppf)PdCl2, Ph3As, K3PO4, aq. DMF, 74%; (3) LiI, py, 125 °C; (4) aq. AcOH, 53% (over 2 steps); (5) DDQ, CH2Cl2, pH 7 buffer, 84%; (6) 2,4,6-trichlorobenzoyl chloride, Et3N, THF, then DMAP, PhMe, 62%. |
In other studies on the amphidinolides, a synthesis of the C11–C29 fragment of amphidinolide F65 has been accomplished using a diastereoselective [3 + 2]-annulation reaction of an allylsilane and ethylglyxolate to produce the key tetrahydrofuran.66 A synthesis of the C19–C26 subunit of amphidinolides B1 and B2 has been completed using a boron-mediated aldol reaction, and a synthesis of the C19–C26 subunit of amphidinolide B3 has also been completed.67 Ghosh has also provided a detailed account of their synthesis of amphidinolide T1.68
![]() | ||
Scheme 19 A comparison of key reactions in the Paterson and Curran dictyostatin-1 syntheses. |
The Paterson synthesis ultimately commences with methyl (3S)-hydroxy-2-methylpropionate (the ‘Roche ester’, 119, Scheme 20). A sequence of 14 relatively routine steps from the Roche ester led to β-ketophosphonate 120, and an 11-step synthesis produced aldehyde 121. These two fragments were coupled by a Horner–Wadsworth–Emmons reaction employing Ba(OH)2 in wet THF73 to give enone 122 in 92% yield. Conjugate reduction of the enone with the Stryker's reagent,74 followed by removal of both PMB groups with DDQ provided an intermediate β-hydroxyketone, which underwent reductione in a 1,3-syn-selective fashion when exposed to Zn(BH4)2 in Et2O to produce triol 123 (>20 : 1 dr).
![]() | ||
Scheme 20 Paterson's intial fragment coupling en route to dictyostatin. Reagents and conditions: (1) 120, Ba(OH)2·8H2O, THF, 1 h, then 121, 40 : 1 THF–H2O, 92%; (2) [Ph3P·CuH]6, PhH, H2O, rt; (3) DDQ, pH 7 buffer, CH2Cl2, 0 °C; (4) Zn(BH4)2, Et2O, −30 °C, 66% (3 steps, >20 : 1 dr). |
Key fragments 124 and 125 were coupled by a Still–Gennari-modified Horner–Wasdworth–Emmons olefination to give α,β-unsaturated ketone 126 with good levels of selectivity (Z/E = 5 : 1, 77%, Scheme 21). Stille coupling of the vinyl iodide with stannane 127 under Liebeskind conditions75 followed by cleavage of the TIPS ester with KF in THF–MeOH gave the seco-acid 128 in 83% overall yield for these two steps. Yamaguchi macrolactonization furnished the macrocycle in 77% yield, and reduction of the C9 ketone under Luche conditions provided the final stereocenter under macrocyclic stereocontrol (70%). Global deprotection of the TBS groups with HCl in methanol gave dictyostatin-1 in 87% yield.
![]() | ||
Scheme 21 Final steps of the Paterson synthesis of dictyostatin-1. Reagents and conditions: (1) K2CO3, [18]-crown-6, PhMe, rt, Z/E = 5 : 1, 77% (from the primary alcohol preceding 124); (2) CuTC, NMP, rt, then KF, THF–MeOH, rt, 83% (2 steps); (3) 2,4,6-trichlorobenzoyl chloride, Et3N, DMAP, PhMe, 60 °C, 77%; (4) NaBH4, CeCl3·7H2O, EtOH, −30 °C, 70%; (5) 3 N HCl, MeOH, rt, 87%. |
Curran's synthesis of dictyostatin is also predicated on the observation that the Roche ester can serve as the starting material for the β-ketophosphonate 129 and alkyne 130 (Scheme 22). The intial subunit coupling was achieved by the reaction of the lithiated acetylide derived from 130 with Weinreb amide 131 to give ynone 132 in 93% yield. The C9 stereocenter was set by reduction with Noyori's Ru(II)-(S,S)-N-p-toluenesulfonyl-1,2-diphenylethylenediamine catalyst with iPrOH as hydrogen donor, and subsequent Lindlar reduction of the alkyne and silylation with TBSOTf gave 133 (91% and 99% for the last two steps).
![]() | ||
Scheme 22 Fragment coupling of the C10–C18 and C3–C9 domains. Reagents and conditions: (1) 130, nBuLi, THF, then 131, 93%; (2) S,S Noyori catalyst (20 mol%), iPrOH, 79%; (3) Lindlar catalyst, H2, PhMe, 91%; (4) TBSOTf, 2,6-lutidine, CH2Cl2, 99%. |
Introduction of the final fragment was achieved by a Horner–Wadsworth–Emmons reaction between aldehyde 134 (obtained from 133 by desilylation with HF·pyridine in 67% yield and Dess–Martin oxidation) employing the Ba(OH)2 system (Scheme 23). Reduction of the enone was achieved by treatment with NiCl2–NaBH4, and the ketone was reduced with NaBH4 to give the desired alcohol 136 in 70% yield (along with 29% of the α-diastereoisomer). This compound was advanced to dictyostatin by a 12-step sequence similar to the approach employed by Paterson that featured Yamaguchi macrolactonization after the installation of the diene.
![]() | ||
Scheme 23 The second key fragment coupling of the Curran synthesis of dictyostatin-1. Reagents and conditions: (1) Ba(OH)2, 129 (see Scheme 21), then 134, THF–H2O, 80% (2 steps from the alcohol); (2) NiCl2, NaBH4, MeOH–THF, 76%; (3) NaBH4, MeOH–THF, 70% (β), 29% (α). |
In other work on dictyostatin, an approach to mixed polyacetate-polypropionates based on the cyclization of (silyloxy)enynes has been reported in the context of a synthesis of the C9–C19 subunit.
Discodermolide also continues to draw attention, and one of the highlights of 2004 in this area is the description of work at Novartis that resulted in the preparation of 60 g of discodermolide. The route is a hybrid of the Smith and Paterson synthetic route and involves 39 steps in totala (26 steps in the longest linear sequence).76 This work has also been reviewed by Mickel.77,78 Paterson has reported a 3rd-generation synthesis that takes advantage of a Still–Gennari-modified HWE reaction that is similar to the one employed en route to dictyostatin-1.79 Several subunit syntheses have been described including: (i) an approach to the C15–C24 subunit employing desymmetrization of a meso dialdehyde with the Haffner–Duthaler crotyltitanium reagent,80 (ii) syntheses of the C1–C8 and C15–C21 subunits by diastereoselective dihydroxylation of dihydropyranones and subsequent α-deoxygenation,81 (iii) a highly diastereoselective and practical route to a lactone that serves as a key building block for the Novartis synthesis82 and (iv) an approach that exploits the pseudosymmetry of the C1–C13 region.83 A series of 7-deoxydiscodermolide analogs in which the lactone was replaced by aromatics were designed, synthesized, and evaluated for cytotoxicity.84 The majority of the compounds in this study displayed low micromolar IC50 values against a variety of cancer cell lines. In other studies a series of analogs (which are minor side products generated during the final stages in the synthesis of (+)-discodermolide) were evaluated for in vitro cytotoxicity against 6 cell lines.85 These compounds showed significant variation of cytotoxicity, and suggest the relevance of the C11 hydroxyl group, the C13 double bond, and the 16S stereochemistry are important for potent cytotoxocity.
![]() | ||
Scheme 24 Key steps of the Stoltz synthesis of dragmacidin F. Reagents and conditions: (1) lithiopyrrole 139, THF, 66% from the acid; (2) Pd(OAc)2, DMSO, tBuOH–AcOH, 74%; (3) KOH, H2O, EtOH, then 6 N HCl, then K2CO3, THF–H2O, 96% (3 steps); (4) (a) TMSI, CH3CN; (b) H2NCN, NaOH, H2O, 82% (2 steps). |
Birman et al.88 and Baran et al.89 have independently described sequences that lead to members of the sceptrin class. The Baran synthesis of sceptrin proceeds in an excellent 24% overall yield from dimethylacetylene dicarbocylate and employs the acid-catalyzed rearrangement of 3-oxaquadricyclane 144 to cyclobutane 145 as the key step (Scheme 25).
![]() | ||
Scheme 25 The key step from the Baran synthesis of sceptrin. Reagents and conditions: (1) H2SO4, MeOH, 24 h, 50%. |
Based on their synthetic studies, Baran has synthesized ageliferin90 in a single step from sceptrin by heating sceptrin in H2O in a microwave (Scheme 26).91 Based on this result, Baran and co-workers have proposed an alternative hypothesis to the long-held [4 + 2]-cycloaddition biogenesis of these compounds and related structures such as axinellamine and palau'amine (Scheme 27).
![]() | ||
Scheme 26 Conversion of sceptrin to ageliferin. Reagents and conditions: (1) H2O, 195 °C, microwave, 1 min, 40% (plus 52% recovered sceptrin) |
![]() | ||
Scheme 27 Baran's proposed sequence for the conversion of dibromosceptrin to axinellamine A in nature. A similar sequence can be drawn for palau'amine. |
Spongidepsin92 has been synthesized by both the Forsyth and Ghosh groups, and the relative and stereochemistry has been established to be as shown.93,94
A synthesis of bistramide A95 that confirms Wipf's stereochemical assignment for bistramide C has been completed by Kozmin.96 The synthesis involves 15 steps in the longest linear sequence and features a nice application of the ring-opening cross-metathesis of a cyclopropene acetal (Scheme 28) as a novel method for the rapid assembly of precursors to spiroacetals. Readily accesible alkene 150 was subjected to 10 mol% Grubbs’ 2nd-generation catalyst in the presence of 1.5 equivalents of cyclopropenone acetal 152, and then aqueous acid to give dienone 151 in 63% over the two steps. Cross-metathesis with alkene 153 then provided 156 in 68% yield. Hydrogenolysis of the benzyl groups and the dienone resulted in acetalization, and subsequent Dess–Martin oxidation gave aldehyde 157 in 53% yield over the two steps. A sequence of three steps led to amino alcohol 158, and completed a very concise approach to this fragment. The synthesis was then completed by PyBOP-mediated acylation of the amine with 154 to produce 159. Removal of the Fmoc protecting group and acylation with activated ester 155 produced bistramide A in 65% over these final two steps.
![]() | ||
Scheme 28 The Kozmin synthesis of bistramide A. Reagents and conditions: (1) 10 mol% Grubbs’ 2nd-generation catalyst, 1.5 equiv. 152, PhH, 60 °C then 1 M H2SO4, MeCN, 63%; (2) 153, 10 mol% Grubbs’ 2nd-generation catalyst, 68%; (3) H2 (80 psi), Pd(OH)2/C; (4) Dess–Martin periodinane, 53% (2 steps); (5) 154, PyBOP, DMF, 85%; (6) Et2NH, DMF; (7) 155, DMF, 20 °C, 65%. |
A large number of other total synthesis of marine natural products were reported in the review period, and papers describing first total syntheses are presented in Table 1. New total syntheses of compounds previously prepared are summarized in Table 2.
Compound | Reference | Notes |
---|---|---|
(+)-Xyloketal D![]() |
Krohn et al.98 | • 5 Steps |
• Non-racemic | ||
• Bioactivity: potent acetylcholine esterase inhibition | ||
Cyercene A ![]() |
Baldwin et al.99 | • 4 Steps from known compound |
• Bioactivity: may act as mediators in tissue regeneration and chemical defence in the mollusc | ||
Plakortone G ![]() |
Nishiyama et al.100 | • 19 Steps |
• Non-racemic | ||
• Bioactivity: cytotoxic | ||
Austrodoric acid ![]() |
Gavagnin et al.101 | • 7 Steps from known compound |
• Non-racemic | ||
• Bioactivity: unknown | ||
(+)-Agelasine D ![]() |
Gundersen et al.102 | • 4 Steps from (+)-manool |
• Non-racemic | ||
• Bioactivity: related compounds are cytotoxic and antimicrobial | ||
Hachijodine B ![]() |
Lee et al.103 | • 10 Steps |
• Non-racemic | ||
• Bioactivity: cytotoxic against the P388 cell line | ||
(18S)-Variabilin ![]() |
Takabe et al.104 | • 11 Steps |
• Non-racemic | ||
• Bioactivity: compounds of this class have shown antiviral and cytotoxic activity | ||
(−)-Presphaerene ![]() |
Lee et al.105 | • 18 Steps |
• Non-racemic | ||
• Bioactivity: none described | ||
Dolastatin 18 ![]() |
Pettit et al.106 | • 5 steps from N-(benzyloxycarbonyl)-N-methylphenylalanine (Z-N-Me PheOH) |
• Non-racemic | ||
• Bioactivity: human cancer cell-growth inhibitor | ||
Barettin ![]() |
Bergman et al.107 | • 3 Steps from known compound |
• Non-racemic | ||
• Bioactivity: not stated | ||
No name ![]() |
Kende et al.108 | • 5 Steps from known compound (Kende) |
Bewley and Fetterolf109 | • 7 Steps from 3,5-dibromotyrosine (Bewley and Fetterolf) | |
• Bioactivity: inhibits the mycobacterial enzyme mycothiol S-conjugate amidase | ||
(+)-Phomopsidin ![]() |
Nakada et al.110 | • 37 Steps |
• Non-racemic | ||
• Bioactivity: inhibitor of microtubule assembly | ||
Phakellistatins 3 and 13 ![]() |
Van Vranken et al.111 | • 3 Steps from linear precursor synthesized on solid phase |
• Non-racemic | ||
• Bioactivity: cytotoxic against the P388 cell line | ||
(−)-Aurilide ![]() |
Suenaga and Kigoshi et al.112 | • 16 Steps |
• Non-racemic | ||
• Bioactivity: cytotoxic | ||
Eurypamide A ![]() |
Nishiyama et al.113 | • 19 Steps from known compound |
• Non-racemic | ||
• Bioactivity: none described, although related compounds are known to exhibit antimicrobial, cytotoxic and enzyme inhibition activity | ||
(±)-Maculalactones A and B ![]() |
Brown and Wong114 | • 6 Steps each |
• Racemic | ||
• Bioactivity: inhibitor of marine herbivore foraging and barnacle settling | ||
(+)-Milnamide A ![]() |
Molinski et al.115 | • 8 Steps from indole |
• Non-racemic | ||
• Bioactivity: disrupts microtubule assembly during cell division and inhibits growth of cultured tumor cells | ||
Subarine ![]() |
Delfourne et al.116 | • 5 Steps from phenanthroline |
• Bioactivity: no significant antitumor activity | ||
(+)-SCH 351448 ![]() |
Lee et al.117 | • 27 Steps |
• Non-racemic | ||
• Bioactivity: activator of low-density lipoprotein receptor promoter with IC50 25 µM | ||
Xestodecalactone A ![]() |
Bringmann et al.118 | • 5 Steps |
• Non-racemic | ||
• Bioactivity: antifungal activity against Candida albicans | ||
Lepadins B, D, E and H ![]() |
Pu and Ma119 | • 21 Steps to lepadins B, E, and H |
• 20 Steps to lepadin D | ||
• Non-racemic | ||
• Bioactivity: significant in vitro cytotoxicity against several human cancer cell lines | ||
(+)-Tricycloclavulone ![]() |
Iguchi et al.120 | • 21 Steps |
• Non-racemic | ||
• Bioactivity: related compounds have antiproliferative activity against human cancer cell lines | ||
Cribostatin 6 ![]() |
Nakahara and Kubo121 | • 10 Steps from known compound |
• Bioactivity: growth inhibitor of cancer cell lines and antibacterial activity | ||
Oscillarin ![]() |
Hanessian et al.122 | • 13 Steps from dimethyl-N-BocGlyOH |
• Non-racemic | ||
• Bioactivity: potent inhibition of thrombin | ||
(+)-Eupenoxide and (+)-Phomoxide ![]() |
Mehta and Roy123 | • 9 Steps from known compound to eupenoxide |
• 8 Steps from known compound to phomoxide | ||
• Non-racemic | ||
• Bioactivity: antibiotic | ||
Basiliskamides A and B ![]() |
Panek et al.124 | • 11 Steps to each compound |
• Non-racemic | ||
• Bioactivity: potent in vivo activity against Candida albicans and Aspergillus fumigatus | ||
Myriaporones 1, 3, and 4 ![]() |
Taylor and Fleming125 and Echavarren126 | • 20 Steps from known compound (Taylor) |
• 14 Steps from known compound (Echavarren) | ||
• Non-racemic | ||
• Bioactivity: inhibits growth of a range of cancer cell lines | ||
Norrisolide ![]() |
Theodorakis et al.127 | • 24 Steps from known compound |
• Non-racemic | ||
• Bioactivity: not described | ||
6-Hydroxy-7-methoxyisoquinolinemethanol ![]() |
Saito et al.128 | • 5 Steps from known compound |
• Bioactivity: not described | ||
Plakortone E ![]() |
Hayes and Kitching129 | • 10 Steps |
• Racemic | ||
• Bioactivity: no significant antitumor activity | ||
Caulibugulones A–E ![]() |
Wipf et al.130 (A–E) and Tamagnan et al.131 (A–D) | • 2–3 Steps each from 5-hydroxyisoquinoline (Wipf) |
• 4–5 Steps (Tamangan) | ||
• Bioactivity: cytotoxic; potent inhibitors of dual specificity phosphatase Cdc25B | ||
Cyclomarin C ![]() |
Wen and Yao132 | • 21 Steps from known compound |
• Non-racemic | ||
• Bioactivity: not described | ||
Aigialomycin D ![]() |
Geng and Danishefsky133 | • 18 Steps from 2-deoxyribose |
• Non-racemic | ||
• Bioactivity: potent antimalarial and antitumor activity | ||
Kalihinol C ![]() |
Wood et al.134 | • 24 Steps |
• Non-racemic | ||
• Bioactivity: related compounds show potent antimalarial activity | ||
(6R,7S)-7-Amino-7,8-dihydro-α-bisabolene ![]() |
Kochi and Ellman135 | • 9 Steps |
• Non-racemic | ||
• Bioactivity: antimicrobial |
Compound | Reference | Compound | Reference |
---|---|---|---|
Helianane | Venkateswaran and Sabui136 | (±)-Pseudopterosins A–F and K–L aglycones | Harrowven and Tyte137 |
(−)-Salicylihalamides A and B | Yadav and Srihari138 | (R)-Strongylodiols A and B | Lee and Baldwin et al.139 |
Cryptophycin-24 | Tripathy and Georg140 | (+)-Cylindricine C | Kibayashi et al.141 |
(−)-Agelastatin A | Hale et al.142 | Siphonarienolone | Negishi et al.143 |
(+)-Anatoxin-a | Brenneman and Martin144 | Haliclorensin and isohaliclorensin | Huang et al.145 |
Lamellarin G trimethyl ether | Handy et al.146 | (−)-Callystatin A | Langille and Panek147 |
Bacillariolide III | Suh et al.148 | Macrosphelides A, B, and E | Nemoto et al.149 |
Caulersin | Bergman et al.150 | (−)-Myclamaide A | Trost et al.151 |
Octalactins A and B | Burton and Holmes et al.152 | (+)-Tanikolide | Borhan and Schomaker153 |
(+)-Pinnatoxin A | Inoue and Hirama et al.154 | Lamellarins K and L | Ruchirawat et al.155 |
(+)-Tanikolide | Masaki et al.156 | 6-Chlorohyellazole | Knölker et al.157 |
Lyngbic Acid | Braddock and Matsuno158 | Hapalosin | Mandai et al.159 and Palomo et al.160 |
Pulchellalactam | Takabe et al.161 | (−)-Dysibetaine | Wardrop and Burge162 |
Bistratamide G | Shin et al.163 | Pulchellalactam | Mangaleswaran and Argade164 |
(−)-Pateamine A | Pattenden et al.165 | (+)-Kalkitoxin | White et al.166 |
(±)-Pinnaic acid and (±)-tauropinnaic acid | Christie and Heathcock167 | (−)-Doliculide | Hanessian et al.168 |
(−)-Flustramine B | MacMillan et al.169 | Peridinin | Katsumura et al.170 |
Petrobactin | Phanstiel et al.171 | (−)-Cycloepoxydon | Mehta and Islam172 |
(−)-α-Kainic Acid | Hoppe and Martinez173 | (−)-Apicularen A | Rizzacasa et al.174 |
(±)-Halichlorine and (±)-pinnaic Acid | Kibayashi et al.175 | (+)-Brasilenyne | Denmark and Yang176 |
Callipeltoside A | Panek and Huang177 | (+)-Acanthodoral | Zhang and Koreeda178 |
(±)-Dibromophakellstatin | Austin et al.179 | Hamigerans A, B and E | Nicolaou et al.180 |
7-Epicyclindrospermopsin | Looper and Williams181 |
This journal is © The Royal Society of Chemistry 2006 |