Abstract
The first asymmetric total synthesis and validation of the structural assignment of des-thiomethyllooekeyolide A (3) is described, featuring a Shiina macrolactonization and a late-stage pyran−hemiketal formation. The eight stereogenic centers of the C16-polyketide chain were installed by sequential aldol and crotylation reactions.
Graphical Abstract

Coral diseases have significantly influenced coral reef ecosystems around the world in the past decades,1 and black band disease (BBD) was the first coral disease observed on reefs of Belize in the western Caribbean by Antonius in 1973.2 BBD is one of the most pervasive and devastating coral diseases,3 and it consists of a cyanobacteria-dominated, sulfide-rich microbial mat. The mat migrates through coral colonies and degrades coral tissue,4 leaving behind the bare coral skeleton. Research indicates that the predominant cyanobacterium in the group that causes BBD disease is Roseofilum reptotaenium.5
Paul and coworkers in 2019 examined secondary metabolites present in the black band layer in situ during BBD from various geographic regions of Caribbean and Pacific coral reefs in an effort to find chemical signals or cytotoxins that could be crucial in the proliferation of R. reptotaenium and the BBD layer.6 From more than a dozen field-collected BBD samples, they obtained looekeyolide A (1) and looekeyolide B (2) (Figure 1) as significant components (∼1 mg g−1 dry wt). Further research indicates that looekeyolide B (2) is an autoxidation product of looekeyolide A (1) and may play a biological role in reducing H2O2 and other reactive oxygen species. These reactive species may occur in the BBD layer as it overgrows and destroys coral tissue.6
Figure 1.
Structures of looekeyolides A and B, and des-thiomethyllooekeyolide A
The cyanobacterium probably produces looekeyolide A (1), which was only stable when exposed to helium gas..6 The stable looekeyolide B (2), on the other hand, exists as a mixture of R and S sulfoxide diastereomers. In order to establish the structure of looekeyolides A and B, looekeyolide B (2) underwent semi-synthetic desulfurization using Raney-Ni to produce. des-thiomethyllooekeyolide A (3). The absolute configuration of all stereogenic centers of des-thiomethyllooekeyolide A (3) was determined using X-ray crystallography and HPLC analysis of its hydrolysate6 (Figure 1).
Structurally, des-thiomethyllooekeyolide A (3) possesses a 20-membered macrocycle that is composed of a highly modified peptidic unit containing a D-leucine, (S)-2-hydroxybutanoic acid, and a C16-polyketide substructure.. The highly oxygenated and methylated C16-polyketide possesses an unstable pyran−hemiketal with eight stereogenic centers and one of which is quaternary. Having a long interest in assignment/revision of marine natural products by total synthesis,7 we initiated the synthetic endeavor of this novel macrocyclic natural product to confirm its structural configuration. Herein, we disclose the first asymmetric total synthesis of des-thiomethyllooekeyolide A (3), enabling the confirmation of its structure assignment.
Our retrosynthetic analysis for des-thiomethyllooekeyolide A (3) is illustrated in Scheme 1. We elected to disconnect the target molecule at two ester linkages to afford amide 4 and polypropionate segment 5. Because of the labile nature of the pyran−hemiketal function, we planned to construct this moiety at the late stage of the synthesis following the Shiina macrolactonization.8 Thus, the C1−C16 segment 5 would be convergently accessible from the aldol coupling of methyl ketone 6 and aldehyde 7. The C1−C10 segment 6 composed of five stereogenic centers would be constructed through a sequence of stereoselective aldol and crotylation reactions. In addition, the C11−C16 subunit 7 could be easily obtained by the Titanium enolate anti-aldol reaction developed by Ghosh et al.9
Scheme 1.
Retrosynthetic Analysis of 3
The construction of des-thiomethyllooekeyolide A (3), commenced with the stereoselective construction of the primary synthetic subtarget (6) (Scheme 2). We elected the titanium-mediated anti-aldol reaction of lactates bearing SuperQuats chiral oxazolidine-2-ones (9), developed by Kobayashi et al10 as a powerful, stereocontrolled entry to the masked chiral diol intermediate. As a result,, the commercial TBS-protected aldehyde 8 was reacted with the titanium enolate produced by treating the SuperQuats derivative 9 with LDA followed by the addition of TiCl(O-i-Pr)3, to afford the aldol adduct 10 in 89% yield with dr >10:1. Subsequent conversion to the fully protected tetraol 11 was accomplished by sequential reductive cleavage of the chiral auxiliary with NaBH4 in THF-H2O, transformation of the primary hydroxyl to its TBDPS ether, and O-methylation of the secondary alcohol (73% overall yield). The subsequent chemoselective removal of the primary TBS ether from 11 to furnish the corresponding primary alcohol was accomplished using catalytic PPTS in ethanol. Treatment of this alcohol with catalytic TEMPO and an excess of trichloroisocyanuric acid (TCCA)11 gave rise to an aldehyde and then immediately subjected to asymmetric Brown crotylation12 to afford the corresponding homoallylic alcohol in 78% yield as a chromatographically separable 4:1 mixture of diastereomers, and the major diastereomer 12 was easily isolated and characterized. Treatment of 12 with PMB-Br and NaH in the presence of TBAI resulted in the detection of just a minute quantity of the unwanted product 13, despite prolonged reaction durations or elevated temperatures. Pleasingly, the PMB ether 13 could be obtained in 74% yield by treatment with KHMDS and PMBBr in the presence of triethylamine at −50 °C, followed by gradual warming to room temperature.13
Scheme 2.
Synthesis of the C1−C10 Segment 6
With intermediate 13 in hand, the next stage was set for the installation of the chiral hydroxyl group at the C3 position. Oxidative cleavage of the terminal olefin 13 with OsO4 and NaIO414 followed by Brown allylation15 delivered the desired product 15 in moderate overall yield with low selectivity (dr < 2:1). We therefore considered alternatives, and to our delight, we found that the diastereoselectivity of the formation of homoallylic alcohol 15 could be increased to 10:1 through a four-step sequence involving oxidative cleavage of the terminal olefin, Grignard addition, Dess-Martin periodinane oxidation,16 and NaBH4 reduction. The relative configuration of the newly generated hydroxyl group at C-3 was later confirmed by the nOe experiments of the PMP acetal 16, which was obtained from 15 by treatment with DDQ in anhydrous DCM solution. Methylation of 15 followed by deprotection of the TBDPS group gave rise to the primary alcohol 17 in 93% yield over two steps. Oxidation of 17 with Dess-Martin periodinane followed by Pinnick-Lindgren-Kraus oxidation17 to the corresponding acid and subsequent coupling to Weinreb’s amine gave 18 in 67% overall yield over three steps. DDQ oxidation effected the removal of the PMB ether, and the resultant alcohol was then reprotected as its TBS ether, and this was followed by addition of methyl lithium to furnish segment 6 in 91% yield (Scheme 2). It is worth noting that to protect 12 with a TBS group at an early stage proved to be completely unworkable in this case as the TBS group underwent 1,3-migration to the C3 hydroxyl group under the subsequent NaH/MeI methylation condition. All other methylation conditions that were tested either did not affect the starting material or resulted in decomposition of the substrate.18
With a reliable route to useful quantities of segment 6 in hand, we then turned our attention to the aldol coupling of 6 with aldehyde 719 (Scheme 3). Surprisingly, the stereoselectivity of this aldol coupling reaction was affected by the reaction conditions. Using LDA or LiHMDS as a base in THF/HMPA resulted in an overwhelming preference for the unwanted diastereomer (20b), with a diastereoselectivity of 20a:20b = 1:10 to 1:15. Literature precedent20 suggested that the selectivity varied depending on the counterion associated with the base employed and was also affected by the chiral substrates. To our delight, when the NaHMDS or KHMDS was employed as a base, the diastereoselectivity was reversed (20a:20b = 2:1 or 4:1, respectively), with the best selectivity for the Felkin diastereomer (20a) being obtained with the potassium enolate. The stereochemical outcomes of the base-promoted aldol coupling reaction upon formation of 20a and 20b, might be interpreted by Felkin-Anh model and Cram-chelate model,20a respectively. Although the minor isomer could not be separated from the diastereomeric mixture of the aldol reaction by silica gel column chromatography at this stage, they were easily separated after the C-11 hydroxyl was protected as its TBS ether followed by removal of the PMB protecting group to furnish 5 in 58% overall yield over two steps. This compound, however, proved to be relatively unstable upon dissolution in NMR grade deuterated chloroform since it was completely transferred to hemiacetal 21 during 13C NMR experiments.21 The NMR spectroscopic analysis of hemiacetal 21 revealed its precise stereochemical configurations of the compound, including the NOE correlation between the methyl group and its adjacent proton, as indicated on its structure (see 21, Scheme 3). Additionally, this study also confirmed the stereochemistry of aldol product 20a.
Scheme 3.
Total Synthesis of des-thiomethyllooekeyolide A (3)
Having established the route to C1−C16 segment 5, we next explored the esterification and macrolactonization to complete the total synthesis. Esterification of alcohol 5 with acid 4 was initially attempted using Yamaguchi22 or Keck protocol;23 However, these reaction conditions also caused formation of the hemiacetal 21 prior to esterification. To circumvent this problem, the more reactive acid chloride 22,24 was employed, and the reaction was performed in the presence of NMM, a catalytic amount of DMAP in dichlorometane, which resulted in 89% yield of ester 23 with minimal hemiacetal formation. Removal of the N-Fmoc group of 23 with diethylamine (DEA) followed by condensation with hydroxyl acid 24 mediated by (7-Azabenzotriazol-1-yloxy)tripyrrolidinophosphonium hexa-fluorophosphate (PyAOP)25 gave rise to peptide 25 in 61% yield. The osmium-mediated dihydroxylation (catalytic in K2OSO4•2H2O) of the terminal alkene was followed by diol cleavage with silica-gel supported sodium periodate26 to afford the desired aldehyde which was immediately oxidized to the carboxylic acid 26 in 63% overall yield using Pinnick-Lindgren-Kraus oxidation conditions. Macrolactonization of seco-acid 26 was initially tried under the popular Yamaguchi conditions. However, in this case, the yield never exceeded 41%. In contrast, macrolactonization with the more reactive Shiina anhydride (2-Methyl-6-nitrobenzoic anhydride, MNBA) in the presence of DMAP induced lactone formation minimizing the formation of side-products and afforded the desired macrolide in 65% yield. Finally, the first total synthesis of des-thiomethyllooekeyolide A (3) was achieved through a one-pot process involving deprotection of silyl with concomitant hemiacetal formation by exposure of the macrolactone to hydrochloric acid, followed by hydrogenolytic removal of the benzyl ether. The proton and carbon spectra of fully synthetic des-thiomethyllooekeyolide A were identical to those of naturally derived material,6 though the spectra of the synthetic material were cleaner. The optical rotation of [α]20D +17.0 (c 0.20, MeOH) was identical in sign to but slightly lower than the reported value of [α]25D +29.3 (c 0.05, MeOH), therefore establishing the absolute stereochemistry as shown for 3.6
In summary, the first asymmetric total synthesis of des-thiomethyllooekeyolide A (3) was achieved in thirty steps over the longest linear sequence and in a 0.3% overall yield starting from readily available aldehyde 8. The key features of the synthesis include a Shiina macrolactonization and a late-stage hemiacetal formation. The eight stereogenic centers of the C16-polyketide chain were introduced stereoselectively through a sequence of aldol and crotylation reactions. The synthetic route we developed is general, efficient, and will allow for the rapid synthesis of other members of the looekeyolide family..
Supplementary Material
ACKNOWLEDGMENT
This work was supported by the Guangdong Department of Education (2021ZDJS097), Guangdong Basic and Applied Basic Research Foundation (2021A1515010188), and the National Institutes of Health (R01CA172310).
Footnotes
Supporting Information Placeholder
Supporting Information
The Supporting Information is available free of charge on the ACS Publications website. Experimental procedures and characterization data for all new compounds (PDF)
The authors declare no competing financial interest.
REFERENCES
- (1).Hoegh-Guldberg O; Mumby PJ; Hooten AJ; Steneck RS; Greenfield P; Gomez E; Harvell CD; Sale PF; Edwards AJ; Caldeira K; Knowlton N; Eakin CM; Iglesias-Prieto R; Muthiga N; Bradbury RH; Dubi A; Hatziolos ME, Coral Reefs Under Rapid Climate Change and Ocean Acidification. Science 2007, 318, 1737–1742. [DOI] [PubMed] [Google Scholar]
- (2).Antonius A, New observations on coral destruction in reefs, Abstract. In Tenth Meeting of the Association of Island Marine Laboratories of the Caribbean: University of Puerto Rico, Mayaguez, PR., 1973; pp 3–3. [Google Scholar]
- (3).Kathryn PS; James WP; Cecilia T, Disease and immunity in Caribbean and Indo-Pacific zooxanthellate corals. Mar. Ecol. Prog. Ser 2004, 266, 273–302. [Google Scholar]
- (4).Richardson LL; Miller AW; Broderick E; Kaczmarsky L; Gantar M; Stanić D; Sekar R, Sulfide, microcystin, and the etiology of black band disease. Dis. Aquat. Org 2009, 87, 79–90. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (5).Casamatta D; Stanić D; Gantar M; Richardson LL, Characterization of Roseofilum reptotaenium (Oscillatoriales, Cyanobacteria) gen. et sp. nov. isolated from Caribbean black band disease. Phycologia 2012, 51, 489–499. [Google Scholar]
- (6).Gunasekera SP; Meyer JL; Ding Y; Abboud KA; Luo D; Campbell JE; Angerhofer A; Goodsell JL; Raymundo LJ; Liu J; Ye T; Luesch H; Teplitski M; Paul VJ, Chemical and Metagenomic Studies of the Lethal Black Band Disease of Corals Reveal Two Broadly Distributed, Redox-Sensitive Mixed Polyketide/Peptide Macrocycles. J. Nat. Prod 2019, 82, 111–121. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (7).(a) Guo Y-A; Zhao M; Xu Z; Ye T, Total Synthesis and Stereochemical Assignment of Actinoranone. Chem. Eur. J 2017, 23, 3572–3576. [DOI] [PubMed] [Google Scholar]; (b) Zhou J; Gao B; Xu Z; Ye T, Total Synthesis and Stereochemical Assignment of Callyspongiolide. J. Am. Chem. Soc 2016, 138, 6948–6951. [DOI] [PubMed] [Google Scholar]; (c) Liu J; Wang L; Zhang J; Xu Z; Ye T, The total synthesis and stereochemical assignment of scytonemin A. Chem. Comm 2016, 52, 1002–1005. [DOI] [PubMed] [Google Scholar]; (d) Qu S; Chen Y; Wang X; Chen S; Xu Z; Ye T, Total synthesis of largamide B. Chem. Comm 2015, 51, 2510–2513. [DOI] [PubMed] [Google Scholar]; (e) Lei H; Yan J; Yu J; Liu Y; Wang Z; Xu Z; Ye T, Total Synthesis and Stereochemical Reassignment of Mandelalide A. Angew. Chem. Int. Ed. Engl 2014, 53, 6533–6537. [DOI] [PubMed] [Google Scholar]; (f) Dai L; Chen B; Lei H; Wang Z; Liu; Xu Z; Ye T, Total synthesis and stereochemical revision of lagunamide A. Chem. Comm 2012, 48, 8697–8699. [DOI] [PubMed] [Google Scholar]; (g) Gao X; Liu Y; Kwong S; Xu Z; Ye T, Total Synthesis and Stereochemical Reassignment of Bisebromoamide. Org. Lett 2010, 12, 3018–3021. [DOI] [PubMed] [Google Scholar]
- (8).Shiina I; Ibuka R; Kubota M, A New Condensation Reaction for the Synthesis of Carboxylic Esters from Nearly Equimolar Amounts of Carboxylic Acids and Alcohols Using 2-Methyl-6-nitrobenzoic Anhydride. Chem. Lett 2002, 31, 286–287. [Google Scholar]
- (9).Ghosh AK; Onishi M, Synthesis of Enantiomerically Pure Anti-Aldols: A Highly Stereoselective Ester-Derived Titanium Enolate Aldol Reaction. J. Am. Chem. Soc 1996, 118, 2527–2528. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (10).Murata Y; Kamino T; Hosokawa S; Kobayashi S, Stereocontrolled preparation of 1,2-diol with quaternary chiral center. Tetrahedron Lett. 2002, 43, 8121–8123. [Google Scholar]
- (11).De Luca L; Giacomelli G; Porcheddu A, A Very Mild and Chemoselective Oxidation of Alcohols to Carbonyl Compounds. Org. Lett 2001, 3, 3041–3043. [DOI] [PubMed] [Google Scholar]
- (12).(a) Brown HC; Bhat KS, Enantiomeric Z- and E-crotyldiisopinocampheylboranes. Synthesis in high optical purity of all four possible stereoisomers of .beta.-methylhomoallyl alcohols. J. Am. Chem. Soc 1986, 108, 293–294. [DOI] [PubMed] [Google Scholar]; (b) Brown HC; Bhat KS; Randad RS, Chiral synthesis via organoboranes. 21. Allyl- and crotylboration of .alpha.-chiral aldehydes with diisopinocampheylboron as the chiral auxiliary. J. Org. Chem 1989, 54, 1570–1576. [Google Scholar]
- (13).Holson EB; Roush WR, Diastereoselective Synthesis of the C(17)−C(28) Fragment (The C−D Spiroketal Unit) of Spongistatin 1 (Altohyrtin A) via a Kinetically Controlled Iodo-Spiroketalization Reaction. Org. Lett 2002, 4, 3719–3722. [DOI] [PubMed] [Google Scholar]
- (14).Yu W; Mei Y; Kang Y; Hua Z; Jin Z, Improved Procedure for the Oxidative Cleavage of Olefins by OsO4−NaIO4. Org. Lett 2004, 6, 3217–3219. [DOI] [PubMed] [Google Scholar]
- (15).Brown HC; Jadhav PK, Asymmetric carbon-carbon bond formation via .beta.-allyldiisopinocampheylborane. Simple synthesis of secondary homoallylic alcohols with excellent enantiomeric purities. J. Am. Chem. Soc 1983, 105, 2092–2093. [Google Scholar]
- (16).Dess DB; Martin JC, A useful 12-I-5 triacetoxyperiodinane (the Dess-Martin periodinane) for the selective oxidation of primary or secondary alcohols and a variety of related 12-I-5 species. J. Am. Chem. Soc 1991, 113, 7277–7287. [Google Scholar]
- (17).(a) Lindgren BO; Nilsson T, Preparation of Carboxylic Acids from Aldehydes (Including Hydroxylated Benzaldehydes) by Oxidation with Chlorite. Acta Chem. Scand 1973, 27, 888–890. [Google Scholar]; (b) Kraus GA; Taschner MJ, Model Studies for the Synthesis of Quassinoids. 1. Construction of the BCE Ring System. J. Org. Chem 1980, 45, 1175–1176. [Google Scholar]; (c) Bal BS; Childers WE; Pinnick HW, Oxidation of α,β-un saturated aldehydes. Tetrahedron 1981, 37, 2091–2096. [Google Scholar]
- (18).Use of MeOTf/2,6-Di-tert-butyl-4-methylpyridine or Ag2O/MeI conditions only resulted in the recovery of starting materials, and Me3OBF4/Proton sponge condition led to decompositon.
- (19).For the synthesis of aldehyde 7, please see SI for details.
- (20).(a) Evans DA; Dart MJ; Duffy JL; Yang MG, A Stereochemical Model for Merged 1,2- and 1,3-Asymmetric Induction in Diastereoselective Mukaiyama Aldol Addition Reactions and Related Processes. J. Am. Chem. Soc 1996, 118, 4322–4343. [Google Scholar]; (b) Mengel A; Reiser O, Around and beyond Cram’s Rule. Chem. Rev 1999, 99, 1191–1224.; Selective examples of aldol coupling of PMB-protected β−hydroxy aldehyde with methyl ketone, see: [DOI] [PubMed] [Google Scholar]; (c) Roush WR; Bannister TD; Wendt MD; Jablonowski JA; Scheidt KA, Studies on the Synthesis of Bafilomycin A1: Stereochemical Aspects of the Fragment Assembly Aldol Reaction for Construction of the C(13)−C(25) Segment. J. Org. Chem 2002, 67, 4275–4283. [DOI] [PubMed] [Google Scholar]; (d) Paquette LA; Duan M; Konetzki I; Kempmann C, A Convergent Three-Component Total Synthesis of the Powerful Immunosuppressant (−)-Sanglifehrin A. J. Am. Chem. Soc 2002, 124, 4257–4270. [DOI] [PubMed] [Google Scholar]; (e) Paterson I; Delgado O; Florence GJ; Lyothier I; Scott JP; Sereinig N, 1,6-Asymmetric Induction in Boron-Mediated Aldol Reactions: Application to a Practical Total Synthesis of (+)-Discodermolide. Org. Lett 2003, 5, 35–38. [DOI] [PubMed] [Google Scholar]; (f) Roush WR; Lane GC, Studies on the Synthesis of Tedanolide: Synthesis of the C(5)−C(21) Segment via a Highly Stereoselective Fragment Assembly Aldol Reaction of a Chiral β,γ-Unsaturated Methyl Ketone. Org. Lett 1999, 1, 95–98. [DOI] [PubMed] [Google Scholar]; (g) Ehrlich G; Hassfeld J; Eggert U; Kalesse M, The Total Synthesis of (+)-Tedanolide. J. Am. Chem. Soc 2006, 128, 14038–14039; [DOI] [PubMed] [Google Scholar]; (h) Hassfeld J; Eggert U; Kalesse M, Synthesis of the C1-C17 Macrolactone of Tedanolide. Synthesis 2005, 2005, 1183–1199. [Google Scholar]
- (21).This transformation could happen even in ethyl acetate.
- (22).Inanaga J; Hirata K; Saeki H; Katsuki T; Yamaguchi M, A Rapid Esterification by Means of Mixed Anhydride and Its Application to Large-ring Lactonization. Bull. Chem. Soc. Jpn 1979, 52, 1989–1993. [Google Scholar]
- (23).Boden EP; Keck GE, Proton-transfer steps in Steglich esterification: a very practical new method for macrolactonization. J. Org. Chem 1985, 50, 2394–2395. [Google Scholar]
- (24).Acid chloride 22 was prepared in situ from the corresponding acid by treating with oxalyl chloride and cat. DMF.
- (25).Albericio F; Cases M; Alsina J; Triolo SA; Carpino LA; Kates SA, On the use of PyAOP, a phosphonium salt derived from HOAt, in solid-phase peptide synthesis. Tetrahedron Lett. 1997, 38, 4853–4856. [Google Scholar]
- (26).Zhong Y-L; Shing TKM, Efficient and Facile Glycol Cleavage Oxidation Using Improved Silica Gel-Supported Sodium Metaperiodate. J. Org. Chem 1997, 62, 2622–2624. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.




