Synthesis and Potent Antimalarial Activity of Kalihinol B - Journal of

Mar 27, 2015 - Synthesis and Potent Antimalarial Activity of Kalihinol B. Showing 1/2: ja5b01152_si_001.pdf. Sorry we could not load your data. figsha...
3 downloads 12 Views 376KB Size
Subscriber access provided by SETON HALL UNIV

Communication

Synthesis and Potent Antimalarial Activity of Kalihinol B Mary Elisabeth Daub, Jacques Prudhomme, Karine Le Roch, and Christopher D. Vanderwal J. Am. Chem. Soc., Just Accepted Manuscript • Publication Date (Web): 27 Mar 2015 Downloaded from http://pubs.acs.org on March 27, 2015

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Journal of the American Chemical Society is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 5

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

Synth es is a n d P o t e nt A ntima l ari a l A cti v ity o f K a l i hin o l B Mary Elisabeth Daub,a Jacques Prudhomme,b Karine Le Roch,b and Christopher D. Vanderwal*a a

b

Department of Chemistry, 1102 Natural Sciences II, University of California, Irvine, CA 92697-2025. Department of Cell Biology & Neuroscience, 900 University Avenue, University of California, Riverside, CA 92521. Supporting Information Placeholder ABSTRACT: Of the 50+ kalihinane diterpenoids re-

ported over the last 30 years, only five had been tested for antimalarial activity, in spite of the fact that kalihinol A is the most potent among the members of the larger family of antimalarial isocyanoterpenes. We have validated a strategy designed to access many of the kalihinanes with a 12-step enantioselective synthesis of kalihinol B, the tetrahydrofuran isomer of kalihinol A (a tetrahydropyran). Kalihinol B shows similarly high potency against chloroquine-resistant Plasmodium falciparum. Many sponge-derived isocyanoterpenes exhibit potent 1 antimalarial activity. In general, they are as active against drug-resistant Plasmodium falciparum (1.2–31 nM for 1–4, Figure 1) as they are against drug-sensitive parasite, and they typically show high selectivity indices (SI) for P. falciparum over human cells. The kalihinanes (3–6 are representative) are a subgroup of these isocyanoterpenes isolated primarily from sponges of the Acanthella genus and characterized by cis- or transdecalin cores bearing pendant tetrahydropyrans or tetrahydrofurans, and a variable set of isonitrile, isocyanate, isothiocyanate, formamide, hydroxyl, and chloride sub2 stituents. H H

NC

H

4

H

H CN

H

H

H

H

1 5

regioselective chlorinolysis; oxidation

NC

10

H

7

H 11 O 15 14

H

OH

O

H

O

CHO

organocatalystcontrolled sequence

14 15

Cl

7

HO

H

O OH

NC

H

NC

Of the more than 50 kalihinanes reported to date, the antimalarial activities of only five members have been reported, and the isonitrile function appears to be critical 1 for high activity. The kalihinanes also demonstrate cy2e,p 2a–c,f,o 2e totoxic, antibacterial, antifungal, anthelmin2d 2j,k,p tic, and antifouling activities. In light of the fact that kalihinol A (3) has been known since 1998 to be an extremely potent antiplasmodial agent (EC50 = 1.2 nM against the drug-resistant FCR-3 strain, the most potent 1c of all isocyanoterpenes tested to date), it is surprising that so few kalihinanes have been studied for antimalarial properties. Laboratory syntheses have been reported by 3 4 Wood (kalihinol C ), Yamada (kalihinene X ), and Mi5a 5b 5b yaoka (kalihinols A and Y, 10-epi-kalihinol I, ); however, no antimalarial evaluations were undertaken 6,7 with the synthetic material. With the increasing threat from drug-resistant P. falciparum strains, including those that do not succumb to artemisinin-based combi8 nation therapies, research into any compounds with documented antimalarial activity and selectivity is clearly valuable.

H 14

8

7: geraniol epoxide H

11 O

NC

9

H

Cl

O

Cl

1: an amphilectane isonitrile W2: 31 nM, SI: 340 H

2: 7,20-diisocyanoadociane W2: 13 nM, SI: 1000

NC

3: kalihinol A FCR-3: 1.2 nM SI: 317

NC

H

H

NC

HO

NC

H

O

H

HO

O H

NC

H

Cl

H

HO

11 O

H 14 15

NC

4: kalihinene FCR-3: 10 nM SI: 4

H NC

H

5: kalihinol B

O H

Cl

5: kalihinol B 6: kalihinol C untested, along with >40 other kalihinanes

Figure 1. Representative antimalarial isocyanoterpenes (W2 and FCR-3 are drug-resistant strains of P. falciparum; SI = Selectivity Index with respect to mammalian cells)

H

annulation or cycloaddition

H O

H

7

H

steps

H

H Cl

10: precursor to THF kalihinanes

Scheme 1. Strategy to access the THF kalihinanes There is a wealth of structural diversity in the kali2 hinane group, and a synthesis that could address multiple naturally occurring kalihinanes as well as synthetic analogues would permit a thorough evaluation of structure-activity relationships. The relative scarcity of mate2 rial from natural sources and the length of the reported chemical approaches to the kalihinanes (3 in ca. 37 steps

ACS Paragon Plus Environment

Journal of the American Chemical Society from 11: O

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

O

H Ph

O

OR

12

OH

O O

O oxone (50%, 89:11 er)

11: R = H 13: R = Ac

from 13: 1. AD-mix-! 2. MsCl, pyr, then K2CO3, MeOH (52%, 2 pots 97:3 er)

OH

MnO2 CH2Cl2

Li2CuCl4 TBSCl (71% + 16% regioisomer)

OHC

OH

H

O

Ph

·TFA N N H 18

N 16 OMe H H

(48–58% 6:1 dr)

OHC 7

MVK, 9 d

11 O

O

hexanes

H

(1:1 at C11 25–30% of 17a)

O

H

(75%)

11 O

O

H

H

Cl

Cl

7

Page 2 of 5

15: incorrect C11 configuration

14

17

9

Cl

1. 19

Cl

2. KOt-Bu t-BuOH (60%)

Cl Li

(2-thienyl)Cu(CN)Li (83%) H 4

HO

NC

H

1. TFAA 2. TMSCN, Sc(OTf)3

5

H NC

H O H

O

H

3. TBAF (18%, 3 steps)

OH

H

H

acetone oxone, NaHCO3

O

(86%)

H

Cl

H

H

+

O

H

OH

MeMgBr

O H

Cl

22

H

H

H

H

Cl

5: kalihinol B 12 steps LLS

OH

20

Cl

21: X-ray

(41% of 20, 44% of 21, from mixture of 10)

H

O

H O H Cl

10: 1.3:1 cis:trans

a

Scheme 2. Enantioselective synthesis of kalihinol B. Yield range shown when material of 97:3 er was used; when material of 89:11 er was used, the yield was 23–25% of 17. 5a

longest linear sequence, LLS; and 6 in ca. 24 steps 3b LLS) were likely impediments to a systematic evaluation of biological activity. From the limited set of bio1c,6 logical testing reported to date, it is clear that the presence of two isonitriles correlates with high antiplasmodial activity, but the specific location of the isonitriles is not critical (compare 3 and 4). To elucidate structure-activity relationships for this family, we aimed to develop a concise strategy that would be applicable to many members of the kalihinane family. As proof of principle, we report in this preliminary communication a short synthesis (12 steps LLS) of (+)-kalihinol B (5) and disclose its potent antimalarial activity. 9 We viewed the “attached ring motif” —in this case the oxygen heterocycle attached to the decalin with its attendant vicinal C7/C11 stereogenic centers—as a significant challenge that could arguably be viewed as the cause of the length of previous kalihinane syntheses. These earlier successes took advantage of a biosynthetically relevant C14(C15)–O bond formation to generate the heterocycle. We viewed the C11–O bond as a strategic disconnection in our plan to address the attached ring problem (Scheme 1). α,β-Unsaturated aldehyde 8, accessed by regioselective chlorinolysis and oxidation of geraniol epoxide (7), will engage in an oxaMichael/Robinson annulation sequence to generate cyclohexenone 9 with the vicinal C7–C11 stereogenic centers. Precedent suggested that pyrrolidine-based secondary amine catalysts of the same enantiomeric series could be used to control the C11 center in the course of 10 the oxa-Michael addition as well as the C7 center in 11 the Michael addition step of the Robinson annulation, thereby presenting the possibility of using a single catalyst for this sequence. Either annulation or cycloaddition processes would permit access to the cis-decalin in key intermediate 10, which should be readily elaborated to 2c kalihinane natural products, including kalihinol B (5).

A key aspect of the endgame involved the installation of both isonitriles in one step. For this finish, we aimed to take advantage of the recent critical advance from the Shenvi laboratory in the stereoinvertive displacement of 12 tertiary trifluoroacetates by TMSCN, a method developed in the context of their exceptional synthesis of am13 philectane 1. Our strategy was executed as illustrated in Scheme 2. Shi epoxidation of geraniol (11) afforded 6,714 epoxygeraniol (7) with moderate enantioselectivity. Alternatively, a two-pot procedure using a Sharpless dihydroxylation of geranyl acetate (13) provided 7 in 15 97:3 er. Regioselective epoxide chlorinolysis using dilithium tetrachlorocuprate in the presence of tert16 butyldimethylchlorosilane afforded desired chlorohydrin 14 in 71% yield along with 16% of its regioisomer (not shown). Allylic oxidation of 14 triggered spontaneous stereoselective cyclization of the secondary alcohol, affording tetrahydrofuran 15 with predominantly the incorrect C11 configuration. However, exposure of 15 to 11 the Gellman prolinol catalyst 16 induced the desired intermolecular Michael reaction, resulting in a 1:1 mixture of Michael adducts epimeric at C11 from which ca. 30% of the desired stereoisomer 17 could be isolated. The catalyst effectively controlled the configuration at C7, and facilitated a partial correction of the C11 center, clearly indicating that the oxa-Michael reaction is re10,17 versible under these reaction conditions. This partial stereochemical correction is a noteworthy victory, and suggests that conditions might eventually be found to 17 favor isomer 17. Complete Robinson annulation could not be accomplished in one step owing to decomposition; aldol condensation of 17 was catalyzed with dia18,19 mine 18 to afford enone 9. Keto-aldehyde 17 proved sensitive to basic conditions, and more typical aldol 20 condensation conditions with hydroxide/alkoxide bases 21,22 23 were less reliable. A Piers-type annulation onto

ACS Paragon Plus Environment

Page 3 of 5

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

enone 9 was accomplished via fully diastereoselective conjugate addition of the cuprate derived from 19 and subsequent intramolecular alkylation using potassium tert-butoxide, delivering trans-decalin 10 as the minor component of a 1:1.3 mixture that is not easily re24 solved. Although this stereochemical outcome is not ideal, the cis-decalin is, in principle, useful for the synthesis of the kalihinenes (see 4). Trans-decalin 20 and its cis-isomer 21 can be accessed in pure form after high-yielding, stereoselective nucleophilic methylation of the mixture of isomers of 10. While trans-decalin 20 failed to crystallize, X-ray crystal structure analysis of cis-decalin 21 indicated that the stereochemistry of the attached ring motif matched that found in the natural product targets. Epoxidation of trans-decalin 20 using dimethyldioxirane (DMDO) yielded the desired α-epoxide (22) as a single diastereomer. At this stage, we looked to Shenvi’s conditions for invertive tertiary trifluoroacetate displacement by trimethylsilyl cyanide for the installation of the isoni12,13 triles. Conversion of the tertiary alcohol of 22 to its trifluoroacetate ester preceded exposure to scandium(III) triflate and trimethylsilyl cyanide, which effected rapid isocyanosilylation of the epoxide and much slower invertive displacement of the trifluoroacetate. Desilylation of the tertiary TMS ether gave kalihinol B (5). The isonitrile introduction result is noteworthy for the two aspects of regiocontrol in the epoxide isocyanolysis. While the Fürst–Plattner principle predicts trans-diaxial nucleophilic opening of the epoxide at the least hindered C5 position, it was not clear whether the isonitrile or 25 nitrile isomer would predominate. Success in this reaction, even with the relatively low overall yield (due largely to competitive elimination of the axial tertiary trifluoroacetate), eliminates a significant number of steps and is critical to the short overall sequence, in which the natural product is obtained in 12 or 13 steps from geraniol. Synthetic (+)-kalihinol B was subjected to the SYBR 26 Green parasite proliferation assay to test for antiplasmodial activity against wild-type P. falciparum (3D7 strain) and chloroquine-resistant parasite (Dd2 strain). It exhibited potent antimalarial activity against each strain (IC50 = 8.4 nM for 3D7 and 4.6 nM for Dd2). We have developed a concise, enantioselective synthesis of (+)-kalihinol B and shown that it is potently toxic toward a strain of drug-resistant malaria parasite. Our synthesis also proves the feasibility of our approach to this family of complex isocyanoterpenes. In principle, this design also permits access to the tetrahydropyrancontaining kalihinanes (such as kalihinol A) via the regioisomer of chlorohydrin 14, a possibility that we are actively investigating. While some transformations in our current synthesis would benefit from improved selectivity, even in its current form our achievement is the most concise synthesis of any of the kalihinanes because nearly all steps generate key C–O, C–Cl, C–C, or C–N bonds. We are currently optimizing the unselective reactions and implementing our strategy in the synthesis of

other members of the family and unnatural analogues in order to gain greater insights into the structural requirements for antiplasmodial activity. A S S O CI AT ED C O N T ENT Supporting Information. Experimental procedures and characterization for all new compounds, X-ray crystallographic structure and information for 21, and entire reference 8b. This material is available free of charge via the Internet at http://pubs.acs.org.

AU THOR INF ORM ATI ON C or resp o n di n g Au t h or

Christopher D. Vanderwal: [email protected] Fu n di n g S o u r c es

No competing financial interests have been declared.

A C KN OW L EDGM ENT We dedicate this work to Profressor Stephen Hanessian on th the occasion of his 80 birthday. We acknowledge the NIH for partial support of this work through grant GM-086483 (C.D.V) and AI-85077 (K.L.R.). M.E.D. was supported by an Allergan Graduate Fellowship. K.L.R. acknowledges support from the UCR AES/CE (Project # CA-R-NEU5048-H). We thank Professor Ryan Shenvi (Scripps) for open communication about their related work, and Professor Sergey Pronin (UC Irvine) for helpful discussions relating to the isonitrile introduction.

RE F EREN C E S 1. Angerhofer, C. K.; Pezzuto, J. M.; König, G. M.; Wright, A. D.; Sticher, O. J. Nat. Prod. 1992, 55, 1787–1789. (b) Wright, A. D.; König, G. M.; Angerhofer, C. K.; Greenidge, P.; Linden, A.; Desqueyroux-Faúndez, R. J. Nat. Prod. 1996, 59, 710–716. (c) Miyaoka, H.; Shimomura, M.; Kimura, H.; Yamada, Y. Tetrahedron 1998, 54, 13467–13474. 2. (a) Chang, C. W. J.; Patra, A.; Roll, D. M.; Scheuer, P. J.; Matsumoto, G. K.; Clardy, J. J. Am. Chem. Soc. 1984, 106, 4644–4646. (b) Patra, A.; Chang, C. W. J.; Scheuer, J. J.; Van Duyne, G. D.; Matsumoto, G. K.; Clardy, J. J. Am. Chem. Soc. 1984, 106, 7981– 7983. (c) Chang, C. W. J.; Patra, A.; Baker, J. A.; Scheuer, P. J. J. Am. Chem. Soc. 1987, 109, 6119–6123. (d) Omar, S.; Albert, C.; Fanni, T.; Crews, P. J. Org. Chem. 1988, 53, 5971–5972. (e) Fusetani, N.; Yasumuro, K.; Kawai, H.; Natori, T.; Brinen, L.; Clardy, J. Tetrahedron Lett. 1990, 31, 3599–3602. (f) Alvi, K. A.; Tenenbaum, L.; Crews, P. J. Nat. Prod. 1991, 54, 71–78. (g) Trimurtulu, G.; Faulkner, D. J. J. Nat. Prod. 1994, 57, 501–506. (h) Braekman, J. C.; Daloze, D.; Gregoire, F.; Popov, S.; Van Soest, R. Bull. Soc. Chim. Belg. 1994, 103, (5-6) 187–191. (i) Rodríguez, J.; Nieto, R. M.; Hunter, L. M.; Diaz, M. C.; Crews, P.; Lobkovsky, E.; Clardy, J. Tetrahedron 1994, 50, 11079–11090. (j) Okino, T.; Yoshimura, E.; Hirota, H.; Fusetani, N. Tetrahedron Lett. 1995, 36, 8637–8640. (k) Hirota, H.; Tomono, Y.; Fusetani, N. Tetrahedron 1996, 52, 2359– 2368. (l) Wolf, D.; Schmitz, F. J. J. Nat. Prod. 1998, 61, 1524–1527. (m) Shimomura, M.; Miyaoka, H.; Yamada, Y. Tetrahedron Lett. 1999, 40, 8015–8017. (n) Clark, R. J.; Stapleton, B. L.; Garson, M. J. Tetrahedron 2000, 56, 3071–3076 (o) Bugni, T. S.; Singh, M. P.; Chen, L.; Arias, D. A.; Harper, M. K.; Greenstein, M.; Maiese, W. M.; Concepción, G. P.; Magalindan, G. C.; Ireland, C. M. Tetrahedron 2004, 60, 6981–6988. (p) Xu, Y.; Li, N.; Jiao, W.-H.; Wang, R.-P.; Peng, Y.; Qi, S.-H.; Song, S.-J.; Chen, W.-S.; Lin, H.-W. Tetrahedron 2012, 68, 2876–2883.

ACS Paragon Plus Environment

Journal of the American Chemical Society

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

3. (a) White, R. D.; Wood, J. L. Org. Lett. 2001, 3, 1825–1827. (b) White, R. D.; Keaney, G. F.; Slown, C. D; Wood, J. L. Org. Lett. 2004, 6, 1123–1126. 4. Miyaoka, H.; Shida, H.; Yamada, N.; Mitome, H.; Yamada, Y. Tetrahedron Lett. 2002, 43, 2227–2230. 5. (a) Miyaoka, H.; Abe, Y.; Sekiya, N.; Mitome, H.; Kawashima, E. Chem. Commun. 2012, 48, 901–903. (b) Miyaoka, H.; Abe, Y.; Kawashima, E. Chem. Pharm. Bull. 2012, 60, 1224–1226. 6. The Wood group tested a number of simplified analogues for antiplasmodial activity: Keaney, G. F. Ph.D. Dissertation, Yale University, 2005. 7. For an excellent recent review on marine isocyanoterpenes, see: Schnermann, M. J; Shenvi, R. A. Nat. Prod. Rep. Advance Article, published online 16 December 2014; doi: 10.1039/C4NP00109E. 8. (a) Turschner, S.; Efferth, T. Mini-Reviews Med. Chem. 2009, 9, 206–214. (b) Ashley, E. A., et al. N. Engl. J. Med. 2014, 371, 411– 423. 9. (a) Overman, L. E.; Pennington, L. D. Can. J. Chem. 2000, 78, 732–738. (b) Overman, L. E.; Velthuisen, E. J. J. Org. Chem. 2006, 71, 1581–1587. 10. (a) McGarraugh, P. G.; Brenner-Moyer, S. E. Org. Lett. 2011, 13, 6460–6463. (b) McGarraugh, P. G.; Johnston, R. C.; MartínezMuñoz, A.; Cheong, P. H.-Y.; Brenner-Moyer, S. E. Chem. Eur. J. 2012, 18, 10742–10752. 11. Chi, Y.; Gellman, S. H. Org. Lett. 2005, 7, 4253–4256. 12. Pronin, S. V.; Reiher, C. A.; Shenvi, R. A. Nature 2013, 501, 195–199. 13. Pronin, S. V.; Shenvi, R. A. J. Am. Chem. Soc. 2012, 134, 19604–19606. 14. Wang, Z.-X.; Shi, Y. J. Org. Chem. 1998, 63, 3099–3104. 15. Surendra, J.; Corey, E. J. J. Am. Chem. Soc. 2008, 130, 8865– 8869. 16. Miyashita, K.; Yoneda, K.; Akiyama, T.; Koga, Y.; Tanaka, M.; Yoneyama, T.; Iwata, C. Chem. Pharm. Bull. 1993, 41, 465–470. 17. The diastereomeric ratio of 15 rapidly reaches equilibrium (1.2:1 favoring the desired isomer) under the conditions used for intermolecular Michael additions with catalyst 16 or achiral catalysts. With other catalysts, we can favor the formation of 17 over other stereoisomers; however, these reactions are not yet efficient and require in-depth study. For further details about these Michael experiments, please see the Supporting Information. 18. Melchiorre, P.; Jørgensen, K. A. J. Org. Chem. 2003, 68, 4151–4157. 19. Houjeiry, T. I.; Poe, S. L.; McQuade, D. T. Org. Lett. 2012, 14, 4394–4397. 20. Chen, K.; Ishihara, Y.; Galán, M. M.; Baran, P. S. Tetrahedron 2010, 66, 4738–4744. 21. Diamine 18 is known to be a poor catalyst for intermolecular Michael additions (ref 18). Therefore, we combined 15, MVK, and catalysts 16 and 18 in the hopes of completing a one-pot Robinson annulation. While Michael product 17 slowly accumulated in the usual 1:1 diastereomeric ratio, cyclohexenone 9 was not observed. 22. Please see the Supporting Information for the optimization of the conversion of 17 into 9 by aldol condensation. 23. Piers, E. Pure Appl. Chem. 1988, 60, 107–114. 24. The 1.3:1 ratio of decalin isomers obtained upon cyclization under the action of KOt-Bu/t-BuOH likely reflects the thermodynamic ratio of 10. Mixtures enriched (by chromatography) in either isomer converge under similar conditions to a 1.2:1 ratio favoring the cis-decalin isomer. After epoxidation of C4–C5 alkene, we have been able to equilibrate to a 2.2:1 trans:cis mixture; however, that sequence of events has proven less attractive than the one shown. 25. Imi, K.; Yanagihara, N.; Utimoto, K. J. Org. Chem. 1987, 52, 1013–1016. 26. Prudhomme, J.; McDaniel, E.; Ponts, N.; Bertani, S.; Fenical, W.; Jensen, P.; Le Roch, K. PLoS One 2008, 3, e2335.

ACS Paragon Plus Environment

Page 4 of 5

Page 5 of 5

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

H

HO

H NC

NC

H O H

kalihinol B 12 steps from geraniol IC50 = 4.6 nM (drug-resistant P. falciparum)

Cl

ACS Paragon Plus Environment