Abstract
The novel bromotriterpene polyethers aplysqualenol A (1) and aplysqualenol B (2) have been isolated from the Caribbean sea slug Aplysia dactylomela collected in Puerto Rico, and their structures and relative configurational assignments established from spectroscopic data aided by quantum mechanical calculations of NMR chemical shifts. Although both these compounds may be conceived as polyoxycyclic derivatives of the C30 squalene skeleton, remarkably 1 and 2 possess an unprecedented C15 to C24 flexible chain of 14S* spatial disposition that contains a unique ether bridge between C16 and C19. Biological activity screening tests revealed that, although aplysqualenol A (1) does not have significant anti-infective properties, it possesses potent antitumoral and antiviral activities.
Keywords: Aplysia dactylomela, Sea slug, Polyether, Antitumor activity, Antiviral activity
Introduction
Previous chemical investigations of the Caribbean sea slug Aplysia dactylomela Rang (order Anaspidea, family Aplysiidae) found worldwide in tropical to warm temperate waters, reported the isolation of two principal groups of compounds, terpenes (monoterpenes, sesquiterpenes, diterpenes and steroids) and nonterpenoid C-15 acetogenins.[1] To date, over 60 terpenoidal natural products have been isolated from A. dactylomela, and they exhibited various biological activities including anticancer, anti-HIV, alguicidal, ichthyotoxic, nematicidal, antiplasmodial, and antibacterial activities.[2] As part of a recent exploratory survey of marine invertebrates for anticancer constituents, A. dactylomela was collected in Cabo Rojo, Puerto Rico (April, 2001), and extracts from this anaspidean mollusk gave evidence of significant cytotoxic activity in the brine shrimp lethality bioassay (BSLT).[3] Thus, further examination of its constituents resulted in the isolation of two cytotoxic and structurally unique organic compounds. Herein, we report the isolation and structure determination of aplysqualenol A (1) and B (2), novel brominated triterpene polyethers with a tetracyclic skeleton, isolated from two specimens of this chemically prolific animal.[4]
The MeOH–CHCl3 (1:1) extract of freeze-dried A. dactylomela (54 g, dry wt) was partitioned between hexane and H2O. The hexane soluble material, which exhibited potent cytotoxicity against Artemia salina,[3] was subjected to fractionation using column chromatography (Silica gel) and polar bonded-phase HPLC to afford aplysqualenol A (1) and aplysqualenol B (2) as colorless oils (0.06% and 0.02% yield, respectively, based on crude extract weight). Aplysqualenol A (1) was evaluated in the National Cancer Institute (NCI) three-cell line, one-dose, primary anticancer assay displaying potent in vitro cytotoxicity against MCF-7 breast cancer, NCI-H460 nonsmall cell lung cancer, and SF-268 CNS cancer (the percent of growth of the treated cells when compared to the untreated cells was approximately 0, 0, and 4%, respectively).

Results and Discussion
The molecular formula of aplysqualenol A (1), [α]D20 +48.1° (c1.6, CHCl3), was determined to be C31H53BrO7 by HRFABMS and 13 C NMR. Treatment of 1 with acetic anhydride and pyridine at 55 °C for 5h gave the corresponding monoacetate 3, C33H55BrO8, νmax 1740 cm−1 and δH 2.05 (3H, s), whose IR spectrum still showed an absorption due to hydroxyl group at 3549 cm−1, thus proving the presence of secondary and tertiary hydroxyl groups in 1. The presence of two hydroxyl groups was further indicated by two D2O exchangeable resonances (δ 2.97 and 3.47) in the 1 H NMR spectrum. The 1 H and 13 C NMR spectral data revealed that aplysqualenol A (1) possessed 7 tertiary methyls, 5 oxygenated quaternary carbons, 6 oxygenated methine carbons, and a brominated methine carbon (δ 58.9). Since aplysqualenol A displayed signals for only one unsaturated bond [δC 144.4 (C) and 113.8 (CH2); δH 4.87 (1H, br s) and 4.93 (1H, br s)], 1 was assumed to comprise four oxacyclic units. The detailed analysis of the 1H–1H COSY, HMQC, and HMBC spectra of 1 showed the presence of the same partial structural units in the molecule (rings A–B–C) as those in thyrsiferol (5) and venustatriol (6), isolated from red algae of the genus Laurencia,[5,6] and aplysiol A (7) and B (8), isolated recently from South China Sea specimens of A. dactylomela.[7] Furthermore, the 1H–1H COSY and HMBC data, summarized in Table 1, allowed the foregoing partial structures (C1 through C15 and C25 through C28) and the remaining ring system D (C16 through C24 and C29 through C31) to be connected, establishing the connectivities of all the carbon atoms in 1 as shown.

Table 1.
1H NMR (500 MHz), 13C NMR (125 MHz), 1H–1H COSY, HMBC, and NOESY spectral data for aplysqualenol A (1) in CDCls.[a]
| Atom | δH, mult, intrgt (J in Hz) | δc, mult[b] | 1H–1H COSY | HMBC[c] | NOESY |
|---|---|---|---|---|---|
| 1 | 1.40, s | 23.6 (CH3) | H3, H25 | H4α, H25, H26 | |
| 2 | 75.0 (C) | H1, H3, H4αβ, H25 | |||
| 3 | 3.88, dd, 1H (4.0, 12.3) | 58.9 (CH) | H4αβ | H1, H4αβ, H5αβ, H25 | H4α, H25 |
| 4α | 2.10, dq, 1H (3.9, 4.0, 4.5, 13.5) | 28.3 (CH2) | H3, H4β, H5αβ | H3, H4β | |
| 4β | 2.22, dq, 1H (3.5, 12.3, 13.1, 13.5) | H3, H4α, H5αβ | H1, H4α, H26 | ||
| 5α | 1.82, m, 1H | 37.1 (CH2) | H4αβ, H5β | H26 | H5β, H7 |
| 5β | 1.53, m, 1H | H4αβ, H5α | H5α, H26 | ||
| 6 | 74.3 (C) | H4αβ, H7, H8α, H26 | |||
| 7 | 3.09, dd, 1H (2.2, 10.9) | 86.9 (CH) | H8αβ | H9β, H26 | H5α, H8α, H9α, H11 |
| 8α | 1.53, m, 1H | 22.9 (CH2) | H7, H8β, H9αβ | H9α | H7, H8β |
| 8β | 1.78, m, 1H | H7, H8α, H9αβ | H8α, H26, H27 | ||
| 9α | 1.53, m, 1H | 37.5 (CH2) | H8αβ, H9β | H27 | H7, H9β, H11 |
| 9β | 1.73, m, 1H | H8αβ, H9α | H9α, H27 | ||
| 10 | 73.0 (C) | H8αβ, H9αβ, H27 | |||
| 11 | 3.01, dd, 1H (3.8, 11.4) | 80.8 (CH) | 12αβ | H9β, H27 | H7, H9α, H13α |
| 12α | 1.66, m, 1H | 24.3 (CH2) | H11, H13αβ | H14 | |
| 12β | 1.53, m, 1H | H11, H13αβ | H14, H27 | ||
| 13α | 1.50, m, 1H | 26.0 (CH2) | H12αβ, H14 | H11, H28 | |
| 13β | 1.65, m, 1H | H12αβ, H14 | |||
| 14 | 3.54, dd, 1H (2.5, 11.6) | 72.5 (CH) | H13αβ | H13αβ, H28 | H12β, H27, H29 |
| 15 | 74.9 (C) | H13αβ, H14, 15-OH, H16, H17αβ, H28 | |||
| 16 | 3.85, dd, 1H (4.2, 9.1) | 79.1 (CH) | H17αβ | H17α, H18, H28 | H17α, H18, H28 |
| 17α | 2.34, ddd, 1H (6.0, 8.9, 14.3) | 35.3 (CH2) | H16, H17β, H18 | H16, H17β, H18 | |
| 17β | 1.96, dd, 1H (3.0, 14.3) | H16, H17α, H18 | H17α, 18-OH, H28 | ||
| 18 | 3.81, dd, 1H (2.0, 5.9) | 75.8 (CH) | H17αβ, 18-OH | H17β, 18-OH, H20β, H29 | H16, H17α, H20α, H30 |
| 19 | 85.7 (C) | H17β, H20αβ, H29 | |||
| 20α | 1.40, m, 1H | 34.1 (CH2) | H20β, H21αβ | H21αβ, H22, H29 | H18, H20β |
| 20 β | 1.30, dd, 1H (4.8, 12.3) | H20α, H21αβ | H20α | ||
| 21αβ | 1.64, m, 1H; 1.56, m, 1H | 28.2 (CH2) | H20αβ, H22 | H29 | |
| 22 | 3.44, dd, 1H (6.5, 6.8) | 86.1 (CH) | H21αβ | H20αβ, H21αβ, H24αβ, H30, OCH3 | H24α, H30, OCH3 |
| 23 | 144.4 (C) | H30 | |||
| 24α | 4.87, br s, 1H | 113.8 (CH2) | H24β, H30 | H22, H30 | H22, H24β |
| 24β | 4.93, br s, 1H | H24α, H30 | H24α, H30 | ||
| 25 | 1.27, s, 3H | 31.0 (CH3) | H1, H3 | H1, H3, | |
| 26 | 1.20, s, 3H | 20.1 (CH3) | H1, H4β, H5β, H8β | ||
| 27 | 1.15, s, 3H | 14.8 (CH3) | H11 | H8β, H9β, H12β, H14 | |
| 28 | 1.17, s, 3H | 18.1 (CH3) | H13α,15-OH, H16, H17β | ||
| 29 | 1.22, s, 3H | 19.4 (CH3) | H14, 18-OH, H21β | ||
| 30 | 1.63, s, 3H | 16.3 (CH3) | H24αβ | H22, H24αβ | H18, H22, H24β, OCH3 |
| OCH3 | 3.19, s, 3H | 56.0 (CH3) | H22 | H22, H30 | |
| 15-OH | 2.97, br s, 1H | H28 | |||
| 18-OH | 3.47, br d, 1H (7.3) | H18 | H17β, H29 |
Spectra were recorded at 25 °C. Chemical shifts are in ppm relative to TMS.
13C NMR multiplicities were obtained from APT experiments.
Protons correlated to carbon resonances in the 13C column. Parameters were optimized for2,3JCH = 6 and 8 Hz.
This unprecedented architecture required further confirmation, which was provided by unambiguous determination of the locations of the two hydroxyl groups in 1. The positions of these pivotal functionalities were initially determined by HMBC correlations (C15/15-OH and C18/18-OH), and later confirmed by a deuterium shift experiment, which revealed that among all oxygenated carbons, only the two signals for C15 and C18 shifted significantly to lower field when measured in CD3OH as compared to CD3OD.[8] Having definitively located the two hydroxyl groups at C15 and C18, we inferred that the other 9 oxycarbons in aplysqualenol A all participate in ether linkages. Thus, the locus of oxolane ring D arising from an ether bridge between C16 and C19, was indicated by the HMBC correlations (C15/H-16, C15/H-17αβ, C16/H-18, C16/H3-28, C19/H-17β, C19/H-20αβ, and C-19/H3-29). The one additional ethereal group (-OCH3) present in 1 was shown to be located at C22 based on the HMBC correlations (C22/-OCH3, C22/H-24αβ, C22/H3-30, and C22/H-21αβ), leaving the bromine atom to reside at C3.
The oxacyclic systems of 1 were also deduced from the specific fragment ions in the EI mass spectrum (Figure 1). The mass spectrum showed fragment ions at m/z 503/505 [M – C7H13O]+, 403/405 [M – C12H21O3]+, 359/361 [M – C14H25O4]+, and 205/207[M – C23H39O6]+ due to cleavage at C19–C20, C15–C16, C14–C15, and C6–C7 bonds, respectively. Hence, the remaining two ether linkages consisted of a 2,7-dioxabicyclo[4.4.0]decane ring as evidenced from the intense (100%) fragment ion at m/z 153 [C9H13O2]+. The presence of an oxolane ether bridge encompassing positions C16 and C19 was corroborated from the rearranged fragment ion at m/z 214 [M – C19H32BrO4 + H]+ resulting from cleavage of the C15-C16 bond. Furthermore, the fragment ion at m/z 257 [M – C17H28BrO3]+ arising from cleavage at C14–C15, confirmed the presence in 1 of the partial structural unit from C15 through C24 (including C28 through C31). Consequently, aplysqualenol A (1) was found to be a new member of squalene-derived bromotriterpene with novel structural features not found in known congeners of this class of natural products.[4,7]
Figure 1.
Diagnostic fragment ions (m/z) of aplysqualenol A (1) detected in the EI mass spectrum.
The assignment of the relative stereochemistry of 1, except at C22, was straightforward. The equatorial orientation of the bromine atom at C3 on the A ring was evident from the J-values with vicinal axial-axial coupling constant of H-3 (12.3 Hz) in the 1 H NMR spectrum of 1. From this point forward, the relative stereochemistries were determined by the NOESY spectrum as summarized in Table 1. NOEs about the ABC-rings were detected between H-3/H3-25, H3-1/H3-26, H-5α/H-7, H3-26/H-8β, H-7/H-11, H-14/H-12β, and, most importantly, H-14/H3-27, respectively, establishing a cis-relationship among these sets of protons. No NOE, however, was detected between H-7/H3-26, H-11/H3-27, or H-11/H-14, thus suggesting that these pairs of protons were trans oriented. In addition, NOEs about the C/D rings were detected between H-13α/H3-28, H-16/H3-28, H-16/H-17α, H-17α/H-18, and H-14/H3-29, suggesting a close spatial proximity among these protons. Moreover, no NOE between H14/H3-28, H-14/H-16, H-16/H3-29, and H-18/H3-29 suggested the stereochemistries of these protons to be trans (Figure 2). The observation that in compound 1 the proton signal for H-14 was resolved as a doublet of doublets with two different coupling constant values (one much larger than the other indicating a chair/chair B–C rings system) confirmed the axial orientation of this methine and, therefore, that the relative stereochemistry at carbon C-14 would be S* for aplysqualenol A (1). As far as we know, this is the first example of a squalene-polyether derivative displaying a C15 to C24 flexible side chain of equatorial spatial disposition.[9] In order to corroborate the relative configuration shown for aplysqualenol A molecular mechanics/dynamics calculations were performed to establish the dominant conformations of 1.[10] The results of the conformational analysis inferred that the secondary alcohol at C18 was fixed with the tertiary hydroxyl oxygen at C15 in a H-bond. In turn, 15-OH is hydrogen bonded to the oxygen atom of oxane ring C. Indeed, the presence of intramolecular hydrogen bonding in aplysqualenol A was supported experimentally by the strong couplings of H-18 and 18-OH observed in the 1H–1H COSY spectrum, the fact that acetylation of 1 did not take place at 25 °C, and the strong NOEs detected between H3-28/15-OH, H-17β/18-OH, and H3-29/18-OH. Furthermore, the coupling constants of H-22 with the diastereotopic protons H2-21 (J22–21α = 6.5 Hz and J22–21β = 6.8 Hz) indicated that free rotation about the C21–C22 bond renders them equivalent, thus strengthening the contention that the secondary alcohol is fixed in a H-bond with 15-OH and not the methoxy-ether at C22. Insofar as the relative configuration for C22 is concerned, it was ultimately assigned as R* on the basis of a combined approach, which included NOESY NMR data acquired with aplysqualenol B (2), an integrated QM (Quantum Mechanical)–NMR approach based on DFT (density functional theory)[11] calculations of 13 C NMR chemical shifts and the analysis of the experimental NMR data of aplysqualenol A (1), as well as biogenetical considerations (vide infra). Figure 2 shows a plausible conformation of rings A–D based on the coupling constants and NOE correlations observed which indicated the relative stereochemistry of aplysqualenol A (1) to be 3R*,6S*,7R*,10S*,11R*,14S*,15R*,16R*,18R*,19S* and 22R*.[12] Sadly, all of our attempts to determine absolute stereochemistry by mixing 1 with a solid matrix-bound auxiliary reagent (MTPA) directly in the NMR tube were unsuccessful.[13,14]
Figure 2.

Plausible conformations of rings A–B–C–D in 1 showing selected NOESY correlations.
Compound 2, aplysqualenol B, was also isolated as a colorless oil, [α]D20 +27.1° (c 1.4, CHCl3). Mass spectral analysis of this metabolite showed a molecular ion consistent with the molecular formula C31H53BrO8 (observed [M + H]+ 633.3003; calculated 633.3002). Assignment of the structure for this metabolite was aided by the considerable spectroscopic analogy to aplysqualenol A (1). The only significant variations in the 1 H NMR spectrum between compound 2 and 1 were the disappearance of the tertiary methyl signal ascribable in 1 to H3-30 (δH 1.63) and the appearance of a pair of AB doublets in 2 at δH 4.21 (1H, J = 13.5 Hz) and 4.09 (1H, J = 14.3 Hz), as well as the downfield shifts in the protons H-22 (δH 3.62, dd, J = 6.5, 6.8 Hz) and H2-24 (δH 5.06, 1H, br s, H-24α and 5.24, 1H, br s, H-24β). These data suggested that the modifications in this compound were mainly located in the C6 side chain appended to ring D. The connectivities observed in 1H–1 H COSY, HMQC, and HMBC experiments (Table 2) made it possible to assign the fragment C20→C24 (including C30) as follows: δC: 34.2 (CH2, C20), 28.6 (CH2, C21), 85.0 (CH, C22), 146.9 (C, C23), 114.0 (CH2, C24) and 62.9 (CH2, C30). These facts can only be explained by the presence in 2 of a primary hydroxyl group at C30. The isolated hydroxymethylene group was connected to the ring D side chain through C23 by HMBC correlations (C30/H-22, C30/H-24αβ, C23/H-30αβ and C24/H-30αβ). In order to confirm the presence of this system, compound 2 gave the 18,30-diacetate 4 by overnight treatment with Ac2O/Py at 25 °C, C35H57BrO10, νmax 3541, 1740 cm−1 and δH 2.05 (3H, s) and 2.10 (3H, s), thus confirming its structure as an aplysqualenol A derivative. As in compound 1, strong NOE correlations established the cis orientation between the methyl group H3-27 (δH 1.15) and the oxymethine proton H-14 (δH 3.54), indicating that the relative configuration at C14 in compound 2 was S*. The relative stereochemistry of the remaining carbons C3, C6, C7, C10, C11, C15, C16, C18, and C19 in compound 2 was established as identical with that observed for aplysqualenol A (1) on the basis of the observed NOE data in the NOESY experiments in CDCl3 (Table 2). The relative stereochemistry of C22 in compound 2 was established as follows. The fact that in aplysqualenol B (2) the H2-30 protons were diastereotopic (they appeared as an isolated AB system indicating rotational constraints along the C23–C30 bond), inferred that the primary alcohol was fixed with the methoxy-ether oxygen in a H-bond. This contention was supported by molecular mechanics calculations, which predicted identical coupling constant values between H-22 with the diastereotopic protons H2-21 (actual values observed: J22–21α = 6.5 Hz and J22–21β = 6.8 Hz) in the most stable conformation of compound 2 (Table 2). The NOE connectivities observed between H-22 with H-30β, H-24α, and, most importantly, H-16, and the observation that in diacetate derivative 4, wherein such H-bonds are no longer attainable, the H2-30 protons appeared as broad singlets centered near δ 4.56.[15] Thus, the relative stereochemistry at C22 in this compound was established as R*. Since these compounds share the same biogenetic pathway, it is highly reasonable that 3R*, 6S*, 7R*, 10S*, 11R*, 14S*, 15R*, 16R*, 18R*, 19S*, 22R* is the most likely configuration for aplysqualenol B (2).
Table 2.
1H NMR (500 MHz), 13C NMR (125 MHz), 1H–1H COSY, HMBC, and NOESY spectral data for aplysqualenol B (2) in CDCls.[a]
| Atom | δH, mult, intrgt (J in Hz) | δc, mult[b] | 1H–1H COSY | HMBC[c] | NOESY |
|---|---|---|---|---|---|
| 1 | 1.39, s | 23.6 (CH3) | H25 | H4β, H25, H26 | |
| 2 | 75.0 (C) | H1, H3, H25 | |||
| 3 | 3.88, dd, 1H (4.0, 12.3) | 58.9 (CH) | H4αβ | H1, H4β, H5β, H25 | H4α, H5α, H25 |
| 4α | 2.09, dq, 1H (3.8, 3.9, 4.1, 13.5) | 28.2 (CH2) | H3, H4β, H5αβ | H3, H4β, H5α | |
| 4β | 2.24, dq, 1H (3.7, 12.8, 13.1, 13.5) | H3, H4α, H5αβ | H1, H4α, H26 | ||
| 5α | 1.81, m, 1H | 37.1 (CH2) | H4αβ, H5β | H26 | H4α, H5β, H7, H26 |
| 5β | 1.52, m, 1H | H4αβ, H5α | H5α, H26 | ||
| 6 | 74.3 (C) | H26 | |||
| 7 | 3.10, dd, 1H (2.2, 10.9) | 86.9 (CH) | H8αβ | H8 α, H9α, H26 | H5α, H8α, H9α, H11 |
| 8α | 1.52, m, 1H | 23.0 (CH2) | H7, H8β, H9αβ | H7, H8β | |
| 8β | 1.78, m, 1H | H7, H8α, H9αβ | H8α, H26, H27 | ||
| 9α | 1.50, m, 1H | 37.5 (CH2) | H8αβ, H9β | H27 | H7, H9β, H11 |
| 9β | 1.72, m, 1H | H8αβ, H9α | H9α, H27 | ||
| 10 | 73.0 (C) | H8α, H9β, H27 | |||
| 11 | 3.01, dd, 1H (3.9, 11.4) | 80.7 (CH) | 12αβ | H9β, H13β, H27 | H7, H9α, H13α |
| 12α | 1.66, m, 1H | 24.3 (CH2) | H11, H12β, H13αβ | H11 | |
| 12β | 1.53, m, 1H | H11, H12α, H13αβ | H27 | ||
| 13α | 1.48, m, 1H | 26.0 (CH2) | H12αβ, H13β, H14 | H11, H28 | |
| 13β | 1.61, m, 1H | H12ab, H13α, H14 | |||
| 14 | 3.54, dd, 1H (2.5, 11.7) | 72.5 (CH) | H13αβ | H13α, H28 | H27 |
| 15 | 75.0 (C) | H14, H17α, H28 | |||
| 16 | 3.84, dd, 1H (4.2, 9.1) | 79.2 (CH) | H17αβ | H17α, H28 | H17α, H22, H28 |
| 17α | 2.33, ddd, 1H (5.9, 9.1, 14.4) | 35.2 (CH2)[d] | H16, H17β, H18 | H16, H17β, H18 | |
| 17β | 1.97, ddd, 1H (1.3, 4.0, 14.4) | H16, H17α, H18 | H17α, H28 | ||
| 18 | 3.80, br m, 1H | 75.9 (CH)[d] | H17αβ, 18-OH | H17β, H29 | H17α, H20α |
| 19 | 85.7 (C) | H17β, H20α, H29 | |||
| 20α | 1.42, m, 1H | 34.2 (CH2) | H20β, H21αβ | H22, H29 | H18, H20β |
| 20 β | 1.31, m, 1H | H20α, H21αβ | H20α | ||
| 21αβ | 1.65, m, 2H | 28.6 (CH2) | H20αβ, H22 | H22 | |
| 22 | 3.62, dd, 1H (6.5, 6.8) | 85.0 (CH) | H21αβ | H21αβ, H24αβ, OCH3 | H16, H24α, H30β, OCH3 |
| 23 | 146.9 (C) | H24αβ, H30αβ | |||
| 24α | 5.06, br s, 1H | 114.0 (CH2) | H24β | H22, H30αβ | H22, H24β, OCH3 |
| 24β | 5.24, br s, 1H | H24α | H24α, H30α | ||
| 25 | 1.27, s, 3H | 31.0 (CH3) | H1 | H1, H3 | |
| 26 | 1.20, s, 3H | 20.1 (CH3) | H5α, H8β | H4β, H5β, H8β | |
| 27 | 1.15, s, 3H | 14.8 (CH3) | H9β, H11 | H8β, H12β, H14 | |
| 28 | 1.17, s, 3H | 18.1 (CH3) | H13α, H16, H17β | ||
| 29 | 1.22, s, 3H | 19.4 (CH3) | |||
| 30α | 4.09, br d, 1H (14.3) | 62.9 (CH2) | H30β | H22, H24αβ | H24β, H30β |
| 30β | 4.21, br d, 1H (13.5) | H30α | H22, H30α | ||
| OCH3 | 3.26, s, 3H | 56.4 (CH3) | H22 | H22, H24α | |
| 15-OH | 3.11, br s, 1H | ||||
| 18-OH | 3.55, br s, 1H | H18 |
Spectra were recorded at 25 °C. Chemical shifts are in ppm relative to TMS.
13C NMR multiplicities were obtained from APT experiments.
Protons correlated to carbon resonances in the 13C column. Parameters were optimized for 2,3JCH = 6 and 8 Hz.
Signal recorded as a broad low-intensity resonance line.
With the aim at strengthening the relative stereochemistry at the 11 stereogenic centers (C3, C6, C7, C10, C11, C14, C15, C16, C18, C19 and C22) of aplysqualenol A (1), we undertook a configurational study by an integrated QM-NMR approach based on DFT calculations of 13 C NMR chemical shifts and the analysis of the experimental NMR data. Critically, 13 C chemical shifts were used to validate the theoretical models, and dipolar coupling correlations derived from 2D-NOESY NMR experiments were used to corroborate the arrangements suggested by QM methods, and to determine the relative configuration of the molecule. As all the proton and carbon values were assigned confidently by 2D-NMR experiments (Table 1), which were in agreement with the proposed structure 1, we confined our study to the four possible diastereomers bearing on C15 and C22 (see stereoisomer models 1a–1d in Figure 3), the two most flexible and thus compromising stereocenters of aplysqualenol A. To determine the relative configuration of these centers the 13 C NMR isotropic shifts of each diastereomer (Figure 3) at its lowest energy configuration were calculated. The minimum energy configuration for each isomer was identified using a Monte Carlo conformational search with the MMFF force field as implemented in the Spartan 04 software package.[16] All diastereomers were further optimized using HF/6-31G(2d,p) level of theory using the Gaussian 03 program.[17] NMR chemical shifts were calculated using the GIAO method at the mPW1PW91/6-31G(2d,p) level of theory. Recently, Bifulco and coworkers reported this level of theory to give reasonable results for the 13 C shift of organic compounds.[7,18] The experimental values were plotted against the calculated shift, and a least-squares fit line was determined. The calculated shifts for each isomer were corrected using the slope and intercept to give scaled 13 C shifts. The difference plots were determined by subtracting the corrected shifts from the experimental chemical shifts.
Figure 3.
Minimum energy configurations of the four stereoisomer models 1a–1d considered for quantum mechanical calculations of 13C NMR isotropic shifts. The hydrogen atoms have been omitted for clarity. Oxygen atoms are indicated in red. Hydrogen bonds are represented as dashed lines between the donor hydrogen and the acceptor atom. Bonds with less than ideal geometry are displayed with a blue tint. The intensity of the colour increases as the bond becomes less ideal.
Diastereomer 1b presented the best 13 C-NMR chemical shift correlation with respect to the theoretical values (Figure 4) with an average of Δδ = 2.9 (Table 3). It also resulted in the conformation with the lowest energy compared with the other three isomers (1a, 1c, and 1d). These theoretical results for diastereomer 1b (15R*,22R*) are in agreement with other previously detailed experimental observations, including the observation that the coupling constants of H-22 with the diastereotopic protons H2-21 are similar suggesting that the C21–C22 bond experiences free rotation and thus the coupling constants are averaged. Had the 18-O H been fixed with the methoxy-ether oxygen at C22 (diastereomers 1c and 1d), two significantly different coupling constant values (one much larger than the other) would have been anticipated. The Δδ values observed for the signals of carbonsnearby the methoxy group at C22 also indicated the R* configuration for this centre, as depicted in formula 1. To summarize, neither diastereomer 1a (15R*,22S*), 1c (15S*,22R*), nor 1d (15S*,22S*) is likely to be the correct structure for aplysqualenol A because of their inherently higher relative energies/Δδ values. These theoretical calculations not only provided information on the R* configuration of C15 and C22, but also corroborated the configurational arrangement of C14 depicted in 1 and determined as described above.
Figure 4.
Deviations from the average of the carbon chemical shifts of distereomers 1a-1d. The x and y axes represent position number and Δδ in parts per million, respectively. To discriminate between stereoisomers 1a and 1b (the two most likely structures for aplysqualenol A on the basis of the calculation results shown in Table 3), a careful analysis was done on individually calculated 13C chemical shifts for carbons C22 and C30, which were expected to experience larger variations upon inversion of configuration at C22. As shown here, very large differences in the Δδ 13C values of 1a and 1b were observed for C22 and C30 (+11.9 vs +4.0 and −7.8 vs −3.0, respectively), suggesting again the exclusion of stereoisomer 1a.
Table 3.
Calculated parameters for the four diastereomers 1a–1d.
| Properties | 1a | 1b | 1c | 1d |
|---|---|---|---|---|
| MAE[a] | 3.1 | 2.9 | 3.9 | 3.5 |
| Energy[b] | −1741.8144 | −1741.8191 | −1741.8057 | −1741.8125 |
| δE[c] | 2.98 | 0 | 8.42 | 4.17 |
| R2[d] | 0.9909 | 0.9982 | 0.9897 | 0.9953 |
Mean absolute error in ppm found for calculated 13C NMR chemical shifts versus 13C experimental values: MAE = Σ [|(δexp – δcalcd)|]/n.
Data in atomic units.
Data in kcal/mol.
Correlation coefficient obtained by a linear fit of the calculated, δcalcd, versus experimental, δexp, 13C NMR chemical shifts.
From a biogenetic viewpoint the polyoxygenated squalene-derived ethers isolated from A. dactylomela could arise from a common precursor, (10R, 11R)-(+)-squalene-10,11-epoxide (9) isolated from Laurencia okamurai.[19] Upon further oxidation, this compound could give rise to (6S, 7S, 10R, 11R, 14S, 15S, 16R, 18R, 19R)-16R-hydroxy-squalene tetraepoxide (10) (such intermediate has not been reported yet). Biogenesis of the A–B–C–D ring system could stem from the concerted cyclizations of four epoxides after formation of a bromonium ion at C2–C3, thus forming the framework of these metabolites. In point of fact, the aberrant S* stereochemistry at carbon C14 in compounds 1 and 2 (indicati n g o verall retention of configuration at that center with respect to 10) indicates that the proposed biogenesis through the cyclization of the squalene tetraepoxide precursor should in fact be concerted. Upon further oxidative metabolism of the remaining olefin followed by O-methylation, aplysqualenol A (1) could be obtained (Scheme 1).[20] Compound 1 would in turn evolve to aplysqualenol B (2) following enzymatic hydroxylation at C30.
Scheme 1.
Proposed biogenetic pathway for aplysqualenol A (1).
Conclusions
Aplysqualenol A (1) and B (2) are novel bromotriterpenes structurally related to thyrsiferol (5) which suggests 1 and 2 to be of dietary origin. Often, opisthobranch mollusks belonging to the order Anaspidea, feed on red and brown algae from which they sequester selected bioactive metabolites that are stored in the digestive gland and secreted in the mucus for defensive purposes.[21] Thus, the presence of aplysqualenol A and B in A. dactylomela suggests a likely involvement of these molecules in the chemical defensive mechanism of the sea slug.[7] Polyether squalene-derived of the type represented by 1 and 2 are currently quite rare, in contrast to thyrsiferol types.[4] Thus, the aplysqualenols represent the only examples of this small family of marine metabolites to display the S* configuration at C14 and to possess an ether linkage between C16 and C19. The unprecedented S* stereochemistry at C14 leads, not only to a conformationally more stable 2,7-dioxabicyclo[4.4.0]decane system (chair/chair B–C rings system) than that present in thyrsiferol (5) and its congeners (chair/twist-boat B–C rings system), but also changes the arrangement and direction of the flexible side chain from a less favored axial position to a more stable equatorial orientation.[22,23]
Upon screening in the NCI’s in vitro antitumor assay consisting of 60 human tumor cell lines aplysqualenol A (1) exhibited inhibitory activity against SNB-19 CNS cancer and T-47D breast cancer with IC50 values of 0.4 and 0.3 μg/mL, respectively. To explore its antiviral properties aplysqualenol A was evaluated against herpes simplex virus type 1 (HSV-1) and type 2 (HSV-2), varicella zoster virus (VZV), human cytomegalovirus (HCMV), and Epstein-Barr virus (EBV). At concentrations above 4 μg/mL compound 1 showed a 90% maximal response (EC90) against HSV-1, HSV-2, HCMV, and VZV viruses (acyclovir was used as a control with EC50 values of 0.95, 0.95, 0.22, and 0.14 μg/mL, respectively). Remarkably, aplysqualenol A was very toxic against the Epstein-Barr virus in VCA Elisa assay (EC90 < 0.08 μg/mL; acyclovir EC50 = 1.1 μg/mL) with no accompanying toxicity seen in the host Daudi cells.[24] Compounds 1 and 2 showed moderate antiplasmodial activity against Plasmodium falciparum with IC50 values of 11 and 18 μg/mL, respectively. In vitro antituberculosis screening of aplysqualenol A (1) against Mycobacterium tuberculosis H37Rv at a concentration of 6.25 μg/mL showed no inhibitory activity.
Experimental Section
General Experimental Procedures
Optical rotations were measured as the sodium line (589 nm) with a Perkin-Elmer Polarimeter Model 243B. FT-IR spectra were measured with a Nicolet Magna 750 FT-IR spectrophotometer as a thin film on a NaCl disc. 1H and 13C NMR spectral data and 1H–1H COSY, NOESY, APT, HMQC, and HMBC experiments were measured in CDCl3 with a Bruker Avance DRX-500 FT-NMR spectrometer. 1H–1H COSY, NOESY, HMQC, and HMBC spectra were measured using standard Bruker pulse sequences. Chemical shifts are given on a δ (ppm) scale with CHCl3(1H, 7.26 ppm) and CDCl3 (13C, 77.0 ppm) as the internal standard. Mass spectra were taken at the Mass Spectrometry Laboratory of the University of Illinois at Urbana–Champaign. Column chromatography (CC) was performed on silica gel (35-75 mesh). TLC analyses were carried out using glass silica gel plates and spots were visualized by exposure to I2 vapors or heating silica gel plates sprayed with 5% H2SO4 in EtOH. All solvents were either spectral grade or distilled from glass prior to use. The percentage yield of each compound is based on the weight of the crude MeOH–CHCl3 extract.
Animal Material
Two large specimens of Aplysia dactylomela (Rang, 1828) (order Anaspidea, superfamily Aplysioidea, family Aplysiidae) were collected on April 28, 2001 in Bahia Salinas, Cabo Rojo, Puerto Rico by hand at 2 feet deep. Each individual was found grazing upon the red alga Laurencia obtusa. A voucher specimen has been deposited at the Department of Chemistry, University of Puerto Rico, Río Piedras, Puerto Rico (deposit number ADPR01-1).
Extraction and Isolation of Aplysqualenol A (1) and B (2) and the Known Compound Elatol (11)
The organisms were freeze-dried and the dry animal (54 g dry weight) was cut into small pieces and extracted with 1:1 CHCl3–MeOH (6 × 1L). The combined organic extracts were concentrated and the residue obtained (39 g) was suspended in water (500 mL) and partitioned between n-hexane, CHCl3, EtOAc and n-butanol. The n-hexane extract (11 g) was chromatographed over silica gel (30 g) using n-hexane→CHCl3→EtOAc→MeOH in increasing polarity as eluant. The 100% CHCl3 eluate (980 mg) was chromatographed over silica gel (30 g) with 15% CHCl3 in n-hexane yielding 16 fractions. Subfraction 8 (55 mg) was subsequently chromatographed over silica gel (3 g) with 25% CHCl3 in n-hexane to afford known elatol (11) (37 mg; 0.1% yield).[25] The fraction (1.2 g) that eluted with CHCl3–EtOAc (8:2) was chromatographed over silica gel (31 g) with 10% n-hexane in CHCl3 to afford 9 fractions. Subfraction 8 (101 mg) was further purified over silica gel (5 g dry-packed) eluting with 2% EtOAc in CHCl3. The least polar fraction (66 mg) was purified further by CC over silica gel (5 g) using 10% EtOAc in n-hexane to yield 8 fractions. Subfraction 7 (44 mg) was subsequently purified by CC over silica gel (3 g) with 10% EtOAc in n-hexane followed by NP-HPLC [Ultrasphere-Cyano 250 × 10 mm, flow rate: 2.0 mL/min, UV detection set at λ = 254 nm] using a combination of 95:5 n-hexane–2-propanol to afford pure aplysqualenol A (1) (24 mg). The portion (1.5 g) eluting with EtOAc–MeOH (8:2) was filtered and purified over a Bio-Beads SX-3 column (toluene) to yield 3 fractions. The second fraction was chromatographed over silica gel (4 g) using 0.5% MeOH in CHCl3 to afford pure aplysqualenol B (2) (7.1 mg).
Aplysqualenol A (1)
colorless oil; [α]D20 +48.1 (c 1.6, CHCl3); IR (film) νmax 3418, 3072, 2984, 2954, 2875, 1738, 1648, 1456, 1379, 1323, 1127, 1098, 905, 761 cm−1; 1H-NMR (CDCl3, 500MHz) and 13C-NMR (CDCl3, 125MHz), see Table 1; EIMS m/z [M – HBr]+ 536 (12), 505 (4), 503 (4), 423 (2), 405 (15), 403 (14), 387 (12), 385 (10), 361 (7), 359 (6), 323 (12), 305 (8), 291 (3), 289 (4), 279 (8), 247 (10), 245 (9), 225 (15), 214 (17), 207 (44), 205 (41), 182 (24), 153 (100), 125 (59), 109 (64), 85 (36); HRFABMS (3-NBA) m/z [M + Na]+ calcd for C31H5379BrO7Na 639.2872, found 639.2858 (Δ 1.4 mDa).
Aplysqualenol B (2)
colorless oil, [α]D20 +27.1 (c 1.4, CHCl3); IR (film) νmax 3408, 2978, 2952, 2862, 1726, 1453, 1381, 1130, 1095, 1059, 1032, 911 cm−1; 1H-NMR (CDCl3, 500MHz) and 13C-NMR (CDCl3, 125MHz), see Table 2; HRFABMS (3-NBA) m/z [M + H]+ calcd for C31H5479BrO8 633.3002, found 633. 3003 (Δ −0.1 mDa).
Elatol (11)
The [α]D20, UV (MeOH), 1H- and 13C-NMR, and HREIMS data were identical in all respects to those previously reported.[25]

Acetylation of Aplysqualenol A (1)
Compound 1 (1.7 mg, 0.003 mmol) was dissolved in a mixture of dry pyridine (500 μL) and acetic anhydride (500 μL) and heated to 55 °C for 5h. The cooled reaction mixture was concentrated in vacuo and the oily residue obtained was purified by CC over silica gel (1 g) using 1% MeOH in CHCl3 to yield pure aplysqualenol A acetate (3) (1.7 mg, quantitative yield). Compound 3: [α]D20 +56.0 (c 1.0, CHCl3); IR (film) νmax 3549, 3071, 2983, 2951, 2858, 2819, 1740, 1454, 1380, 1242, 1125, 1099, 1029, 921 cm−1; 1H-NMR (CDCl3, 500MHz) and 13C-NMR (CDCl3, 125MHz), see Table 4 as Supporting Information; HREIMS m/z [M – HBr]+ calcd for C33H54O8 578.3819, found 578.3826 (Δ −0.7 mDa), 578 (18), 405 (7), 403 (7), 387 (5), 385 (4), 361 (5), 359 (5), 323 (8), 256 (11), 224 (26), 207 (42), 205 (42), 153 (59), 125 (72), 120 (94), 118 (100), 117 (57).
Acetylation of Aplysqualenol B (2)
Aplysqualenol B (3.0 mg; 0.005 mmol) was dissolved in a mixture of dry pyridine (500 μL) and acetic anhydride (500 μL) and then stirred at 25 °C overnight. The reaction mixture was concentrated in vacuo and the oil obtained was purified by CC over silica gel (1.0 g) with 0.5% MeOH in CHCl3 to afford aplysqualenol B diacetate (4) (3.0 mg; 88% yield). Compound 4: colorless oil; [α]D20 +44.0 (c 0.5, CHCl3); IR (film) νmax 3541, 3467, 3087, 2987, 2951, 2869, 2859, 2821, 1740, 1655, 1459, 1440, 1380, 1242, 1121, 1097, 1054, 1029, 922 cm−1; 1H-NMR (CDCl3, 300MHz) and 13C-NMR (CDCl3, 75MHz), see Table 5 as Supporting Information; HRESIMS m/z [M + H]+ calcd for C35H5879BrO10 717.3213, found 717.3216 (Δ −0.3 mDa).
Computational Method
Initially, a Monte Carlo (MC) multiple-minimum search with the MMFFs force field was conducted to make a full exploration of the conformational space for all the four possible stereoisomers involving the C15 and C22 stereocenters of aplysqualenol A (1). The molecular mechanics MC conformational search employs an algorithm as implemented in Spartan 04 software package. All the stereoisomers thus obtained were further subjected to ab initio quantum chemical geometry optimizations at the HF/6-31G(2d,p) level of theory. The ab initio calculations were carried out on a Linux AMD64 1.8 GHz using the Gaussian 03 program package.[17] The Gaussian 03 uses an expansion of molecular orbitals in atomic-centered Gaussian basis sets such as the double-zeta plus polarization 6-31G(2d,p). NMR shielding tensors were computed with the gauge-independent atomic orbital (GIAO) method at the hybrid density-functional mPW1PW91/6-31G(2d,p) level of theory.
Biological Screening Assays
For a general description of the approach used by the NIAID’s Antimicrobial Acquisition and Coordinating Facility (AACF) for determining antiviral activity and toxicity for herpes viruses, visit: http://niaid-aacf.org/protocols/Herpes.htm. Anticancer activity screening by the Developmental Therapeutics Program (DTP) of the National Cancer Institute is conducted following this general protocol: most of the compounds screened have no antiproliferative activity (up to 85%). In order to avoid screening inactive compounds across all the cell lines, a prescreen is done using 3 highly sensitive cell lines (breast MCF-7, lung NCI-H640, CNS SF-268). Antiproliferative activity must be seen in these cell lines in order to continue to the 60 cell line panel. The 60 different human tumor lines are incubated with 5 different doses of compound and a sulforhodamine blue (SRB) assay is performed after 48 hours to determine cytotoxicity. From the 5 point curve, the following concentrations are extrapolated: GI50 (inhibits growth by 50%), TGI (totally inhibits growth), LC50 (kills 50% of cells). For the specific screening methods from the DTP website, visit: http://www.dtp.nci.nih.gov/branches/btb/ivclsp.html. Compounds shown to have anticancer activity in cell lines within the NCI 60 panel may then move on to animal trials and if successful, may eventually move on to be tested in clinical trials. Additional experimental details for our primary in vitro antimicrobial assays against Mycobacterium tuberculosis and Plasmodium falciparum have been previously described.[26,27]
Supplementary Material
Acknowledgments
We thank Dr. Mark W. Miller (UPR Institute of Neurobiology) and Dr. Ileana I. Rodríguez (UPR-RP) for providing logistic support during the collection of A. dactylomela. The assistance of Juan A. Santana during molecular modeling studies is gratefully acknowledged. The National Cancer Institute (NCI), the National Institute of Allergy and Infectious Diseases (NIAID), the Tuberculosis/Antimicrobial Acquisition & Coordinating Facility (TAACF), and the Institute for Tropical Medicine and Health Sciences (Panama) provided in vitro cytotoxicity, antiviral, antituberculosis, and antimalarial activity data, respectively. Highresolution EI, ESI, and FAB mass spectral determinations were provided by the Mass Spectroscopy Laboratory of the University of Illinois at Urbana-Champaign. B. V. thanks the UPR-RISE Fellowship Program and the Puerto Rico Cancer Center for financial support. This work was partially supported by the NIH-SCORE Program (Grant S06GM08102) of the University of Puerto Rico.
Footnotes
Supporting information for this article is available on the WWW under http://www.eurjoc.org/ or from the author.
Supporting Information: 1H and 13C NMR spectra and representative 2D NMR data (1H–1H COSY, HMQC, HMBC, and NOESY) for compounds 1 and 2. Complete 1H and 13C NMR spectral data for aplysqualenol A acetate (3) (Table 4) and aplysqualenol B diacetate (4) (Table 5). Atom coordinates for the minimum energy structures of stereoisomer models 1a–1d, as well as statistical information regarding computational results (such as the calculated 13C chemical shifts, the corrected values, and Δδ) are summarized in Tables 6-9.
References
- [1].Blunt JW, Copp BR, Hu W-P, Munro MHG, Northcote PT, Prinsep MR. Nat. Prod. Rep. 2009;26:170–244. doi: 10.1039/b805113p. and previous articles in this series. [DOI] [PubMed] [Google Scholar]
- [2].a) Wessels M, Konig GM, Wright AD. J. Nat. Prod. 2000;67:920–928. doi: 10.1021/np9905721. [DOI] [PubMed] [Google Scholar]; b) Dias T, Brito I, Moujir L, Paiz N, Darias J, Cueto M. J. Nat. Prod. 2005;68:1677–1679. doi: 10.1021/np050240y. [DOI] [PubMed] [Google Scholar]; c) Manzo E, Ciavatta ML, Gavagnin M, Puliti R, Mollo E, Guo YW, Mattia CA, Mazzarella L, Cimino G. Tetrahedron. 2005;61:7456–7460. [Google Scholar]
- [3].Ferrigni NR, Meyer BN, Putnam JE, Jacobsen LB, Nichols DE, McLaughlin JL. J. Med. Plant Res. 1982;45:31–34. [PubMed] [Google Scholar]
- [4].Fernández JJ, Souto ML, Norte M. Nat. Prod. Rep. 2000;17:235–246. doi: 10.1039/a909496b. [DOI] [PubMed] [Google Scholar]
- [5].Blunt JW, Hartshorn MP, McLennan TJ, Munro MHG, Robinson WT, Yorke SC. Tetrahedron Lett. 1978:69–72. [Google Scholar]
- [6].Sakemi S, Higa T, Jefford CW, Bernardinelli G. Tetrahedron Lett. 1986;27:4287–4290. [Google Scholar]
- [7].Manzo E, Gavagnin M, Bifulco G, Cimino P, Di Micco S, Ciavatta ML, Guo YW, Cimino G. Tetrahedron. 2007;63:9970–9978. [Google Scholar]
- [8].Pfeffer PE, Valentine KM, Parrish FW. J. Am. Chem. Soc. 1979;101:1265–1274. [Google Scholar]
- [9].Squalene-derived polyethers having the usual R* configuration at C14 (i.e. compounds 5–8) also display H-14 as a doublet of doublets with two markedly different coupling constant values, consistent with a chair/twist-boat B–Crings system. Notwithstanding, the abnormal relative stereochemistry at C14 in 1 is substantiated by noticeable variations in 13C NMR chemical shifts about the 2,7-dioxabicyclo[4.4.0]decane ring. For instance, in 1 C27 resonates at 14.8 ppm, whereas all related congeners with the usual C14(R*) configuration show this signal above 20 ppm. This modification in stereochemistry also causes C11 to shift downfield (i.e., from 76.3 ppm in 7 to 80.8 ppm in compound 1). A reversal in relative stereochemistry at C14 also causes the nearby C28 methyl in 1 to resonate at 18.1 ppm, rather than 22.9 or 23.0 ppm as seen in 5 and 7, respectively, in spite of the fact that all three compounds have the C15(R*) configuration.
- [10].The minimum energy configuration for 1 was identified using a Monte Carlo conformational search with the MMFF force field as implemented in Spartan 04 software package (Wavefunction Inc., Irvine, CA).
- [11].a) Hohenberg P, Kohn W. Phys. Rev. 1964;136:B864–B871. [Google Scholar]; b) Kohn W, Sham LJ. Phys. Rev. 1965;140:A1133–A1138. [Google Scholar]
- [12].We attempted to confirm the proposed structure of aplysqualenol A (1) by X-ray crystallography upon chemical transformation into a crystalline derivative. Thus, after heating a mixture of 1 and 4-bromophenyl isocyanate in toluene for 3 h, the desired carbamate was obtained in excellent yield, but it failed to crystallize in all of the solvents tried.
- [13].Porto S, Durán J, Seco JM, Quiñoá E, Riguera R. Org. Lett. 2003;5:2979–2982. doi: 10.1021/ol034853i. [DOI] [PubMed] [Google Scholar]
- [14].The procedure is based on the binding of MTPA or MPA to a resin in such a way that, when attacked by the chiral molecule (i.e., compound 1), the reagent part acts as an electrophile and liberates the corresponding ester derivative into the solution while the solid matrix behaves as the leaving group. The NMR spectra of the expected derivatives are then obtained without any separation, workup, or manipulation.
- [15].The observation that acetylation at C18 and C30 of aplysqualenol B takes place readily at 25 °C points to the fact that the hydrogen bonds between the donor hydrogens and the acceptor atoms in 2 have less than ideal geometry.
- [16].Spartan 04 for Macintosh. Wavefunction Inc.; Irvine, CA: [Google Scholar]
- [17].Frisch MJ, Trucks GW, Schlegel HB, Scuseria GE, Robb MA, Cheeseman JR, Montgomery JA, Jr., Vreven T, Kudin KN, Burant JC, Millam JM, Iyengar SS, Tomasi J, Barone V, Mennucci B, Cossi M, Scalmani G, Rega N, Petersson GA, Nakatsuji H, Hada M, Ehara M, Toyota K, Fukuda R, Hasegawa J, Ishida M, Nakajima T, Honda Y, Kitao O, Nakai H, Klene M, Li X, Knox JE, Hratchian HP, Cross JB, Adamo C, Jaramillo J, Gomperts R, Stratmann RE, Yazyev O, Austin AJ, Cammi R, Pomelli C, Ochterski JW, Ayala PY, Morokuma K, Voth GA, Salvador P, Dannenberg JJ, Zakrzewski VG, Dapprich S, Daniels AD, Strain MC, Farkas O, Malick DK, Rabuck AD, Raghavachari K, Foresman JB, Ortíz JV, Cui Q, Baboul AG, Clifford S, Cioslowski J, Stefanov BB, Liu G, Liashenko A, Piskorz P, Komaromi I, Martin RL, Fox DJ, Keith T, Al-Laham MA, Peng CY, Nanayakkara A, Challacombe M, Gill PMW, Johnson B, Chen W, Wong MW, González C, Pople JA. Gaussian 03: Revision C.02. Gaussian, Inc.; Wallingford CT: 2004. [Google Scholar]
- [18].a) Cimino P, Gómez-Paloma L, Duca D, Riccio R, Bifulco G. Magn. Reson. Chem. 2004;42:S26–S33. doi: 10.1002/mrc.1410. [DOI] [PubMed] [Google Scholar]; b) Bifulco G, Dambruoso P, Gómez-Paloma L, Riccio R. Chem. Rev. 2007;107:3744–3779. doi: 10.1021/cr030733c. [DOI] [PubMed] [Google Scholar]
- [19].a) Kigoshi H, Ojika M, Shizuri Y, Niwa H, Yamada K. Tetrahedron Lett. 1982;23:5413–5414. [Google Scholar]; b) De Napoli L, Fattorusso E, Magno S, Mayol L. Phytochemistry. 1982;21:782–784. [Google Scholar]
- [20].Alternatively, methanolysis at the C22–C23 epoxide could explain the presence of the C22-OCH3 group implying that the genesis of this functionality in 1 and 2 could be artifactual.
- [21].Rogers CN, De Nys R, Charlton TS, Steinberg PD. J. Chem. Ecol. 2000;26:721–743. [Google Scholar]
- [22].Differences in cytotoxicity detected among compounds of this class have been related to the different conformations adopted by the C14 flexible side chain, see: Norte M, Fernández JJ, Souto ML, Gavin JA, García-Grávalos MD. Tetrahedron. 1997;53:3173–3178.
- [23].Parallel configurational studies indicate that 1, with an equatorial arrangement of the C15 to C24 side chain, is about 8.84 kcal/mol more stable than its corresponding C14-epimer.
- [24].Primary Epstein-Barr virus infection in childhood is usually mild but in later life usually manifests as infectious mononucleosis.
- [25].a) Sims JJ, Lin GHY, Wing RM. Tetrahedron Lett. 1974:3487–3490. [Google Scholar]; b) González AG, Darias J, Díaz A, Fourneron JD, Martín JD, Pérez C. Tetrahedron Lett. 1976:3051–3054. [Google Scholar]
- [26].Collins LA, Franzblau SG. Antimicrob. Agents Chemother. 1997;41:1004–1009. doi: 10.1128/aac.41.5.1004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [27].Corbett Y, Herrera L, González J, Cubilla L, Capson T, Colley PD, Kursar TA, Romero LI, Ortega-Barria E. J. Trop. Med. Hyg. 2004;70:119–124. [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.




