Qian Xing‡
ac,
Li-She Gan‡b,
Xiao-Feng Mou‡a,
Wei Wanga,
Chang-Yun Wangad,
Mei-Yan Weia and
Chang-Lun Shao*ad
aKey Laboratory of Marine Drugs, The Ministry of Education of China, School of Medicine and Pharmacy, Ocean University of China, Qingdao 266003, People's Republic of China. E-mail: shaochanglun@163.com
bCollege of Pharmaceutical Sciences, Zhejiang University, Hangzhou 310058, People's Republic of China
cHuantai Food and Drug Administration, Zibo 256400, People's Republic of China
dLaboratory for Marine Drugs and Bioproducts, Qingdao National Laboratory for Marine Science and Technology, Qingdao 266200, People's Republic of China
First published on 22nd February 2016
Marine life forms are an important source of structurally diverse and biologically active natural products. As a unique case of both enantiomeric and atropisomeric isomers being present in one marine natural product structure, two new dichlorinated diphenylmethanes containing both point and axial chirality, (±)-pestalachlorides E (1) and F (2), were isolated from a marine-derived Pestalotiopsis (ZJ-2009-7-6) fungus. Both of them showed potent antifouling activities against the larval settlement of the barnacle Balanus amphitrite at nontoxic concentrations with EC50 values of 1.65 and 0.55 μg mL−1, respectively, and antifouling activity was detected for the first time for this class of metabolites.
700 new marine natural compounds have been obtained from marine organisms collected from 1300 genera across 28 phyla.1,2 Marine natural products have proven to be rich sources of structurally novel and biologically active compounds that have become significant chemical entities for drug discovery.3,4 To date, there are seven marine natural products and thirteen marine natural products inspired compounds that are FDA-approved agents or in clinical trial.5
Marine-derived microorganisms have proved to be a rich and important source of drug candidates.1,6 Dichlorinated benzophenones (diphenylmethanes) represent a rare family of marine natural products that only six in the class have been reported to date, including pestalone,7 desmethyl pestalone,8 and pestalachlorides A–D.9,10 As part of a continuing program to evaluate bioactive marine natural products as potential drug leads from marine invertebrate derived microorganisms,11–14 further chemical investigation15,16 of laboratory cultures of the Pestalotiopsis fungus (ZJ-2009-7-6) resulted in the isolation of two new chlorinated enantiomeric diphenylmethanes, (±)-pestalachlorides E (1) and F (2) (Fig. 1). They both exist as a pair of separable enantiomers, which in turn display additional atropisomerism. The planar structures and relative configurations were first identified by comprehensive analysis of spectroscopic data, including 2D NMR and single crystal X-ray diffraction data. After chiral HPLC separation of (+)-1/(−)-1 and (+)-2/(−)-2, the inseparable atropisomeric features of (+)-1 and (−)-1, (+)-2 and (−)-2 caused by restricted rotation of the C8–C9 bond, were studied theoretically as M and P isomers by potential energy surface scans and energy barrier calculations at the B3LYP/6-31G level. Furthermore, DFT and TDDFT calculations of their CD and NMR spectra allowed the assignments of their absolute configurations. The isolation, structure elucidation, chiral resolution, and theoretical studies of the atropisomeric features of compounds 1 and 2 (Fig. 1) are presented herein, along with their potent antifouling activities against the larval settlement of barnacle Balanus amphitrite.
:
1. Close examination of the NMR data allowed the interpretation and assignments of all protons to both isomers. For the major isomer, the existence of one hydrogen-bonded phenolic proton (δH 11.86, 1H, s), one aldehyde group (δH 9.58, 1H, s), one aromatic proton (δH 6.25, 1H, s), one methoxy group (δH 3.15, 3H, s), one aromatic methyl group (δH 2.42, 3H, s), two methine protons (δH 5.27 and 2.65), two methylene protons (δH 3.21 and 2.87), and two methyl groups (δH 1.26 and 1.20) were observed. The 13C NMR and DEPT spectra showed 21 carbon signals, including one aldehyde group (δC 193.8) and twelve aromatic carbons representing two phenyl rings, which took up nine of the ten degrees of unsaturation. The structure of 1 was then suggested to contain a third ring. Moreover, the position of the hydroxyisopropyl group was located at C-2′ in ring B supported by the HMBC correlations from H-1′ and H-2′ to C-3′, and from H-4′ and H-5′ to C-2′. Finally, these HMBC correlations from H-8 to C-9, C-10, and C-14, and from H-2′ to C-9 enabled the establishment of the C-8–C-9 linkage.
The structural architecture of 1 was then established by comprehensive analysis of 2D NMR spectroscopic data, especially 1H–1H COSY and HMBC correlations. The aromatic ring A and the five-membered ring B were established by 1H–1H COSY correlations of H-8/H-2′ and H-2′/H-1′ together with HMBC correlations from H-8 to C-2, C-6, and C-7, from H-1′ to C-5, C-6, and C-7, and from H-2′ to C-6 and C-7 (Fig. 1). According to the above data, the planar structure was established as shown in 1. Furthermore, the relative configuration of 1 was assigned as trans on the basis of the NOESY correlations between H-8 and the H3-4′ and H3-5′ of the two methyls on the side chain. For the minor component of 1, the same structure can be concluded. The most distinguishable differences in 1H NMR of the two isomers are the chemical shifts of the two methoxy groups at C-14 (CH3-16), indicating the possible atropisomerism in 1. It should be mentioned that the 13C NMR data of the minor component of 1 were not observed because of the minor quantity (Table 2).
| Position | 1 | 2 | ||
|---|---|---|---|---|
| Major | Minor | Major | Minor | |
| 1 | 9.58 (1H, s) | 9.73 (1H, s) | 9.49 (1H, s) | 9.60 (1H, s) |
| 4 | 6.25 (1H, s) | 6.17 (1H, s) | 6.21 (1H, s) | 6.28 (1H, s) |
| 8 | 5.27 (1H, d, 5.5) | 5.45 (1H, d, 5.5) | 5.11 (1H, d, 6.0) | 5.33 (1H, brs) |
| 15 | 2.42 (3H, s) | 2.42 (3H, s) | 2.43 (3H, s) | 2.43 (3H, s) |
| 16 | 3.15 (3H, s) | 4.13 (3H, s) | 3.89 (3H, s) | 3.17 (3H, s) |
| 1′ | 3.21 (1H, dd, 16.3, 10.1) | 3.29 (1H, overlapped) | 2.82 (1H, dd, 17.2, 7.5) | 2.82 (1H, dd, 17.2, 7.5) |
| 2.87 (1H, dd, 16.3, 4.2) | 2.87 (1H, overlapped) | 3.29 (1H, overlapped) | 3.29 (1H, overlapped) | |
| 2′ | 2.65 (1H, ddd, 10.1, 5.5, 4.2) | 2.65 (1H, overlapped) | 3.47 (1H, overlapped) | 3.27 (1H, overlapped) |
| 4′ | 1.26 (3H, s) | 1.26 (3H, s) | 4.72 (2H, s) | 4.72 (2H, s) |
| 5′ | 1.20 (3H, s) | 1.20 (3H, s) | 1.81 (3H, s) | 1.78 (3H, s) |
| 3-OH | 11.86 (H, brs) | 12.04 (H, brs) | 11.92 (1H, s) | 11.90 (1H, s) |
| 5-OH | — | — | 9.74 (1H, s) | 9.89 (1H, s) |
| 10-OH | — | — | 8.13 (1H, s) | 8.66 (1H, s) |
| Position | 1 | 2 | |
|---|---|---|---|
| Major | Major | Minor | |
| 1 | 193.8, CH | 192.9, CH | 193.6, CH |
| 2 | 111.8, C | 112.0, C | 112.1, C |
| 3 | 166.2, C | 165.8, C | 166.1, C |
| 4 | 101.7, CH | 101.4, CH | 101.6, CH |
| 5 | 162.7, C | 161.9, C | 162.3, C |
| 6 | 125.1, C | 123.9, C | 124.1, C |
| 7 | 152.9, C | 151.9, C | 152.3, C |
| 8 | 44.4, CH | 47.3, CH | 46.8, CH |
| 9 | 129.9, C | 126.7, C | 128.1, C |
| 10 | 149.2, C | 151.0, C | 151.2, C |
| 11 | 119.1, C | 119.0, C | 118.0, C |
| 12 | 134.7, C | 134.7, C | 134.8, C |
| 13 | 121.4, C | 120.5, C | 121.4, C |
| 14 | 155.4, C | 153.1, C | 155.5, C |
| 15 | 18.2, CH3 | 18.2, CH3 | 19.7, CH3 |
| 16 | 60.5, CH3 | 62.3, CH3 | 60.4, CH3 |
| 1′ | 32.2, CH2 | 34.6, CH2 | 34.7, CH2 |
| 2′ | 59.0, CH | 54.1, CH | 54.9, CH |
| 3′ | 73.6, C | 148.0, C | 148.4, C |
| 4′ | 28.2, CH3 | 112.3, CH2 | 110.4, CH2 |
| 5′ | 26.0, CH3 | 18.2, CH3 | 18.2, CH3 |
Compound 2 (ref. 18) was also obtained as a colorless crystal in acetone. Its HRESIMS data showed a molecular formula of C21H20Cl2O5 and one more degree of unsaturation than that in 1. The 1H and 13C NMR spectra of 2 also displayed duplicated signals in a ratio of approximately 5
:
4. Compared to the 1H NMR spectrum of 1, a singlet for a terminal double bond at δH 4.72 (2H, s) was exhibited replacing that methyl group in 1, which was also confirmed by two olefinic carbon signals resonating at δC 148.4 and 110.4 in its 13C NMR spectrum. Thus the gross structure of 2 was concluded as the 3′,4′-dehydration product of pestalachloride E (1). The presence of the C-3′–C-4′ double bond was further supported by the HMBC correlations from H-5′ to C-3′ and C-4′ (Fig. 1). Furthermore, the relative configuration between H-8 and H-2′ in 2 was also determined as trans on the basis of the NOESY data. Therefore, the structure of 2 was established and named as pestalachloride F. Similar to 1, the major and minor components in 2 may also be atropisomers.
Fortunately, by slow crystallization in acetone, single crystals of 1 suitable for X-ray diffraction analysis were obtained. The structure and the trans relative configuration of 1 was then confirmed by the single-crystal X-ray diffraction data19 (Fig. 2a). Further analysis of the X-ray data of 1 revealed that it possesses a centrosymmetric space group C2/c, indicating a racemic nature.
![]() | ||
| Fig. 2 (a) Perspective ORTEP drawing for 1. (b) View of 2-D sheet structure. The broken lines represent the hydrogen bonds [O–H⋯O and C–H⋯O (blue)]. | ||
In order to fully assign the NMR data and clarify the racemic atropisomer nature of 1 and 2, HPLC analysis of 1 and 2 on a chiral column (Kromasil 5-TBB) were carried out firstly. Two distinct chromatographic peaks with a ratio of 1
:
1 were found and isolated from both 1 and 2. Subsequently, chemical shifts of the most distinguishable signal of H3-16 can be rationally explained by the shielding and deshielding zones of the benzene ring A (Fig. 2a) based on the X-ray structure. As shown in Fig. 2a for the M isomer X-ray 3D structure of 1, the methoxy group situates directly over the shielded area of aromatic ring A. Meanwhile, intramolecular hydrogen bonds between OH-10 and OH-3′, which can be observed in X-ray data (Fig. 2b), may also stabilize this M isomer. Thirdly, variable temperature NMR experiments has shown that these minor atropisomers can be transformed to the corresponding major one at 75 °C in DMSO-d6 (Fig. S13†). Because the M isomer in (−)-1 and the P isomer in (+)-1 are enantiomers, full assignments still cannot be done without the assignment of absolute configuration.
The absolute configurations of (+)-1, (−)-1, (+)-2, and (−)-2 as well as the inseparable P and M atrop-diastereomers were studied theoretically. Conformational analysis was carried out for 1 and 2 via Monte Carlo searching using molecular mechanism with MMFF94 force field in the SPARTAN 08 software package. A 1D PES scan on the dihedral angle of 10-9-8-2′ in representing conformer of M isomers of 1 and 2 were then performed by modredundant optimization at the semi-empirical AM1 level in Gaussian 09 software package.20 The resulted transitional state conformers and all the lowest energy conformers were re-optimized using DFT at the B3LYP/6-31+G(d) level in gas phase. All the ΔG (relative Gibbs energy barrier) for M–P conversion is greater than 20 kcal mol−1, which confirmed the existence of M/P isomers at room temperature (Fig. 4).21 Subsequently, TD DFT calculations of their ECD spectra showed that the inseparable M and P atrop-diastereomers have similar theoretical ECD spectra, which gave the possibility for the determination of the absolute configurations at C-8 and C-2′ of the chiral column isolates. As shown in Fig. 3, the calculated 8S, 2′R configuration of compounds 1 and 2 showed first positive and second negative Cotton effects, and correspondingly, the experimental ECD spectra of (−)-1 and (−)-2 also exhibited curves with the same pattern, which indicated the absolute configuration of (−)-1 and (−)-2 as 8S,2′R. The 8R,2′S configuration was then assigned to compounds (+)-1 and (+)-2.
![]() | ||
| Fig. 3 Experimental ECD spectra (black lines) and calculated ECD spectra of M and P atropisomers (red and blue lines) of 1 and 2. | ||
As indicated by the 1H NMR integration values of the M and P atrop-diastereomers, their ratio are 6
:
1 in (−)-1 and (+)-1, and 5
:
4 in (−)-2 and (+)-2 (Table 1). The major and minor components in the atrop-diastereomers were assigned based on their relative Gibbs free energy and Boltzmann distribution theory, as well as NMR theoretical studies of the four representative lowest energy conformers. To predict the ratio of M and P diastereomers in (−)-1 and (−)-2, all the calculated lowest energy M and P conformers were weighted together in the Boltzmann function by their relative Gibbs free energy and calculated separately in two groups of M and P, the results showed a ratio of P
:
M = 2.73
:
1 for (−)-1, and P
:
M = 1
:
2.64 for (−)-2. Based on the above conclusion, the major component in (−)-1 is the P isomer, while the M isomer is the major component in (−)-2. The assignments were further confirmed by NMR calculation. As indicated in Table S1,† the theoretical chemical shifts at CH3-16 of four representative lowest energy conformers, (8S,2′R)-1M-C3, (8S,2′R)-1P-C8, (8S,2′R)-2M-C1, and (8S,2′R)-2P-C1 were again in good agreements with the experimental data (Fig. S28–31†). Because the M isomer in (−)-1 and P isomer in (+)-1 are enantiomers, and (+)-1 and (−)-1 show the same 1H NMR spectra, the ratio of P and M atropisomers in (+)-1 should be just on the contrary of (−)-1. The same conclusion also can be drawn for (+)-2. The ratio of P and M atropisomers in 2 is obviously different from that in 1, which may be caused by the loss of intramolecular hydrogen bond.
Compounds (±)-1 and (±)-2 were also evaluated for antifouling activity against the larval settlement of the barnacle Balanus amphitrite.22 Both of them showed potent antifouling activities at nontoxic concentrations with EC50 values of 1.65 and 0.55 μg mL−1 (SeaNine 211 as a positive control, IC50 = 1.23 μg mL−1, LC50/EC50 = 20.3), respectively, which were much lower than the standard requirement of an EC50 of 25 μg mL−1 established by the U.S. Navy program as an efficacious level for natural antifouling agents. To the best of our knowledge, this is the first report of antifouling activities for this class of metabolites. Furthermore, a compound with therapeutic ratio (LC50/EC50) > 15 is often considered to be a nontoxic antifouling compound.23 Compounds (±)-1 and (±)-2 have high therapeutic ratios LC50/EC50 > 30.3 and 18.2, respectively, which were higher than 15, suggesting that they might be regarded as environmentally benign antifouling agents.
:
1). The organic extracts were combined and concentrated under vacuum to afford a dry crude extract.
:
1) and purified repeatedly by semi-preparative HPLC eluted with 68% of MeOH/H2O at a flow rate of 2.0 mL min−1 to give 1 (3.2 mg). Compound 1 was then obtained by chiral preparative HPLC eluting with 85% of n-hexane/isopropanol at a flow rate of 2.0 mL min−1 to give (+)-1 (1.7 mg) and (−)-1 (1.5 mg). Similarly, 2 was separated by chiral preparative HPLC eluting with 95% of n-hexane/isopropanol at a flow rate of 2.0 mL min−1 to give (+)-2 (0.9 mg) and (−)-2 (1.0 mg).
ε) 209 (4.13), 292 (3.86), 341 (3.19) nm; IR (KBr) νmax 3398, 2925, 2855, 2369, 1630, 1381, 1282, 1140 cm−1; 1H NMR and 13C NMR, see Tables 1 and 2; ESI-MS m/z 441.2 [M + H]+, 423.1 [M + H − H2O]+; HRESI-MS m/z 441.0857 [M + H]+ (calcd for C21H23O6Cl2, 441.0866). (+)-Pestalachloride E [(+)-(1)]: [α]26D +24.0 (c 0.065, MeOH). (−)-Pestalachloride E [(−)-(1)]: [α]26D −20.0 (c 0.065, MeOH).
ε) 210 (3.81), 292 (3.52), 337 (3.03) nm; IR (KBr) νmax 3484, 2364, 2342, 1631, 1377, 1289, 728 cm−1; 1H NMR and 13C NMR, see Tables 1 and 2; ESI-MS m/z 423.14 [M + H]+; HRESI-MS m/z 423.0759 [M + H]+ (calcd for C21H21O5Cl2, 423.0761). (+)-Pestalachloride F [(+)-(2)]: [α]26D +18.0 (c 0.045, MeOH). (−)-Pestalachloride F [(−)- (2)]: [α]26D −15.3 (c 0.045, MeOH).Footnotes |
| † Electronic supplementary information (ESI) available. CCDC 1052426. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c6ra00374e |
| ‡ These authors contributed equally. |
| This journal is © The Royal Society of Chemistry 2016 |