Next Article in Journal
Petromylidenes A–C: 2-Alkylidene Bile Salt Derivatives Isolated from Sea Lamprey (Petromyzon marinus)
Previous Article in Journal
Anti-Obesity Effects of Collagen Peptide Derived from Skate (Raja kenojei) Skin Through Regulation of Lipid Metabolism
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

α-Glucosidase Inhibitors: Diphenyl Ethers and Phenolic Bisabolane Sesquiterpenoids from the Mangrove Endophytic Fungus Aspergillus flavus QQSG-3

1
School of Chemistry, Sun Yat-Sen University, Guangzhou 510275, China
2
South China Sea Bio-Resource Exploitation and Utilization Collaborative Innovation Center, School of Marine Sciences, Sun Yat-Sen University, Guangzhou 510006, China
3
State Key Laboratory of Applied Microbiology, Southern China, Guangdong Institute of Microbiology, Guangzhou 510075, China
4
CAS Key Laboratory of Tropical Marine Bio-Resources and Ecology, Guangdong Key Laboratory of Marine Materia, RNAM Center for Marine Microbiology, South China Sea Institute of Oceanology, Chinese Academy of Sciences, Guangzhou 510301, China
*
Authors to whom correspondence should be addressed.
Mar. Drugs 2018, 16(9), 307; https://doi.org/10.3390/md16090307
Submission received: 20 August 2018 / Revised: 27 August 2018 / Accepted: 29 August 2018 / Published: 1 September 2018

Abstract

:
Two new diphenyl ethers (1 and 2) and four new phenolic bisabolane sesquiterpenoids (36), together with five known related derivatives, were isolated from the culture of the endophytic fungus Aspergillus flavus QQSG-3 obtained from a fresh branch of Kandelia obobata, which was collected from Huizhou city in the province of Guangdong, China. The structures of compounds 16 were determined by analyzing NMR and HRESIMS data. The absolute configurations of 5 and 6 were assigned by comparing their experimental ECD spectra with those reported for similar compounds in the literature. All isolates were evaluated for their α-glucosidase inhibitory activity, of which compounds 3, 5, 10, and 11 showed strong inhibitory effects with IC50 values in the range of 1.5–4.5 μM.

Graphical Abstract

1. Introduction

Chemical investigations on mangrove endophytic fungi have generated considerable attention from natural product researchers due to their unique ecosystem, which leads to the isolation of secondary metabolites with diverse structures and excellent biological activities. The genus Aspergillus, a ubiquitous fungus, is recognized as a rich source of biomolecules for constructing novel skeletons for drug discovery, including terpenoids [1,2,3], cyclopeptides [4,5], alkaloids [6], butenolides [7], coumarins [8], and quinones [9]. As a class of terpenoids, their diverse biological activities gave rise to phenolic bisabolane sesquiterpenoids getting more attention, including antibacterial activities [10,11], acetylcholinesterase inhibition [12], cytotoxicity [13], and antioxidant activities [14].
As part of our efforts to discover potent and new α-glucosidase inhibitors [15,16,17,18], Aspergillus flavus QQSG-3 was chemically investigated, which was isolated from a fresh branch of Kandelia obovata collected from Huizhou, Guangdong, China. The EtOAc extract afforded two new diphenyl ethers (1 and 2), four new phenolic bisabolane sesquiterpenoids (36), and five known compounds (711) (Figure 1). The α-glucosidase inhibitory activities of all isolates were tested. Compounds 3, 5, 10, and 11 exhibited strong inhibitory effects compared to the positive control acarbose. In this paper, we describe the purification, structure elucidation, and bioactivities of these compounds.

2. Results

Compound 1 was obtained as a colorless oil. The molecular formula was deduced to be C14H14O4 based on the protonated molecular ion peak at m/z 247.0962 ([M + H]+, calcd. for 247.0964) from HRESIMS, indicating eight degrees of unsaturation. The 1H NMR (Table 1) showed five aromatic protons (δH 6.43 (1H, d, J = 2.6 Hz), 6.41 (1H, d, J = 2.6 Hz), 6.22 (1H, dd, J = 3.6, 2.1 Hz), 6.34 (1H, brs), and 6.34 (1H, brs)), and two methyl groups (δH 2.24 (3H, s) and 2.22 (3H, s)). The 13C NMR spectrum (Table 1) displayed 12 aromatic carbons (δC 159.6, 156.5, 149.4, 144.0, 141.0, 138.5, 125.5, 114.1, 111.1, 110.5, 105.5, and 102.3) and two methyl carbons (δC 21.6, 15.8). The 1D NMR data compared to the known compound cordyol C [19] indicated the same diphenyl ether skeleton was present in compound 1. The main difference was the obvious upfield shift of H-2 (δH 6.43) and C-7 (δC 15.8), which suggested the different substituted location of the hydroxyl group (4-OH in 1 and 2-OH in cordyol C). Combined with HMBC correlations of H-6/C-1, C-2, C-4, C-5, and C-7; and H-2/C-1, C-3, and C-4 (Figure 2), the 2,3,5-trihydroxytoluene unit was established. The 3,5-dihydroxytoluene unit was supposed based on the 1H NMR signals of δH 6.22 (dd, J = 3.6, 2.1 Hz, H-2′), 6.34 (brs, H-4′), and 6.34 (brs, H-6′), which was further confirmed by the correlations from H3-7′ to C-3′, C-4′, and C-5′; from H-2′/H-4′ to C-3′; and from H-2′/H-6′ to C-1′. Thus, compound 1 was identified as shown in Figure 1.
Compound 2 was isolated as a colorless oil. Its molecular formula was established as C19H21O8 on the basis of HRESIMS data (m/z 377.1245 [M − H] (calcd. for 377.1242)). Compound 2 contained two structural moieties (a diphenyl ether moiety similar to 1 and an erythritol moiety) deduced from the analysis of 1D and 2D NMR signals. The weak four-bond HMBC correlations from H-2 and H-6 to C-8 (Figure 2) revealed the replacement of OH-4 in 1 with the carbonyl carbon, C-8 (δC 172.4). This was further confirmed by the downfield shift of C-3 (δC 164.9) and upfield shift of C-4 (δC 109.9) caused by the electron-donating conjugative effect of C-8. The erythritol moiety was confirmed by the HSQC data and COSY correlations from H2-9 to H-10, H-10 to H-11, and H-11 to H2-12 (Figure 2). The moiety has a linkage with C-8 that was verified by the key correlation from H2-9 to C-8 in HMBC. However, acid hydrolysis of compound 2 could not be performed due to its limited quantity, and the absolute configuration was not assigned.
Compound 3 was a pale yellow oil with the molecular formula C22H28O3 established from HRESIMS at m/z 339.1963 [M − H] (calcd. for 339.1965). The 1H NMR spectrum displayed characteristic aromatic protons of 1,2,4-trisubstituted benzene (δH 6.88 (d, J = 7.7 Hz), 6.63 (dd, J = 7.7, 1.4 Hz), 6.60 (d, J = 1.4 Hz)), 1,2,3,5-tetrasubstituted benzene (δH 6.47 (d, J = 1.3 Hz), 6.34 (dd, J = 1.3 Hz)), and a trisubstituted double bond (δH 5.38 (td, J = 7.2, 1.3 Hz)) (Table 2). The key HMBC correlations from H2-15 to C-4, C-5, C-6, C-16, C-17, and C-21 revealed that two benzene moieties were connected to each other via a methylene (C-15) between C-5 and C-20 (Figure 2). The 1H–1H COSY cross peaks from H-8 to H-13 established the only spin coupling system, which could be constructed to the side chain based on the HMBC correlations between H3-14 to C-2, C-7, and C-8 (Figure 2). H2-9 displayed a cross-peak with the olefinic proton H-8, which, as well as the correlations from H3-14 to C-7 and C-8 in HMBC, established a 6-methyl-2-heptenyl unit. This unit was attached at C-2 on the basis of the HMBC correlations of H-8/C-2. The E-geometry for the trisubstituted double bond was deduced from the signal enhancement of H2-9 upon irradiation of H3-14 in the NOEDIFF experiment [20] (Supporting Information Figures S21 and S22). Therefore, the configuration of 3 was determined.
Compound 4 was isolated as a colorless oil. Analysis of its HRESIMS data indicated compound 4 had the molecular formula C15H22O3, m/z 249.1498 [M − H] (calcd. for 249.1496). 1H and 13C NMR spectra were similar to those of compound 8, except for the downfield shift for C-11 (δC 71.4). These data suggested that the hydroxyl had a linkage with C-11. This deduction was further supported by the HMBC correlations from H3-12, H2-10, and H2-9 to C-11 (Figure 2). The E-geometry for the double bond was also deduced from the signal enhancement of H2-9 upon irradiation of H3-14 in the NOEDIFF experiment [20] (Supporting Information Figure S29).
The molecular formula of compound 5 was established as C30H46O5 by the HRESIMS ion at m/z 485.3271 [M − H] (calcd. for 485.3272). The 13C NMR spectrum displayed fifteen carbon signals that were close to those of sydonol, 9. The major difference was that the chemical shift of C-15 (C-15′) was changed from δC 64.8 in 9 (numbered as C-15) to δC 72.6 in compound 5 (Table 3). These data indicated that compound 5 was a dimeric analogue of sydonol 9. Moreover, the HMBC correlation from H2-15 to C-15 (C-15′) also supported the dimer construction. The ECD spectrum showed one negative Cotton effect (CE) at 209 nm and two positive CEs near 227 and 279 nm (Supporting Information Figure S36), which were identical to that of peniciaculin B [21]. Thus, the absolute configuration of compound 5 was speculated to be 7S, 7′S, which is the same as peniciaculin B.
Compound 6 was isolated as a colorless oil and had the molecular formula of C15H22O3, determined by HRESIMS data m/z 249.1497 [M − H] (calcd. for 249.1496) with five degrees of unsaturation. The 1D NMR and HSQC data displayed three aromatic protons at δH 7.14 (1H, d, J = 7.9 Hz), 6.78 (1H, dd, J = 7.9, 1.1 Hz), and 6.75 (1H, d, J = 1.1 Hz); three methyl protons at δH 1.50 (3H, s), 1.03 (3H, d, J = 6.7 Hz), and 0.93 (3H, d, J = 6.7 Hz); three methylene protons at δH 4.50 (2H, s); 2.14 (1H, m), 2.40 (1H, m); and 1.71 (1H, m), 1.86 (1H, m); two methine protons at δH 1.75 (1H, m) and 3.65 (1H, m); six aromatic carbons at δC 155.8, 142.8, 132.0, 127.2, 119.0, and 116.0; three methyl carbons at δC 29.5, 19.5, and 18.9; three methylene carbons at δC 64.9, 39.5, and 29.7; two methine carbons at δC 86.4 and 34.4; and one oxygenated quaternary carbon at δC 87.2 (Table 2). The NMR data of 6 were similar to those of the known compound (7R,10S)-7,10-epoxysydonic acid [22], except for the absence of the carboxyl carbon in 13C NMR and the presence of additional hydroxymethyl protons, H2-15 (δH 4.50), in the 1H NMR, revealing that the carboxylic was reduced into a hydroxymethyl group in 6. Analysis of the key correlations of H-10/C-7 in the HMBC data indicated that C-7 was linked to C-10 via an ether bond (Figure 2). C-14 and H-10 were oriented on the same side of the furan ring, which could be established based on the NOE correlations of H-10/H3-14. The absolute configuration of 6 was speculated as 7R,10S on the basis of the similar Cotton effects in the ECD spectrum compared (Supporting Information Figure S44) with that of (7R,10S)-7,10-epoxysydonic acid.
In addition, the structures of diorcinol (7) [23], (E)-5-(hydroxymethyl)-2-(6′-methylhept-2’-en-2’-yl) phenol (8) [24], sydonol (9) [25], peniciaculin A (10) [21], and expansol D (11) [26] were identified by comparing their NMR data with those reported in the literature.
The isolated compounds (111) were evaluated for their inhibitory activities against α-glucosidase (Table 4). The results displayed that compounds 3, 5, 10, and 11 were strong inhibitors with the IC50 values of 4.5, 3.1, 1.5, and 2.3 μM, respectively. Moreover, the activities of compounds 1 and 2 were better than that of acarbose (used as a positive control).

3. Experimental Section

3.1. General Experimental Procedures

Optical rotations were recorded in MeOH on an MCP 300 (Anton Paar, Shanghai, China) polarimeter at 25 °C. UV data were obtained on a PERSEE TU-1900 spectrophotometer (Persee, Beijing, China). IR spectra were measured in KBr on a Nicolet Nexus 670 spectrophotometer (Nicolet, Madison, WI, USA). CD data were measured on a J-810 spectropolarimeter (JASCO, Tokyo, Japan). The NMR spectra were performed on a Bruker Avance 500 spectrometer (1H/500 MHz, 13C/125 MHz, Bruker Bio Spin Corporation, Bellerica, MA, USA). All chemical shifts (δ) are given in ppm and coupling constants (J) are given in Hz. HRESIMS data were recorded on a Thermo Fisher Scientific Q-TOF mass spectrometer (Thermo Fisher Scientific, Waltham, MA, USA). Column chromatography (CC) was carried out on silica gel (200–300 mesh, Qingdao Marine Chemical Factory, Qingdao, China) and Sephadex LH-20 (Amersham Pharmacia, Piscataway, NJ, USA). Thin-layer chromatography (TLC) was performed on silica gel plates (Qingdao Huang Hai Chemical Group Co., G60, F-254, Qingdao, China). Phenomenex Luna (Phenomenex, Torrance, CA, USA) C18 column (250 × 10 mm, 5 μm) was used for semipreparative HPLC.

3.2. Fungal Materials

The fungus QQSG-3 investigated in this study was isolated from a fresh branch of the mangrove plant Kandelia obovata collected from Huizhou in Guangdong Province, China, in August 2015. Upon analysis of ITS sequence data of rDNA, the strain was identified as Aspergillus flavus, which had 99% sequence identity to that of Aspergillus flavus (GenBank FJ011545.1). A voucher specimen has been deposited at the Guangdong Microbial Culture Center (patent depository number GDMCC 60380).

3.3. Extraction and Isolation

The fungus Aspergillus flavus QQSG-3 was grown in 60 1000-mL Erlenmeyer flasks at 27 °C for 28 days, containing autoclaved rice solid-substrate medium composed of 50 g rice and 50 mL 3‰ saline water. After incubation, the mycelia were extracted with EtOAc and the extract was concentrated to yield 11.3 g of residue under reduced pressure. Then, the residue was eluted by a gradient of petroleum ether/EtOAc from 9:1 to 0:10 on silica gel CC (480 × 110 mm) and divided into ten fractions (Fr.1–Fr.10). Fr.3 (812 mg) was further eluted on silica gel CC (280 × 20 mm) by 3:7 (petroleum ether/EtOAc) to give compound 8 (147 mg) and five fractions (Fr.3.1–Fr.3.5). Fr.3.1 (215 mg) was subjected to Sephadex LH-20 CC (300 × 25 mm) and eluted with MeOH to obtain compound 1 (3.8 mg), 7 (107 mg) and 9 (10.8 mg). Fr.3.4 (7 mg) was purified by semipreparative RP-HPLC (MeOH/H2O, 70/30; 1.0 mL/min) to afford compound 4 (1.5 mg, tR = 17.5 min) and 6 (1.2 mg, tR = 20.3 min). Fr.4 (78 mg) was eluted (by petroleum ether/EtOAc, 4:7) to obtain seven fractions (Fr.4.1–Fr.4.7). Fr.4.2 (3.8 mg) was applied to Sephadex LH-20 CC (300 × 25 mm) and was eluted with MeOH/CD2Cl2 (1:1) to yield compounds 2 (1.1 mg) and 5 (1.3 mg). Fr.4.5 (11.5 mg) was further purified by CC over silica gel (200 × 15 mm) using MeOH/CD2Cl2 (1:40) to furnish compound 3 (1.4 mg), 10 (3.1 mg), and 11 (2.5 mg), respectively.
  • Compound 1: colorless oil; UV (MeOH) λmax (logε): 220 (4.23), 281 (3.47) nm; IR (KBr) νmax: 3371, 2934, 1614, 1476, 1307, 1139, 1046, 840 cm−1; 1H and 13C NMR data, see Table 1; HRESIMS m/z 247.0962 [M + H]+ (calcd. for 247.0964).
  • Compound 2: colorless oil; UV (MeOH) λmax (logε): 216 (5.12), 264 (3.55), 305 (1.36) nm; IR (KBr) νmax: 3362, 2915, 1672, 1590, 1461, 1306, 1266, 1168 cm−1; 1H and 13C NMR data, see Table 1; HRESIMS m/z 377.1245 [M − H] (calcd. for 377.1242).
  • Compound 3: pale yellow oil; UV (MeOH) λmax (logε): 213 (4.98), 285 (2.16) nm; IR (KBr) νmax: 3468, 2956, 1623, 1493, 1412, 1306, 1184 cm−1; 1H and 13C NMR data, see Table 2; HRESIMS m/z 339.1963 [M − H] (calcd. for 339.1965).
  • Compound 4: colorless oil; UV (MeOH) λmax (logε): 233 (4.27), 285 (3.63) nm; IR (KBr) νmax: 3414, 2931, 1550, 1317, 846 cm−1; 1H and 13C NMR data, see Table 2; HRESIMS m/z 249.1498 [M − H] (calcd. for 249.1496).
  • Compound 5: colorless oil; [ α ] D 25 + 3.2 (c 0.18, MeOH); UV (MeOH) λmax (logε): 203 (4.05), 221 (3.51), 280 (1.17) nm; IR (KBr) νmax: 3387, 2948, 1379, 1282, 772 cm−1; ECD λmax (∆ε): 209 (−1.20), 227 (+1.29), 279 (+1.65); 1H and 13C NMR data, see Table 3; HRESIMS m/z 485.3271 [M − H] (calcd. for 485.3272).
  • Compound 6: colorless oil; [ α ] D 25 + 22.3 (c 0.05, MeOH); UV (MeOH) λmax (logε): 202 (4.32), 218 (3.98), 279 (3.55) nm; IR (KBr) νmax: 3224, 2932, 1388, 1022, 872 cm−1; ECD λmax (∆ε): 210 (−0.20), 222 (+1.94), 278 (+0.98); 1H and 13C NMR data, see Table 2; HRESIMS m/z 249.1497 [M − H] (calcd. for 249.1496).

3.4. Biological Assays

α-Glucosidase was assayed according to standard procedures [27,28]. Compounds 111 and acarbose (positive control) were dissolved in DMSO, and enzyme (0.4 units/mL) and substrate (p-nitrophenyl-α-glucopyranoside, 5 mM) were dissolved in phosphate buffer solution (PBS, 100 mM, pH 7). Ten microliters of testing materials (in triplicate) were incubated for 10 min with 20 μL of enzyme stock solution and 60 μL of PBS. After incubation, 10 μL of substrate was added and incubated for 20 min at 37 °C. Absorbance at 405 nm was then determined.
The inhibitory activity of the isolates was determined as a percentage in comparison to a blank (DMSO) according to the following equation:
% α GHY = ( 1 A 405 t A 405 c ) × 100 %
where %αGHY is the percentage of inhibition, A405t is the corrected absorbance of the compound under testing (A405end − A405initial), and A405c is the absorbance of the blank (A405endblank − A405initialblank). The IC50 values of compounds were calculated by the nonlinear regression analysis.

4. Conclusions

In summary, two new diphenyl ether derivatives, 1 and 2, and four new phenolic bisabolane sesquiterpenoids, 36, along with five known compounds were isolated from the mangrove endophytic fungus Aspergillus flavus QQSG-3. Bioactivity assays displayed that phenolic bisabolane sesquiterpenoid derivatives 3, 5, 10, and 11 exhibited strong inhibitory activity against α-glucosidase, suggesting that they might have potential to be developed as α-glucosidase inhibitors. This is the first report of α-glucosidase inhibitory activity of diphenyl ethers and phenolic bisabolane sesquiterpenoids. Our findings enrich the diversity of their family and bioactivities.

Supplementary Materials

The following are available online at https://www.mdpi.com/1660-3397/16/9/307/s1, Figures S1–S44: The HRESIMS, NMR, and ECD spectra of compounds 16.

Author Contributions

Y.W. contributed to isolation and wrote the paper; Y.W. and Y.C. contributed to extraction, characterization, and biological testing of all the compounds; T.Y., W.C., X.H. and Y.P. contributed to supplementary NMR data analysis; Z.S. and Z.L. guided the whole experiment and revised the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (81741162, 2187713, 21472251), GDME (2018C004), the key project of the Natural Science Foundation of Guangdong Province (2016A040403091), and the Special Promotion Program for Guangdong Provincial Ocean and Fishery Technology (A201701C06).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, J.F.; Wei, X.Y.; Qin, X.C.; Tian, X.P.; Liao, L.; Li, K.M.; Zhou, X.F.; Yang, X.W.; Wang, F.Z.; Zhang, T.Y.; et al. Antiviral Merosesquiterpenoids Produced by the Antarctic Fungus Aspergillus ochraceopetaliformis SCSIO 05702. J. Nat. Prod. 2016, 79, 59–65. [Google Scholar] [CrossRef] [PubMed]
  2. Tan, Y.H.; Yang, B.; Lin, X.P.; Luo, X.W.; Pang, X.Y.; Tang, L.; Liu, Y.H.; Li, X.J.; Zhou, X.F. Nitrobenzoyl Sesquiterpenoids with Cytotoxic Activities from a Marine-Derived Aspergillus ochraceus Fungus. J. Nat. Prod. 2018, 81, 92–97. [Google Scholar] [CrossRef] [PubMed]
  3. Liu, Z.M.; Chen, Y.; Chen, S.H.; Liu, Y.Y.; Lu, Y.J.; Chen, D.N.; Lin, Y.C.; Huang, X.S.; She, Z.G. Aspterpenacids A and B, Two Sesterterpenoids from a Mangrove Endophytic Fungus Aspergillus terreus H010. Org. Lett. 2016, 18, 1406–1409. [Google Scholar] [CrossRef] [PubMed]
  4. Ma, Y.M.; Liang, X.A.; Zhang, H.C.; Liu, R. Cytotoxic and Antibiotic Cyclic Pentapeptide from an Endophytic Aspergillus tamarii of Ficus carica. J. Agric. Food Chem. 2016, 64, 3789–3793. [Google Scholar] [CrossRef] [PubMed]
  5. Zheng, J.K.; Xu, Z.H.; Wang, Y.; Hong, K.; Liu, P.P.; Zhu, W.M. Cyclic Tripeptides from the Halotolerant Fungus Aspergillus sclerotiorum PT06-1. J. Nat. Prod. 2010, 73, 1133–1137. [Google Scholar] [CrossRef] [PubMed]
  6. Liu, L.; Bao, L.; Wang, L.; Ma, K.; Han, J.J.; Yang, Y.L.; Liu, R.X.; Ren, J.W.; Yin, W.B.; Wang, W.Z.; et al. Asperorydines A–M: Prenylated Tryptophan-Derived Alkaloids with Neurotrophic Effects from Aspergillus oryzae. J. Org. Chem. 2018, 83, 812–822. [Google Scholar] [CrossRef] [PubMed]
  7. Wang, C.; Guo, L.; Hao, J.J.; Wang, L.P.; Zhu, W.M. α-Glucosidase Inhibitors from the Marine-Derived Fungus Aspergillus flavipes HN4-13. J. Nat. Prod. 2016, 79, 2977–2981. [Google Scholar] [CrossRef] [PubMed]
  8. Li, W.S.; Xiong, P.; Zheng, W.X.; Zhu, X.W.; She, Z.G.; Ding, W.J.; Li, C.Y. Identification and Antifungal Activity of Compounds from the Mangrove Endophytic Fungus Aspergillus clavatus R7. Mar. Drugs 2017, 15, 259. [Google Scholar] [CrossRef] [PubMed]
  9. Myobatake, Y.; Takemoto, K.; Kamisuki, S.; Inoue, N.; Takasaki, A.; Takeuchi, T.; Mizushina, Y.; Sugawara, F. Cytotoxic Alkylated Hydroquinone, Phenol, and Cyclohexenone Derivatives from Aspergillus violaceofuscus Gasperini. J. Nat. Prod. 2014, 77, 1236–1240. [Google Scholar] [CrossRef] [PubMed]
  10. Liu, S.; Dai, H.F.; Konuklugil, B.; Orfali, R.S.; Lin, W.H.; Kalscheuer, R.; Liu, Z.; Proksch, P. Phenolic bisabolanes from the sponge-derived fungus Aspergillus sp. Phytochem. Lett. 2016, 18, 187–191. [Google Scholar] [CrossRef]
  11. Li, D.; Xu, Y.; Shao, C.L.; Yang, R.Y.; Zheng, C.J.; Chen, Y.Y.; Fu, X.M.; Qian, P.Y.; She, Z.G.; de Voogd, N.J.; et al. Antibacterial Bisabolane-Type Sesquiterpenoids from the Sponge-Derived Fungus Aspergillus sp. Mar. Drugs 2012, 10, 234–241. [Google Scholar] [CrossRef] [PubMed]
  12. Fujiwara, M.; Yagi, N.; Miyazawa, M. Acetylcholinesterase Inhibitory Activity of Volatile Oil from Peltophorum dasyrachis Kurz ex Bakar (Yellow Batai) and Bisabolane-Type Sesquiterpenoids. J. Agric. Food Chem. 2010, 58, 2824–2829. [Google Scholar] [CrossRef] [PubMed]
  13. Lu, Z.Y.; Zhu, H.J.; Fu, P.; Wang, Y.; Zhang, Z.H.; Lin, H.P.; Liu, P.P.; Zhuang, Y.B.; Hong, K.; Zhu, W.M. Cytotoxic Polyphenols from the Marine-Derived Fungus Penicillium expansum. J. Nat. Prod. 2010, 73, 911–914. [Google Scholar] [CrossRef] [PubMed]
  14. Takamatsu, S.; Hodges, T.W.; Rajbhandari, I.; Gerwick, W.H.; Hamann, M.T.; Nagle, D.G. Marine Natural Products as Novel Antioxidant Prototypes. J. Nat. Prod. 2003, 66, 605–608. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Liu, Z.M.; Xia, G.P.; Chen, S.H.; Liu, Y.Y.; Li, H.X.; She, Z.G. Eurothiocin A and B, Sulfur-Containing Benzofurans from a Soft Coral-Derived Fungus Eurotium rubrum SH-823. Mar. Drugs 2014, 12, 3669–3680. [Google Scholar] [CrossRef] [PubMed]
  16. Liu, Y.Y.; Wu, Y.N.; Zhai, R.; Liu, Z.M.; Huang, X.S.; She, Z.G. Altenusin derivatives from mangrove endophytic fungus Alternaria sp. SK6YW3L. RSC Adv. 2016, 6, 72127–72132. [Google Scholar] [CrossRef]
  17. Chen, S.H.; Liu, Z.M.; Lu, Y.J.; Xia, G.P.; He, L.; She, Z.G. New depsidones and isoindolinones from the mangrove endophytic fungus Meyerozyma guilliermondii (HZ-Y2) isolated from the South China Sea. Beilstein J. Org. Chem. 2015, 11, 1187–1193. [Google Scholar] [CrossRef] [PubMed]
  18. Liu, Y.Y.; Chen, S.H.; Liu, Z.M.; Xia, G.P.; He, L.; She, Z.G. Bioactive Metabolites from Mangrove Endophytic Fungus Aspergillus sp. 16-5B. Mar. Drugs 2015, 13, 3091–3102. [Google Scholar] [CrossRef] [PubMed]
  19. Bunyapaiboonsri, T.; Yoiprommarat, S.; Intereya, K.; Kocharin, K. New Diphenyl Ethers from the Insect Pathogenic Fungus Cordyceps sp. BCC 1861. Chem. Pharm. Bull. 2007, 55, 304–307. [Google Scholar] [CrossRef] [PubMed]
  20. Trisuwan, K.; Rukachaisirikul, V.; Kaewpet, M.; Phongpaichit, S.; Hutadilok-Towatana, N.; Preedanon, S.; Sakayaroj, J. Sesquiterpene and Xanthone Derivatives from the Sea Fan-Derived Fungus Aspergillus sydowii PSU-F154. J. Nat. Prod. 2011, 74, 1663–1667. [Google Scholar] [CrossRef] [PubMed]
  21. Li, X.D.; Li, X.M.; Xu, G.M.; Zhang, P.; Wang, B.G. Antimicrobial Phenolic Bisabolanes and Related Derivatives from Penicillium aculeatum SD-321, a Deep Sea Sediment-Derived Fungus. J. Nat. Prod. 2015, 78, 844–849. [Google Scholar] [CrossRef] [PubMed]
  22. Wang, P.; Yu, J.H.; Zhu, K.K.; Wang, Y.Y.; Cheng, Z.Q.; Jiang, C.S.; Dai, J.G.; Wu, J.; Zhang, H. Phenolic bisabolane sesquiterpenoids from a Thai mangrove endophytic fungus, Aspergillus sp. xy02. Fitoterapia 2018, 127, 322–327. [Google Scholar] [CrossRef] [PubMed]
  23. Takenaka, Y.; Tanahashi, T.; Nagakura, N.; Hamada, N. Phenyl Ethers from Cultured Lichen Mycobionts of Graphis scripta var. serpentina and G. rikuzensis. Chem. Pharm. Bull. 2003, 51, 794–797. [Google Scholar] [CrossRef]
  24. Sumarah, M.W.; Kesting, J.R.; Sørensen, D.; Miller, J.D. Antifungal metabolites from fungal endophytes of Pinus strobus. Phytochemistry 2011, 72, 1833–1837. [Google Scholar] [CrossRef] [PubMed]
  25. Nukina, M.; Sato, Y.; Ikeda, M.; Sassa, T. Sydonol, a New Fungal Morphogenic Substance Produced by an Unidentified Aspergillus sp. Agric. Biol. Chem. 1981, 45, 789–790. [Google Scholar] [CrossRef]
  26. Wang, J.F.; Lu, Z.Y.; Liu, P.P.; Wang, Y.; Li, J.; Hong, K.; Zhu, W.M. Cytotoxic Polyphenols from the Fungus Penicillium expansum 091006 Endogenous with the Mangrove Plant Excoecaria agallocha. Planta Med. 2012, 78, 1861–1866. [Google Scholar] [PubMed]
  27. Rivera-Chávez, J.; Figueroa, M.; González, M.D.C.; Glenn, A.E.; Mata, R. α-Glucosidase Inhibitors from a Xylaria feejeensis Associated with Hintonia latif lora. J. Nat. Prod. 2015, 78, 730–735. [Google Scholar] [CrossRef] [PubMed]
  28. Rivera-Chávez, J.; González-Andrade, M.; González, M.D.C.; Glenn, A.E.; Mata, R. Thielavins A, J and K: α-Glucosidase inhibitors from MEXU 27095, an endophytic fungus from Hintonia latiflora. Phytochemistry 2013, 94, 198–205. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Chemical structures of compounds 111.
Figure 1. Chemical structures of compounds 111.
Marinedrugs 16 00307 g001
Figure 2. Key COSY (bold line) and HMBC (arrow) correlations of compounds 16.
Figure 2. Key COSY (bold line) and HMBC (arrow) correlations of compounds 16.
Marinedrugs 16 00307 g002
Table 1. 1H and 13C NMR spectroscopic data (500/100 MHz) for 1 and 2 a.
Table 1. 1H and 13C NMR spectroscopic data (500/100 MHz) for 1 and 2 a.
No.12
δCδHδCδH
1149.4 163.6
2105.56.43, d (2.6)103.76.23, s
3144.0 164.9
4138.4 109.9
5125 144.6
6114.16.41, d (2.6)113.16.35, s
715.82.22, s24.02.53, s
8 172.4
9 68.44.40, dd (6.6, 11.6)
4.62, dd (2.8, 11.6)
10 71.13.89, td (2.8, 6.8)
11 73.73.63, m
12 64.63.65, dd (5.3, 8.0)
3.77, dd (5.7, 7.7)
1′159.6 159.9
2′102.36.22, dd (3.6, 2.1)105.76.27, s
3′156.5 157.4
4′111.16.34, brs113.66.47, s
5′141.1 142.2
6′110.56.34, brs113.46.35, s
7′21.62.24, s21.52.26, s
aδ in ppm, J in Hz, 1 in CDCl3, 2 in methanol-d4.
Table 2. 1H and 13C NMR spectroscopic data (500/100 MHz) for 3, 4, and 6a.
Table 2. 1H and 13C NMR spectroscopic data (500/100 MHz) for 3, 4, and 6a.
No.346
δCδHδCδHδCδH
1154.9 155.2 155.8
2131.6 133.2 132.0
3130.56.88, d (7.7)130.36.99, d (7.7)127.27.14, d (7.9)
4121.16.63, dd (7.7, 1.4)119.16.75, d (7.7)119.06.78, dd (7.9, 1.1)
5142.5 142.3 142.8
6116.86.60, d (1.4)115.06.77, s116.06.75, d (1.1)
7135.6 135.4 87.2
8130.15.38, td (1.3, 7.2)130.85.43, td (1.1, 7.1)39.52.14, 2.40, m
927.22.15, m24.52.25, m29.71.71, 1.86, m
1040.01.33, m44.31.60, m86.43.65, m
1128.91.61, m71.4 34.41.75, m
1223.00.93, d (6.6)29.21.23, s18.90.93, d (6.7)
1323.00.93, d (6.6)29.21.23, s19.51.03, d (6.7)
1417.31.94, s17.21.96, s29.51.50, s
1536.13.80, s65.04.5, s64.94.50, s
16129.2
17141.9
18145.9
19114.86.47, d (1.3)
20129.7
21122.96.34, d (1.3)
2220.92.13, s
aδ in ppm; J in Hz; 3, 4, and 6 in methanol-d4.
Table 3. 1H and 13C NMR spectroscopic data (500/100 MHz) for 5a.
Table 3. 1H and 13C NMR spectroscopic data (500/100 MHz) for 5a.
No.δCδHNo.δCδH
1 (1′)157.0 9 (9′)23.01.27, m
2 (2′)131.9 10 (10′)40.51.13, m
3 (3′)127.77.10, d (7.9)11 (11′)29.01.48, m
4 (4′)119.86.78, d (7.9)12 (12′)23.00.82, d (6.6)
5 (5′)139.5 13 (13′)23.00.82, d (6.6)
6 (6′)117.16.76, s14 (14′)29.01.58, s
7 (7′)78.0 15 (15′)72.64.44, s
8 (8′)44.01.77, 1.88, m
aδ in ppm, J in Hz, 5 in methanol-d4.
Table 4. α-Glucosidase inhibitory activities.
Table 4. α-Glucosidase inhibitory activities.
Compounds1234567891011Acarbose a
IC50 (μM)165.2129.94.5-3.1-532.5--1.52.3840.2
- means no activity; a positive control.

Share and Cite

MDPI and ACS Style

Wu, Y.; Chen, Y.; Huang, X.; Pan, Y.; Liu, Z.; Yan, T.; Cao, W.; She, Z. α-Glucosidase Inhibitors: Diphenyl Ethers and Phenolic Bisabolane Sesquiterpenoids from the Mangrove Endophytic Fungus Aspergillus flavus QQSG-3. Mar. Drugs 2018, 16, 307. https://doi.org/10.3390/md16090307

AMA Style

Wu Y, Chen Y, Huang X, Pan Y, Liu Z, Yan T, Cao W, She Z. α-Glucosidase Inhibitors: Diphenyl Ethers and Phenolic Bisabolane Sesquiterpenoids from the Mangrove Endophytic Fungus Aspergillus flavus QQSG-3. Marine Drugs. 2018; 16(9):307. https://doi.org/10.3390/md16090307

Chicago/Turabian Style

Wu, Yingnan, Yan Chen, Xishan Huang, Yahong Pan, Zhaoming Liu, Tao Yan, Wenhao Cao, and Zhigang She. 2018. "α-Glucosidase Inhibitors: Diphenyl Ethers and Phenolic Bisabolane Sesquiterpenoids from the Mangrove Endophytic Fungus Aspergillus flavus QQSG-3" Marine Drugs 16, no. 9: 307. https://doi.org/10.3390/md16090307

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop