Next Article in Journal
Pharmaceutical Standardization and Physicochemical Characterization of Traditional Ayurvedic Marine Drug: Incinerated Conch Shell (Shankha Bhasma)
Next Article in Special Issue
Diphenyl Ethers from a Marine-Derived Aspergillus sydowii
Previous Article in Journal
Unstable Tetramic Acid Derivatives from the Deep-Sea-Derived Fungus Cladosporium sphaerospermum EIODSF 008
Previous Article in Special Issue
Anthracimycin B, a Potent Antibiotic against Gram-Positive Bacteria Isolated from Cultures of the Deep-Sea Actinomycete Streptomyces cyaneofuscatus M-169
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Potential Antiviral Xanthones from a Coastal Saline Soil Fungus Aspergillus iizukae

1
College of Pharmacy, Binzhou Medical University, Yantai 264003, China
2
Beijing Higher Institution Engineering Research Center of Food Additives and Ingredients, Beijing Key Laboratory of Flavor Chemistry, Beijing Laboratory for Food Quality and Safety, Beijing Technology and Business University, Beijing 100048, China
*
Authors to whom correspondence should be addressed.
Mar. Drugs 2018, 16(11), 449; https://doi.org/10.3390/md16110449
Submission received: 27 October 2018 / Revised: 13 November 2018 / Accepted: 14 November 2018 / Published: 15 November 2018
(This article belongs to the Special Issue Isolation and Structure Elucidation of Marine Secondary Metabolites)

Abstract

:
Five new (15) and two known xanthones (6 and 7), one of the latter (6) obtained for the first time as a natural product, together with three known anthraquinones, questin, penipurdin A, and questinol, were isolated from the coastal saline soil-derived Aspergillus iizukae by application of an OSMAC (one strain many compounds) approach. Their structures were determined by interpretation of nuclear magnetic resonance (NMR) and high-resolution electrospray ionization mass spectroscopy (HRESIMS) data, as well as comparison of these data with those of related known compounds. Antiviral activity of xanthones 17 was evaluated through the cytopathic effect (CPE) inhibition assay, and compound 2 exhibited distinctly strong activity towards influenza virus (H1N1), herpes simplex virus types 1 (HSV-1) and 2 (HSV-2) with IC50 values of 44.6, 21.4, and 76.7 μM, respectively, which indicated that it was worth to further investigate it as a potential lead compound. The preliminary structure-activity relationship of the xanthones is discussed.

1. Introduction

Aspergillus iizukae is present in various environments, such as leaves of Silybum marianum [1], sponges [2], guts of a cricket species Gryllus testaceus [3], earthworm casts and different kinds of soil [4]. Recently, Özkaya et al found that the ethyl acetate extract of a sponge-associated A. iizukae showed inhibitory effects against the aquaculture pathogens Lactococcus garvieae and Vagococcus salmoninarum [2]. Additionally, a novel flavin adenine dinucleotide-dependent glucose dehydrogenase was separated from the culture of A. iizukae, which was capable of catalyzing the oxidation of glucose to glucono-δ-lactone [5]. Nevertheless, there are only a couple of reports on its metabolites. Up to now, four new and three known aromatic butenolides were isolated from the insect symbiont A. iizukae [3]. It was also reported that the endophytic A. iizukae from S. marianum could produce silybin A, silybin B and isosilybin A [1].
In our continuing search for new biologically active metabolites from fungi [6,7], the prolific fungus A. iizukae was isolated from coastal saline soil in Kenli, China. Previously, two new diphenyl derivatives and nine known compounds were obtained from it in a shaken fermentation, and six of them contain chloride atoms, that demonstrating the strain was able to utilize chlorine in the biosynthesis of its metabolites [8]. In order to obtain more halogen-containing compounds, the working strain was fermented statically in liquid culture medium supplemented with sodium bromide applying an OSMAC strategy. Chemical exploration of the fungal extract led to the isolation of seven xanthones (17) (Figure 1), including five new xanthones (15), one new natural product (6) previously reported as a semisynthetic compound, and a known one (7), along with three known anthraquinones. Herein, the isolation and structure elucidation of the new metabolites, along with the antiviral activity of 17 are reported.

2. Results

Compound 1 was obtained as a yellow amorphous powder. Its molecular formula C16H11ClO6, was established by the high-resolution electrospray ionization mass spectroscopy (HRESIMS) m/z 335.0316 [M + H]+, revealing eleven degrees of unsaturation. The infrared (IR) spectrum revealed the presence of hydroxy (3219 cm−1), conjugated ketone carbonyl (1704 cm−1), conjugated ester carbonyl (1645 cm−1), and aromatic (1608, 1574, 1493 and 1437 cm−1) groups. The ultraviolet (UV) spectrum maxima at 358, 312, 290 and 237 nm showed typical absorptions of xanthones [9]. The 1H NMR spectrum (Table 1) presented two phenolic hydroxy groups at δH 13.00 (br s, 1-OH) and 11.63 (br s, 6-OH), an aromatic singlet at δH 7.12 (s, H-4), two doublets of meta-coupled aromatic protons at δH 6.96 (d, J = 2.1 Hz, H-5) and 6.89 (d, J = 2.1 Hz, H-7) , one methoxy group at δH 3.88 (s, Me-13), and one aromatic methyl group at δH 2.46 (s, Me-11). The 13C NMR data (Table 1) displayed sixteen carbon signals, including two conjugated carbonyl groups at δC 178.9 and 168.1, twelve aromatic carbon signals (three methines and four oxygenated quaternary carbons) in the region δC 103.3–164.4, one methoxy group at δC 52.7 and one methyl group at δC 20.8, which were clearly supported by the heteronuclear single quantum coherence (HSQC) data. The above information suggested 1 to be a chlorinated derivative of methyl-(l,6-dihydroxy-3-methylxanthone)-8-carboxylate (7) [10], which was also isolated from this culture. The positions of the carbomethoxy and hydroxy groups in ring B were confirmed by the correlations from H-7 to C-5, C-12 and C-9a, and from H-5 to C-6, C-7, C-9a and C-10a in the heteronuclear multiple bond correlation (HMBC) data (Figure 2). The chemical shift (δC 178.9) of the carbonyl (C-9) indicated the presence of a hydrogen-bonded hydroxyl group at C-1 [11,12]. Additional HMBC correlations from 1-OH to C-1, C-2 and C-1a, and from H3-11 to C-2, C-3 and C-4, placed the chlorine atom on C-2. Therefore, the structure of 1 was determined as methyl-(2-chloro-l,6-dihydroxy-3-methylxanthone)-8-carboxylate.
Compound 2 was isolated as a light-yellow powder. The HRESIMS data (m/z 335.0319 [M + H]+) gave the same molecular formula C16H11ClO6 as 1. The IR and UV data of 2 resembled those of 1, indicating that they have same scaffold. Detailed analysis and comparison of their 1H and 13C NMR data revealed a slight difference in the chemical shifts of 1H and 13C atoms in the rings A, especially those of C-1 and C-4a, while chemical shifts in the rings B were almost identical between 1 and 2. The aforementioned data suggested that the chlorine atom was at C-4 in 2, which was confirmed by the HMBC correlations (Figure 2) from H3-11 to C-2, C-3 and C-4, and from 1-OH to C-1, C-2 and C-1a. The other HMBC correlations shown in Figure 2 supported that the structure of 2 is methyl-(4-chloro-l,6-dihydroxy-3-methylxanthone)-8-carboxylate.
Compound 3, a light-yellow solid, was isolated in small amount, and showed a protonated ion in its HRESIMS at m/z 349.0476 [M + H]+, indicating a molecular formula of C17H13ClO6 having one CH2 unit more than that of 2. The 1H and 13C NMR data were very similar to those of 2, except for the appearance of an additional methoxyl group and the disappearance of a hydroxyl group. Therefore, 3 was a methylated derivative of 2. The lower resonance frequencies of C-9 (δC 172.9) suggested that the methoxyl was at C-1 [13]. This hypothesis was proved by the HMBC correlations (Figure 2) from H3-14 to C-1, and from H-2 to C-1, C-4 and C-1a. Based on these evidences, the structure of 3 was established as methyl-(4-chloro-6-hydroxy-1-methoxy-3-methylxanthone)-8-carboxylate.
Compound 4 was obtained as an orange powder. The molecular formula was determined as C17H14O6 from the HRESIMS m/z 313.0727 [M − H], showing the absence of chlorine atoms and one hydrogen atom more than that of 3. In the 1H NMR spectrum, four broad singlets (δH 6.94, 6.87 6.82 and 6.73, 1H, s, each) were observed. Careful comparison of the NMR spectra data between 3 and 4 (Table 1 and Table 2) revealed that it was a dechlorinated derivative of 3. This hypothesis was confirmed by analysis of the HMBC correlations (Figure 2). Therefore, the structure of 4 was elucidated as methyl-(6-hydroxy-1-methoxy-3-methylxanthone)-8-carboxylate.
Compound 5 was isolated as a pale-yellow powder and had the molecular formula C15H9ClO6 as determined by the HRESIMS m/z 319.0013 [M − H]. Therefore, 5 has one CH2 unit less than 2. In its IR spectrum the broad absorption at 3300−2600 cm−1, along with the absorption at 1694 cm−1, revealed the presence of a carboxylate functionality in 5. Its NMR data were very similar to those of 2, except for the absence of a methoxy group and the presence of a broad singlet of a hydroxy group in 5. In the HMBC spectrum, the correlations from H-11 to C-2, C-3 and C-4, and from H-2 to C-1, C-4, C-11 and C-1a were observed, which unequivocally established the substitution of ring A. In spite of no HMBC correlations observed from H-6 and H-7 to any carbons, the ring B and the chemical shift assignment of corresponding carbons and protons could be achieved by comparison of its 1H and 13C data with those of 6 and calyxanthone [14]. Interestingly, in the 13C NMR spectrum, the resonance intensity of carbon atoms in ring B was much weaker than that of carbon atoms in ring A (Figure S30). Compound 5 was identified as 4-chloro-1,6-dihydroxy-3-methylxanthone-8-carboxylic acid.
Compound 6 was isolated as an orange-red powder and has low solubility in methanol. Its molecular formula C15H8Cl2O6 was determined on the basis of the HRESIMS m/z 352.9604 [M − H]. The relative height of the typical isotopic ion peak (at m/z 354.9570 [M − H]) was approximately two thirds of that of the quasi-molecular ion peak (Figure S40), suggesting the existence of two chlorine atoms. Its UV spectrum showed characteristic absorption bands of xanthones. In the 1H NMR spectrum, the meta-coupled aromatic protons at δH 6.95 (d, J = 2.2 Hz, H-5) and 6.86 (d, J = 2.1 Hz, H-7) of ring B were supported by the HMBC correlations (Figure 2) from H-5 to C-6, C-7, C-9a and C-10a, and from H-7 to C-5, C-12 and C-9a, which was confirmed by similar 1H and 13C chemical shifts values to those of calyxanthone [14]. The downfield shift of the carbonyl implied the hydrogen bonded phenolic hydroxyl at δH 13.22 (br s, 1-OH) attached to C-1. Additionally, the HMBC correlations from the aromatic methyl protons to C-2, C-3 and C-4, and from 1-OH to C-2, demonstrated that chlorine atoms were both at C-4 and C-2, respectively. Ultimately, the structure of 6 was established to be 2,4-dichloro-1,6-dihydroxy-3-methylxanthone-8-carboxylic acid. Compound 6 was previously described as a synthetic intermediate using only UV and IR data [15]. This is the first report of its isolation from a natural extract.
The known compounds were identified as methyl-(l,6-dihydroxy-3-methylxanthone)-8-carboxylate (7) [14], questin [8], penipurdin A [16], and questinol [17] by comparison of their NMR data with those reported in the literatures.
Xanthones 17 were screened for their antiviral activity against H1N1, HSV-1 and HSV-2 using the CPE inhibition assay (Table 3). Compounds 1, 2 and 7 exhibited anti-H1N1 activity with IC50 values of 133.4, 54.6 and 140.4 μM, respectively, while the others were inactive (ribavirin was used as the positive control, IC50 101.4 μM). Compounds 1, 2 and 7 showed a strong anti-HSV-1 activity with IC50 values of 55.5, 21.4 and 75.7 μM, respectively, and the other compounds showed a moderate anti-HSV-1 activity compared with the positive control (acyclovir, IC50 150.2 μM). Compounds 2 and 7 also possessed a strong anti-HSV-2 effect with IC50 values of 76.7 and 95.4 μM (acyclovir as the positive control, IC50 128.6 μM), respectively.

3. Materials and Methods

3.1. General Experimental Procedures

Instrumentation used to acquire UV, IR, HRESIMS, optical rotation, and 1D and 2D NMR spectra and to perform column chromatography has been previously described [6,7,8] A TU-1091 spectrophotometer (Beijing Purkinje General Instrument Co., Ltd., Beijing, China) was used to measure the UV spectra in MeOH. An attenuated total reflection (ATR) method was employed to record the infrared spectra on a Nicolet 6700 spectrophotometer (Thermo Fisher Scientific, Madison, WI, USA). An Autopol V Plus digital polarimeter (Rudolph Research Analytical, Hackettstown, NJ, USA) was used to measure optical rotation. Detailed 1D and 2D NMR spectra were recorded on a Bruker AV-400 or Bruker AVIII 500 spectrometers (Bruker Biospin Group, Karlsruhe, Germany) with tetramethylsilane as an internal reference. A 1200RRLC-6520 Accurate-Mass Q-TOF LC/MS mass spectrometer (Agilent Technologies, Ltd., Palo Alto, CA, USA) was used to acquire HRESIMS spectra. HPLC purification was carried on a SHIMADZU LC-6AR (Shimadzu Corporation, Kyoto, Japan) Liquid Chromatograph equipped with an SPD-20A diode array detector, using an ODS column (HyperClone 5 μm ODS (C18), 120 Å, 250 mm × 10 mm, Phenomenex; Shim-pack GIS, 5 μm C18, 250 mm × 10 mm, Shimadzu. 4 mL/min).

3.2. Fungal Material

Aspergillus iizukae KL33 (GenBank accession numbers: HQ717800) was isolated from coastal saline soil in Kenli, Shandong Province of China, in August 2008. The strain was deposited at the Department of Chemistry, Binzhou Medical University, Yantai.

3.3. Fermentation and Extraction

A. iizukae KL33 was transferred aseptically to fresh PDA culture plates, and incubated at 28 °C for one week. Emerging fungal colonies were transferred into 500 mL Erlenmeyer flasks containing 180 mL of culture medium composed of glucose (20 g), maltose (10 g), mannitol (10 g), peptone (10 g), corn syrup (3 g), KH2PO4 (0.5 g), MgSO4·7H2O (0.3 g), sodium glutamate (10 g), sodium bromide (10 g), water (1 L, half seawater and half tap water) and statically fermented at room temperature for 6 weeks.
40 L of the fermentation broth was separated into mycelium and filtrate through cheesecloth. The filtrate was extracted three times with ethyl acetate. The mycelium was extracted with methanol for three times. The methanol solution was concentrated under reduced pressure to give an aqueous solution. The aqueous solution was extracted three more times with ethyl acetate. Both the ethyl acetate solutions were concentrated under reduced pressure to give a crude extract (65 g).

3.4. Purification

The crude extract (65 g) was fractionated into ten fractions (Frs 1−10) on a silica gel (200−300 mesh) column chromatography using a gradient of petroleum ether/chloroform (2:1, 1:1, 1:2, and 0:1, v/v), followed by chloroform/methanol (100:1, 50:1, 10:1, and 0:1, v/v). Fr. 2 was passed through an ODS column (25−40 μm, Merck, Darmstadt, Germany) using a sequential mixture of MeOH and H2O as eluent from 20% to 100% to obtain nine fractions (Frs 2.1−2.9). Similarly, Fr. 5 was fractionated into eight fractions (Frs 5.1−5.8). Fr. 2.5 and Fr. 2.3 were purified by semipreparative HPLC on an ODS column (shim-pack GIS) with MeOH/0.2% trifluoroacetic acid (TFA) aqueous solution (v/v) (4:1, v/v; 4 mL/min) as the mobile phase to yield 1 (20.3 mg, tR 17.5 min) and 2 (18.6 mg, tR 20.5 min), respectively. Fr. 4 was fractionated into three fractions (Frs 4.1−4.3) by an ODS column chromatograph (MeOH/H2O, 7:3, v/v). Fr. 4.1 was chromatographied on a silica gel column chromatography (CHCl3/MeOH, 50:1, v/v), and then purified by semipreparative HPLC on an ODS column (Phenomenex) eluted with MeOH/0.2% TFA aqueous solution (v/v) (70:30, v/v, 4 mL/min) to give 3 (10.3 mg, tR 11.9 min) and 4 (7.3 mg, tR 7.2 min). Fr. 6 was passed through ODS column (MeOH/H2O, 1:1, v/v), Sephadex LH-20 (MeOH), and semipreparative HPLC [shim-pack GIS, MeOH/0.2% TFA aqueous solution (v/v) (70:30, v/v; 4 mL/min)] successively to yield 5 (15.4 mg, tR 22.8 min). Fr. 5.5, Fr. 2.4 and Fr. 5.4 were chromatographed on Sephadex LH-20 columns eluted with MeOH to obtain 6 (18.9 mg), 7 (17.5 mg), 9 (3.7 mg), respectively. Fr. 3 was purified by ODS (MeOH/H2O, 7:3, v/v) and Sephadex LH-20 columns (MeOH) to obtain 8 (17.3 mg). Fr. 5.3 was applied on Sephadex LH-20 (MeOH) column and further purified by semipreparative HPLC on an ODS column (Phenomenex) with MeOH/0.2% TFA aqueous solution (v/v) (55:45, v/v; 4 mL/min) as the eluting solvent to afford 10 (17.0 mg, tR 12.1 min).
Compound 1: a yellow amorphous powder; UV (MeOH) λmax (log ε) 358 (3.92), 312 (4.09), 290 (3.97), 237 (4.46), 204 (4.29) nm; IR (ATR) νmax 3219, 1704, 1645, 1608, 1574, 1493, 1437, 1414, 1381, 1273, 1231, 1176, 1146, 1024, 949, 848, 816, 775 cm−1; HRESIMS m/z 335.0316 [M + H]+ (calculated for C16H12ClO6, 335.0317). 1H and 13C NMR data: see Table 1.
Compound 2: a light-yellow powder; UV (MeOH) λmax (log ε): 360 (3.98), 307 (4.12), 272 (4.09), 253 (4.31), 235 (4.51), 204 (4.31) nm; IR (ATR) νmax 3094, 2960, 1683, 1646, 1599, 1511, 1473, 1441, 1371, 1261, 1175, 1152, 907, 882, 811, 766 cm−1; HRESIMS m/z 335.0319 [M + H]+ (calculated for C16H12ClO6, 335.0317). 1H and 13C NMR data: see Table 1.
Compound 3: a light-yellow solid; UV (MeOH) λmax (log ε): 347 (3.59), 297 (3.74), 250 (3.97), 235 (4.12), 203 (4.12) nm; IR (ATR) νmax 3203, 1740, 1560, 1584, 1480, 1436, 1325, 1225, 1151, 903, 886, 815, 770 cm−1; HRESIMS m/z 349.0476 [M + H]+ (calculated for C17H14ClO6, 349.0473). 1H and 13C NMR data: see Table 1.
Compound 4: an orange powder; UV (MeOH) λmax (log ε): 339 (3.38), 300 (3.53), 290 (3.55), 246 (3.78), 234 (3.93), 203 (3.98) nm; IR (ATR) νmax 3392, 1740, 1678, 1620, 1607, 1437, 1197, 1135, 842, 822, 802, 724 cm−1; HRESIMS m/z 313.0727 [M − H] (calculated for C17H13O6, 313.0718). 1H and 13C NMR data: see Table 2.
Compound 5: a pale-yellow powder; UV (MeOH) λmax (log ε): 355 (3.81), 306 (4.21), 250 (4.40), 237 (4.53), 204 (4.37) nm; IR (ATR) νmax 3101, 1694, 1644, 1607, 1558, 1506, 1475, 1239, 1149, 896, 821, 717 cm−1; HRESIMS m/z 319.0013 [M − H] (calculated for C15H8ClO6, 319.0015). 1H and 13C NMR data: see Table 2.
Compound 6: an orange-red powder; UV (MeOH) λmax (log ε): 361 (3.96), 312 (4.41), 240 (4.73), 203 (4.67) nm; IR (ATR) νmax 3081, 1691, 1610, 1575, 1498, 1432, 1218, 1156, 913, 883, 828, 778 cm−1. HRESIMS m/z 352.9604 [M − H] (calculated for C15H7Cl2O6, 352.9625). 1H and 13C NMR data: see Table 2.

3.5. Antiviral Activity

The antiviral activity against influenza A virus (H1N1) was carried out by CPE inhibition assay as previously reported [18,19]. First, confluent MDCK cell monolayers and influenza virus (A/Puerto Rico/8/34 (H1N1), PR/8) were incubated together at 37 °C for 1 h. Then, the cells were treated with different test compounds after removing the virus dilution. After incubating at 37 °C for 48 h, the cells were fixed with 4% formaldehyde of 100 μL for 20 min at room temperature. Later on, the cells were stained with 0.1% crystal violet for 30 min after removal of the formaldehyde. Finally, the plates were washed and dried, followed by the measurement of the intensity of crystal violet staining for each well at 570 nm in a microplate reader (Bio-Rad, USA). Ribavirin was used as the positive control.
The anti-herpes simplex virus types 1 (HSV-1) and 2 (HSV-2) activity of 17 on Vero cells were conducted in the same way as described above [20], and acyclovir (ACV) was applied as the positive control.

4. Conclusions

In summary, the culture based on OSMAC strategy of the fungus A. iizukae yielded seven xanthones, including five new and one isolated for the first time as a natural product. Among them, five metabolites contain chlorine, and their structures were different from those previously reported from A. iizukae. From this study, it is clear that the OSMAC strategy is still a powerful tool in producing new metabolites from microorganisms. Compound 2 exhibited a strong antiviral activity against H1N1, HSV-1 and HSV-2 with IC50 values of 44.6, 21.4, and 76.7 μM, respectively, compared with the positive controls.
The results of antiviral activity of 17 indicated that the hydroxy group at C-1 and the methyl carboxylate group at C-8 essentially contributed to the anti-H1N1, anti-HSV-1 and anti-HSV-2 activities, and the position of the chlorine atom in ring A would affect the antiviral activities. Additionally, it seemed that methylation of the hydroxy group at C-1 or replacement of methyl carboxylate at C-8 by carboxylic acid, to a large extent, lower the antiviral effect.
Xathones have attracted considerable interest for their promising biological activities and the interesting structural scaffold, which could be modified by various substituents [21,22]. Our finding suggests that 2 might be a potential anti-H1N1 lead candidate, worthy of a further pharmacological exploration.

Supplementary Materials

The following are available online at https://www.mdpi.com/1660-3397/16/11/449/s1, Figure S1: 1H NMR spectrum (400 MHz) of compound 1 in DMSO-d6, Figure S2: 13C NMR spectrum (100 MHz) of compound 1 in DMSO-d6, Figure S3: HSQC spectrum of compound 1 in DMSO-d6, Figure S4: HMBC spectrum of compound 1 in DMSO-d6, Figure S5: HRESIMS spectrum of compound 1, Figure S6: IR spectrum of compound 1, Figure S7: UV spectrum of compound 1 in MeOH, Figure S8: 1H NMR spectrum (400 MHz) of compound 2 in DMSO-d6, Figure S9: 13C NMR spectrum (100 MHz) of compound 2 in DMSO-d6, Figure S10: HSQC spectrum of compound 2 in DMSO-d6, Figure S11: HMBC spectrum of compound 2 in DMSO-d6, Figure S12: HRESIMS spectrum of compound 2, Figure S13: IR spectrum of compound 2, Figure S14: UV spectrum of compound 2 in MeOH, Figure S15: 1H NMR spectrum (500 MHz) of compound 3 in DMSO-d6, Figure S16: 13C NMR spectrum (125 MHz) of compound 3 in DMSO-d6, Figure S17: HSQC spectrum of compound 3 in DMSO-d6, Figure S18: HMBC spectrum of compound 3 in DMSO-d6, Figure S19: HRESIMS spectrum of compound 3, Figure S20: IR spectrum of compound 3, Figure S21: UV spectrum of compound 3 in MeOH, Figure S22: 1H NMR spectrum (400 MHz) of compound 4 in DMSO-d6, Figure S23: 13C NMR spectrum (100 MHz) of compound 4 in DMSO-d6, Figure S24: HSQC spectrum of compound 4 in DMSO-d6, Figure S25. HMBC spectrum of compound 4 in DMSO-d6, Figure S26: HRESIMS spectrum of compound 4, Figure S27: IR spectrum of compound 4, Figure S28: UV spectrum of compound 4 in MeOH, Figure S29: 1H NMR spectrum (400 MHz) of compound 5 in DMSO-d6, Figure S30: 13C NMR spectrum (100 MHz) of compound 5 in DMSO-d6, Figure S31: HSQC spectrum of compound 5 in DMSO-d6, Figure S32: HMBC spectrum of compound 5 in DMSO-d6, Figure S33: HRESIMS spectrum of compound 5, Figure S34: IR spectrum of compound 5, Figure S35: UV spectrum of compound 5 in MeOH, Figure S36: 1H NMR spectrum (400 MHz) of compound 6 in DMSO-d6, Figure S37: 13C NMR spectrum (100 MHz) of compound 6 in DMSO-d6, Figure S38: HSQC spectrum of compound 6 in DMSO-d6, Figure S39: HMBC spectrum of compound 6 in DMSO-d6, Figure S40: HRESIMS spectrum of compound 6, Figure S41: IR spectrum of compound 6, Figure S42: UV spectrum of compound 6 in MeOH.

Author Contributions

H.-H.K. contributed to the fermentation, purification, structural elucidation and manuscript preparation. H.-B.Z. and M.-J.Z. performed partial purification and IR and UV data acquisition experiments. H.R. guided the antivirus activity assay. W.-Z.L., L.-Y.M. and D.-S.L. participated in the design of the research and revised the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (No. 31270082) and Natural Science Foundation of Shandong Province, China (No. Y2008B17, 2018GSF119020).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. El-Elimat, T.; Raja, H.A.; Graf, T.N.; Faeth, S.H.; Cech, N.B.; Oberlies, N.H. Flavonolignans from Aspergillus iizukae, a fungal endophyte of Milk Thistle (Silybum marianum). J. Nat. Prod. 2014, 77, 193–199. [Google Scholar] [CrossRef] [PubMed]
  2. Özkaya, F.C.; Peker, Z.; Camas, M.; Camas, A.S.; Altunok, M. Marine fungi against aquaculture pathogens and induction of the activity via co-culture. Clean-Soil Air Water 2017, 45, 1700238. [Google Scholar] [CrossRef]
  3. Li, L.J.; Li, T.X.; Kong, L.Y.; Yang, M.H. Antioxidant aromatic butenolides from an insect-associated Aspergillus iizukae. Phytochem. Lett. 2016, 16, 134–140. [Google Scholar] [CrossRef]
  4. Hubka, V.; Nováková, A.; Kolařík, M.; Jurjević, Ž.; Peterson, S.W. Revision of Aspergillus section Flavipedes: seven new species and proposal of section Jani sect. nov. Mycologia 2014, 107, 169–208. [Google Scholar] [CrossRef] [PubMed]
  5. Nishio, K.; Kojima, Y.; Yamaguchi, S. Novel Flavin Adenine Dinucleotide-Dependent Glucose Dehydrogenase from Aspergillus iizukae for Highly Glucose-Specific and pH-Stable Performance in Glucose Sensor. Patent No. WO2017077924A1, 11 May 2017. [Google Scholar]
  6. Ma, L.Y.; Liu, D.S.; Li, D.G.; Huang, Y.L.; Kang, H.H.; Wang, C.H.; Liu, W.Z. Pyran rings containing polyketides from Penicillium raistrickii. Mar. Drugs 2017, 15, 2. [Google Scholar] [CrossRef] [PubMed]
  7. Liu, D.S.; Huang, Y.L.; Li, C.M.; Ma, L.Y.; Pan, X.H.; Ferreira, D.; Liu, W.Z. A new sesquiterpenoid derivative from the coastal saline soil fungus Aspergillus fumigates. Rec. Nat. Prod. 2016, 10, 708–713. [Google Scholar]
  8. Liu, D.S.; Yan, L.; Ma, L.Y.; Huang, Y.L.; Pan, X.H.; Liu, W.Z. Diphenyl derivatives from coastal saline soil fungus Aspergillus iizukae. Arch. Pharm. Res. 2015, 38, 1038–1043. [Google Scholar] [CrossRef] [PubMed]
  9. Wang, Y.; Wang, S.S.; Zhou, D.H.; Xiao, H.B. Spectral characteristics of phytoxanthones. Nat. Prod. Res. Dev. 2002, 14, 85–90. [Google Scholar] [CrossRef]
  10. Ngan, N.T.T.; Quang, T.H.; Kim, K.W.; Kim, H.J.; Sohn, J.H.; Kang, D.G.; Lee, H.S.; Kim, Y.C.; Oh, H. Anti-inflammatory effects of secondary metabolites isolated from the marine-derived fungal strain Penicillium sp. SF-5629. Arch. Pharm. Res. 2017, 40, 328–337. [Google Scholar] [CrossRef] [PubMed]
  11. Li, H.L.; Li, X.M.; Liu, H.; Meng, L.H.; Wang, B.G. Two new diphenylketones and a new xanthone from Talaromyces islandicus EN-501, an endophytic fungus derived from the marine red alga Laurencia okamurai. Mar. Drugs 2016, 14, 223. [Google Scholar] [CrossRef] [PubMed]
  12. Wang, Y.C.; Zheng, Z.H.; Liu, S.C.; Zhang, H.; Li, E.; Guo, L.D.; Che, Y.S. Oxepinochromenones, furochromenone, and their putative precursors from the endolichenic fungus Coniochaeta sp. J. Nat. Prod. 2010, 73, 920–924. [Google Scholar] [CrossRef] [PubMed]
  13. Healy, P.C.; Hocking, A.; Tran-Dinh, N.; Pitt, J.I.; Shivas, R.G.; Mitchell, J.K.; Kotiw, M.; Davis, R.A. Xanthones from a microfungus of the genus Xylaria. Phytochemistry 2004, 65, 2373–2378. [Google Scholar] [CrossRef] [PubMed]
  14. Siddiqui, I.N.; Zahoor, A.; Hussain, H.; Ahmed, I.; Ahmad, V.U.; Padula, D.; Draeger, F.; Schulz, B.; Meier, K.; Steinert, M.; et al. Diversonol and blennolide derivatives from the endophytic fungus Microdiplodia sp.: absolute configuration of diversonol. J. Nat. Prod. 2011, 74, 365–373. [Google Scholar] [CrossRef] [PubMed]
  15. Barton, D.H.R.; Scott, A.I. The constitutions of geodin and erdin. J. Chem. Soc. 1958, 1767–1772. [Google Scholar] [CrossRef]
  16. Xue, J.H.; Fu, Y.; Wu, P.; Xu, L.X.; Huang, R.M.; Wei, X.Y.; Li, H.X. Two new anthraquinones from the soil fungus Penicillium purpurogenum SC0070. J. Antibiot. 2015, 68, 598–599. [Google Scholar] [CrossRef] [PubMed]
  17. Zhao, D.; Cao, F.; Guo, X.J.; Zhang, Y.R.; Kang, Z.J.; Zhu, H.J. Antibacterial indole alkaloids and anthraquinones from a sewage-derived fungus Eurotium sp. Chem. Nat. Compd. 2018, 54, 399–401. [Google Scholar] [CrossRef]
  18. Peng, J.X.; Zhang, X.M.; Du, L.; Wang, W.; Zhu, T.J.; Gu, Q.Q.; Li, D.H. Sorbicatechols A and B, antiviral sorbicillinoids from the marine-derived fungus Penicillium chrysogenum PJX-17. J. Nat. Prod. 2014, 77, 424–428. [Google Scholar] [CrossRef] [PubMed]
  19. Peng, J.X.; Zhang, X.M.; Wang, W.; Zhu, T.J.; Gu, Q.Q.; Li, D.H. Austalides S–U, new meroterpenoids from the sponge-derived fungus Aspergillus aureolatus HDN14–107. Mar. Drugs 2016, 14, 131. [Google Scholar] [CrossRef] [PubMed]
  20. Pertino, M.W.; Petrera, E.; Alché, L.E.; Schmeda-Hirschmann, G. Synthesis, antiviral and cytotoxic activity of novel terpenyl hybrid molecules prepared by click chemistry. Molecules 2018, 23, 1343. [Google Scholar] [CrossRef] [PubMed]
  21. Shagufta; Ahmad, I. Recent insight into the biological activities of synthetic xanthone derivatives. Eur. J. Med. Chem. 2016, 116, 267–280. [Google Scholar] [CrossRef] [PubMed]
  22. Santos, C.M.M.; Freitas, M.; Fernandes, E. A comprehensive review on xanthone derivatives as α-glucosidase inhibitors. Eur. J. Med. Chem. 2018, 157, 1460–1479. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Structures of 17.
Figure 1. Structures of 17.
Marinedrugs 16 00449 g001
Figure 2. Selected key HMBC correlations of 16.
Figure 2. Selected key HMBC correlations of 16.
Marinedrugs 16 00449 g002
Table 1. 1H and 13C NMR data for 13 (DMSO-d6).
Table 1. 1H and 13C NMR data for 13 (DMSO-d6).
Position1 a2 a3 b
δC, typeδH (J in Hz)δC, typeδH (J in Hz)δC, typeδH (J in Hz)
1155.8, C 158.6, C 158.4, C
2113.8, C 112.3, CH6.83, s109.3, CH7.05, s
3145.7, C 145.7, C 144.2, C
4108.7, CH 7.12, s109.7, C 111.0, C
5103.3, CH6.96, d (2.1)103.3, CH6.97, brs103.3, CH6.91, d (2.2)
6164.4, C 164.5, C 163.3, C
7113.1, CH6.89, d (2.1)113.4, CH6.89, brs113.3, CH6.78, d (2.2)
8135.0, C 135.0, C 135.8, C
9178.9, C 178.9, C 172.9, C
1120.8, CH32.46, s20.6, CH32.43, s21.2, CH32.49, s
12168.1, C 168.1, C 169.3, C
1352.7, CH33.88, s52.7, CH33.88, s52.9, CH33.84, s
14 56.8, CH33.88, s
1a106.6, C 107.0, C 112.5, C
4a153.1, C 150.4, C 152.4, C
9a108.7, C 108.7, C 112.2, C
10a157.7, C 157.4, C 156.3, C
HO-6 11.63, brs 11.25, brs
HO-1 13.00, s 12.29, brs
a 1H (400 MHz) and 13C (100 MHz); b 1H (500 MHz) and 13C (125 MHz).
Table 2. 1H and 13C NMR data for 46 (1H 400 MHz and 13C 100 MHz in DMSO-d6).
Table 2. 1H and 13C NMR data for 46 (1H 400 MHz and 13C 100 MHz in DMSO-d6).
Position456
δC, typeδH (J in Hz)δC, typeδH (J in Hz)δC, typeδH (J in Hz)
1159.8, C 158.7, C 154.6, C
2107.7, CH6.82, s112.3, CH6.85 s114.4, C
3146.5, C 145.6, C 142.6, C
4109.4, CH6.94, s109.6, C 110.6, C
5102.8, CH6.87, s102.8, CH6.93, s102.7, CH6.95, d (2.2)
6162.5, C 164.6, C 164.9, C
7112.2, CH6.73, s112.3, CH6.85 s113.1, CH6.86, d (2.1)
8135.4, C 137.4 137.2, C
9172.7, C 179.0, C 178.7, C
1121.8, CH32.42, s20.6, CH32.44, s
12169.1, C 168.9, C 168.8, C
1352.4, CH33.84, s 18.5, CH32.56, s
1456.1, CH33.87, s
1a109.2, C 107.0, C 107.1, C
4a156.9, C 150.4, C 149.0, C
9a112.2, C 108.3, C 108.0, C
10a156.1, C 157.5, C 157.5, C
HO-6 12.44, s
HO-1 12.44, s 13.22, brs
Table 3. Antiviral activity of xanthones 17 against H1N1, HSV-1 and HSV-2.
Table 3. Antiviral activity of xanthones 17 against H1N1, HSV-1 and HSV-2.
CompoundsIC50 (μM)Acyclovir cRibavirin c
1234567
H1N1 a133.444.6>200>200>200>200140.4 101.4
HSV-1 b55.521.4139.4157.7183.3144.475.7150.2
HSV-2 b175.576.7>200163.3>200>20095.4128.6
a Tested on MDCK cells; b Tested on Vero cells; c Positive control.

Share and Cite

MDPI and ACS Style

Kang, H.-H.; Zhang, H.-B.; Zhong, M.-J.; Ma, L.-Y.; Liu, D.-S.; Liu, W.-Z.; Ren, H. Potential Antiviral Xanthones from a Coastal Saline Soil Fungus Aspergillus iizukae. Mar. Drugs 2018, 16, 449. https://doi.org/10.3390/md16110449

AMA Style

Kang H-H, Zhang H-B, Zhong M-J, Ma L-Y, Liu D-S, Liu W-Z, Ren H. Potential Antiviral Xanthones from a Coastal Saline Soil Fungus Aspergillus iizukae. Marine Drugs. 2018; 16(11):449. https://doi.org/10.3390/md16110449

Chicago/Turabian Style

Kang, Hui-Hui, Huai-Bin Zhang, Mei-Jia Zhong, Li-Ying Ma, De-Sheng Liu, Wei-Zhong Liu, and Hong Ren. 2018. "Potential Antiviral Xanthones from a Coastal Saline Soil Fungus Aspergillus iizukae" Marine Drugs 16, no. 11: 449. https://doi.org/10.3390/md16110449

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