Next Article in Journal
Ciona intestinalis as a Marine Model System to Study Some Key Developmental Genes Targeted by the Diatom-Derived Aldehyde Decadienal
Next Article in Special Issue
Identification and Bioactivity of Compounds from the Fungus Penicillium sp. CYE-87 Isolated from a Marine Tunicate
Previous Article in Journal
Solvent Separating Secondary Metabolites Directly from Biosynthetic Tissue for Surface-Assisted Laser Desorption Ionisation Mass Spectrometry
Previous Article in Special Issue
Territrem and Butyrolactone Derivatives from a Marine-Derived Fungus Aspergillus Terreus
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A New Cyclic Hexapeptide and a New Isocoumarin Derivative from the Marine Sponge-Associated Fungus Aspergillus similanensis KUFA 0013

1
ICBAS-Instituto de Ciências Biomédicas Abel Salazar, Universidade do Porto, Rua de Jorge Viterbo Ferreira, 228, 4050-313 Porto, Portugal
2
Interdisciplinary Centre of Marine and Environmental Research (CIIMAR), Rua dos Bragas 289, 4050-123 Porto, Portugal
3
Centro de Química Medicinal da Universidade do Porto (CEQUIMED-UP), Laboratório de Química Orgânica, Departamento de Ciências Químicas, Faculdade de Farmácia, Rua de Jorge Viterbo Ferreira, 228, 4050-313 Porto, Portugal
4
Department of Plant Pathology, Faculty of Agriculture, Kasetsart University, Bangkok 10240, Thailand
5
Departamento de Química & QOPNA, Universidade de Aveiro, 3810-193 Aveiro, Portugal
*
Author to whom correspondence should be addressed.
Mar. Drugs 2015, 13(3), 1432-1450; https://doi.org/10.3390/md13031432
Submission received: 27 January 2015 / Revised: 4 March 2015 / Accepted: 9 March 2015 / Published: 17 March 2015
(This article belongs to the Special Issue Bioactive Compounds from Marine Fungi)

Abstract

:
A new isocoumarin derivative, similanpyrone C (1), a new cyclohexapeptide, similanamide (2), and a new pyripyropene derivative, named pyripyropene T (3) were isolated from the ethyl acetate extract of the culture of the marine sponge-associated fungus Aspergillus similanensis KUFA 0013. The structures of the compounds were established based on 1D and 2D NMR spectral analysis, and in the case of compound 2 the stereochemistry of its amino acid constituents was determined by chiral HPLC analysis of the hydrolysate by co-injection with the d and l amino acids standards. Compounds 2 and 3 were evaluated for their in vitro growth inhibitory activity against MCF-7 (breast adenocarcinoma), NCI-H460 (non-small cell lung cancer) and A373 (melanoma) cell lines, as well as antibacterial activity against reference strains and the environmental multidrug-resistant isolates (MRS and VRE). Only compound 2 exhibited weak activity against the three cancer cell lines, and neither of them showed antibacterial activity.

1. Introduction

In recent years, there has been an increasing interest in marine-derived fungi as a target source of bioactive marine natural products because many consider them among the world’s greatest untapped resources for new biodiversity as well as chemodiversity [1,2,3]. Moreover, through established culture methods, the compounds can be produced in quantity needed for medicinal chemistry development, clinical trials and even marketing. Among the marine fungal strains investigated, the fungi of the genus Aspergillus are the most prolific source of bioactive secondary metabolites, including sterols [4], cerebrosides [5], sesquiterpenoids [6,7], sesterterpenoids [8,9], diterpenoids [10], meroterpenoids [11], anthraquinone derivatives [12,13], nucleoside derivatives [14], indole alkaloids [15,16,17], prenylated indole alkaloids [18,19,20,21], quinazolinone alkaloids [22,23], pyrrolidine alkaloids [8], and cyclic peptides [24,25,26,27,28].
In our ongoing search for new natural products with antibacterial and anticancer activities produced by the marine-derived fungi of the genera Neosartorya and Aspergillus, we have recently reported the isolation of new isocoumarins similanpyrones A and B, a new chevalone (chevalone E), and a new natural product pyripyrone S; besides the previously reported chevalone B and C, a meroterpenoid S14-95, and pyripyropene E, from the crude ethyl acetate extract of the undescribed marine sponge-associated fungus Aspergillus similanensis KUFA 0013 [29]. Reexamination of the fractions remaining from the previous study of this fungus led to the isolation of a new 8-hydroxy-3-methylisocoumarin derivative, which we have named similanpyrone C (1), a new cyclic hexapeptide, similanamide (2), and a new pyripyropene analog, pyripyropene T (3) (Figure 1). Hydrolysis of compound 2, followed by HPLC analysis of its hydrolysate using a chiral column, led to the elucidation of its amino acid constituents. Compounds 2 and 3 were evaluated for their antibacterial activity as well their cytotoxicity against three human cancer cell lines.
Figure 1. Secondary metabolites from Aspergillus similanensis KUFA0013.
Figure 1. Secondary metabolites from Aspergillus similanensis KUFA0013.
Marinedrugs 13 01432 g001

2. Results and Discussion

Compound 1 was isolated as pale yellow viscous mass, and its molecular formula C19H20O6 was established on the basis of the (+)-HRESIMS m/z 345.1342 [M + H]+, indicating ten degrees of unsaturation. The IR spectrum showed absorption bands for hydroxyl (3443 cm−1), ketone carbonyl (1730 cm−1), conjugated lactone carbonyl (1683 cm−1), olefin (1647 cm−1) and aromatic (1625, 1572 cm−1) groups. The 13C NMR (Supplementary Information, Figure S3), DEPTs and HSQC spectra (Table 1, Supplementary Information, Figure S4) revealed the presence of one ketone carbonyl (δC 205.3), one conjugated ester carbonyl (δC 166.8), six quaternary sp2C 160.0, 158.3, 153.4, 136.7, 110.0 and 99.6), two methine sp2C 104.4 and 103.0), one ketal (δC 102.9), two sp3 methine (δC 67.2 and 33.9), two sp3 methylene (δC 48.3 and 47.2), and three methyl (δC 21.6, 19.4 and 15.9) carbons. The 1H NMR spectrum (Supplementary Information, Figure S1) revealed, besides a singlet of the hydrogen bonded hydroxyl proton at δH 11.35, one doublet at δH 6.13 (J = 0.9 Hz) and one singlet at δH 6.26, two multiplets at δH 4.15 and δH 1.98, two double double doublets at δH 2.48 (J = 14.7, 2.9, 1.9 Hz) and δH 2.28 (J = 14.7, 11.3, 0.7 Hz), two double doublets at δH 2.81 (J = 16.8, 5.6 Hz) and δH 2.55 (14.0, 1.9), two broad doublets at δH 2.81 (J = 14.0 Hz) and δH 2.57 (J = 16.8 Hz), one methyl singlet at δH 2.24, and two methyl doublets at δH 1.23 (J = 6.2 Hz) and δH 1.21 (J = 6.2 Hz).
Table 1. 1H and 13C NMR (CDCl3, 500.13 MHz and 125.8 MHz) and HMBC assignment for 1.
Table 1. 1H and 13C NMR (CDCl3, 500.13 MHz and 125.8 MHz) and HMBC assignment for 1.
PositionδC, TypeδH, (J in Hz)COSYHMBCNOESY
1166.8, CO-
3153.4, C-
4104.4, CH6.13, d (0.9)CH3-3C-3, 5, 8aCH3-3
4a136.7, C-
5103.0, CH6.26, s C-4, 6, 7, 8a
6158.3, C-
7110.0, C-
8160.0, C-
8a99.6, C-
23.6, CH22.81, dd (16.8, 5.6)H-9β, H-10C-6, 7, 10, 11H-9β, H-10, CH3-10
2.57, brd (16.8)H-9α, H-10C-7, 10, 11H-9α, H-10, CH3-10
1033.9, CH1.98, mH-9α, H-9β, CH3-10 H-9α,H-9β, CH3-10
11102.9, C-
12α47.2, CH22.55, dd (14.0, 1.9)H-12βC-10, 11, 13H-12β
12β 2.81, brd (14.0)H-12αC-10, 11, 13H-12α
13205.3, CO-
14α48.3, CH22.28, ddd (14.7, 11.3, 0.7)H-14β, H-15C-13, 15H-14β, CH3-15
14β 2.48, ddd (14.7, 2.9, 1.9)H-14α, H-15C-13H-14β, CH3-15
1567.2, CH4.15, mH-14α, H-14β, CH3-15 H-14β, CH3-15
CH3-319.4, CH32.24, sH-4C-3, 4H-4
CH3-1015.9, CH31.23, d (6.2)H-10C-9, 10, 11H-9β, H-10, H-12α, 2β
CH3-1521.6, CH31.21, d (6.2)H-15C-14, 15H-14α, H-14β, H-15
OH-8-11.35, s C-7, 8, 8a
Analysis of the 1H, 13C NMR, HSQC and HMBC spectra (Table 1) revealed the presence of a 6,7-disubstituted 3-methyl-1H-isochromen-1-one nucleus, similar to that of similanpyrone B [29]. Thus, another portion of the molecule consisted of one ketone (δC 205.3), one ketal (δC 102.9), one methine (δH 1.98, m; δC 33.9), one oxymethine (δH 4.15, m; δC 67.2), three methylene (δH 2.81, dd, J = 16.8, 5.6 Hz and δH 2.57, brd, J = 16.8 Hz, δC 23.6; δH 2.81, brd, J = 14.0 Hz and 2.55, dd, J = 14.0, 1.9 Hz; δC 47.2; δH 2.48, ddd, J = 14.7, 2.9, 1.9 Hz and 2.28, ddd, J = 14.7, 11.3, 0.7 Hz, δC 48.3), two methyl (δH 1.21, d, J = 6.2 Hz; δC 21.6 and δH 1.23, d, J = 6.2 Hz, δC 15.9) groups. That this portion was 2,10-dimethyl-1,7-dioxaspiro[5.5]undec-8-en-4-one was evidenced by the COSY correlations (Table 1, Supplementary Information, Figure S2) of H2-9 (δH 2.81, dd, J = 16.8, 5.6 Hz and 2.57, brd, J = 16.8 Hz) to H-10 (δH 1.98, m), of H-10 to CH3-10 (δH 1.23, d, J = 6.2 Hz), of H2-14 (δH 2.48, ddd, J = 14.7, 2.9, 1.9 Hz and 2.28, ddd, J = 14.7, 11.3, 0.7 Hz) to H-15 (δH 4.15, m), and of H-15 to CH3-15 (δH 1.21, d, J = 6.2 Hz), as well as by the HMBC cross peaks (Table 1, Supplementary Information, Figure S5) of H2-9 to C-10 and C-11 (δC 102.9), of H2-12 (δH 2.81, brd, J = 14.0 Hz and 2.55, dd, J = 14.0, 1.9 Hz) to C-10, 11, 13 (δC 205.3), and of H-14 to C-13 and 15, respectively (Figure 2). That the 1,7-dioxaspiro[5.5]undec-8-en-4-one ring system was fused with the 3-methyl-1H-isochromen-1-one nucleus, through C-8 and C-9 of the methyldihydropyran ring of the former and C-6 and C-7 of the latter, was supported by the HMBC correlations of H2-9 to C-6, 7 (Table 1, Supplementary Information, Figure S5). Literature search revealed that 1 is a new compound, and in compliance with our previous work, we have therefore named it similanpyrone C. Since 1 was isolated as viscous mass, it was not possible to obtain suitable crystals for the X-ray analysis. Consequently, the absolute configuration of C-10, C-11 and C-15 is still undetermined.
Figure 2. Key HMBC correlations of compound 1.
Figure 2. Key HMBC correlations of compound 1.
Marinedrugs 13 01432 g002
In an attempt to determine the stereochemistry of C-10 and C-15, we have sorted out the coupling constants of both H2-9 and H2-14. The fact that H2-9 appeared as a double doublet at δH 2.81 with a germinal coupling of 16.8 Hz, and a small coupling at δH 5.6 Hz, typical of axial-equatorial coupling, and a broad doublet at δH 2.57 with a germinal coupling constant of 16.8 Hz, we concluded that CH3-10 was in the β-axial position of the half-chair conformation of the tetrahydropyran ring. This was corroborated by the correlations of H-10, which was in α equatorial position, to both H-9β (δH 2.57, brd, J = 16.8 Hz) and H-9α (δH 2.81, dd, J = 16.8, 5.6 Hz) in the NOESY spectrum (Table 1, Supplementary Information, Figure S6). On the contrary, one of H-14 appeared as a double double doublet at δH 2.28, with a germinal coupling of 14.7 Hz and a diaxial coupling of 11.3 Hz, as well as a small coupling (long range) of 0.7 Hz, while another appeared also as a double double doublet at δH 2.48, with a germinal coupling of 14.7 Hz and an axial-equatorial coupling of 2.9 Hz, as well as a small coupling (long range) of 1.9 Hz. These patterns of couplings revealed that CH3-15 was in α equatorial position of the chair conformation of the tetrahydro-4H-pyran-4-one ring. This analysis was corroborated by the NOESY spectrum (Table 1, Supplementary Information, Figure S6), which exhibited a correlation only between H-15 and H-14β (δH 2.48, ddd, J = 14.7, 2.9, 1.9 Hz) and not between H-15 and H-14α (δH 2.28, ddd, J = 14.7, 11.3, 0.7 Hz). This assignment was also supported by the higher chemical shift value of H-14β than that of H-14α since the former is under the anisotropic deshielding of the C-13 carbonyl group. Based on the same reasoning, we assigned the broad doublet at δH 2.81 (J = 14.0 Hz) as H-12β, and the double doublet at δH 2.55 (J = 14.0, 1.9 Hz) as H-12α. Consequently, the relative configuration of C-10 and C-15 was tentatively assigned as 10S* and 15R*. The relative configuration of C-11 was tentatively assigned as 11S* based on the fact that the NOESY spectrum (Table 1, Supplementary Information, Figure S6) did not exhibit any correlation between H-10 and H-15. According to the molecular model, when the configuration of C-11 is S*, the substituents on C-11 are arranged in a way that H-10 and H-15 are pointing toward the opposite directions. On the contrary, the R* configuration of C-11 would have H-10 and H-15 close enough to give a strong NOESY correlation.
Similanpyrone C (1) can be assumed to be derived from the acetate-malonate pathway (Scheme 1). Cyclization and enolization of the pentaketide (I) leads to the formation of 6,8-dihydroxy-3-methylisocoumarin (II), which, after Claisen condensation with the tetraketide (III), gives rise to IV. Enolization of the side chain, together with a formation of the hemiketal by the phenolic hydroxyl group on C-6 of the isocoumarin nucleus and the ketone carbonyl of the side chain, leads to the formation of a hemiketal V. Formation of the ketal and methylation by SAM in the side chain finally gives rise to similanpyrone C (1).
Compound 2 was isolated as pale yellow viscous mass, and its molecular formula C34H52N6O6 was established on the basis of the (+)-HRESIMS m/z 641.4053 [M + H]+, indicating twelve degrees of unsaturation. The IR spectrum showed absorption bands for amine (3335 cm−1), carbonyl (1682, 1644 cm−1) and aromatic (1594, 1519 cm−1). The 13C NMR (Supplementary Information, Figure S8), DEPTs and HSQC spectra (Table 2, Supplementary Information, Figure 10S) revealed the presence of six amide carbonyls (δC 174.3, 174.2, 170.7, 170.2, 169.3, 168.9), two quaternary sp2C 137.0, 122.7), four methine sp2C 131.7, 127.1, 123.9, 123.4), eight methine sp3C 65.1, 61.4, 59.3, 50.9, 47.9, 29.9, 25.5, 24.4), six methylene sp3C 52.5, 37.8, 36.2, 28.1, 27.4, 24.5) and eight methyl (δC 37.9, 23.3, 23.2, 22.1, 21.7, 19.8, 18.4, 16.2) carbons. The 1H NMR spectrum (Table 2, Supplementary Information, Figure S7) revealed, besides four NH signals at δH 7.43, d (J = 7.4 Hz), 7.64, d (J = 9.8 Hz), 8.02, d (J = 7.9 Hz) and 9.41, brs, the signals of the aromatic protons of the 1,2-disubstituted benzene ring at δH 7.20, dd (J = 7.7, 1.5 Hz), 7.13, ddd (J = 7.9, 7.9, 1.0 Hz), 7.47, ddd, (J = 7.9, 7.9, 1.6 Hz) and 8.29, d (J = 8.3 Hz). That the 1,2-disubstituted benzene ring belonged to the anthranilic acid residue was corroborated by the HMBC correlations of the NH signal at δH 9.41, brs to the carbon signal at δC 123.9 (C-6), and of the double doublet at δH 7.20 (J = 7.7, 1.5 Hz, H-3) to the carbons at δC 131.7 (C-5), 137.0 (C-7) and 170.2 (CO-1) (Table 2, Figure 3, Supplementary Information, Figure S11). The anthranilic acid residue was linked to the valine residue, through the amino group of the former and the carboxyl group of the latter, since the HMBC spectrum (Supplementary Information, Figure S11) showed correlations of the NH signal at δH 9.41, brs to the carbonyl carbon at δC 170.7 (C-8), of the methine proton at δH 4.32 dd, J = 7.4, 3.3 Hz (H-9) to the methine carbon at δC 29.9 (C-10), the methyl carbon at δC 16.2 (C-12) and C-8, and of the NH signal at δH7.43, d (J = 7.5 Hz) to C-9 (δC 59.3) and C-10 (Table 2, Figure 3). The presence of the leucine residue was supported by the coupling system from CH-14 (δH 4.57, m; δC 50.9) through CH3-17 (δH 0.97, d, J = 6.5 Hz; δC 23.3) and CH3-18 (δH 0.88, d, J = 6.4 Hz; δC 21.1), and of NH at δH 8.02 d (J = 7.9 Hz) to H-14, as observed in the COSY spectrum (Table 2, Supplementary Information, Figure S9), as well as by the HMBC correlations of the NH signal at δH 8.02 d (J = 7.9 Hz) to C-14 (Table 2, Figure 3). That the valine residue was linked to the leucine residue was supported by the HMBC cross peak between the NH signal of the former (δH7.43, d, J = 7.5 Hz) to the signal of the carbonyl carbon (δC174.2, C-13) of the latter. In turn, the leucine residue was linked to the alanine residue, as evidenced by the HMBC cross peaks of the NH signal of the former to the carbonyl carbon signal (δC174.3, C-19) of the latter, and of the proton signal at δH 4.82, dd, J = 9.7, 7.3 Hz (H-20) to C-19 and the methyl carbon at δC 18.4 (C-21), as well as by the COSY cross peaks of H-20 to CH3-21 (δH 1.29, d, J = 7.3 Hz), and of H-20 to NH at δH 7.64, d (J = 7.9 Hz). The presence of the N- methyl leucine moiety was evidenced by the coupling system from H-23 (δH 3.49, dd, J = 9.0, 4.7 Hz) through CH3-26 (δH 0.97d, J = 6.5; δH 23.2) and CH3-27 (δH 0.99, d, J = 6.5; δH 22.1), as observed in the COSY spectrum (Table 2, Supplementary Informatscheion, Figure S9), as well as by the HMBC correlations of H-23 to C-22 (δC 169.3) and CH3-28 (δC 37.9). Finally, the presence of the pipecolic acid residue was supported by the coupling system from CH-30 (δH 3.71, dd, J = 11.3, 2.5; δC 61.4) through CH2-34 (δH 3.16, dd, J = 13.2, 2.3; 4.14, dd, J = 14.4, 2.4; δC 52.5), as observed in the COSY spectrum (Table 2, Supplementary Information, Figure S9). Since both H-30 and CH3-28 gave HMBC cross peaks to C-29 (δC 168.9), the pipecolic acid residue was linked to the N-methyl leucine residue through the carboxyl group of the former and a nitrogen atom of the latter. Since 2 presents twelve degrees of unsaturation, the nitrogen atom of the piperidine ring of the pipecolic acid residue was linked to the carbonyl of the anthranilic acid residue. The proposed structure was supported by the NOESY correlations which showed cross peaks of NH at δH 9.41, brs to H-9, CH3-12, of NH at δH 7.43 (d, J = 7.5 Hz) to H-9, CH3-11 (δH 1.06, d, J = 6.9 Hz), CH3-12, H-14, of NH at δH 8.02 (d, J = 7.9 Hz) to H-14, H-15 (δH 2.02, m; 1.77, m), CH3-17, of NH at δH 7.64 (d, J = 7.9 Hz) to CH3-21, H-23, CH3-28, of H-3 to H-34β (δH 4.14, dd, J = 14.4, 2.4 Hz), and of H-30 to CH3-28 (Table 2, Supplementary Information, Figure S12). Combining this information, it was possible to conclude that 2 was cyclo (anthranilic acid-Val-Leu-Ala-N-methyl-Leu-pipecolic acid).
Scheme 1. Proposed biogenesis of similanpyrone C (1).
Scheme 1. Proposed biogenesis of similanpyrone C (1).
Marinedrugs 13 01432 g004
Table 2. 1H and 13C NMR (CDCl3, 500.13 MHz and 125.8 MHz) and HMBC assignment for 2.
Table 2. 1H and 13C NMR (CDCl3, 500.13 MHz and 125.8 MHz) and HMBC assignment for 2.
PositionδC, TypeδH, (J in Hz)COSY HMBCNOESY
Anthranilic acid1170.2, CO-
2122.7, C-
3127.1, CH7.20, dd (7.7, 1.5)H-4C-1, 5, 7H-34
4123.4, CH7.13, ddd (7.9, 7.9, 1.0)H-3, 5C-2, 6
5131.7, CH7.47, ddd (7.9, 7.9, 1.6)H-4, 6C-3, 7
6123.9, CH8.29, d (8.3)H-5C-2, 4H-12
7137.0, C-
NH-9.41, brs C-6, 7, 8NH (Val), H-9, 12
Val8170.7, CO-
959.3, CH4.32, dd (7.4, 3.3)H-10, NHC-8, 10, 11, 12H-10, 11
1029.9, CH2.68, mH-9, 11, 12 H-9, 11, 12
1119.8, CH31.06, d (6.9)H-10C-9, 10, 12
1216.2, CH30.94, d (7.0)H-10C-9, 10, 11
NH-7.43, d (7.5)H-9C-9, 10, 13H-9, 11, 12, 14
Leu13174.2, CO-
1450.9, CH4.57, mH-15, NH H-15, 18
1536.2, CH22.02, m; 1.77, mH-14, 16
1624.4, CH1.77, mH-15, 17, 18
1723.3, CH30.97, d (6.5)H-16C-15, 16, 18
1821.7, CH30.88, d (6.4)H-16C-15, 16, 17
NH-8.02, d (7.9)H-14C-13, 19NH (Ala), H-14, 15, 17
Ala19174.3, CO-
2047.9, CH4.82, dd (9.7, 7.3)H-21, NHC-19, 21H-21
2118.4, CH31.29, d (7.3)H-20C-19, 20
NH-7.64, d (7.9)H-20C-22C-21, 23, 28
N-Me Leu22169.3, CO-
2365.1, CH3.49, dd (9.0, 4.7)H-24C-22, 24, 28, 29
2437.8, CH21.95, m; 2.20, mH-23, 25
2525.5, CH1.65, mH-24, 26, 27
2623.2, CH30.97, d (6.5)H-25C-24, 25, 27
2722.1, CH30.99, d (6.5)H-25C-24, 25, 26
2837.9, CH33.20, s C-23, 2923, 30, 32, 34α
Pipecolic acid29168.9, CO-
3061.4, CH3.71, dd (11.3, 2.5)H-31C-1, 29H-34α
3128.1, CH22.05, mH-30, 32
3224.5, CH22.07, mH-31, 33
3327.4, CH21.56, mH-32, 34
34α52.5, CH23.16, dd (13.2 2.3)H-33 H-34β
34β 4.14, dd (14.4, 2.4) H-34α
Figure 3. Key HMBC correlations of compound 2.
Figure 3. Key HMBC correlations of compound 2.
Marinedrugs 13 01432 g003
The 1H and 13C NMR data of compound 2 resembled those of PF1171C, a cyclic hexapeptide isolated from extracts of the unidentified ascomycete OK-128 fermented with okara by Kai et al. [30], and later by a total synthesis by Masuda et al. [31]; however, its value of optical rotation was quite different from that of PF1171C [30,31]. Moreover, PF1171C was reported as white solid (mp 138–140 °C) [31], while compound 2 is pale yellow viscous mass. This observation suggested that compound 2 and PF1171C could be diastereomers.
The stereochemistry of the amino acid residues of compound 2 was determined by chiral HPLC analysis of its acidic hydrolysate, using appropriate d and l amino acids standards, according to the previously described method [32]. The enantioseparations of the standard amino acids were successfully performed with the Chirobiotic T column under reversed-phase elution conditions [33]. Actually, the teicoplanin selector of this column has several characteristic features that make it suitable for amino acid analysis [33,34,35]. The elution order of the enantiomers of all the standard amino acids was confirmed by injecting the solutions of the racemic or enantiomeric mixtures and then each enantiomer separately at a flow rate of 1 mL/min (Supplementary Information, Figure S18). As predicted, the D enantiomer was always more strongly retained than the corresponding L enantiomer on the Chirobiotic T column [33]. Based on mix HPLC analyses of the acidic hydrolysate with standard amino acids (co-injection) (Supplementary Information, Figure S19 and Table S1), compound 2 was elucidated unambiguously as cyclo (anthranilic acid-l-Val-d-Leu-l-Ala-N-methyl-l-Leu-d-pipecolic acid). Although the amino acid sequence of compound 2 is the same as that of PF1171C, the stereochemistry of its amino acid constituents is different from that of the amino acids constituent of PF1171C. While PF1171C contains d-Ala, l-Leu, d-Val, and l-pipecolic acid, compound 2 contains l-Ala, d-Leu, l-Val and d-pipecolic acid. Thus compound 2 is a new compound, and we have named it similanamide.
It is interesting to note that Kai et al. [30] has firstly assigned the stereochemistry of the amino acid constituents of PF1171C using the reversed phase HPLC analysis of the Marfey derivatives of the amino acids. Since the retention times for the Marfey derivatives of d-Ala (19.4 min) and l-Ala (20.0 min) were too close for resolution, they had wrongly assigned d-Ala for l-Ala. However, in our chiral HPLC analysis using the Chirobiotic T column under reversed-phase elution conditions, not only the retention time of l-Ala (7.16 min) was very different from that of d-Ala (9.36 min), but also the retention times of the d/l pair of other amino acid constituents (Supplementary Information, Table S1).
So far, only few cyclohexapeptides have been reported from marine-derived microorganisms. Wu et al. [36] have isolated two cyclohexapeptides, nocardiamides A and B, from the culture broth of the marine-derived actinomycete Nocardiopsis sp. CNX037, isolated from sediment. Cai et al. [37] have isolated two anti-Mycobacterium tuberculosis cyclohexapeptides from a Streptomyces hygroscopicus Strain, while Song et al. [38] reported isolation of three new cyclopeptides, desotamides B–D, from the deep South China Sea-derived Streptomyces scopuliridis SCSIO ZJ46. To the best of our knowledge, compound 2 is the first cyclopeptide containing d-pipecolic acid residue ever isolated from marine fungi.
Compound 3 was isolated as pale yellow viscous mass, and its molecular formula C29H33NO8 was established on the basis of (+)-HRESIMS m/z 524.2287 [M + H]+, indicating fourteen degrees of unsaturation. The IR spectrum showed absorption bands for hydroxyl (3418 cm−1), ester carbonyl (1732 cm−1), conjugated ester carbonyl (1667cm−1), olefin (1643 cm−1), aromatic (1557, 1507 cm−1). The 13C NMR (Supplementary Information, Figure S15), DEPTs and HSQC spectra (Table 3, Supplementary Information, Figure S16) revealed the presence of two ester carbonyl (δC 170.1 and 169.8), one conjugated carbonyl (δC 161.6), five quaternary sp2C 160.2, 156.8, 146.1, 126.9, 100.3), six methine sp2C 151.3, 146.5, 132.8, 123.9, 109.4, 98.7), one oxyquaternary sp3C 85.9), two oxymethine sp3C 75.7 and 72.8), one oxymethylene sp3C 64.4), two quaternary sp3C 40.3, 38.4), one methine sp3C 40.7), three methylene sp3C 35.2, 27.3, 22.9), and five methyl (δC 23.8, 20.8, 20.5, 20.1 and 12.7) carbons. The general feature of 13C and 1H spectra of compound 3 closely resembled those of pyripyropene S, previously isolated from the same fungus [29]. Analysis of the 1H (Supplementary Information, Figure S13), 13C, HSQC and HMBC spectra (Table 3, Supplementary Information, Figure S17) revealed the presence of the ring system comprising of the decahydronaphthalene fused, on C-5 and C-6, with the 2H,5H-pyrano[4,3-b]pyran-5-one, which connected to the pyridine ring through C-6′ of the former and C-3″ of the latter, similar to pyripyropene S [29]. However, there were only two acetoxyl groups (δC 170.1, CO; δC 20.5, CH3, δH 2.00, s and δC 169.8, CO; δC 20.8, CH3; δH 2.00, s) in compound 3. That the acetoxyl groups were on C-1 and C-11 was supported by the fact that the chemical shift values of H-1 (δH 4.64, t, J = 8.5 Hz) and H-11 (δH 3.75, s) were very similar to those of pyripyropene S, whereas the chemical shift of H-7 (δH 3.85, dd, J = 10.6, 4.2 Hz) was nearly 1.4 ppm less than that of pyripyropene S. On the other hand, the 13C chemical shift values of C-6 (δH 85.9) and C-8 (δC 27.4) of compound 3 were 2.00 and 3.00 ppm, respectively, higher than those of the corresponding carbons in pyripyropene S, while the 13C chemical shift value of C-7 (δC 75.5) of compound 3 was 2.00 ppm lower than that of C-7 of pyripyropene S. Since H-7 appeared as a double doublet with coupling constants of 10.6 and 4.2 Hz, the position of the hydroxyl group on C-7 was β. Thus, compound 3 is 7-deacetylpyripyropene S. In order to prove the stereochemistry of compound 3, the NOESY experiment was carried out. As the NOESY spectrum (Table 3) clearly exhibited correlations of CH3-15 to CH3-12, but not to H-1 and H-9; of CH3-12 to CH3-14 and CH3-15, and of CH3-14 to CH3-12, but not to H-7 (Table 3, Supplementary Information, Figure S18), the stereochemistry of compound 3 is the same as that of pyripyropene S [29], i.e., 1S*, 4R*, 6S*, 7S*, 9R*, 10R*. Since it is a new compound we have named it pyripyropene T.
Table 3. 1H and 13C NMR (DMSO, 300.13 MHz and 75.47 MHz) and HMBC assignment for 3.
Table 3. 1H and 13C NMR (DMSO, 300.13 MHz and 75.47 MHz) and HMBC assignment for 3.
PositionδC, TypeδH, (J in Hz)COSY HMBC NOESY
172.8, CH4.64, t (8.5)H-2
222.9, CH21.79, m H-1, 3
335.2, CH21.98, m H-2
438.4, C -
5146.1, C-
685.9, C -
775.7, CH3.85, dd (10.6, 4.2)H-8
827.3, CH21.70, m H-7
940.7, CH1.48, m H-8
1040.3, C -
1164.4, CH23.75, s C-1, 9H3-15
1223.8, CH31.19, s C-3, 4, 5H3-14, 15
13109.4, CH6.16, s C-4, 6, 2″, 4″
1420.1, CH31.45, s C-5, 6, 7H3-12
1512.7, CH30.84, s C-1, 9, 10, 11H2-11, H3-12
2′161.6, C-
3′100.3, C-
4′160.2, C-
5′98.7, CH7.11, s C-3′, 4′, 6′, 3″
6′156.8, C-
2″146.5, CH9.0, d (1.7)H-4″C-3″, 6″H-5′
3″126.9, C-
4″132.8, CH8.25, dt (8.7, 2.2)H-2″, 5″ H-5′, 5″
5″123.9, CH7.54, dd (7.9, 4.8)H-4″, 6″C-3″H-4″, 6″
6″151.3, CH8.68, dd (4.8, 1.5)H-2″, 5″C-2″, 4″H-5″
OAc-1170.1, CO-
20.5, CH32.00, s CO (Ac)
OAc-11169.8, CO-
20.8, CH32.00, s CO (Ac)
Since compound 1 was isolated in a very small amount, only compounds 2 and 3 were evaluated for their cytotoxicity and antibacterial activity. Compounds 2 exhibited weak in vitro growth inhibitory activity, by Sulforhodamine B (SRB) assay [39], against the MCF-7 (breast adenocarcinoma, GI50 = 125 ± 0), NCI-H460 (non-small cell lung cancer, GI50 = 117.50 ± 3.55) and A373 (melanoma, GI50 = 115 ± 7.07) cell lines. Compounds 2 and 3 were also tested for their antibacterial activity against four reference strains (Staphylococcus aureus, Bacillus subtilis, Escherichia coli and Pseudomonas aeruginosa), as well as the environmental multidrug-resistant isolates, according to the previously described method [40], and neither of them showed activity (MIC values higher than 256 µg/mL).

3. Experimental Section

3.1. General Procedure

Melting points were determined on a Bock monoscope and are uncorrected. Optical rotations were determined on an ADP410 Polarimeter (Bellingham + Stanley Ltd., Tunbridge Wells, Kent, UK). Infrared spectra were recorded in a KBr microplate in a FTIR spectrometer Nicolet iS10 from Thermo Scientific (Waltham, MA, USA.) with Smart OMNI-Transmission accessory (Software 188 OMNIC 8.3). 1H and 13C-NMR spectra were recorded at ambient temperature on a Bruker AMC instrument (Bruker Biosciences Corporation, Billerica, MA, USA) operating at 500.13 and 125.8 MHz or at 300.13 and 75.4 MHz, respectively. High-resolution mass spectra were measured with a Waters Xevo QToF mass spectrometer (Waters Corporations, Milford, MA, USA) coupled to a Waters Aquity UPLC system. A Merck (Darmstadt, Germany) silica gel GF254 was used for preparative TLC, and a Merck Si gel 60 (0.2–0.5 mm) was used for analytical chromatography.

3.2. Extraction and Isolation

Isolation and identification of the fungus as well as fractionation of the crude extract of the culture of A. similanensis KUFA0013 have been previously described by us [29]. Frs 185–196 were combined (654 mg) and purified by TLC (Si gel, CHCl3:Me2CO:HCO2H, 97:3:0.1) to give 7.4 mg of 1. Frs 310–327 were combined (1.19 g), applied on a Sephadex H-20 column (10 g) and eluted MeOH, wherein ten sfrs of 1 mL were collected. Sfrs 1–7 were combined and purified by TLC (Si gel, CHCl3:Me2CO:HCO2H, 19:1:0.01) to give 108 mg of 2. Frs 336–345 were combined (165 mg) and purified by TLC (Si gel, CHCl3:Me2CO:HCO2H, 17:3:0.01) to give additional 60 mg of 2. Frs 354–398 were combined (1.15 g), applied on a Sephadex LH-20 column (10 g) and eluted with MeOH, wherein thirty four sfrs of 1 mL were collected. Sfrs 7–15 were combined (150 mg) and purified by TLC (Si gel, CHCl3:Me2CO:HCO2H, 4:1:0.01) to give 67 mg of pyripyropene S [29]. Frs 435–443 were combined (377 mg), applied on a Sephadex LH-20 column (10 g) and eluted with a mixture of 1:1 v/v of CHCl3: MeOH, wherein fourteen sfrs of 1 mL were collected. Sfrs 8–11 (62 mg) were combined and purified by TLC (Si gel, CHCl3:Me2CO:HCO2H, 19:1:0.01) to give 35 mg of 3.

3.2.1. Similanpyrone C (1)

Pale yellow viscous mass; [α]D20 = −80.0 (c 0.01, CHCl3); λmax (log ε) 239 (4.57), 245(4.59), 332 (2.10) nm; IR (KBr) νmax 3443, 2923, 2852, 1730, 1683, 1647, 1625, 1572, 1508, 1457, 1429, 1352, 1251 cm−1; 1H and 13C NMR (Table 1); HRESIMS m/z 345.1342 (M + H)+ (calculated for C19H21O6, 345.1338).

3.2.2. Similanamide (2)

Pale yellow viscous mass; [α]D20 = +30.3 (CHCl3, c 0.03), IR (KBr) νmax 3335, 3054, 2958, 2870, 1682, 1644, 1594, 1519, 1449, 1292 cm−1; 1H and 13C NMR (Table 2); HRESIMS m/z 641.4053 (calculated for C34H53N6O6, 641.4027).

3.2.3. Pyripyropene T (3)

Pale yellow viscous mass; [α]D20 = +106 (CHCl3, c 0.03), IR (KBr) νmax 3418, 2949, 1732, 1667, 1643, 1557, 1507, 1480, 1246, 1028 cm−1; 1H and 13C NMR (Table 3); HRESIMS m/z 524.2287 (M + H)+, calcd for C29H34NO8, 524.2284.

3.3. Amino Acids Analysis of Acidic Hydrolysate of Compound 2

3.3.1. Acid Hydrolysis

The stereochemistry of the amino acids was determined by analysis of the acidic hydrolysate from compound 2. Compound 2 (5.0 mg) was dissolved in 6 N HCl (5 mL) and heated at 110 °C, in a furnace, for 24 h in a sealed glass tube. After cooling to room temperature, the solution was dried under N2 for 24 h, reconstituted in methanol for HPLC-MS (200 µL), filtered through a 4 mm PTFE Syringe Filter F2504-4 of 0.2 µm pore size (Thermo Scientific, Mumbai, India), and then analyzed by HPLC equipped with a chiral column.

3.3.2. Chiral HPLC Analysis

The HPLC system consisted of Shimadzu LC-20AD pump, equipped with a Shimadzu DGV-20A5 degasser, a Rheodyne 7725i injector fitted with a 20 µL loop, and a SPD-M20A DAD detector (Kyoto, Japan). Data acquisition was performed using Shimadzu LCMS Lab Solutions software, version 3.50 SP2. The chiral column used in this study was Chirobiotic T (15 cm × 4.6 mm I.D., particle size 5 µm) manufactured by ASTEC (Whippany, NJ, USA). The mobile phase composition was MeOH:H2O:CH3CO2H (70:30:0.02, v/v/v), all were LC-MS grade solvents obtained from Sigma-Aldrich Co (St. Louis, MO, USA). The flow rate was 0.5 mL/min and the UV detection wavelength was 210 nm. Analyses were performed at room temperature in an isocratic mode.
All standards of racemic amino acids and pure amino acid enantiomers were purchased from Sigma-Aldrich Co (St. Louis, MO, USA). The elution order of the enantiomers of all the standards amino acids was confirmed by injecting the solutions of the racemic or enantiomeric mixtures, and then each enantiomer separately (or only l- amino acid in the case of N-methyl leucine) at a flow rate of 1 mL/min or 0.5 mL/min. Working solutions of single enantiomeric amino acids were prepared by dissolution in MeOH at the concentration of 1 mg/mL (10 µL sample injection), while the enantiomeric mixtures were prepared by mixing equal aliquots of each enantiomer (20 µL sample injection). Mix HPLC analyses of the acidic hydrolysate with standard amino acids (co-injection) confirmed the stereochemistry of the amino acids of compound 2.

4. Conclusions

Following our first report of the isolation of new isocoumarin derivatives and merotepenoids from the ethyl acetate crude extract of the culture of the undescribed marine sponge-associated fungus Aspergillus similanensis KUFA 0013, we have reexamined its remaining column fractions and have isolated a new isocoumarin derivative containing an unusual 1,7-dioxaspiro-undecenone moiety, together with a new cyclohexapeptide and a new pyripyropene analog. Although several cyclopeptides have been reported from many fungi of the genus Aspergillus, this is the first report of isolation of cyclohexapeptide from the marine-derived fungus. The fact that these new cyclohexapeptide and pyripyropene analog did not exhibit relevant antibacterial and the in vitro growth inhibitory activities on human cancer cell lines does not mean that they are void of other interesting biological activities. In order to prove this hypothesis, it is necessary to explore their potential in a broader biological or pharmacological assay system.

Supplementary Files

Supplementary File 1

Acknowledgments

This work was partially supported by the Project MARBIOTECH (reference NORTE-07-0124-FEDER-000047) within the SR&TD Integrated Program MARVALOR—Building research and innovation capacity for improved management and valorization of marine resources, supported by the Programa Operacional Regional do Norte (ON.2—O Novo Norte), the European Regional Development Fund, and FCT-Fundação para a Ciência e a Tecnologia under the project CEQUIMED-PEst-OE/SAU/UI4040/2014. We thank Michael Lee of the Department of Chemistry, Leicester University (UK), for providing the HRESIMS. C.P. thanks the Faculty of Pharmaceutical Sciences, Burapha University, Thailand for her scholarship to the University of Porto. We thank Júlia Bessa for technical support.

Author Contributions

Chadaporn Prompanya performed isolation, purification and structure elucidation of some compounds; Carla Fernandes and Sara Cravo performed HPLC analysis of amino acids by chiral column; Tida Dethoup isolated, identified, cultured the fungi, and prepared the crude extract; Artur M.S. Silva provided 1D and 2D NMR spectra. Madalena M.M. Pinto, and Anake Kijjoa conceived, designed the research, elucidated the structure of the compounds and wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bugni, T.S.; Ireland, C.M. Marine-derived fungi: A chemically and biologically diverse group of microorganisms. Nat. Prod. Rep. 2004, 21, 143–163. [Google Scholar] [CrossRef] [PubMed]
  2. Saleem, M.; Ali, M.S.; Hussain, S.; Jabbar, A.; Ashraf, M.; Lee, Y.S. Marine natural products of fungal origin. Nat. Prod. Rep. 2007, 24, 1142–1152. [Google Scholar] [CrossRef] [PubMed]
  3. Rateb, M.E.; Ebel, R. Secondary metabolites of fungi from marine habitats. Nat. Prod. Rep. 2011, 28, 290–344. [Google Scholar] [CrossRef] [PubMed]
  4. Liu, X.-H.; Miao, F.-P.; Liang, X.-R.; Ji, N.-Y. Ergosteroid derivatives from an algicolous strain of Aspergillus ustus. Nat. Prod. Res. 2014, 28, 1182–1186. [Google Scholar] [CrossRef] [PubMed]
  5. Yang, G; Sandjo, L.; Yun, K.; Leutou, A.S.; Kim, G.-D.; Choi, H.D.; Kang, S.J.; Hong, J.; Son, B.W. Flavusides A and B, antibacterial cerebrosides from the marine-derived fungus Aspergillus flavus. Chem. Pharm. Bull. 2011, 59, 1174–1177. [Google Scholar]
  6. Sun, L.-L.; Shao, C.-L.; Chen, J.–F.; Guo, Z.-Y.; Fu, X.-M.; Chen, M.; Chen, Y.-Y.; Li, R.; de Voogd, N.J.; She, Z.-G.; Lin, Y.-C.; Wang, C.-Y. New bisabolane sesquiterpenoids from a marine-derived fungus Aspergillus sp. isolated from the sponge Xestospongia testudinaria. Bioorg. Med. Chem. Lett. 2012, 22, 1326–1329. [Google Scholar] [CrossRef] [PubMed]
  7. Kitano, M.; Yamada, T.; Amagata, T.; Minoura, K.; Tanaka, R.; Numatra, A. Novel pyridino-α-pyrone sesquiterpene type pileotin prduced by a sea urchin-derived Aspergillus sp. Tetrahedron Lett. 2012, 53, 4192–4194. [Google Scholar] [CrossRef]
  8. Liu, H.-B.; Edrada-Ebel, R.; Wang, Y.; Schulz, B.; Draeger, S.; Müller, W.E.G.; Wray, V.; Lin, W.-H.; Proksch, P. Ophiobolin sesterterpenoids and pyrrolidine alkaloids from the sponge-derived fungus Aspergillus ustus. Helv. Chim. Acta 2011, 94, 623–631. [Google Scholar] [CrossRef]
  9. Zhang, D.; Fukuzawa, S.; Satake, M.; Li, X.; Kuranaga, T.; Niitsu, A.; Yoshizawa, K.; Tachibana, T. Ophiobolin O and 6-epi-ophiobolin O, two new cytotoxic sesterterpenes from the marine-derived fungus Aspergillus sp. Nat. Prod. Commun. 2012, 7, 1411–1414. [Google Scholar] [PubMed]
  10. Sun, H.-F.; Li, X.-M.; Cui, C.-M.; Gao, S.-S.; Li, C.-S.; Huang, C.-G.; Wang, B.-G. Asperolides A–C, tetranorlabdane diterpenoids from the marine alga-derived endophytic fungus Aspergillus wentii EN-48. J. Nat. Prod. 2012, 75, 148–152. [Google Scholar] [CrossRef] [PubMed]
  11. Zhou, Y.; Debbab, A.; Wray, V.; Lin, W.; Schulz, B.; Trepos, R.; Pile, C.; Hellio, C.; Proksch, P.; Aly, A.H. Marine bacterial inhibitors from the sponge-derived fungus Aspergillus sp. Tetrahedron Lett. 2014, 55, 2789–2792. [Google Scholar] [CrossRef] [Green Version]
  12. Zhang, Y.; Li, X.-M.; Wang, B.-G. Anthraquinone derivatives produced by marine-derived fungus Aspergillus versicolor EN-7. Biosci. Biotechnol. Biochem. 2012, 76, 1774–1776. [Google Scholar] [CrossRef] [PubMed]
  13. Huang, H.; Wang, F.; Luo, M.; Chen, Y.; Song, Y.; Zhang, W.; Zhang, S.; Ju, J. Halogenated anthraquinones from the marine-derived fungus Aspergillus sp. SCSIO F063. J. Nat. Prod. 2012, 75, 1346–1352. [Google Scholar] [CrossRef]
  14. Chen, M.; Fu, X.-M.; Kong, C.-J.; Wang, C.-Y. Nucleoside derivatives from the marine-derived fungus Aspergillus versicolor. Nat. Prod. Res. 2014, 28, 895–900. [Google Scholar] [CrossRef] [PubMed]
  15. Cai, S.; Kong, X.; Wang, W.; Zhou, H.; Zhu, T.; Li, D.; Gu, Q. Aspergilazine A, a diketopiperazine dímer with a rare N-1 to C-6 linkage, from a marine-derived fungus Aspergillus taichungensis. Tetrahedron Lett. 2012, 53, 2615–2617. [Google Scholar] [CrossRef]
  16. He, F.; Han, Z.; Peng, J.; Qian, P.-Y.; Qi, S.-H. Antifouling indole alkaloids from two marine-derived fungi. Nat. Prod. Commun. 2013, 8, 329–332. [Google Scholar] [PubMed]
  17. Ji, I.-Y.; Liu, Z.-H.; Miao, F.-P.; Qiao, M.-F. Aspeverin, a new alkaloid from algicolous strain of Aspergillus versicolor. Org. Lett. 2013, 15, 2327–2329. [Google Scholar] [CrossRef] [PubMed]
  18. Tsukamoto, S.; Umaoka, H.; Yoshikawa, K.; Ikeda, T.; Hirota, H. Notoamide O, a structurally unprecedented prenylated índole alkaloid, and notamides P-R from a marine-derived fungus, Aspergillus sp. J. Nat. Prod. 2010, 73, 1438–1440. [Google Scholar] [CrossRef] [PubMed]
  19. Wang, Y.; Lin, C.-L.; Bai, J.; Zhang, L.-M.; Wu, X.; Zhang, L.; Pei, Y.-H.; Jing, Y.-K.; Hua, H.-M. 2,5-Diketopiperazines from the marine-derived fungus Aspergillus fumigatus YK-7. Chem. Biodivers. 2012, 9, 385–393. [Google Scholar] [CrossRef] [PubMed]
  20. Song, F.; Liu, X.; Guo, H.; Ren, B.; Chen, C.; Piggott, A.M.; Yu, K.; Gao, H.; Wang, Q.; Liu, M.; et al. Brevianamides with antitubercular potential from a marine-derived isolate of Aspergillus versicolor. Org. Lett. 2012, 14, 4770–4773. [Google Scholar] [CrossRef] [PubMed]
  21. Chen, M.; Shao, C.-L.; Fu, X.-M.; Xu, R.-F.; Zheng, J.-J.; Zhao, D.-L.; he, Z.-G.; Wang, C.-Y. Bioactive indole alkaloids and phenyl ether derivatives from a marine-derived Aspergillus sp. fungus. J. Nat. Prod. 2013, 76, 547–553. [Google Scholar] [CrossRef] [PubMed]
  22. Peng, J.; Zhang, X.-Y.; Tu, Z.-C.; Xu, X.-Y.; Qui, S.-H. Alkaloids from the deep-sea-derived fungus Aspergillus westerdijkiae DFFSCS013. J. Nat. Prod. 2013, 76, 983–987. [Google Scholar] [CrossRef] [PubMed]
  23. He, F.; Bao, J.; Zhang, X.-Y.; Tu, Z.-C.; Shi, Y.-M.; Qi, S.-H. Asperterrestide A, a cytotoxic tetrapeptide from the marine-derived fungus Aspergillus terreus SCSGAF0162. J. Nat. Prod. 2013, 76, 1182–1186. [Google Scholar] [CrossRef] [PubMed]
  24. Zhou, L.-N.; Gao, H.-Q.; Cai, S.-X.; Zhu, T.-J.; Gu, Q.-Q.; Li, D.-H. Two new cyclic pentapeptides from the marine-derived fungus Aspergillus versicolor. Hel. Chim. Acta 2011, 96, 1065–1070. [Google Scholar] [CrossRef]
  25. Zhuang, Y.; Teng, X.; Wang, Y.; Liu, P.; Wang, H.; Li, J.; Li, G.; Zhu, W. Cyclopeptides and polyketides from coral-associated fungus, Aspergillus versicolor LCJ-5-4. Tetrahedron 2011, 67, 7085–7089. [Google Scholar] [CrossRef]
  26. Liu, S.; Shen, Y. A new cyclic peptide from the marine fungal strain Aspergillus sp. AF119. Chem. Nat. Compd (Engl. Transl.) 2011, 47, 786–788. [Google Scholar] [CrossRef]
  27. Bao, J.; Zhang, X.-Y.; Xu, X.-Y.; He, F.; Nong, X.-H.; Qi, S.-H. New cyclic tripeptides and asteltoxins from gorgonian-derived fungus Aspergillus sp. SCSGAF0076. Tetrahedron 2013, 69, 2113–2117. [Google Scholar] [CrossRef]
  28. Ebada, S.S.; Fischer, T.; Hamacher, A.; Du, F.-Y.; Roth, Y.O.; Kassack, M.U.; Wang, B.-G.; Roth, E.H. Psychrophilin E, a new cyclopeptide, from co-fermentation of two marine alga-derived fungi of the genus Aspergillus. Nat. Prod. Res. 2014, 28, 776–781. [Google Scholar] [CrossRef] [PubMed]
  29. Prompanya, C.; Dethoup, T.; Bessa, L.J.; Pinto, M.M.M.; Gales, L.; Costa, P.M.; Silva, A.M.S.; Kijjoa, A. New isocoumarin derivatives and meroterpenoids from the marine sponge-associated fungus Aspergillus similanensis sp. nov. KUFA 0013. Mar. Drugs 2014, 12, 5160–5173. [Google Scholar] [CrossRef] [PubMed]
  30. Kai, K.; Yoshikawa, H.; Kuo, Y.-H.; Akiyama, K.; Hayashi, H. Determination of absolute structure of cyclic peptides, PF1171A and PF1171C, from unidentified ascomycete OK-128. Biosci. Biotechnol. Biochem. 2010, 74, 1309–1311. [Google Scholar] [CrossRef] [PubMed]
  31. Masuda, Y.; Tanaka, R.; Kai, K.; Ganesan, A.; Doi, T. Total synthesis and biological evaluation of PF1171A, C, F and G, cyclic hxapeptides with insecticidal activity. J. Org. Chem. 2014, 79, 7844–7853. [Google Scholar] [CrossRef] [PubMed]
  32. Huang, H.; She, Z.; Lin, Y.; Vrijmoed, L.L.P.; Lin, W. Cyclic peptide from an endophytic fungus obtained from a mangrove leaf (Kandelia candel). J. Nat. Prod. 2007, 70, 1696–1699. [Google Scholar] [CrossRef] [PubMed]
  33. Berthod, A.; Liu, Y.; Bagwill, C.; Armstrong, D.W. Facile liquid chromatographic enantioresolution of native amino acids and peptides using a teicoplanin chiral stationary phase. J. Chromatogr. A 1996, 731, 123–127. [Google Scholar] [CrossRef] [PubMed]
  34. Péter, A.; Török, G.; Armstrong, D.W. High-performance liquid chromatographic separation of enantiomers of unusual amino acids on a teicoplanin chiral stationary phase. J. Chromatogr. A 1998, 793, 283–296. [Google Scholar] [CrossRef] [PubMed]
  35. Kučerová, G.; Vozka, J.; Kalílová, K.; Geryx, R.; Plecita, D.; Pajponova, T.; Tesařová, E. Enantioselective separation of unusual amino acids by high performance liquid chromatography. Sep. Purif. Technol. 2013, 119, 123–128. [Google Scholar] [CrossRef]
  36. Wu, Z.-C.; Li, S.; Nam, S.-J.; Liu, Z.; Zhang, C. Nocardiamides A and B, two cyclohexapeptides from the marine-derived Actinomycete nocardiopsis sp. CNX037. J. Nat. Prod. 2013, 76, 694–701. [Google Scholar] [CrossRef] [PubMed]
  37. Cai, G.; Napolitano, J. G.; McAlpine, J. B.; Wang, Y.; Jaki, B.U.; Suh, J.-V.; Yang, S.H.; Lee, I.-A.; Franzblau, S.G.; Pauli, G.F.; Cho, S. Hytramycins V and I, Anti-mycobacterium tuberculosis hexapeptides from a Streptomyces hygroscopicus strain. J. Nat. Prod. 2013, 76, 2009–2018. [Google Scholar] [CrossRef] [PubMed]
  38. Song, Y.; Li, Q.; Liu, X.; Chen, Y.; Zhang, Y.; Sun, A.; Zhang, W.; Zhang, J.; Ju, J. Cyclic hexapeptides from the deep South China Sea-derived Streptomyces scopuliridis SCSIO ZJ46 active against pathogenic Gram-positive bacteria. J. Nat. Prod. 2014, 77, 1937–1941. [Google Scholar] [CrossRef]
  39. Kijjoa, A.; Wattanadilok, R.; Campos, N.; Maria São José Nascimento, M.S.J.; Pinto, M.; Herz, W. Anticancer activity evaluation of kuanoniamines A and C isolated from the marine sponge Oceanapia sagittaria, collected from the Gulf of Thailand. Mar. Drugs 2007, 5, 6–22. [Google Scholar] [CrossRef] [PubMed]
  40. Gomes, N.M.; Bessa, L.J.; Buttachon, S.; Costa, P.M.; Buaruang, J.; Tida Dethoup, T.; Silva, A.M.S.; Kijjoa, A. Antibacterial and antibiofilm activities of tryptoquivalines and meroditerpenes isolated from the marine-derived fungi Neosartorya paulistensis, N. laciniosa, N. tsunodae, and the soil fungi N. fischeri and N. siamensis. Mar. Drugs 2014, 12, 822–839. [Google Scholar] [CrossRef] [PubMed]

Share and Cite

MDPI and ACS Style

Prompanya, C.; Fernandes, C.; Cravo, S.; Pinto, M.M.M.; Dethoup, T.; Silva, A.M.S.; Kijjoa, A. A New Cyclic Hexapeptide and a New Isocoumarin Derivative from the Marine Sponge-Associated Fungus Aspergillus similanensis KUFA 0013. Mar. Drugs 2015, 13, 1432-1450. https://doi.org/10.3390/md13031432

AMA Style

Prompanya C, Fernandes C, Cravo S, Pinto MMM, Dethoup T, Silva AMS, Kijjoa A. A New Cyclic Hexapeptide and a New Isocoumarin Derivative from the Marine Sponge-Associated Fungus Aspergillus similanensis KUFA 0013. Marine Drugs. 2015; 13(3):1432-1450. https://doi.org/10.3390/md13031432

Chicago/Turabian Style

Prompanya, Chadaporn, Carla Fernandes, Sara Cravo, Madalena M. M. Pinto, Tida Dethoup, Artur M. S. Silva, and Anake Kijjoa. 2015. "A New Cyclic Hexapeptide and a New Isocoumarin Derivative from the Marine Sponge-Associated Fungus Aspergillus similanensis KUFA 0013" Marine Drugs 13, no. 3: 1432-1450. https://doi.org/10.3390/md13031432

Article Metrics

Back to TopTop