Next Article in Journal
Enhanced Photosynthetic Pigment Production Using a Scaled-Up Continuously Circulated Bioreactor
Next Article in Special Issue
New Cyclopiane Diterpenes and Polyketide Derivatives from Marine Sediment-Derived Fungus Penicillium antarcticum KMM 4670 and Their Biological Activities
Previous Article in Journal
Metabolic Profiling, Antiviral Activity and the Microbiome of Some Mauritian Soft Corals
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Lecanicilliums A–F, Thiodiketopiperazine-Class Alkaloids from a Mangrove Sediment-Derived Fungus Lecanicillium kalimantanense

1
CAS Key Laboratory of Tropical Marine Bio-Resources and Ecology, Guangdong Key Laboratory of Marine Materia Medica, South China Sea Institute of Oceanology, Chinese Academy of Sciences, Guangzhou 510301, China
2
University of Chinese Academy of Sciences, Beijing 100049, China
3
The Marine Biomedical Research Institute, Guangdong Medical University, Zhanjiang 524023, China
*
Author to whom correspondence should be addressed.
Mar. Drugs 2023, 21(11), 575; https://doi.org/10.3390/md21110575
Submission received: 2 October 2023 / Revised: 21 October 2023 / Accepted: 24 October 2023 / Published: 31 October 2023
(This article belongs to the Special Issue Natural Products Isolated from Marine Sediment)

Abstract

:
Six new thiodiketopiperazine-class alkaloids lecanicilliums A–F were isolated from the mangrove sediment-derived fungus Lecanicillium kalimantanense SCSIO41702, together with thirteen known analogues. Their structures were determined by spectroscopic analysis. The absolute configurations were determined by quantum chemical calculations. Electronic circular dichroism (ECD) spectra and the structure of Lecanicillium C were further confirmed by a single-crystal X-ray diffraction analysis. Lecanicillium A contained an unprecedented 6/5/6/5/7/6 cyclic system with a spirocyclic center at C-2′. Biologically, lecanicillium E, emethacin B, and versicolor A displayed significant cytotoxicity against human lung adenocarcinoma cell line H1975, with IC50 values of 7.2~16.9 μM, and lecanicillium E also showed antibacterial activity against four pathogens with MIC values of 10~40 μg/mL. Their structure–activity relationship is also discussed.

1. Introduction

Natural products containing sulfur exhibit significant structural diversities and diverse biological activities [1,2]. Among them, thiodiketopiperazine, as an interesting subgroup, is an important class of secondary metabolites of fungi, especially species of Aspergillus fumigatus, A. terreus, A. flavus, A. niger, Trichoderma virens, T. viride, and Penicillium terlikowskii [3]. The structures of thiodiketopiperazines are diverse [4,5], of which the spirocyclic thiodiketopiperazines are rare in nature [5,6]. Many thiodiketopiperazines have been reported to display a wide range of biological properties, including antiproliferative, cytotoxic, antibacterial, antifungal, antiviral, and anti-angiogenic activities [7,8]. For examples, brocazine G displayed potent antibacterial activity against Staphylococcus aureus with a MIC value of 0.25 μg/mL and strong cytotoxicity against human ovarian cancer cells (IC50 = 0.66 μM) [8]. Penicisulfuranols A–C showed cytotoxicity towards human cervical carcinoma cell lines and human promyelocytic leukemia cells with IC50 values of 0.1–3.9 μM [9]. Penicibrocazine C exhibited antibacterial activity against Micrococcus luteus with MIC value of 0.25 μg/mL, which was stronger than the positive control, chloromycetin (MIC = 2.0 μg/mL) [10]. Numerous structurally unique and potential variety activities of the thiodiketopiperazines have drawn much attention from synthetic chemists and pharmacologists.
The fungal genus Lecanicillium (formerly Verticillium), including L. fusisporum, L. psalliotae, and L. lecanii, are pathogenic fungi that are widely used as biological pesticides [11,12,13]. The secondary metabolisms of these fungi are rarely reported [11]. Only a few compounds belonging to indolosesquiterpenoids, phenopicolinic acid derivatives, tetracyclic diterpenoids, pregnanes, and cyclic lipodepsipeptides have been isolated from the genus of Lecanicillium up to now [14,15,16]. The fungus L. kalimantanense SCSIO41702 was isolated from a mangrove sediment sample collected from Hainan province. There was no report about the secondary metabolisms of L. kalimantanense. In order to explore novel, natural compounds from fungi, we investigated the secondary metabolisms of L. kalimantanense SCSIO41702, which led to the isolation of six new thiodiketopiperazine-class alkaloids, lecanicilliums A-F (16), together with thirteen known analogues (Figure 1): emethacin B (7) [17], bisdethiobis(methylsulfanyl)acetylaranotin (8) [8], bisdethiobis(methylsulfanyl)deacetylaranotin (9) [8], bisdethiobis(methylsulfanyl)-aranotin (10) [8], 8-acetyl-bisdethiobis(methylsulfanyl)apoaranotin (11) [18], bisdethiobis(methylsulfanyl)acetylapoaranotin (12) [8], bisdethiobis(methylsulfanyl)apoaranotin (13) [19], bisdethiobis(methylsulfanyl)deacetylapoaranotin (14) [8], haematocin (15) [8], versicolor A (16) [20], 12,12a-anhydro-desacetyl-bis-dethio-7a, 14a-di-(methylmercapto)-apoaranotin (17) [21], emestrin H (18) [22], and citriperazine B (19) [23]. These compounds were evaluated for their cytotoxicity, toxicity against brine shrimps, and antibacterial activity. Herein, we report the isolation and structural elucidation as well as the biological activities of these compounds.

2. Results and Discussion

Lecanicillium A (1) was obtained as a white powder with the molecular formula C18H16N2O5S as determined by HRESIMS ([M + H]+ m/z 373.0851) and NMR data. The 1H NMR spectrum (Table 1) displayed the presence of two active hydrogens at δH 10.06 (1H, s), 5.31 (1H, s), four aromatic hydrogens at δH 7.26 (1H, d, J = 7.5 Hz), 7.14 (1H, t, J = 8.0 Hz), 6.92 (1H, t, J = 7.5 Hz), 6.81 (1H, d, J = 8.0 Hz), two olefinic methines at δH 6.00 (1H, d, J = 6.5 Hz), 4.82 (1H, ddd, J = 7.8, 6.5, 1.2 Hz), four tertiary methine groups at δH 6.61 (1H, d, J = 6.5 Hz), 4.86 (1H, ddd, J = 7.8, 6.5, 1.2 Hz), 3.87 (1H, d, J = 6.5 Hz), 3.06 (1H, t, J = 6.5 Hz), and two methylenes at δH 3.94 (1H, d, J = 16.4 Hz), 3.26 (1H, d, J = 16.4 Hz), 2.71 (1H, dd, J = 12.7, 6.5 Hz), 2.21 (1H, d, J = 12.7 Hz). The 13C NMR spectrum (Table 2) displayed 18 carbon signals including two methylenes, ten methines (six aromatic/olefinic and three oxygenated or heteroatomic), and six nonprotonated carbons (two diagnostic amide carbonyl, two olenfinic and two heteroatomic). These NMR data of 1 (Table 1 and Table 2) showed similarity to those of spirobrocazines A–C [5], which suggested that 1 was a thiodiketopiperazine alkaloid. Detailed analysis of HSQC, HMBC and COSY spectra (Figure 2) proved that 1 contained the same structural parts of rings A–C as those of spirobrocazines A–C. However, the COSY spectrum (Figure 2) showing correlations from H-4 (δH 3.06, 1H, t, J = 6.5 Hz) to Hb-3 [δH 2.71 (1H, dd, J = 12.7, 6.5 Hz)], H-5 (δH 3.87, 1H, d, J = 6.5 Hz) and H-9 (δH 6.61, 1H, d, J = 6.5 Hz), from H-5 to H-6 (δH 4.86, 1H, ddd, J = 7.8, 6.5, 1.2 Hz), and from H-7 (δH 4.82, 1H, ddd, J = 7.8, 6.5, 1.2 Hz) to H-6 and H-8 (δH 6.00, 1H, d, J = 6.5 Hz), and the HMBC spectrum (Figure 2) showing correlations from H-4 to C-2 (δC 70.8), C-3 (δC 48.2), C-6 (δC 52.5), from OH-5 to C-5 (δC 69.9), from H-6 to C-2, C-4 (δC 46.3), C-5, C-7, C-8 (δC 144.9), from H-8 (δH 6.00, 1H, d, J = 6.5 Hz) to C-6, C-7 (δC 100.6), C-9 (δC 92.9), from H-9 to C-2, C-5, C-8, and from H-3 to C-2, C-4, C-9, suggested the presence and assignment of D/E rings in 1 as shown. Combining the molecular formula and the chemical shifts of C-2 and C-6/H-6, it was reasonable to infer a sulfide bond between C-2 and C-6 to form a sulfide six-membered-ring. Thus, the planar structure of 1 was determined as shown.
The relative configuration of 1 was elucidated by analysis of the NOESY spectrum (Figure S6). The NOE correlations of H-4 with OH-5 and H-9, H-9 with H-3 (δH 2.71) and H-4, and H-6 with H-4 (Figure 3), indicated their cofacial β-orientations. The NOE correlation of H-5 with Ha-3 (δH 2.21) indicated their cofacial α-orientations. The NOE correlations from H-5 to H2-3′ revealed their cofacial orientations and the CH2-3′ group at the axial position. The coupling constant of JH-7-H-8 = 6.5 Hz indicated that the double bond was a Z-configuration. The absolute configuration of 1 was further determined by electronic circular dichroism (ECD) calculations. The calculated ECD spectrum of (2R, 4S, 5S, 6S, 9R, 2′S)-1 showed similarity to the experimental ECD spectrum of 1 (Figure 4), which confirmed that the absolute configuration of 1 was 2R, 4S, 5S, 6S, 9R, 2′S.
A possible biosynthetic pathway of 1 is shown in Figure S66 [5,8]. The biosynthetic pathway of 1 likely starts with the cyclodipeptide (I) composed of two phenylalanines followed by oxidation to afford the key intermediate II, which could be transferred to III by dehydration and oxidation. Then cyclization and oxidation of III via intermediate IV could produce V. Finally, V could be translated into compound 1 by further oxidation, cyclization, and sulfurization.
Lecanicillium B (2) was obtained as a white powder with the molecular formula C19H18N2O5S as determined by HRESIMS ([M + H]+ m/z 387.1009) and NMR data. The 1H and 13C NMR spectra of 2 (Table 1 and Table 2) showed similarity to those of spirobrocazine A [5]. The obvious difference between them was the chemical shift changes of C-4, C-5, C-6, C-7, C-8, and C-9 (δC 134.1 (C), 120.3 (CH), 124.0 (CH), 131.6 (CH), 75.1 (CH), and 70.4 (CH) in spirobrocazine A [5], and correspondingly δC 109.1 (C), 136.7 (CH), 137.9 (CH), 111.0 (CH), 71.2 (CH) and 64.1 (CH) in 2, respectively). The COSY spectrum (Figure 2) showing the sequential correlations of H-6/H-7/H-8/H-9, and the HMBC spectrum (Figure 2) showing correlations from H-5 (δH 6.69, 1H, t, J = 2.4 Hz) to C-3 (δC 40.2), C-4 (δC 109.1), C-6 (δC 137.9), C-9 (δC 64.1), from H-6 (δH 6.29, 1H, dd, J = 8.2, 2.4 Hz) to C-5 (δC 136.7), C-7 (δC 111.0), C-8 (δC 71.2), from H-8 (δH 4.46, 1H, ddt, J = 7.7, 6.7, 2.1 Hz) to C-7, C-9, from OH-8 (δH 5.28, d, J = 6.7 Hz) to C-8, and from H-3 [δH 2.91 (1H, dt, J = 15.3, 1.2 Hz), 3.14 (1H, dt, J = 15.3, 2.2 Hz)] to C-2 (δC 69.2), C-4, C-5, C-9, suggested that the E ring in 2 was a seven-membered 4,5-dihydrooxepine ring, as shown, instead of a six-membered ring. The coupling constant of JH-6-H-7 = 8.2 Hz indicated that the double bond was a Z-configuration. In addition, the NOESY sepctrum of 2 (Figure S15) also showed great similarity to that of spirobrocazine A. The NOE correlation of OH-8 with H-9 indicated their cofacial β-orientations, while the NOE correlations of H-8 with H3-10 and Ha-3′ (δH 4.01) indicated their cofacial α-orientations and the CH2-3′ group at the axial position (Figure 3). Furthermore, the calculated ECD spectrum of (2R, 8S, 9S, 2′S)-2 showed similarity to the experimental ECD spectrum of 2 (Figure 4), indicating the absolute configuration of 2, as shown.
Lecanicillium C (3) was isolated as a white powder with the molecular formula C24H26N2O9S2 as determined by HRESIMS ([M + Na] + m/z 573.0969) and NMR. The 1H and 13C NMR spectra of 3 (Table 1 and Table 2) showed similarity to those of bisdethiobis(methylsulfanyl)acetylapoaranotin [8]. The obvious difference between them was the additional presence of three tertiary methine signals (δH 3.83 (1H, d, J = 2.7 Hz), 3.67 (1H, dd, J = 4.3, 2.7 Hz), 3.34 (1H, dd, J = 4.3, 1.2 Hz); δC 54.1, 53.6, 47.9) and one oxygenated nonprotonated carbon (δC 60.9), and the disappearance of the signals for two double bonds in 3. Detailed analysis of HSQC, HMBC and COSY spectra (Figure 2, Figures S21–S23) proved that 3 contained the same structural parts of rings B–E as those of bisdethiobis(methylsulfanyl)acetylapoaranotin. Furthermore, the COSY spectrum (Figure 2) showed correlations from H-5′ (δH 3.83, 1H, d, J = 2.7 Hz) to H-6′ (δH 3.67, 1H, dd, J = 4.3, 2.7 Hz), from H-6′ to H-7′ (δH 3.34, 1H, dd, J = 4.3, 1.2 Hz), and from H-8′ (δH 5.79, 1H, dd, J = 9.1, 1.2 Hz) to H-7′ and H-9′ (δH 4.34, 1H, d, J = 9.1 Hz), and the HMBC spectrum (Figure 2) showed correlations from H-5′ to C-4′ (δC 60.9) and C-6′ (δC 47.9), from H-6′ to C-4′, C-5′ (δC 53.6), from H-7′ to C-8′ (δC 72.6), C-9′ (δC 57.5), from H-8′ to C-9′, C-10′ (δC 169.4), from H-9′ to C-1 (δC 163.5), C-4′, C-5′, C-8′, from H-3′ [δH 2.85 (1H, d, J = 14.4 Hz), 2.43 (1H, d, J = 14.4 Hz)] to C-1′ (δC 163.6), C-2′ (δC 71.7), C-4′, C-5′, and from H3-11′ to C-10′. Combining with the molecular formula, these above data suggested that the A ring was a six-membered ring with an acetyl group attached at C-8′ and two oxygenated three-membered rings between C-4′ and C-5′, and between C-6′ and C-7′, respectively. The small coupling constants of JH-5′-H-6′ (2.7 Hz), JH-6′-H-7′ (4.3 Hz), and JH-7′-H-8′ (1.2 Hz) suggested a cis-diaxial relationship between H-5′ and H-6′, between H-6′ and H-7′, and between H-7′ and H-8′, respectively. NOE correlations of H-5′ with H-6′, H-6′ with H-7′ and H-7′ with H-8′ also indicated H-5′, H-6′, H-7′, and H-8′ on the same face. The large coupling constants of JH-8-H-9 (8.1 Hz) and JH-8′-H-9′ (9.1 Hz) suggested a trans-diaxial relationship between H-8 and H-9, and between H-8′ and H-9′, respectively. Furthermore, the NOE correlations of H-3 (δH 3.25) with H-8 and H3-12 indicated the assignment of α-orientation for H-8 and H3-12, NOE correlations of H-3′ (δH 2.43) with H-5′ and H3-12′ indicated α-orientation for H3-12′ and H-5′, and NOE correlations of H-9 with H-9′ indicated β-orientation for H-9 and H-9′ (Figure 3). Thus, the structure of 3 was inferred as shown. The absolute configuration of 3 was determined as 2R, 8S, 9S, 2′R, 4′S, 5′R, 6′R, 7′R, 8′R, 9′R by a single crystal X-ray diffraction analysis using Cu Kα radiation (Figure 5), which was supported by the calculated ECD spectrum of (2R, 8S, 9S, 2′R, 4′S, 5′R, 6′R, 7′R, 8′R, 9′R)-3 showing agreement with the experimental ECD spectrum of 3 (Figure 4).
Lecanicillium D (4) was isolated as a white solid with the molecular formula C24H26N2O9S2 as determined by HRESIMS ([M + NH4]+ m/z 568.1408). The 1H and 13C NMR spectra of 4 (Table 1 and Table 2) showed similarity to those of 3. The obvious difference between them was the presence of an α, β-unsaturated ketone group (δH 6.98 (1H, d, J = 10.3 Hz), 6.09 (1H, d, J = 10.3 Hz); δC 191.5, 150.2, 125.8) and the disappearance of three tertiary methine signals in 4. The HMBC spectrum (Figure 2) showing correlations from H-5′ (δH 6.98, 1H, d, J = 10.3 Hz) to C-7′ (δC 191.5) and C-9′ (δC 69.1), from H-6′ (δH 6.09, 1H, d, J = 10.3 Hz) to C-8′ (δC 74.8) and C-4′ (δC 75.5), from H-8′ (δH 5.73, 1H, d, J = 11.2 Hz) to C-7′ and C-10′, from H-3′ [δH 2.94 (1H, d, J = 15.4 Hz), 3.16 (1H, dd, J = 15.4, 2.1 Hz)] to C-4′ and C-5′, and from H3-11′ to C-10′, and the COSY spectrum (Figure 2) showing correlations from H-5′ to H-6′, and from H-8′ to H-9′, revealed the existence of an α, β-unsaturated ketone instead of two oxygenated three-membered rings in 4. The NOE correlation of H-9 with H3-11 suggested the assignment of α-oriention for H-8 and β-oriention for H-9, and the NOE correlations of H-3 (δH 2.92) with H-9, and H-3 (δH 3.20) with H3-12 indicated H3-12 was α-oriented. The 11.2 Hz coupling constant between H-8′ and H-9′ and the NOE correlation of H-9′ with H3-11′ suggested the assignment of α-oriention for H-8′ and β-oriention for H-9′. The NOE correlation of OH-4′ with H-8′ indicated their cofacial α-orientations. The NOE correlations of H-3′ (δH 2.94) with H-8′ and H3-12′ indicated H3-12′ was also α-oriented. The NOE correlations of H-8 with OH-4′ indicated that H-8 and OH-4′were α-orientated. The absolute configuration of 4 was further determined by ECD calculation. The great similarity between the experimental ECD spectrum of 4 and the calculated ECD spectrum of (2R, 8S, 9S, 2′R, 4′S, 8′R, 9′R)-4 (Figure 4) indicated the absolute configuration of 4 as shown.
Lecanicillium E (5) was obtained as a white powder with the molecular formula C24H26N2O8S3 as determined by HRESIMS ([M + NH4]+ m/z 584.1189). The 1H and 13C NMR spectra of 5 (Table 1 and Table 2) showed great similarity to those of bisdethiobis(methylthio)acetylaranotin [8]. The only obvious difference between them was the down-field shifts of H3-12′ (δH 2.47, 1H, s), C-12′ (δC 24.1) and C-2′ (δC 74.9) in 5. Combined with the molecular formula, these data indicated a disulfide methyl attached at C-2′ instead of a sulfide methyl in 5. Combined with the coupling constants of JH-8′-H-9′ and JH-8-H-9 (both 8.2 Hz), the NOE correlations of H-8 with H3-12, H-9 with H3-11, H-8′ with H3-12′, and H-9′ with H3-11′ indicated that H-8, H-8′, H3-12′, H3-12 were α-oriented, and H-9, H-9′, H3-11′, H3-11 were β-oriented (Figure S65). The absolute configuration of 5 was determined as shown by comparing the experimental ECD spectrum of 5 and the calculated ECD spectrum of (2R, 8S, 9S, 2′R, 8′S, 9′S)-5 (Figure 4), which was further supported by the great similarity of the ECD spectra of 5 with bisdethiobis(methylthio)acetylaranotin [24].
Lecanicillium F (6) was isolated as a white solid with the molecular formula C19H20N2O2S as determined by HRESIMS ([M + H]+ m/z 341.1316). The 1H and 13C NMR spectra of 6 (Table 1 and Table 2) exhibited great similarity to those of emethacin B [17]. The only obvious difference between them was the disappearance of a sulfide methyl and the additional presence of a methine (δC 54.8, δH 3.26) instead of a nonprotonated carbon in 6. The COSY correlations of δH 3.26 with H-4 and NH-1′ (δH 8.23), and the HMBC correlations from δH 3.26 to C-2 (δC 166.2), C-4 (δC 37.2), C-5 (δC 135.9), from NH-1 (δH 8.66) to C-2′ (δC 164.8), C-3′ (δC 68.2), C-4′ (δC 44.4), and from NH-1′ to C-2, C-3′ (Figure S63), suggested the methine was C-3 (δC 54.8, δH 3.26) in 6. The NOE correlations of H-3 with H3-11′, and H-4 (δH 2.70) with H-4′ (δH 3.28) (Figure S65) indicated the cofacial orientations for H-3 and H3-11′, and H-4 (δH 2.70) and H-4′ (δH 3.28), respectively. The similarity between the experimental ECD spectrum of 6 and the calculated ECD spectrum of (3S,3′R)-6 (Figure 4) indicated the absolute configuration of 6 as shown.
Compounds 219 were tested for their cytotoxicity towards human lung adenocarcinoma cell line H1975 and human hepatocellular carcinoma cell line HepG-2, and toxicity towards brine shrimps. The cytotoxicity results (Table 3) showed that 5, 7, and 16 displayed significant cytotoxicity against H1975, with IC50 values of 7.2~16.9 μM, 4, 11, 17, and 18 displayed mild cytotoxicity against H1975, with IC50 values of 35.2~71.5 μM; only 18 had mild cytotoxicity against HepG-2, with an IC50 value of 41.2 μM. The results indicated that the disulfide bond unit in 5 was an active group for its cytotoxicity, which is consistent with the conclusions reported in the literatures [25,26]. A comparison of the structures and cytotoxicities of 6, 7, and 19 suggested that lacking thiomethyl and benzene could significantly decrease the cytotoxicity of this type of alkaloids. In addition, a comparison of the structures and cytotoxicity of 2, 3, 4, 11, and 16 suggests the skeleton of A/B/C ring fragment also could significantly affect the cytotoxicity of this type of alkaloids. Furthermore, brine shrimp lethality assays (Table 3) showed that only 4 and 15 exhibited medium toxicity, with TC50 values of 40–50 μM, towards brine shrimps.
Compounds 119 were also evaluated for their antibacterial activity towards eight pathogens: Bacillus subtilis, Micrococcus luteus, Escherichia coli, Staphylococcus aureus, S. aureus MRSA, Streptococcus agalactiae, S. iniae, and Pseudomonas aeruginosa. Antibacterial assays (Table 3) exhibited that only 5 had significant antibacterial activity against B. subtilis, M. luteus, S. agalactiae, and S. iniae, with MIC values of 10~40 μg/mL. Other compounds showed mild or no obvious antibacterial activity. The results indicated that a disulfide bond unit at C-2′ was crucial for the antibacterial activity of this type of alkaloids.

3. Experimental Section

3.1. General Experimental Procedure

UV spectra were measured using a UV-2600 spectrophotometer (Shimadzu). IR spectra were obtained on an IR Affinity-1 Fourier transform infrared spectrophotometer (Shimadzu, Kyoto, Japan). ECD spectra were acquired on a Chirascan circular dichroism spectrometer (Applied Photophysics Ltd., Graz, Austria). Optical rotations were recorded using a MCP 500 polarimeter (Anton Paar). Melting points were recorded with a digital display microscopic melting point instrument (SGW X-5). NMR data were acquired with a Bruker AVANCE III HD 700 MHz NMR spectrometer (Bruker) with TMS as reference. HRESIMS spectroscopic data were obtained on a MaXis quadrupole-time-of-flight mass spectrometer (Bruker, Karlsruhe, Germany). Preparative reversed-phase HPLC was performed on a Shimadzu LC-20A preparative liquid chromatography system with a YMC-Pack ODS column (250 × 20 mm, S-5 μm, 12 nm). Sephadex LH-20 (GE Healthcare) was used for the chromatographic column (CC). RP-MPLC (reversed-phase-medium pressure preparative liquid chromatography) was carried out using the CHEETAH MP200 system (Agela Technologies, Tianjin, China) and Claricep Flash columns filled with ODS (40-63 μm, YMC). Silica gel (200–300 mesh) for CC and GF254 for TLC were purchased from Yantai Jiangyou Silica Gel Development Co., Ltd. Sea salts were commercially obtained from Guangzhou Hai Li Aquarium Technology Co., Ltd., Guangzhou, China.

3.2. Fungal Material

The fungus Lecanicillium kalimantanense was isolated from a mangrove sediment sample collected in the Bailu park, Sanya city, Hainan province. The strain was identified as Lecanicillium kalimantanense by internally transcribed spacer (ITS) region sequence data of the rDNA and given the Genbank accession number KM264285. The fungus L. kalimantanense was deposited in the RNAM Center, South China Sea Institute of Oceanology, Chinese Academy of Science.

3.3. Fermentation and Extraction

The spores of the fungus L. kalimantanense were added to 5 × 500 mL Erlenmeyer flasks, each containing 200 mL potato dextrose (PD) medium, and fermented for 3 days at 28 °C. Then 3 mL of spore suspension was transferred into 267 × 1 L Erlenmeyer flasks, each containing 300 mL culture media (glucose 1%, D-mannitol 2%, maltose 2%, corn meal 0.05%, monosodium glutamate 1%, KH2PO4 0.05%, MgSO4·7H2O 0.03%, yeast extract 0.3%, sea salt 3%). Static fermentation was performed for 26 days at 28 °C. After fermentation, the broth and mycelia were separated with gauze. The broth was extracted with EtOAc to obtain crude extract (28.9 g). The mycelia were extracted three times with acetone, and further extracted three times with EtOAc to yield a crude extract (58.3 g). Then the two crude extracts (28.9 g and 58.3 g) were combined for further isolation.

3.4. Isolation and Purification

The combined crude extract (87.2 g) was fractionated on a normal-phase column using a stepped gradient elution with CH2Cl2/MeOH (v/v, 100:0, 98:2, 95:5, 90:10, 85:15, 80:20, 70:30, 50:50, 0:100) to obtain eight fractions (Fr.1–Fr.8). Fr.3 (10.9 g) was separated with Sephadex LH-20 eluting with CH2Cl2/MeOH (1:1) to obtain eleven subfractions (Fr.3.1–Fr.3.11). Fr.3.3 was further purified by HPLC eluting with MeOH/H2O (v/v 7:3, 3 mL/min) to give 12 (9.0 mg, tR = 13.0 min). Fr.3.4 was further purified by HPLC eluting with MeOH/H2O (v/v 6:4, 3 mL/min) to give 8 (3.5 mg, tR = 29.0 min), 10 (11.0 mg, tR = 22.0 min), and 13 (9.0 mg, tR = 19.0 min). Fr.3.5 was further purified by HPLC eluting with MeOH/H2O (v/v 55:45, 3 mL/min) to give 5 (1.7 mg, tR = 43.0 min), 7 (1.1 mg, tR = 35.5 min), and 15 (2.6 mg, tR = 44.0 min). Fr.3.9 was further purified by HPLC eluting with MeOH/H2O (v/v 7:3, 3 mL/min) to give 9 (20.0 mg, tR = 10.0 min). Then Fr.3.6–Fr.3.8 were combined and further separated by ODS column eluting with MeOH/H2O (v/v 10:90–100:0) to obtain subfractions (Fr.3.6.1–Fr.3.6.22). Fr.3.6.6 was purified by HPLC eluting with MeOH/H2O (v/v 42.5:57.5, 2.5 mL/min) to give 6 (2.3 mg, tR = 32.0 min). Fr.3.6.8 was purified by HPLC eluting with MeOH/H2O (v/v 55:45, 2.5 mL/min) to give 3 (1.6 mg, tR = 26.0 min), 2 (1.5 mg, tR = 30.0 min), and 4 (2.6 mg, tR = 24.0 min). Fr.3.2.10 was purified by HPLC eluting with MeOH/H2O (v/v 58.5:41.5, 3 mL/min) to give 14 (3.0 mg, tR = 33.0 min). Fr.3.6.12 was purified by HPLC eluting with MeOH/H2O (v/v 59:41, 3 mL/min) to give 11 (6.5 mg, tR = 40.0 min) and 18 (3.4 mg, tR = 45.0 min). Fr.3.6.15 was further purified by HPLC eluting with MeOH/H2O (v/v 6:4, 2.5 mL/min) to give 16 (0.8 mg, tR = 75.0 min) and 17 (1.4 mg, tR = 68.0 min). Fr.4 (8.6 g) was separated with Sephadex LH-20 eluting with CH2Cl2/MeOH (1:1) to obtain eleven subfractions (Fr.4.1–Fr.4.3). Fr.4.2 was further separated by ODS column eluting with MeOH/H2O (v/v 10:90–100:0) to obtain subfractions (Fr.4.2.1–Fr.4.2.33). Fr.4.2.8 was purified by HPLC eluting with MeOH/H2O (v/v 9:11, 3 mL/min) to give 19 (2.0 mg, tR = 20.0 min). Fr.4.2.10 was purified by HPLC eluting with MeOH/H2O (v/v 9:11, 3 mL/min) to give 1 (1.5 mg, tR = 26.0 min).
Lecanicillium A (1). White powder; [α ] D 25 −86 (c 0.10, MeOH); UV (MeOH) λmax (log ε) 212 (4.30), 277 (3.70), 283 (3.60) nm; ECD (0.33 mM, MeOH) λmax (∆ε) 200 (−9.00), 223 (16.86), 258 (−6.02) nm; IR (film) νmax 3746, 3433, 3267, 2922, 2855, 2365, 2322, 2257, 1676, 1458, 1389, 1236, 1190, 1109, 1084, 1022, 995, 862, 760, 644, 596 cm−1; 1H and 13C NMR data, Table 1 and Table 2; HR-ESIMS m/z 373.0851 [M + H]+ (calcd for C18H17N2O5S, 373.0853).
Lecanicillium B (2). White powder; [α ] D 25 −168 (c 0.10, MeOH); UV (MeOH) λmax (log ε) 210 (4.30), 276 (3.53), 283 (3.47) nm; ECD (0.32 mM, MeOH) λmax (∆ε) 200 (21.64), 201 (34.30), 229 (−34.46) nm; IR (film) νmax 3341, 2922, 2841, 1676, 1649, 1545, 1514, 1460, 1410, 1113, 1018, 671, 598 cm−1; 1H and 13C NMR data, Table 1 and Table 2; HR-ESIMS m/z 387.1009 [M + H]+ (calcd for C19H19N2O5S, 387.1009).
Lecanicillium C (3). colorless crystals; mp 172–174 °C; [α ] D 25 −213 (c 0.10, CH3CN); UV (CH3CN) λmax (log ε) 206 (4.50) nm; ECD (0.45 mM, CH3CN) λmax (∆ε) 200 (−24.56), 201 (−17.74), 203 (−25.16), 214 (−11.98), 227 (−31.84), 247 (−3.68), 257 (−6.35) nm; IR (film) νmax 3861, 3744, 3618, 2922, 2853, 1738, 1674, 1514, 1377, 1236, 1194, 1138, 1034, 972, 731, 648, 602 cm−1; 1H and 13C NMR data, Table 1 and Table 2; HR-ESIMS m/z 573.0969 [M + Na]+ (calcd for C24H26N2NaO9S2, 573.0972).
Lecanicillium D (4). White powder; [α ] D 25 −206 (c 0.10, MeOH); UV (MeOH) λmax (log ε) 207 (4.35) nm; ECD (0.45 mM, MeOH) λmax (∆ε) 200 (3.58), 226 (−33.79), 247 (−8.64), 256 (−10.55) nm; IR (film) νmax 3331, 2945, 2920, 2841, 1740, 1659, 1535, 1514, 1460, 1398, 1113, 1018, 671, 598 cm−1; 1H and 13C NMR data, Table 1 and Table 2; HR-ESIMS m/z 568.1408 [M + NH4]+ (calcd for C24H30N3O9S2, 568.1418).
Lecanicillium E (5). White powder; [α ] D 25 −282 (c 0.10, MeOH); UV (MeOH) λmax (log ε) 207 (4.36) nm; ECD (0.44 mM, MeOH) λmax (∆ε) 200 (9.37), 202 (−2.30), 225 (−61.07), 247 (−17.03), 256 (−19.43) nm; IR (film) νmax 3347, 2951, 2924, 2843, 1734, 1670, 1375, 1302, 1236, 1192, 1130, 1018, 669, 598 cm−1; 1H and 13C NMR data, Table 1 and Table 2; HR-ESIMS m/z 584.1189 [M + NH4]+ (calcd for C24H30N3O8S3, 584.1190).
Lecanicillium G (6). White powder; [α ] D 25 −5 (c 0.10, MeOH); UV (MeOH) λmax (log ε) 206 (4.08) nm; ECD (0.44 mM, MeOH) λmax (∆ε) 200 (−19.33), 201 (16.40), 204 (21.49), 228 (−10.86) nm; IR (film) νmax 3354, 2943, 2833, 1670, 1653, 1558, 1541, 1506, 1472, 1456, 1418, 1115, 1020, 667, 599 cm−1; 1H and 13C NMR data, Table 1 and Table 2; HR-ESIMS m/z 341.1316 [M + H]+ (calcd for C19H21N2O2S, 341.1318).

3.5. X-ray Crystallographic Analysis of 3

Crystallographic data were collected on a Rigaku MicroMax 007 diffractometer (Rigaku Corporation, Tokyo, Japan) equipped with Cu Kα radiation and a graphite monochromator. The structure was solved by direct methods with the SHELXTL and refined by full-matrix, least-squares techniques. Crystallographic data for 3 were deposited with the Cambridge Crystallographic Data Centre as supplementary publication number CCDC 2298244.
Crystal data for 3: C24H26N2O9S2·CH3OH, FW = 582.63; colorless crystal from MeOH; crystal size = 0.15 × 0.12 × 0.1 mm3; T = 100.00 (10) K; monoclinic, space group P21 (no. 4); unit cell parameters: a = 7.14328(5) Å, b = 16.25526(10) Å, c = 11.41675(7) Å, α = 90°, β = 93.4877(6)°, γ = 90°, V = 1323.211(14) Å3, Z = 2, Dcalc = 1.462 g/cm3, F (000) = 612.0, μ (CuKα)= 2.357 mm−1; 25678 reflections measured (7.758° ≤ 2θ ≤ 148.752°), 5292 unique (Rint = 0.0323, Rsigma = 0.0216), which were used in all calculations. The final R1 was 0.0242 (I > 2σ(I)) and wR2 was 0.0641 (all data). Flack parameter = −0.001(4).

3.6. ECD Calculations

The ECD calculations for 16 were performed using the Gaussian 16 program package. The procedures were the same as described in our previous studies [27]. The Spartan 14 program (Wavefunction Inc., Tokyo, Japan) was used for calculating the molecular Merck force field (MMFF). Density functional theory (DFT) and time-dependent density functional theory (TDDFT) calculations were performed using the Gaussian 16 program package. An MMFF model was used for the conformational search, then the conformers with lower relative energies (<10 kcal/mol) were subjected to geometry optimization with the DFT method at the B3LYP/6-31g (d) level in MeOH or CH3CN. Vibrational frequency calculations were carried out at the same level to evaluate their relative thermal (ΔE) and free energies (ΔG) at 298.15 K. The geometry optimized conformers were further calculated at the M062X/def2TZVP level and the solvent (MeOH or CH3CN) effects were taken into consideration by using SMD. The optimized conformers with a Boltzmann distribution of more than 1% population were further subjected to ECD calculation, which were performed by TDDFT methodology at the PBE1PBE/6-311g(d) level. The number of excited states was 60 for 1, 2, and 6, 85 for 3 and 4, and 102 for 5. The ECD spectra were generated by the software SpecDis using a Gaussian band shape with 0.20–0.30 eV exponential half-width from dipole-length dipolar and rotational strengths. The equilibrium population of each conformer at 298.15 K was calculated from its relative free energies using Boltzmann statistics. The calculated spectra of compounds 16 were generated from the low-energy conformers according to the Boltzmann weighting of each conformer in MeOH solution for 1, 2, 4, 5, and 6 and CH3CN for 3, respectively.

3.7. Cytotoxicity

Cytotoxic activity was evaluated using the human lung adenocarcinoma cell line H1975 and human hepatocellular carcinoma cell line HepG-2, using the CCK-8 method as described in our previous study [28]. Briefly, each of the test compounds was dissolved in DMSO and further diluted to give final concentrations of 80, 40, 20, 10, 5, 2.5, and 1.25 μM, respectively. H1975 cells or HepG-2 cells (5 × 103 cells/plate) were seeded in 96-well plates and treated with compounds at the indicated concentration for 24 hours and then 10 μL CCK-8 reagent was added to each well. The plates were incubated at 37 °C for another 4 hours. Finally, the optical density was measured at a wavelength of 450 nm with a Bio-Rad microplate reader. Dose–response curves were generated, and the IC50 values were calculated from the linear portion of log dose–response curves.

3.8. Brine Shrimp Lethality Bioassays

The methods were the same as those used in our previous studies [27].

3.9. Antibacterial Assays

The micro broth dilution method [29] was used to evaluate the antibacterial activities of compounds 119 against the growth of eight common pathogens: Bacillus subtilis BS01, Micrococcus luteus, Escherichia coli ATCC 25922, Staphylococcus aureus ATCC 6538, S. aureus MRSA, Streptococcus agalactiae ATCC 13813, S. iniae, and Pseudomonas aeruginosa ATCC 9027 in 96-well polystyrene plates. Vancomycin and ciprofloxacin were used as positive controls. Briefly, wells containing 100 μL bacteria dilutions (OD600 = 0.01) in Luria–Bertani (LB) medium were supplemented with different concentrations of 119 (80, 40, 20, 10, 5, 2.5, 1.25, and 0.625 μg/mL), vancomycin, and ciprofloxacin (40, 20, 10, 5, 2.5, 1.25, 0.625, and 0.3125 μg/mL), respectively. The MICs were determined after 24 hours’ incubation at 37 °C with the tested compounds.

4. Conclusions

In summary, six new thiodiketopiperazine-class alkaloids, lecanicilliums A–F (16), together with 13 known analogues (719) were isolated from the mangrove sediment-derived fungus L. kalimantanense SCSIO41702. Lecanicillium A contained an unprecedented 6/5/6/5/7/6 cyclic system with a spirocyclic center at C-2′. Biologically, lecanicillium E, emethacin B, and versicolor A displayed significant inhibitory activity against H1975, with IC50 values of 7.2~16.9 μM, and lecanicillium E also showed antibacterial activities against four pathogens, with MIC values of 10~40 μg/mL. The cytotoxicity and antibacterial activity results indicated that the disulfide bond unit at C-2′ was crucial for the activity of this kind of alkaloids. This finding further clarified the chemical structure diversity of thiodiketopiperazine-class alkaloids, and the structural diversity and biological activities of thiodiketopiperazine-class alkaloids may be worthy of further studies.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/md21110575/s1: the 1D NMR, 2D NMR, HRESIMS, IR, and UV spectra of compounds 16, details of ECD calculation of compounds 16, X-ray crystallographic data for compound 3.

Author Contributions

Performing experiments, data analysis, and writing—original draft preparation, L.-F.Z. and J.L.; testing the activities, L.-X.L. and C.-N.Y.; formal analysis, X.L.; resources, data analysis, writing—review and editing, supervision, and funding acquisition, S.-H.Q. All authors have read and agreed to the published version of the manuscript.

Funding

The authors are grateful for the financial support provided by the National Natural Science Foundation of China (82173709 and 82373756), Key Science and Technology Project of Hainan Province (ZDKJ202018), and the Marine Economy Development Project of Guangdong Province (GDNRC [2023]43).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

The original data presented in the study are included in the article/Supplementary Material; further inquiries can be directed to the corresponding author.

Acknowledgments

The authors also appreciate the analytical facility center (Zhi-Hui Xiao, Xiao-Hong Zheng, Ai-Jun Sun, Xuan Ma, and Yun Zhang) of the South China Sea Institute of Oceanology, Chinese Academy of Sciences, for acquiring NMR, HRESIMS data, experimental ECD data, and X-ray crystallography of these compounds.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Gardiner, D.M.; Waring, P.; Howlett, B. The epipolythiodioxopiperazine (ETP) class of fungal toxins: Distribution, mode of action, functions and biosynthesis. J. Microbiol. 2005, 151, 1021–1032. [Google Scholar] [CrossRef]
  2. Welch, T.R.; Williams, R.M. Epidithiodioxopiperazines. occurrence, synthesis and biogenesis. Nat. Prod. Rep. 2014, 31, 1376–1404. [Google Scholar] [CrossRef]
  3. Ellen, M.F.; Barbara, J.H. Biosynthetic gene clusters for epipolythiodioxopiperazines in filamentous fungi. Mycol. Res. 2008, 112, 162–169. [Google Scholar]
  4. Wang, J.M.; Ding, G.Z.; Fang, L.; Dai, J.G.; Yu, S.S.; Wang, Y.H.; Chen, X.G.; Ma, S.G.; Qu, J.; Xu, S.; et al. Thiodiketopiperazines Produced by the Endophytic Fungus Epicoccum nigrum. J. Nat. Prod. 2010, 73, 1240–1249. [Google Scholar] [CrossRef]
  5. Meng, L.H.; Wang, C.Y.; Mandi, A.; Li, X.M.; Hu, X.Y.; Kassack, M.U.; Kurtan, T.; Wang, B.G. Three Diketopiperazine Alkaloids with Spirocyclic Skeletons and One Bisthiodiketopiperazine Derivative from the Mangrove-Derived Endophytic Fungus Penicillium brocae MA-231. Org. Lett. 2016, 18, 5304–5307. [Google Scholar] [CrossRef]
  6. Niu, S.W.; Liu, D.; Shao, Z.Z.; Proksch, P.; Lin, W.H. Eutypellazines N−S, new thiodiketopiperazines from a deep sea sediment derived fungus Eutypella sp. with anti-VRE activities. Tetrahedron Lett. 2017, 58, 3695–3699. [Google Scholar] [CrossRef]
  7. Liu, H.; Fan, J.; Zhang, P.; Hu, Y.C.; Liu, X.Z.; Li, S.M.; Yin, W.B. New insights into the disulfide bond formation enzymes in epidithiodiketopiperazine alkaloids. Chem. Sci. 2021, 12, 4132. [Google Scholar] [CrossRef]
  8. Guo, C.J.; Yeh, H.H.; Chiang, Y.M.; Sanchez, J.F.; Chang, S.L.; Bruno, K.S.; Wang, C.C.C. Biosynthetic Pathway for the Epipolythiodioxopiperazine Acetylaranotin in Aspergillus terreus Revealed by Genome-Based Deletion Analysis. J. Am. Chem. Soc. 2013, 135, 7205–7213. [Google Scholar] [CrossRef] [PubMed]
  9. Zhu, M.L.; Zhang, X.M.; Feng, H.M.; Dai, J.J.; Li, J.; Che, Q.; Gu, Q.Q.; Zhu, T.J.; Li, D.H. Penicisulfuranols A–F, Alkaloids from the Mangrove Endophytic Fungus Penicillium janthinellum HDN13-309. J. Nat. Prod. 2017, 80, 71–75. [Google Scholar] [CrossRef] [PubMed]
  10. Meng, L.H.; Zhang, P.; Li, X.M.; Wang, B.G. Penicibrocazines A–E, Five New Sulfide Diketopiperazines from the Marine-Derived Endophytic Fungus Penicillium brocae. Mar. Drugs 2015, 13, 276–287. [Google Scholar] [CrossRef] [PubMed]
  11. Xu, X.Y.; Tan, Y.H.; Gao, C.H.; Liu, K.; Tang, Z.Z.; Lu, C.J.; Li, H.Y.; Zhang, X.Y.; Liu, Y.H. New 3-Acyl Tetramic Acid Derivatives from the Deep-Sea-Derived Fungus Lecanicillium fusisporum. Mar. Drugs 2022, 20, 255. [Google Scholar] [CrossRef]
  12. Poolsawas, S.; Napasintuwong, O. Speed of hybrid maize adoption in Thailand: An application of duration analysis. J. ISSAAS 2019, 25, 95–103. [Google Scholar]
  13. Su, L.; Zhu, H.; Guo, Y.X.; Du, X.P.; Guo, J.G.; Zhang, L.; Qin, C. Lecanicillium coprophilum (Cordycipitaceae, Hypocreales), a new species of fungus from the feces of Marmota monax in China. Phytotaxa 2019, 387, 055–062. [Google Scholar] [CrossRef]
  14. Ishidoh, K.I.; Kinoshita, H.; Igarashi, Y.; Ihara, F.; Nihira, T. Cyclic lipodepsipeptides verlamelin A and B, isolated from entomopathogenic fungus Lecanicillium sp. J. Antibiot. 2014, 67, 459–463. [Google Scholar] [CrossRef] [PubMed]
  15. Wang, Z.Y.; Sang, X.N.; Sun, K.; Huang, S.D.; Chen, S.S.; Xue, C.M.; Ban, L.F.; Li, Z.L.; Hua, H.M.; Peia, Y.H.; et al. Lecanicillones A–C, three dimeric isomers of spiciferone A with a cyclobutane ring from an entomopathogenic fungus Lecanicillium sp. PR-M-3. RSC Adv. 2016, 6, 82348. [Google Scholar] [CrossRef]
  16. Wang, Z.Y.; Chen, S.S.; Sang, X.N.; Pan, H.Q.; Li, Z.L.; Hua, H.M.; Han, A.H.; Bai, J. Lecanicillolide, an α-pyrone substituted spiciferone from the fungus Lecanicillium sp. PR-M-3. Tetrahedron Lett. 2017, 58, 740–743. [Google Scholar] [CrossRef]
  17. Kawahara, N.; Nozawa, K.; Nakajim, S.; Yamazaki, M.; Kawai, K. Sulfur-containing Dioxopiperazine Derivatives from Emericella heterothallica. Heterocycles 1989, 29, 397–402. [Google Scholar]
  18. Wang, L.S.; Zong, S.K.; Wu, P.P.; Sun, X.L.; Wu, C.Z.; Wang, H.T.; Zhu, M.L. Two new alkaloids produced by Dothideomycetes sp. BMC-101 isolated from the Magnolia grandiflora. J. Asian Nat. Prod. Res. 2022, 25, 731–740. [Google Scholar] [CrossRef]
  19. Haritakun, R.; Rachtawee, P.; Komwijit, S.; Nithithanasilp, S.; Isaka, M. Highly Conjugated Ergostane-Type Steroids and Aranotin-Type Diketopiperazines from the Fungus Aspergillus terreus BCC 4651. Helv. Chim. Acta 2012, 95, 308–313. [Google Scholar] [CrossRef]
  20. He, T.; Wang, Y.D.; Du, L.Q.; Li, F.G.; Hu, Q.F.; Cheng, G.G.; Wang, W.G. Overexpression of Global Regulator LaeA Induced Secondary Metabolite Production in Aspergillus versicolor 0312. Rec. Nat. Prod. 2020, 14, 387–394. [Google Scholar] [CrossRef]
  21. Neuss, N.; Nagarajan, R.; Molloy, B.B.; Huckstep, L.L. Aranotin and related metabolites.II.Isolation, characterization, and structures of two new metabolites. Tetrahedron Lett. 1968, 9, 4467–4471. [Google Scholar] [CrossRef]
  22. Li, Y.; Yue, Q.; Krausert, N.M.; An, Z.Q.; Gloer, J.B.; Bills, G.F. Emestrins: Anti-Cryptococcus Epipolythiodioxopiperazines from Podospora australis. J. Nat. Prod. 2016, 79, 2357–2363. [Google Scholar] [CrossRef] [PubMed]
  23. Yurchenko, A.N.; Berdyshev, D.V.; Smetanina, O.F.; Ivanets, E.V.; Zhuravleva, O.I.; Rasin, A.B.; Khudyakova, Y.V.; Popov, R.S.; Dyshlovoy, S.A.; Amsberg, G.V.; et al. Citriperazines A-D produced by a marine algaederived fungus Penicillium sp. KMM 4672. Nat. Prod. Res. 2020, 34, 1118–1123. [Google Scholar] [CrossRef] [PubMed]
  24. Wu, J.S.; Shi, X.H.; Yao, G.S.; Shao, C.L.; Fu, X.M.; Zhang, X.L.; Guan, H.S.; Wang, C.Y. New Thiodiketopiperazine and 3,4-Dihydroisocoumarin Derivatives from the Marine-Derived Fungus Aspergillus terreus. Mar. Drugs 2020, 18, 132. [Google Scholar] [CrossRef]
  25. Cai, C.L.; Waring, P. Redox sensitive epidithiodioxopiperazines in biological mechanisms of toxicity. Redox Rep. 2000, 5, 257–264. [Google Scholar]
  26. Orfali, R.S.; Aly, A.H.; Ebrahim, W.; Abdel-Aziz, M.S.; Muller, W.E.G.; Lin, W.H.; Daletos, G.; Proksch, P. Pretrichodermamide C and N-methylpretrichodermamide B, two new cytotoxic epidithiodiketopiperazines from hyper saline lake derived Penicillium sp. Phytochem. Lett. 2015, 11, 168–172. [Google Scholar] [CrossRef]
  27. Liang, X.; Huang, Z.H.; Ma, X.; Zheng, Z.H.; Zhang, X.X.; Lu, X.H.; Qi, S.H. Mycotoxins as inhibitors of protein tyrosine phosphatases from the deep-sea-derived fungus Aspergillus puniceus SCSIO z021. Bioorganic Chem. 2021, 107, 104571. [Google Scholar] [CrossRef]
  28. Liu, C.M.; Yao, F.H.; Lu, X.H.; Zhang, X.X.; Luo, L.X.; Liang, X.; Qi, S.H. Isoquinoline Alkaloids as Protein Tyrosine Phosphatase Inhibitors from a Deep-Sea-Derived Fungus Aspergillus puniceus. Mar. Drugs 2022, 20, 78. [Google Scholar] [CrossRef]
  29. Wang, J.; Nong, X.H.; Amin, M.; Qi, S.H. Hygrocin C from marine-derived Streptomyces sp. SCSGAA 0027 inhibits biofilm formation in Bacillus amyloliquefaciens SCSGAB0082 isolated from South China Sea gorgonian. Appl. Microbiol. Biotechnol. 2018, 102, 1417–1427. [Google Scholar] [CrossRef]
Figure 1. Structures of compounds 119 isolated from L. kalimantanense SCSIO41702.
Figure 1. Structures of compounds 119 isolated from L. kalimantanense SCSIO41702.
Marinedrugs 21 00575 g001
Figure 2. Key HMBC and COSY correlations of compounds 14.
Figure 2. Key HMBC and COSY correlations of compounds 14.
Marinedrugs 21 00575 g002
Figure 3. Key NOESY correlations of compounds 14.
Figure 3. Key NOESY correlations of compounds 14.
Marinedrugs 21 00575 g003
Figure 4. Comparison of experimental and calculated ECD spectra of 1, 2, 46 in MeOH and 3 in CH3CN.
Figure 4. Comparison of experimental and calculated ECD spectra of 1, 2, 46 in MeOH and 3 in CH3CN.
Marinedrugs 21 00575 g004
Figure 5. ORTEP plot of the X-ray crystallographic data for 3.
Figure 5. ORTEP plot of the X-ray crystallographic data for 3.
Marinedrugs 21 00575 g005
Table 1. 1H NMR data of compounds 16 (700 MHz, δH in ppm, J in Hz) in DMSO-d6.
Table 1. 1H NMR data of compounds 16 (700 MHz, δH in ppm, J in Hz) in DMSO-d6.
No.123456
32.21 (d, 12.7),
2.71 (dd, 12.7, 6.5)
2.91 (dt, 15.3, 1.2),
3.14 (dt, 15.3, 2.2)
2.96 (dt, 15.2, 1.2)
3.25 (dt, 15.2, 2.3)
2.92 (dd, 16.1, 2.1)
3.20 (d, 16.1)
2.96 (dt, 15.4, 1.2)
3.24 (dt, 15.4, 2.3)
3.26 (overlap, m)
43.06 (t, 6.5)----2.70 (dd, 13.6, 5.2), 3.05 (dd, 13.6, 4.5)
53.87 (d, 6.5)6.69 (t, 2.4)6.76 (t, 2.6)6.81 (t, 2.6)6.77 (t, 2.2)-
64.86 (ddd, 7.8, 6.5, 1.2)6.29 (dd, 8.2, 2.4)6.43 (dd, 8.3, 2.6)6.43 (dd, 8.2, 2.6)6.42 (dd, 8.2, 2.0)7.19 (overlap, m)
74.82 (ddd, 7.8, 6.5, 1.2)4.79 (dd, 8.2, 2.1)4.70 (dd, 8.3, 2.2)4.69 (dd, 8.2, 2.2)4.72 (dd, 8.2, 2.0)7.26 (overlap, m)
86.00 (d, 6.5)4.46 (ddt, 7.7, 6.7, 2.1)5.57 (dt, 8.1, 2.2)5.52 (dt, 8.2, 2.2)5.54 (dt, 8.2, 2.1)7.18 (overlap, m)
96.61 (d, 6.5)4.74 (dq, 7.7, 2.1)4.92 (dq, 8.1, 2.2)5.00 (dq, 8.2, 2.2)4.99 (dq, 8.2, 2.1)7.26 (overlap, m)
10-2.29 (s)---7.19 (overlap, m)
11--2.01 (s)1.97 (s)1.95 (s)-
12--2.13 (s)2.18 (s)2.15 (s)-
3′3.26 (d, 16.4),
3.94 (d, 16.4)
3.25 (d, 16.6),
4.01 (d, 16.6)
2.43 (d, 14.4)
2.85 (d, 14.4)
2.94 (d, 15.4)
3.16 (dd, 15.4, 2.1)
3.30 (overlap, m)
3.45 (dt, 15.9, 1.3)
-
4′-----2.84 (d, 13.1),
3.28 (d, 13.1)
5′7.26 (d, 7.5)7.26 (d, 7.5)3.83 (d, 2.7)6.98 (d, 10.3)6.86 (t, 2.2)-
6′6.92 (t, 7.5)6.91 (t, 7.5)3.67 (dd, 4.3, 2.7)6.09 (d, 10.3)6.43 (dd, 8.2, 2.0)7.12 (overlap, m)
7′7.14 (t, 8.0)7.14 (t, 8.0)3.34 (dd, 4.3, 1.2)-4.68 (dd, 8.2, 2.0)7.24 (overlap, m)
8′6.81 (d, 8.0)6.79 (d, 8.0)5.79 (dd, 9.1, 1.2)5.73 (d, 11.2)5.52 (dt, 8.2, 2.2)7.26 (overlap, m)
9′--4.34 (d, 9.1)4.92 (dd, 11.2, 1.5)4.95 (dq, 8.2, 2.2)7.24 (overlap, m)
10′-----7.12 (overlap, m)
11′--1.96 (s)2.07 (s)1.97 (s)1.37 (s)
12′--2.25 (s)2.16 (s)2.47 (s)-
NH10.06 (s)9.77 (s)----
1-NH-----8.66 (s)
1′-NH-----8.23 (d, 2.0)
5-OH5.31 (s)-----
8-OH-5.28 (d, 6.7)----
4′-OH---6.18 (s)--
Table 2. 13C NMR data of compounds 16 (175 MHz, δC in ppm) in DMSO-d6.
Table 2. 13C NMR data of compounds 16 (175 MHz, δC in ppm) in DMSO-d6.
No.123456
1165.2, C166.4, C163.5, C164.3, C163.7, C-
270.8, C69.2, C70.1, C70.8, C70.7, C166.2, C
348.2, CH240.2, CH238.7, CH240.0, CH239.0, CH254.8, C
446.3, CH109.1, C110.7, C110.4, C110.9, C37.2, CH2
569.9, CH136.7, CH137.0, CH137.6, CH137.1, CH135.9, C
652.5, CH137.9, CH139.9, CH139.8, CH139.9, CH130.5, CH
7100.6, CH111.0, CH105.3, CH105.8, CH105.5, CH128.0, CH
8144.9, CH71.2, CH71.2, CH71.5, CH71.2, CH126.6, CH
992.9, CH64.1, CH59.8, CH59.6, CH59.6, CH128.0, CH
10-13.3, CH3169.6, C169.4, C169.4, C130.5, CH
11--20.7, CH320.8, CH320.9, CH3-
12--13.8, CH314.1, CH313.9, CH3-
1′162.8, C163.4, C163.6, C163.1, C164.3, C-
2′91.8, C93.0, C71.7, C69.2, C74.9, C164.8, C
3′38.3, CH238.4, CH239.0, CH248.0, CH237.6, CH268.2, C
4′125.4, C125.5, C60.9, C75.5, C110.6, C44.4, CH2
5′124.5, CH124.5, C53.6, CH150.2, CH137.3, CH135.0, C
6′121.3, CH121.2, CH47.9, CH125.8, CH139.9, CH130.3, CH
7′128.1, CH128.0, CH54.1, CH191.5, C105.5, CH128.0, CH
8′109.1, CH109.1, CH72.6, CH74.8, CH70.7, CH127.1, CH
9′156.6, C156.7, CH57.5, CH69.1, CH60.1, CH128.0, CH
10′--169.4, C168.9, C169.5, C130.3, CH
11′--20.6, CH320.3, CH320.7, CH311.2, C
12′--14.3, CH313.8, CH324.1, CH3-
Table 3. Antibacterial activities, cytotoxicity, and toxicity of 119.
Table 3. Antibacterial activities, cytotoxicity, and toxicity of 119.
Comp.Antibacterial Activity (MIC: μg/mL)Cytotoxicity
(IC50)
against
(μM)
Toxicity (TC50 in μg/mL)
B. subtilisM.
luteus
E.
coli
S.
aureus
MRSAS. agalactiaeP. aeruginosaS.
iniae
H1975HepG-2Brine Shrimp
160>80>80>80>80>80>80>80-->80
28060>80>80>80>80>8070>80>80>80
3>80>80>80>80>80>80>80>80>80>80>80
450>80>80>80>80>80>804071.5>8080
5104080>80>8025>801516.3>8040
6>80>80>80>80>80>80>80>80>80>80-
7>80>80>80>80>80>80>80>8016.9>80-
8>80>80>80>80>80>80>80>80>80>80-
9>80>80>80>80>80>80>80>80>80>80>80
10>80>80>80>80>80>80>80>80>80>80>80
1160>80>80>80>80>80>805069.3>80>80
12>8050>80>80>8050>80>80>80>8050
13>80>80>80>80>80>80>80>80>80>80-
14>80>80>80>80>80>80>8040>80>80>80
1580>80>80>80>80>80>80>80>80>80>80
1670>80>80>80>8070>80>807.2>80-
1780>80>80>80>8070>803035.2>80-
1880>80>80>80>80>80>804050.641.2>80
198070>80>80>80>80>80>80>80>80-
VMN-0.6--2.5------
CFN0.6-0.60.6-50.62.5---
VMN: Vancomycin; CFN: Ciprofloxacin; “-”: Not tested.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhong, L.-F.; Ling, J.; Luo, L.-X.; Yang, C.-N.; Liang, X.; Qi, S.-H. Lecanicilliums A–F, Thiodiketopiperazine-Class Alkaloids from a Mangrove Sediment-Derived Fungus Lecanicillium kalimantanense. Mar. Drugs 2023, 21, 575. https://doi.org/10.3390/md21110575

AMA Style

Zhong L-F, Ling J, Luo L-X, Yang C-N, Liang X, Qi S-H. Lecanicilliums A–F, Thiodiketopiperazine-Class Alkaloids from a Mangrove Sediment-Derived Fungus Lecanicillium kalimantanense. Marine Drugs. 2023; 21(11):575. https://doi.org/10.3390/md21110575

Chicago/Turabian Style

Zhong, Lin-Fang, Juan Ling, Lian-Xiang Luo, Chang-Nian Yang, Xiao Liang, and Shu-Hua Qi. 2023. "Lecanicilliums A–F, Thiodiketopiperazine-Class Alkaloids from a Mangrove Sediment-Derived Fungus Lecanicillium kalimantanense" Marine Drugs 21, no. 11: 575. https://doi.org/10.3390/md21110575

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