Next Article in Journal
The Applicability of Chromatographic Retention Modeling on Chiral Stationary Phases in Reverse-Phase Mode: A Case Study for Ezetimibe and Its Impurities
Next Article in Special Issue
First Gallium and Indium Crystal Structures of Curcuminoid Homoleptic Complexes: All-Different Ligand Stereochemistry and Cytotoxic Potential
Previous Article in Journal
Inflammation and Psoriasis: A Comprehensive Review
Previous Article in Special Issue
Mechanisms of NMDA Receptor Inhibition by Sepimostat—Comparison with Nafamostat and Diarylamidine Compounds
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mono- and Dimeric Sorbicillinoid Inhibitors Targeting IL-6 and IL-1β from the Mangrove-Derived Fungus Trichoderma reesei BGRg-3

1
School of Chemistry, Sun Yat-sen University, Guangzhou 510275, China
2
Guangdong Key Laboratory of Animal Conservation and Resource Utilization, Institute of Zoology, Guangdong Academy of Sciences, Guangzhou 510260, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2023, 24(22), 16096; https://doi.org/10.3390/ijms242216096
Submission received: 23 October 2023 / Revised: 3 November 2023 / Accepted: 7 November 2023 / Published: 8 November 2023
(This article belongs to the Special Issue Natural Products and Synthetic Compounds for Drug Development)

Abstract

:
Four new sorbicillinoids, named trichodermolide E (1), trichosorbicillin J (2), bisorbicillinolide B (3), and demethylsorbiquinol (5), together with eight known compounds (4, 612), were isolated from the cultures of the mangrove-derived fungus Trichoderma reesei BGRg-3. The structures of the new compounds were determined by analyzing their detailed spectroscopic data, while the absolute configurations were further determined through electronic circular dichroism calculations. Snatzke’s method was additionally used to determine the absolute configurations of the diol moiety in 1. In a bioassay, compounds 7 and 10 performed greater inhibitory activities on interleukin-6 and interleukin-1β than the positive control (dexamethasone) at the concentration of 25 μM. Meanwhile, compounds 5 and 6 showed potent effects with stronger inhibition than dexamethasone on IL-1β at the same concentration.

Graphical Abstract

1. Introduction

The mangrove is a complex ecosystem occurring in tropical and subtropical intertidal estuarine zones [1]. The special geographical location of mangroves provides a special environment with high salt, hypoxia, frequent tides, and strong solar radiation, which nourishes a diverse group of microorganisms. Among them, mangrove-derived fungi play an essential role in creating structurally unique and diverse bioactive secondary metabolites, which attract the significant attention of organic chemists and pharmacologists [1,2].
Sorbicillinoids are a family of polyketides characterized by a cyclic hexaketide nucleus with a sorbyl side chain [3]. Diverse and complex sorbicillinoids have been isolated mainly from marine fungi Aspergillus, Penicillium, Trichoderma, Phialocephala, etc. [4,5,6,7]. Sorbicillinoids can be classified into monomers, bisorbicillinoids, trisorbicillinoids, and hybrid sorbicillinoids according to their structures. The unique structures of sorbicillinoids lead to various biological activities, including cytotoxic, antibacterial, antifungal, anti-inflammatory, phytotoxic, and α-glucosidase inhibitory activities [3,8,9,10,11], which makes them one of the important lead compounds for drug development.
In our ongoing research on secondary metabolites from mangrove-derived fungi [1], four new sorbicillinoids, named trichodermolide E (1), trichosorbicillin J (2), bisorbicillinolide B (3), and demethylsorbiquinol (5), were isolated from the endophytic fungus Trichoderma reesei BGRg-3, which came from the root of the mangrove plant Avicennia marina at Nansha mangrove reserve in Guangzhou, China. Meanwhile, eight known compounds, including bisorbicillinolide (4) [12], sorbiquinol (6) [13], 13-hydroxy-trichodermolide (7) [6], 14-hydroxybislongiquinolide (8) [14], demethyltrichodimerol (9), trichodimerol (10) [15], (+)-(R)-vertinolide (11) [16], and saturnispols H (12) [10], were also obtained from the fungus BGRg-3 (Figure 1). The separation process, structural characteristics, and biological activities of the isolated sorbicillinoids are reported.

2. Results and Discussion

2.1. Structure Identification

Trichodermolide E (1) was isolated as a yellow oil. Its molecular formula was confirmed as C18H22O7 by using high resolution electrospray ionization mass spectroscopy (HR-ESIMS) (m/z 351.1433 [M+H]+, calcd. for C18H23O7, 351.1438, Figure S1) and NMR data, showing that eight degrees of unsaturation were required. The 1H NMR spectrum (Table 1) displayed four olefinic protons (δH 7.25 (dd, J = 15.6, 9.6 Hz, H-13), 6.28 (overlap, H-14, H-15), and 6.18 (d, J = 15.6 Hz, H-12)), two methine groups (δH 5.38 (dd, J = 3.1, 3.1 Hz, H-17) and 3.06 (t, J = 5.2 Hz, H-9)), two methylene groups (δH 3.98 (dd, J = 12.3, 2.8 Hz, H-18), 3.88 (dd, J = 12.3, 3.1 Hz, H-18), 3.61 (dd, J = 17.3, 5.3 Hz, H-10) and 2.86 (dd, J = 17.3, 5.3 Hz, H-10)), and three methyl groups (δH 1.86 (d, J = 5.0 Hz, H3-16), 1.85 (s, H3-7), and 1.42 (s, H3-8)). Then, 18 carbon signals in the 13C NMR spectrum (Table 1) were classified into three methyls (δC 24.2, 18.8 and 11.0), two methylenes (δC 62.4 and 35.2), two methines (δC 81.8 and 45.2), two quaternary carbons (δC 76.4 and 69.4), six olefinic carbons (δC 151.2, 144.4, 141.4, 132.2, 131.7 and 129.0), two carbonyls (δC 202.9, and 201.3), and one ester group (δC 175.4) using DEPT (distortionless enhancement by polarization transfer) spectra. Based on the HSQC (heteronuclear single quantum coherence) spectrum, the connectivity between protons and their associated carbons was confirmed (Table 1).
1H-1H COSY (proton-proton correlation spectroscopy) correlations (Figure 2) proved that there was a long side chain (from H-9 to H-10, from H-12 to H3-16) and a short one (from H-17 to H-18). Based on the HMBC (heteronuclear multiple bond correlation) correlations (Figure 2) from H-12 to C-11, from H-13 to C-11, and from H-10 to C-11, a sorbyl side chain was constructed for 1. The HMBC correlations from H3-7 to C-3, C-4, and C-5, from H3-8 to C-5, C-6, and from H-17 to C-1, C-2, C-3, and C-4 showed the presence of a heterocycle. Other HMBC correlations from H-9 to C-1, C-2, C-6, and C-8 exhibited the connection of the sorbyl side chain and the heterocycle. Thus, the constitutional formula of compound 1 was identified.
The relative configuration of compound 1 was determined using coupling constants and a NOESY (nuclear overhauser effect spectroscopy) spectrogram (Figure 3). The double bonds of the side chain were supposed to be E geometries according to the coupling constants (15.6 Hz for JH-12/H-13) and the NOESY correlations of H-13/H-15 and H-12/H-14. Due to the heterocycle skeleton, the configuration of C-2 and C-6 was assumed to be 2S, 6S, or 2R, 6R. Additional NOESY correlations of H3-8/H-9 and H3-7/H-9 indicated that the relative configuration of the skeleton was 2R*, 6R*, 9R*.
The absolute configuration of compound 1 was confirmed based on the ECD (electronic circular dichroism) calculation and Snatzke’s method [17,18]. In order to determine the absolute configuration of the skeleton, ECD calculations were performed in B3LYP functional with the 6-311+G(d,p) basis set, using the B3LYP/6-31G(d) optimized geometries which were found through the conformer distribution calculation in the MMFF force field. Supported by the comparison of the experimental ECD spectrum with the calculated one (Figure 4a), the absolute configuration of the skeleton was confirmed to be 2R, 6R, 9R. Then, Snatzke’s method was used to determine the absolute configuration of acyclic 1,2-diols. Based on the Mo2(OAc)4-induced ECD measurement (Figure 4b), the 17S configuration was confirmed by the positive cotton effect at 360 nm [18]. Thus, the absolute configuration of compound 1 was defined as 2R, 6R, 9R, and 17S.
Trichosorbicillin J (2) was expected to have a molecular formula of C15H18O6, confirmed through the HR-ESIMS (m/z 293.1034 [M−H], calcd. for C15H17O6, 293.1031, Figure S9), and was obtained as a yellow oil. The 13C NMR spectrum (Table 1) showed 15 carbons, and they were classified via the DEPT spectrum into one methyl (δC 7.6), three methylenes (δC 46.2, 36.8 and 29.3), one methine (δC 68.8), two olefinic carbons (δC 148.8, 124.3), six aromatic carbons (δC 164.4, 164.0, 130.8, 114.2 and 112.3), one carbonyl (δC 204.7), and one carboxyl (δC 171.8). Then, 14 protons were connected with their associated carbons according to the HSQC spectrum (Table 1). The COSY correlations (Figure 2) showed the existence of a side chain from C-8 to C-13. The COSY correlation of H-4/H-5, together with the coupling constant (8.9 Hz for JH-4/H-5), showed the ortho arrangement of them. The HMBC correlation (Figure 2) from H-12 to C-14 proved the existence of the carboxyl (C-14, δC 171.8). Based on the HMBC correlation from H3-15 to C-1, C-2, C-3, from H-4 to C-2, C-3 and from H-5 to C-1, C-6, C-7, a sorbicillinoid skeleton was built for 2. Other HMBC correlations (from H-8 and H-9 to C-7) exhibited the connection of the side chain and the sorbicillinoid skeleton. Comparing the specific rotation with the similar compound isotrichosorbicillin E ( α D 20 = +8.1 (c 0.07, MeOH)) [19], the configuration of 2 ( α D 20 = +4.9 (c 0.1, MeOH)) was identified as 9R.
Bisorbicillinolide B (3) showed the appearance of a red oil and its molecular formula was C28H36O8 based on the HR-ESIMS (m/z 499.2346 [M−H], calcd. for C28H35O8, 499.2326, Figure S17). According to the 13C NMR spectrum (Table 2), 28 carbon signals were found and were then classified into six methyls, four methylenes, three methines, six quaternary carbons, four alkene carbons, one enol carbon, one carboxyl, and three carbonyls based on the DEPT and HSQC spectra, thus revealing a similar structure to the known compound bisorbicillinolide (4) [12]. Analyzing the NMR data (Table 2), in association with the indices of hydrogen deficiency, 3 was assumed to be structurally related to known compound 4 with a difference in the reduced Δ2′- and Δ2″-double bonds. The COSY correlations (Figure 2) proved the existence of two side chains from H-2′ to H-6′ and from H-2″ to H-6″. According to the HMBC correlations, together with the remaining COSY correlations (Figure 2), the constitutional formula of 3 was allowed to be confirmed.
The double bonds in the side chains were both proved to be E geometries according to the NOESY signals (Figure 3) of H-4′/H-6′ and H-4″/H-6″. The NOESY correlations of H3-15/H-(10-OH), H3-15/H3-13, H3-15/H-8, H3-15/H-11, and H-8/H-13 indicated that C-15, 10-OH, C-13, H-8, and H-11 were syn-oriented, partially showing that the relative configuration of 3 would be 4R*, 8S*, 9S*, 10S*, 11R*. The NOESY correlations of H-8/H-2″, H3-14/H-2′, respectively determined the relative configuration of C-5 and C-7. An additional NOESY correlation of H-1/H-15 proved that the relative configuration of C-1 was 1S*. Eventually, the relative configuration of 3 would be 1S*, 4R*, 5R*, 7S*, 8S*, 9S*, 10S*, 11R*. To determine the absolute configuration of 3, ECD calculations were performed. The comparison between the experimental spectrum (Figure 5) with the calculated one could give the absolute configuration of compound 3 as 1S, 4R, 5R, 7S, 8S, 9S, 10S, 11R.
Demethylsorbiquinol (5) showed its molecular formula as C27H30O7, confirmed through HR-ESIMS (m/z 465.1925 [M−H], calcd. for C27H29O7, 465.1908, Figure S25). Its 13C NMR data showed 27 carbons, including five methyls, three methines, two quaternary carbons, seven alkene carbons, six aromatic carbons, one enol carbon, and three carbonyls. A similar compound, sorbiquinol (6) [13], was obtained in the same fraction. A comparison of the NMR data is shown in Table 3, showing the absence of a methyl in 5.
Key COSY and HMBC correlations (Figure 2) were then used to prove the structure of 5. The NOESY correlations (Figure 3) of H-2′/H-4′, H-3′/H-5′, and H-9/H3-11, together with the coupling constants (14.9 Hz for JH-2′/H-3′, 14.9 Hz for JH-4′/H-5′, and 15.0 Hz for JH-9/H-10), confirmed the E geometries of all double bonds. According to the NOESY correlations of H3-19/H-8 and H-1/H-8, the relative configuration was partially assumed to be 1S*, 4R*, 8R*. Additional NOESY correlations of H-1/H-18 and H-7/H-9 confirmed the relative configuration to be 1S*, 4R*, 7R*, 8R*. The configuration 6S* was determined based on the NOESY correlation of H3-20/H-2′. Thus, the relative configuration of 5 was assumed to be 1S*, 4R*, 6S*, 7R*, 8R*, which was then confirmed to be its absolute configuration through ECD calculations (Figure 5).

2.2. Anti-Inflammatory Activities

The bioactivities of 12 sorbicillinoids (compounds 112) were evaluated by measuring the effects of each compound on IL-6- and IL-1β-induced luciferase expression in RAW264.7 cells, testing under the concentration of 25 μM for 4 h. Their cytotoxicity was also evaluated at the same concentration level. The results are shown in Table 4. Among them, compounds 2, 57, and 10 exhibited inhibition against IL-6 and IL-1β. Compounds 7 and 10 presented their remarkable anti-inflammatory activities, respectively, with 47% and 67% inhibition of IL-6, 85% and 87% inhibition of IL-1β, which were even more effective than the dexamethasone. Compound 5 also showed its potent inhibition of IL-6 and IL-1β (27% and 58%, respectively), which was mildly weaker than the similar compound 6. However, compound 5 exhibited the advantage of less cytotoxicity. Moreover, compound 2 presented a moderate inhibition of both targets with little cytotoxicity.

3. Materials and Methods

3.1. General Experimental Procedures

The NMR spectra were obtained from Bruker Avance 400 MHz and 600 MHz spectrometers (Karlsrule, Germany) at room temperature. The HR-ESIMS spectra were tested on a ThermoFisher LTQ-Or-bitrap-LC-MS spectrometer (Palo Alto, CA, USA). Optical rotation data were collected from an MCP300 (Anton Paar, Shanghai, China). UV spectra were recorded on a Shimadzu UV-2600 spectrophotometer (Shimadzu, Kyoto, Japan). The ECD experiments were processed on a J-810 spectropolarimeter (JASCO, Tokyo, Japan). Silica gel (200–300 mesh, Marine Chemical Factory, Qingdao, China) and Sephadex LH-20 (Amersham Pharmacia, Piscataway, NJ, USA) were used to perform column chromatography (CC). Semi-preparative HPLC (Ultimate 3000 BioRS, Thermo Scientific, Bremen, Germany) was used to purify the compounds.

3.2. Fungal Material

Fungus BGRg-3 was isolated from the root of mangrove plant Avicennia marina, which was collected from the Nansha Mangrove National Nature Reserve in Guangdong Province, China. The strain was then identified as Trichoderma reesei based on the analysis of its ITS sequence (deposited in GenBank, accession no OR353740). The fungus is now deposited at Sun Yat-sen University, China.

3.3. Fermentation

The fungus Trichoderma reesei BGRg-3 was cultivated on solid cultured medium in 1 L erlenmeyer flasks, containing 50 g of rice and 80 mL of 3‰ saline water. A total of 100 flasks were fermented for 30 days at room temperature.

3.4. Extraction and Purification

When the fermentation was carried out, the mycelia, as well as the rice medium, was extracted four times with MeOH and EtOAc. Eventually, a 95.5 g extract was collected (Figure S37). Then, the crude extract was eluted with a gradient elution with petroleum ether (PE) and EtOAc (from 9:1 to 0:10) on silica gel CC to obtain six fractions (Fr. A–F).
Fr.A was first fractionated on a Sephadex LH-20 column with MeOH/CH2Cl2 (1:1) to afford 2 fractions (Fr.A.1 to Fr.A.2). Fr.A.1 was then subjected to silica gel CC eluting with CH2Cl2/MeOH (100:1) to give 3 (10.3 mg). Fr.A.2 was applied to silica gel CC with CH2Cl2/MeOH (60:1) to afford a mixture of 5 (8.4 mg) and 6 (21.4 mg), which was then separated via RP-HPLC with MeOH/H2O (80:20). Fr.B was successively subjected to a Sephadex LH-20 column with MeOH/CH2Cl2 (1:1) and silica gel CC with CH2Cl2/MeOH (60:1) to give 4 (7.2 mg). Fr.C was also fractionated on a Sephadex LH-20 column with MeOH/CH2Cl2 (1:1) to afford 2 fractions (Fr.C.1 to Fr.C.2). Fr.C.1 was purified on RP-HPLC with MeOH/H2O (70:30) to afford 1 (5.3 mg). Then, 7 (6.6 mg) was obtained from Fr.C.2 by using silica gel CC eluting with CH2Cl2/MeOH (40:1). Fr.D was successively subjected to a Sephadex LH-20 column with MeOH/CH2Cl2 (1:1) and silica gel CC with CH2Cl2/MeOH (40:1) to give 9 (9.7 mg), 10 (11.2 mg), and 12 (13.5 mg). Each of them was purified via RP-HPLC with MeCN/H2O (75:25) and a Sephadex LH-20 column with MeOH. Fr.E was separated into Fr.E.1 and Fr.E.2 on Sephadex LH-20 column with MeOH/CH2Cl2 (1:1). Fr.E.1 was purified via RP-HPLC with MeCN/H2O (70:30) to give 2 (6.7 mg). Fr.E.2 was then subjected to normal-phase HPLC with n-hexane/2-propanol (75:25) to afford 8 (7.5 mg). Fr.F was successively subjected to Sephadex LH-20 column with MeOH/CH2Cl2 (1:1) and silica gel CC with CH2Cl2/MeOH (30:1) to give 11 (7.9 mg).
Trichodermolide E (1): yellow oil (5.3 mg). α D 20 = +18.1 (c 0.1, MeOH); UV (MeOH) λmax (log ε): 238 (2.42), 271 (2.52); ECD (MeOH) λmaxε): 238 (+4.12), 264 (+1.35), 342 (−0.91), see Figure S33; 1H NMR (CD3OD, 600 MHz), 13C NMR (CD3OD, 600 MHz) and other NMR data, see Table 1 and Figures S2–S8; HR-ESIMS m/z 351.1433 [M+H]+ (calcd. for C18H23O7, 351.1438), see Figure S1.
Trichosorbicillin J (2): yellow oil (6.7 mg). α D 20 = +4.9 (c 0.1, MeOH); UV (MeOH) λmax (log ε): 213 (2.92), 286 (2.65), see Figure S34; 1H NMR (CD3OD, 400 MHz), 13C NMR (CD3OD, 400 MHz) and other NMR data, see Table 1 and Figures S10–S16; HR-ESIMS m/z 293.1034 [M−H] (calcd. for C15H17O6, 293.1031), see Figure S9.
Bisorbicillinolide B (3): red oil (10.2 mg). α D 20 = +203.9 (c 0.1, MeOH); UV (MeOH) λmax (log ε): 306 (2.27); ECD (MeOH) λmaxε): 214 (−5.03), 247 (+3.41), 286 (−1.42), 333 (+7.75), see Figure S35; 1H NMR (CDCl3, 400 MHz), 13C NMR (CDCl3, 400 MHz) and other NMR data, see Table 2 and Figures S18–S24; HR-ESIMS m/z 499.2326 [M−H] (calcd. for C28H35O8, 499.2346), see Figure S17.
Demethylsorbiquinol (5): yellow oil (8.4 mg). α D 20 = +223.3 (c 0.1, MeOH); UV (MeOH) λmax (log ε): 213 (2.62), 237 (2.42), 287 (252), 338 (2.65), 360 (2.63); ECD (MeOH) λmaxε): 214 (−1.37), 239 (+4.97), 289 (−12.18), 306 (−9.47), 342 (+11.88), see Figure S36; 1H NMR (CDCl3, 400 MHz), 13C NMR (CDCl3, 400 MHz) and other NMR data, see Table 3 and Figures S26–S32; HR-ESIMS m/z 465.1925 [M−H] (calcd. for C27H29O7, 465.1908), see Figure S25.

3.5. ECD Calculation

Conformational analyses were carried out via random searching with an MMFF force field using Spartan’14 (v1.1.0) [20]. Subsequently, the conformers were re-optimized with density functional theory (DFT) methods at the B3LYP/6-31+G (d, p) level, solvated in methanol, with the CPCM method using the Gaussian 09 [21]. Then, TDDFT calculations were performed at the B3LYP/6-311+G(d, p) level (nstates = 50) in methanol. The ECD spectra were obtained by overlapping Gaussian functions (σ = 0.3 eV). To obtain the final spectra, the simulated spectra of each conformer were averaged weighted with Boltzmann distribution and their relative Gibbs free energy, using the program SpecDis (v1.64) [22]. The theoretical ECD spectra of the corresponding enantiomers were obtained through direct inversions of the ECD spectra of the abovementioned conformers.

3.6. Absolute Configurations of the 17,18-Diol Moiety in 1

A modified version of Snatzke’s method was performed according to the published literature [17,18]. Mixtures of 1:1.2 diol-Mo2(OAc)4 for 1 were subjected to ECD measurement at a compound concentration of 1.0 mg/mL in DMSO for each. After mixing, the first ECD was recorded immediately, and the time evolution was surveyed until stationary (about 10 min after mixing). The inherent ECD was subtracted. In the induced ECD spectra, the observed signs of the diagnostic bands at around 310 and 400 nm were correlated to the absolute configuration of the 17,18-diol moiety.

3.7. Anti-Inflammatory Assay

The real-time PCR method was performed to measure the expression of the LPS-induced inflammatory factors IL-1β and IL-6 in RAW264.7 cells. Cells were randomly divided into control, stimulation, and administration groups, with 3 wells in each group. The control groups were cultured with a complete culture medium, the stimulation groups were stimulated with LPS (final concentration in 1 µg/mL), and the administration groups were stimulated with the mixture of LPS (final concentration in 1 µg/mL) and test compounds (final concentration in 25 µM). After culturing in an incubator for 4 h, reverse transcription was performed. Real-time PCR was performed to quantitatively analyze the products. The difference in CT values between the target gene and the internal reference gene GADPH (ΔCT) was used to perform the quantitative detection of genes in the samples. Results were given with the 2−△△CT values of the same gene in different samples, which were then converted into inhibition ratios.

4. Conclusions

Our work discovered four new sorbicillin-based derivatives from the secondary metabolites of Trichoderma reesei BGRg-3, derived from the mangrove plant Avicennia marina. Compound 1 exhibited a special skeleton that was rare in the sorbicillinoid family. Its 1D and 2D NMR data gave the relative configuration, and its absolute configuration was deduced based on ECD calculations and Snatzke’s method. Compound 2 was a derivative of sorbicillinol. Based on the comparison with similar compounds, constitutional formulas of compounds 3 and 5 were confirmed. Their absolute configurations were also determined through ECD calculations. In the bioactive evaluation, known compounds 7 and 10 exhibited the highest inhibition of IL-6 and IL-1β among the tested compounds, even more effective than dexamethasone. Compounds 5 and 6 also presented potent inhibition of IL-1β and moderate inhibition of IL-6. In addition, compound 6 showed the advantage of little cytotoxicity, which makes it possible to develop a nontoxic anti-inflammatory drug. Compound 8 exhibited remarkable cytotoxicity against RAW264.7 cells at the test concentration level; therefore further research on its cytotoxicity against other cells is expected.

Supplementary Materials

The supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms242216096/s1. References [23,24] are cited in the Supplementary Materials.

Author Contributions

Conceptualization, Z.S.; methodology, Z.S. and H.Y.; software, Y.L.; investigation, Y.L.; resources, T.C., Q.T. and Y.L.; data curation, Y.L. and H.O.; writing—original draft preparation, Y.L.; writing—review and editing, T.C. and Z.S.; visualization, Y.L. and T.C.; supervision, B.S.; project administration, B.W. and Z.S. All authors have read and agreed to the published version of the manuscript.

Funding

We thank the Guangdong Marine Economy Development Special Project (GDNRC[2022]35, GDNRC[2023]39), the National Natural Science Foundation of China (U20A2001, 21877133, 82104228), the Key-Area Research and Development Program of Guangdong Province (2020B1111030005), and the GDAS Special Project of Science and Technology Development (Grant No. 2021GDASYL–20210103057) for their generous support.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Acknowledgments

The authors sincerely thank X.M. and the South China Sea Institute of Oceanology, Chinese Academy of Sciences, for the collection of experimental ECD spectra, UV spectra, and specific rotation data.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chen, S.; Cai, R.; Liu, Z.; Cui, H.; She, Z. Secondary metabolites from mangrove-associated fungi: Source, chemistry and bioactivities. Nat. Prod. Rep. 2022, 39, 560–595. [Google Scholar] [CrossRef]
  2. Carroll, A.R.; Copp, B.R.; Davis, R.A.; Keyzers, R.A.; Prinsep, M.R. Marine natural products. Nat. Prod. Rep. 2023, 40, 275–325. [Google Scholar] [CrossRef]
  3. Harned, A.M.; Volp, K.A. The sorbicillinoid family of natural products: Isolation, biosynthesis, and synthetic studies. Nat. Prod. Rep. 2011, 28, 1790–1810. [Google Scholar] [CrossRef]
  4. Corral, P.; Esposito, F.P.; Tedesco, P.; Falco, A.; Tortorella, E.; Tartaglione, L.; Festa, C.; D’Auria, M.V.; Gnavi, G.; Varese, G.C.; et al. Identification of a sorbicillinoid-producing Aspergillus strain with antimicrobial activity against Staphylococcus aureus: A new polyextremophilic marine fungus from Barents sea. Mar. Biotechnol. 2018, 20, 502–511. [Google Scholar] [CrossRef]
  5. Ding, W.; Wang, F.; Li, Q.; Xue, Y.; Dong, Z.; Tian, D.; Chen, M.; Zhang, Y.; Hong, K.; Tang, J. Isolation and Characterization of Anti-Inflammatory Sorbicillinoids from the Mangrove-Derived Fungus Penicillium sp. DM815. Chem. Biodivers. 2021, 18, e2100229. [Google Scholar] [CrossRef]
  6. Rehman, S.U.; Yang, L.; Zhang, Y.; Wu, J.; Shi, T.; Haider, W.; Shao, C.; Wang, C. Sorbicillinoid derivatives from sponge-derived fungus Trichoderma reesei (HN-2016-018). Front. Microbiol. 2020, 11, 1334. [Google Scholar] [CrossRef]
  7. Zhang, Z.; He, X.; Che, Q.; Zhang, G.; Zhu, T.; Gu, Q.; Li, D. Sorbicillasins A–B and scirpyrone K from a deep-sea-derived fungus, Phialocephala sp. FL30r. Mar. Drugs 2018, 16, 245. [Google Scholar] [CrossRef] [PubMed]
  8. Hou, X.; Zhang, X.; Xue, M.; Zhao, Z.; Zhang, H.; Xu, D.; Lai, D.; Zhou, L. Recent advances in sorbicillinoids from fungi and their bioactivities (Covering 2016–2021). J. Fungi 2022, 8, 62. [Google Scholar] [CrossRef]
  9. Xie, C.; Zhang, D.; Lin, T.; He, Z.; Yan, Q.; Cai, Q.; Zhang, X.; Yang, X.; Chen, H. Antiproliferative sorbicillinoids from the deep-sea-derived Penicillium allii-sativi. Front. Microbiol. 2021, 11, 636948. [Google Scholar] [CrossRef]
  10. Meng, J.; Cheng, W.; Heydari, H.; Wang, B.; Zhu, K.; Konuklugil, B.; Lin, W. Sorbicillinoid-based metabolites from a sponge-derived fungus Trichoderma saturnisporum. Mar. Drugs 2018, 16, 226. [Google Scholar] [CrossRef]
  11. Deng, J.; Dai, H.; Wang, Y.; Chen, H.; Tan, Z.; Mei, W. Isolation and identification of the fungus Aspergillus sp. HNWSW-20 from Chinese agarwood and its secondary metabolites. Chin. J. Trop. Crops 2018, 39, 1618–1624. [Google Scholar]
  12. Abe, N.; Murata, T.; Hirota, A. Novel oxidized sorbicillin dimers with 1,1-diphenyl-2-picrylhydrazyl-radical scavenging activity from a fungus. Biosci. Biotechnol. Biochem. 1998, 62, 2120–2126. [Google Scholar] [CrossRef]
  13. Andrade, R.; Ayer, W.A.; Trifonov, L.S. The metabolites of Trichoderma longibrachiatum. Part II The structures of trichodermolide and sorbiquinol. Can. J. Chem. 1996, 74, 371–379. [Google Scholar] [CrossRef]
  14. Li, J.; Chen, T.; Yu, J.; Jia, H.; Chen, C.; Long, Y. New Sorbicillinoids from the Mangrove Endophytic Fungus Trichoderma reesei SCNU-F0042. Mar. Drugs 2023, 21, 442. [Google Scholar] [CrossRef]
  15. Abe, N.; Murata, T.; Hirota, A. Novel DPPH radical scavengers, bisorbicillinol and demethyltrichodimerol, from a fungus. Biosci. Biotechnol. Biochem. 1998, 62, 661–666. [Google Scholar] [CrossRef]
  16. McMullin, D.R.; Renaud, J.B.; Barasubiye, T.; Sumarah, M.W.; Miller, J.D. Metabolites of Trichoderma species isolated from damp building materials. Can. J. Microbiol. 2017, 63, 621–632. [Google Scholar] [CrossRef]
  17. Di Bari, L.; Pescitelli, G.; Pratelli, C.; Pini, D.; Salvadori, P. Determination of absolute configuration of acyclic 1,2-diols with Mo2(OAc)4. 1. Snatzke’s method revisited. J. Org. Chem. 2001, 66, 4819–4825. [Google Scholar] [CrossRef]
  18. Wang, Q.; Bao, L.; Yang, X.; Liu, D.; Guo, H.; Dai, H.; Song, F.; Zhang, L.; Guo, L.; Li, S.; et al. Ophiobolins P–T, five new cytotoxic and antibacterial sesterterpenes from the endolichenic fungus Ulocladium sp. Fitoterapia 2013, 90, 220–227. [Google Scholar] [CrossRef]
  19. Zhang, P.; Deng, Y.; Lin, X.; Chen, B.; Li, J.; Liu, H.; Chen, S.; Liu, L. Anti-inflammatory mono- and dimeric sorbicillinoids from the marine-derived fungus Trichoderma reesei 4670. J. Nat. Prod. 2019, 82, 947–957. [Google Scholar] [CrossRef]
  20. Spartan’ 14; Wavefunction Inc.: Irvine, CA, USA, 2013.
  21. Frisch, M.J.T.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09, Revision A.1; Gaussian, Inc.: Wallingford, CT, USA, 2009.
  22. Bruhn, T.H.Y.; Schaumlöffel, A.; Bringmann, G. SpecDis, version 1.53; University of Wuerzburg: Wurzburg, Germany, 2011.
  23. Kahlert, L.; Bassiony, E.F.; Cox, R.J.; Skellam, E.J. Diels–Alder Reactions During the Biosynthesis of Sorbicillinoids. Angew. Chem. Int. Edit. 2020, 59, 5816–5822. [Google Scholar] [CrossRef]
  24. Ren, S.; Zeng, Y.; Wang, Q.; Lin, Q.; Yin, X.; Chen, S.; Wang, M.; Liu, L.; Gao, Z. Major Facilitator Superfamily Transporter Participates in the Formation of Dimeric Sorbicillinoids Pigments. J. Agric. Food Chem. 2023, 71, 12216–12224. [Google Scholar] [CrossRef]
Figure 1. Structure of compounds 112.
Figure 1. Structure of compounds 112.
Ijms 24 16096 g001
Figure 2. Key HMBC and COSY correlations of 13 and 5.
Figure 2. Key HMBC and COSY correlations of 13 and 5.
Ijms 24 16096 g002
Figure 3. Key NOESY correlations of compounds 1, 3, and 5.
Figure 3. Key NOESY correlations of compounds 1, 3, and 5.
Ijms 24 16096 g003
Figure 4. (a) Experimental and calculated ECD spectrum of 1; (b) Mo2(OAc)4-induced ECD spectrum of 1.
Figure 4. (a) Experimental and calculated ECD spectrum of 1; (b) Mo2(OAc)4-induced ECD spectrum of 1.
Ijms 24 16096 g004
Figure 5. (a) Experimental and calculated ECD spectrum of 3; (b) experimental and calculated ECD spectrum of 5.
Figure 5. (a) Experimental and calculated ECD spectrum of 3; (b) experimental and calculated ECD spectrum of 5.
Ijms 24 16096 g005
Table 1. 1H and 13C NMR data of 1 and 2 (δ (ppm)).
Table 1. 1H and 13C NMR data of 1 and 2 (δ (ppm)).
Position1 (CD3OD)2 (CD3OD)
δCδH (J in Hz)δCδH (J in Hz)
1175.4 (C) 164.0 (C)
269.4 (C) 114.2 (C)
3151.2 (C) 164.4 (C)
4132.2 (C) 108.0 (CH)6.4 (1H, d, J = 8.9)
5202.9 (C) 130.8 (CH)7.62 (1H, d, J = 8.9)
676.4 (C) 112.3 (C)
711.0 (CH3)1.85 (3H, s)204.7 (C)
824.2 (CH3)1.42 (3H, s)46.2 (CH2)3.13 (1H, dd, J = 15.6, 8.0)3.01 (1H, dd, J = 15.6, 4.4)
945.2 (CH)3.06 (1H, t, J = 5.2)68.8 (CH)4.2 (1H, tt, J = 8.4, 4.4)
1035.2 (CH2)3.61 (1H, dd, J = 17.3, 5.2)2.86 (1H, dd, J = 17.3, 5.2)36.8 (CH2)1.69 (2H, m)
11201.3 (C) 29.3 (CH2)2.36 (2H, m)
12129.0 (CH)6.18 (1H, d, J = 15.6)148.8 (CH)6.92 (1H, dt, J = 15.4, 6.9)
13144.4 (CH)7.25 (1H, dd, J = 15.6, 9.6)124.3 (CH)5.87 (1H, d, J = 15.4)
14131.7 (CH)6.28 (overlap)171.8 (C)
15141.4 (CH)6.28 (overlap)7.6 (CH3)2.04 (3H, s)
1618.8 (CH3)1.86 (3H, d, J = 5.0)
1781.8 (CH)5.38 (1H, dd, J = 3.1, 3.1)
1862.4 (CH2)3.98 (1H, dd, J = 12.3, 3.1)3.88 (1H, dd, J = 12.3, 3.1)
Table 2. 1H and 13C NMR data of 34 (δ (ppm)).
Table 2. 1H and 13C NMR data of 34 (δ (ppm)).
Position3 (CDCl3)4 (CDCl3)
δCδH (J in Hz)δCδH (J in Hz)
144.2 (CH)3.25 (1H, d, J = 5.1)44.4 (CH)3.38 (1H, d, J = 5.4)
2108.0 (C) 107.3 (C)
3185.3 (C) 194.5 (C)
465.3 (C) 68.0 (C)
573.5 (C) 73.5 (C)
6210.2 (C) 210.3 (C)
784.2 (C) 84.1 (C)
856.5 (CH)3.62 (1H, d, J = 5.1)56.9 (CH)3.73 (1H, d, J = 5.4)
994.2 (C) 93.8 (C)
1088.5 (C) 88.3 (C)
1141.2 (CH)2.68 (1H, q, J = 7.7)41.1 (CH)2.66 (1H, q, J = 8.0)
12178.5 (C) 177.9 (C)
1310.0 (CH3)1.24 (3H, s)10.3 (CH3)1.36 (3H, s)
1417.4 (CH3)1.23 (3H, s)18.1 (CH3)1.23 (3H, s)
1524.2 (CH3)1.32 (3H, s)24.3 (CH3)1.39 (3H, s)
1613.8 (CH3)1.19 (3H, d, J = 7.7)13.7 (CH3)1.21 (3H, d, J = 8.0)
1′195.2 (C) 177.3 (C)
2′36.4 (CH2)2.50, 2.32 (2H, m)119.0 (CH)6.14 (1H, d, J = 14.8)
3′26.8 (CH2)2.28 (2H, m)144.6 (CH)7.39 (1H, dd, J = 14.8, 10.4)
4′129.4 (CH)5.37–5.52 (m)131.0 (CH)6.1–6.5 (m)
5′126.5 (CH)5.37–5.52 (m)141.5 (CH)6.1–6.5 (m)
6′18.0 (CH3)1.61 (3H, d, J = 6.3)18.9 (CH3)1.90 (3H, d, J = 5.2)
1″212.2 (C) 199.4 (C)
2″40.9 (CH2)2.89, 2.55 (2H, m)123.8 (CH)6.46 (1H, d, J = 14.8)
3″26.2 (CH2)2.19 (2H, m)147.9 (CH)7.42 (1H, dd, J = 14.8, 11.2)
4″128.7 (CH)5.28 (1H, m)130.3 (CH)6.21 (1H, dd, J = 15.2, 11.2)
5″127.1 (CH)5.37–5.52 (m)145.0 (CH)6.36 (1H, dq, J = 15.2, 7.6)
6″18.0 (CH3)1.59 (3H, d, J = 6.4)19.0 (CH3)1.89 (3H, d, J = 7.6)
Table 3. 1H and 13C NMR data of 56 (δ (ppm)).
Table 3. 1H and 13C NMR data of 56 (δ (ppm)).
Position5 (CDCl3)6 (CDCl3)
δCδH (J in Hz)δCδH (J in Hz)
147.1 (CH)3.31 (1H, d, J = 1.8)47.0 (CH)3.32 (1H, d, J = 1.5)
2106.7 (C) 106.8 (C)
3198.0 (C) 198.0 (C)
463.2 (C) 63.2 (C)
5211.5 (C) 211.5 (C)
675.9 (C) 75.7 (C)
746.7 (CH)4.27 (1H, dd, J = 6.5, 1.8)46.7 (CH)4.28 (1H, dd, J = 6.5, 1.5)
846.4 (CH)3.23 (1H, dd, J = 10.0, 6.5)46.5 (CH)3.26 (1H, dd, J = 10.0, 6.5)
9128.2 (CH)5.02 (1H, ddq, J = 15.0, 10.0, 1.8)128.6 (CH)5.02 (1H, ddq, J = 15.0, 10.0, 1.5)
10130.8 (CH)5.44 (1H, dq, J = 15.0, 6.5)130.6 (CH)5.45 (1H, dq, J = 15.0, 7.0)
1117.9 (CH3)1.61 (3H, dd, J = 6.5, 1.8)17.8 (CH3)1.60 (3H, dd, J = 6.5, 1.5)
12202.0 (C) 202.1 (C)
13112.8 (C) 112.0 (C)
14164.2 (C) 161.9 (C)
15103.5 (CH)6.34 (1H,s)110.7 (C)
16161.2 (C) 159.0 (C)
17116.4 (C) 115.0 (C)
18132.4 (CH)7.70 (1H, s)129.0 (CH)7.59 (1H, s)
1910.1 (CH3)1.16 (3H, s)10.0 (CH3)1.16 (3H, s)
2024.5 (CH3)1.21 (3H, s)24.3 (CH3)1.20 (3H, s)
2115.6 (CH3)2.25 (3H,s)16.0 (CH3)2.26 (3H,s)
22--7.5 (CH3)2.12 (3H,s)
1′169.0 (C) 168.8 (C)
2′117.2 (CH)5.57 (1H, d, J = 14.9)117.2 (CH)5.53 (1H, d, J = 15.0)
3′142.8 (CH)7.19 (1H,dd, J = 14.9, 10.7)142.4 (CH)7.17 (1H,dd, J = 15.0, 10.5)
4′130.9 (CH)5.97 (1H, m)130.8 (CH)5.93 (1H, ddq, J = 15.0, 10.0, 1.5)
5′139.7 (CH)6.08 (1H, dq, J = 14.9, 6.9)139.4 (CH)6.08 (1H, dq, J = 15.0, 7.0)
6′18.9 (CH3)1.82 (3H, d, J = 6.9)18.9 (CH3)1.82 (3H, dd, J = 7.0, 1.5)
Table 4. The cytotoxicity and the inhibitory effects on IL-6 and IL-1β of the tested compounds.
Table 4. The cytotoxicity and the inhibitory effects on IL-6 and IL-1β of the tested compounds.
Raw 264.7IL-6 Inhibition (%) 1IL-1β Inhibition (%) 1Viability of Cell (%) 1
Compound25 μM
1NANA79.85 ± 2.07
245.55 ± 9.2221.63 ± 5.7896.40 ± 6.26
3NANA98.39 ± 4.72
4NANA97.88 ± 2.84
526.76 ± 3.4457.77 ±14.38106.09 ± 0.99
634.86 ± 5.5575.08 ± 6.8086.72 ± 2.83
746.68 ± 11.8084.80 ± 4.4992.05 ± 1.86
8NANA55.44 ± 4.23
9NANA83.02 ± 0.59
1066.64 ± 11.3687.05 ± 1.94117.24 ± 4.72
11NANA83.96 ± 0.70
12NANA84.94 ± 1.81
DEX 236.19 ± 3.7548.19 ± 9.50NT
1 Data are presented as the mean ± SD. 2 Positive control; NA: no activity (inhibition ≤ 20%); NT: 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

Liu, Y.; Chen, T.; Sun, B.; Tan, Q.; Ouyang, H.; Wang, B.; Yu, H.; She, Z. Mono- and Dimeric Sorbicillinoid Inhibitors Targeting IL-6 and IL-1β from the Mangrove-Derived Fungus Trichoderma reesei BGRg-3. Int. J. Mol. Sci. 2023, 24, 16096. https://doi.org/10.3390/ijms242216096

AMA Style

Liu Y, Chen T, Sun B, Tan Q, Ouyang H, Wang B, Yu H, She Z. Mono- and Dimeric Sorbicillinoid Inhibitors Targeting IL-6 and IL-1β from the Mangrove-Derived Fungus Trichoderma reesei BGRg-3. International Journal of Molecular Sciences. 2023; 24(22):16096. https://doi.org/10.3390/ijms242216096

Chicago/Turabian Style

Liu, Yufeng, Tao Chen, Bing Sun, Qi Tan, Hui Ouyang, Bo Wang, Huijuan Yu, and Zhigang She. 2023. "Mono- and Dimeric Sorbicillinoid Inhibitors Targeting IL-6 and IL-1β from the Mangrove-Derived Fungus Trichoderma reesei BGRg-3" International Journal of Molecular Sciences 24, no. 22: 16096. https://doi.org/10.3390/ijms242216096

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