Next Article in Journal
Probing the Interactions of Sulfur-Containing Histidine Compounds with Human Gamma-Glutamyl Transpeptidase
Next Article in Special Issue
Monarubins A–C from the Marine Shellfish-Associated Fungus Monascus ruber BB5
Previous Article in Journal
Plocabulin Displays Strong Cytotoxic Activity in a Personalized Colon Cancer Patient-Derived 3D Organoid Assay
Previous Article in Special Issue
Novel Microcystins from Planktothrix prolifica NIVA-CYA 544 Identified by LC-MS/MS, Functional Group Derivatization and 15N-labeling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Four New Insecticidal Xanthene Derivatives from the Mangrove-Derived Fungus Penicillium sp. JY246

1
Key Laboratory of Tropical Medicinal Resource Chemistry of Ministry of Education, Hainan Normal University, Haikou, Hainan 571127, China
2
Key Laboratory of Tropical Medicinal Plant Chemistry of Hainan Province, College of Chemistry and Chemical Engineering, Hainan Normal University, Haikou, Hainan 571127, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Submission received: 28 September 2019 / Revised: 12 November 2019 / Accepted: 13 November 2019 / Published: 20 November 2019

Abstract

:
Four new xanthene derivatives, penicixanthenes A–D (14), and one known compound 5 were isolated from a marine mangrove endophytic fungus Penicillium sp. JY246 that was obtained from the stem of Ceriops tagal. Their structures were determined by detailed NMR, MS spectroscopic data, modified Mosher’s method, and calculated electronic circular dichroism data. All of the isolated compounds were examined for insecticidal activity. Compounds 2 and 3 showed growth inhibition activity against newly hatched larvae of Helicoverpa armigera Hubner with the IC50 values 100 and 200 μg/mL, respectively, and compounds 1, 3, and 4 showed insecticidal activity against newly hatched larvae of Culex quinquefasciatus with LC50 values of 38.5 (±1.16), 11.6 (±0.58), and 20.5 (±1) μg/mL, respectively. The four xanthene derivatives have the potential to be developed as new biopesticides.

1. Introduction

Fungal secondary metabolites have always been considered an important source for drug discovery due to their diverse chemical structures and bioactivities [1]. Among them, Penicillium are recognized as important producers of structurally unusual natural products, especially of terpenes with pharmaceutical potential as illustrated with chrysogenester, an anti-inflammatory meroterpenoid-type derivative [2], shearilicine, a cytotoxic indole-diterpenoid possessing a rare carbazole unit [3,4], and the penerpenes, unusual indole-terpenoids which have shown potent protein tyrosine phosphatase inhibitory activity [5]. Insecticidal agents were also reported for penicianstinoid A, an austinoid-like meroterpenoid [6], as well as plant regulators for the dongtingnoids, diterpenoid glycosides which revealed promising seed-germination-promoting activities [7]. Accordingly, strains of Penicillium have gained considerable attention due to their ability to produce unusual secondary metabolites and have proved to be a prolific source of bioactive compounds.
In our search for new bioactive compounds from fungal sources [6,8,9], the fungus Penicillium sp. JY246 was isolated from the mangrove’s stem Ceriops tagal, collected from the South China Sea. The EtOAc extract of the fermentation broth showed significant activity against newly hatched larvae of Helicoverpa armigera Hubner. Chemical investigation of the fungus fermentation’s organic extract resulted in the isolation of four new xanthene derivatives (14), along with one known compound 5 (Figure 1). Herein, the isolation, structure elucidation, and insecticidal activity of these compounds are described.

2. Results and Discussion

2.1. Structure Elucidation

Compound 1 was obtained as a white amorphous powder, and its molecular formula was determined to be C16H14O3 based on HRESIMS (Figure S8), accounting for 10 degrees of unsaturation. Analysis of the 1H NMR spectrum (Table 1, Figure S1) displayed six olefinic proton signals at δH 7.05 (dd, J = 8.4, 8.4 Hz, H-6), 7.04 (dd, J = 8.0, 8.0 Hz, H-3), 6.58 (dd, J = 8.0, 0.8 Hz, H-4), 6.55 (dd, J = 8.0, 0.8 Hz, H-2), 6.52 (dd, J = 8.4, 1.2 Hz, H-5), and 6.50 (dd, J = 8.4, 1.2 Hz, H-7); two methine signals at δH 4.64 (m, H-11) and 3.99 (dd, J = 12.4, 4.0 Hz, H-9); one methylene group at δH 3.17 (ddd, J = 13.2, 4.0, 1.6 Hz, H-10b) and 1.79 (dt, J = 12.4, 6.4 Hz, H-10a); and one methyl signal at δH 1.39 (d, J = 6.4 Hz, H-12). Interpretation of the 13C NMR and DEPT spectra (Table 2, Figures S2 and S3) revealed 16 carbon signals, attributable to twelve sp2-hybridized carbons (δC 156.9, 154.3, 152.3, 150.1, 127.9, 127.8, 110.4, 110.2, 109.3, 109.1, 107.0 and 106.5) indicating the presence of two benzene rings, two sp3 methine groups (δC 71.7 and 22.3), one sp3 methylene group (δC 33.0), and one methyl group (δC 20.7). The detailed structure of 1 was identified by interpretation of the COSY (Figure S6) and HMBC spectra (Figure 2 and Figure S5). The key HMBC correlations from H-9 to C-4a/C-4b, and further—H-10b to C-8a/C-8b, H-11 to C-8/C-9, and H3-12 to C-10/C-11—indicate that C-9 linked the two benzene rings at C-8a/C-8b, and the HMBC correlations from H-3 to C-4a and H-6 to C-4b indicated the presence of one oxygen bridge between C-4a and C-4b. So, the two benzene rings were connected through C-9 and oxygen-bridge to build up the skeleton of 9-methylene-9H-xanthene [10]. Furthermore, C-11 linked to C-8 through the oxygen bridge was deduced by chemical shift at C-8 (δC 154.3) and C-11 (δC 71.7). Hereto, the planar structure of 1 was elucidated (Figure 1).
The relative configuration of 1 was revealed by the NOESY experiment (Figure 3 and Figure S7). The NOESY correlations of H-9 to H3-12 indicated that H-9 and H3-12 were on the contrary side of the H-11. The absolute configuration of 1 was determined by comparing experimental and calculated electronic circular dichroism (ECD) spectra for the truncated model (9R, 11S)-1 and the truncated model (9S, 11R)-1 using time-dependent density-functional theory (TDDFT). The DFT reoptimization of the initial Merck molecular force field (MMFF) minima was performed at the B3LYP/6 − 31 + g (d, p) level with a conductor-like polarizable continuum model (CPCM) solvent model for MeOH [11]. The theoretical spectrum of 1 showed the same Cotton effect with the experimental plot recorded in MeOH (Figure 4), which supported that the absolute configuration was 9S, 11R. Thus, the completed structure of 1 was elucidated as depicted in Figure 1, and was named penicixanthene A.
Compound 2 was isolated as a colorless amorphous powder. The molecular formula of 2 was determined to be C20H22O5 (10 degrees of unsaturation) by HRESIMS (Figure S16). The 1H and 13C NMR data (Table 1 and Table 2, Figures S9 and S10) revealed that 2 belongs to the 9-methylene-9H-xanthene class [10], and suggested a close structural relationship to 1. The obvious differences in 1H NMR spectrum were the disappearance of two aromatic protons signals at δH 6.50/H-5 and δH 7.05/H-6 in 2. In addition, in the 13C NMR spectra, the C-5/C-6 signals moved downfield δC 110.2/127.9 in 1 vs. δC 110.7/157.3 in 2, indicating that the disappearance of two aromatic protons were respectively replaced by a butan-1-one unit at C-5 and a hydroxy group at C-6 in 2, which was further supported by HMBC (Figure S13). The existence of the butan-1-one unit was confirmed by HMBC correlations from H-14 to C-5/C-13/C-16, H-15 to C-13/C-14, and H-16 to C-14/C-15.
Furthermore, the oxygen-bridge between C-11 to C-8 in 1 was broken in 2. The C-8 signal moved high-field significantly at δC 134.4 in 2 vs. δC 154.3 in 1 in the 13C NMR spectrum, and the H-8 signal at δH 6.69 in 2 in the 1H NMR spectrum, indicating that oxygen function at C-8 in 1 was replaced by an aromatic proton in 2, which was confirmed by HMBC correlations (Figure 2) of H-7 to C-5/C-8a, H-8 to C-4b/C-6/C-9, H-9 to C-4a/C-4b, H-10 to C-8a/C-8b, H-11 to C-9, and H3-12 to C-10/C-11. The 1H-1H COSY (Figure S14) and HMBC spectra allowed the complete assignment of 2. The absolute configuration of C-9 in 2 was resolved by comparing experimental and calculated ECD spectra using TDDFT (Figure 4) [11]. The absolute configuration of C-11 was determined by making MTPA esters of 2 [12], and the differences in 1H NMR (Figures S33 and S35) chemical shifts between (S)- and (R)-MTPA esters (Δδ = δSδR) (Figure 5) were calculated to assign the absolute configuration of C-11 to be R. Thus, the absolute configuration of 2 was established as 9R, 11R, and it was named penicixanthene B.
Compound 3, isolated as a colorless amorphous powder, was evidenced to have a molecular formula of C20H22O5 (10 degrees of unsaturation) from its HRESIMS data (Figure S24). The similarities in the NMR data (Figures S17 and S18) for 3 and 2 suggested that 3 was structurally similar to 2, with the main differences being the chemical shifts of C-8b/C-9/C-10 (Table 2) and the coupling constants of H-9/H-10 (Table 1), and as indicated by the small coupling constant J9,10 = 1.2 Hz and 2D NMR data in 3 (vs. the coupling constant J9,10 = 8.0 Hz in 2) [13], implying that the H-9 orientation of 3 was different from that of 2. The absolute configuration of C-9 in 3 was established by comparison of experimental and calculated electronic circular dichroism (ECD) data [11]. The experimental spectrum of 3 showed the opposite Cotton effect with the experimental results of 2 recorded in MeOH (Figure 4), which supported that the absolute configuration of C-9 in 3 was 9S. The absolute configuration of C-11 was determined by making MTPA esters of 3 [12]. The differences in 1H NMR (Figures S37 and S39) chemical shifts between (S)- and (R)-MTPA esters (Δδ = δSδR) (Figure 5) were calculated to assign the absolute configuration of C-11 to be R. Thus, the absolute configuration of 3 was assigned as 9S, 11R, and the structure of 3 was defined as penicixanthene C.
Compound 4 was obtained as white amorphous powder, and the molecular formula was deduced to be C21H24O5 on the basis of HRESIMS (Figure S32), implying 10 degrees of unsaturation. Analysis of the 1H NMR spectrum (Table 1, Figure S25) displayed signals corresponding to five olefinic protons signal at δH 7.24 (dd, J = 8.0, 8.0 Hz, H-3), 7.06 (d, J = 8.0 Hz, H-4), 6.81 (d, J = 8.0 Hz, H-2), 6.22 (d, J = 8.4 Hz, H-8), and 6.05 (d, J = 8.4 Hz, H-7); two methine signals at δH 4.72 (m, H-12) and 4.63 (m, H-9); four methylene groups at δH 3.12 (t, J = 7.2 Hz, H-14), 2.22 (m, H-10b), 1.81 (m H-10a), 1.74 (m, H-11), and 1.72 (m, H-15); and two methyl groups at δH 3.57 (s, H-17) and 0.99 (t, J = 7.2 Hz, H-16). The 13C NMR spectrum (Table 2, Figure S26) displayed 21 signals, which were classified by DEPT (Figure S27) and HMQC spectra (Figure S28) as twelve sp2-hybridized carbons (δC 162.1, 159.7, 158.3, 141.6, 135.6, 128.9, 128.3, 125.4, 123.3, 110.9, 110.8, and 105.8) indicating the presence of two benzene rings, two sp3 methine groups (δC 67.9 and 32.3), four sp3 methylene groups (δC 47.2, 28.2, 23.9, and 19.2), and two methyl groups (δC 59.9 and 14.3). The above NMR spectroscopic data indicated the presence of a structure similar to nodulisporin F [13] with two subunits. One of the subunits was similar to 5 [14], isolated from the same source. The other subunit showed signals corresponding to the naphthalene part. The locations of connection between the two units at C-8a/C-9 were evidenced from HMBC correlations (Figure 2, Figure S29) of H-9 to C-4a/C-4b/C-8/C-8a/C-8b, as well as from H-8 to C-9. Based on the above data, the naphthalene part of 4 was linked to a 1-(2,6-dihydroxyphenyl)-butan-1-one moiety in para-position to the phenol group (at C-9). In addition, the C-4a and C-12 in 4 were replaced by a methoxyl group and a hydroxy group, respectively.
The relative configuration of 4 was based on the NOESY (Figure S31) correlations, as indicated in Figure 3. The NOESY correlations of H-9 to H-10a, H-11a, and H-12, as well as H-10b to H-11b, indicated that H-9, H-10a, H-11a, and H-12 were on the contrary side of the H-10b and H-11b. The absolute configuration of 4 was elucidated as 9R, 12R by comparing its experimental ECD spectrum to the calculated spectrum (Figure 4) [11]. The structure of 4 was assigned, named as penicixanthene D.

2.2. Biological Activity

Compounds 15 were examined for insecticidal activity. In the test, we set up three parallel trials. Compounds 2 and 3 showed growth inhibition activities against newly hatched larvae of Helicoverpa armigera Hubner with the IC50 values 100 and 200 μg/mL, respectively. Azadirachtin was used as positive control with the IC50 value of 25 μg/mL. Compounds 1, 3, and 4 showed insecticidal activity against newly hatched larvae of Culex quinquefasciatus with LC50 values of 38.5 (±1.16), 11.6 (±0.58) and 20.5 (±1) μg/mL, respectively. Azadirachtin was used as positive control with the LC50 value of 8.8 (±0.58) μg/mL (Table 3).

3. Materials and Methods

3.1. General Experimental Procedures

Optical rotations were measured on a JASCO P-1020 digital polarimeter (JASCO, Tokyo, Japan). IR spectra were recorded on a Thermo Nicolet 6700 (using KBr disks) spectrophotometer (Thermo, Madison, USA). 1D and 2D NMR spectra were measured on a Bruker AV-400 spectrometer with TMS as the internal standard (Bruker Corporation, Fällanden, Switzerland) (The PULPROG for testing NOESY was noesygpphpp, the temperature was 300 K, and the mix time of NOESY was 0.3 s). HRESIMS spectra were obtained on a Q-TOF Ultima Global GAA076 LC mass spectrometer (Waters, Milford, MA, USA). Preparative HPLC was used for an Agilent 1260 prep-HPLC system with an Agilent Eclipse XDB-C18 column (250 mm × 9.4 mm, 7 µm, Agilent Corporation, Santa Clara, CA, USA). The other experimental procedures were performed as reported previously [6].

3.2. Fungal Materials

The fungal strain Penicillium sp. JY246 was isolated from the stem of mangrove Ceriops tagal, collected in the South China Sea in July 2016. The fungus was identified according to its morphological characteristics and by comparison of the internal transcribed spacer (ITS) sequence amplification, primer pair ITS1 and ITS4 and sequencing of the ITS region. The sequence data has been submitted to GenBank with the accession number MK050979, and identified as Penicillium.
The fungal strain was cultivated in 20 L potato glucose liquid medium (15 g of glucose and 30 g of sea salt in 1 L of potato infusion, in 1 L Erlenmeyer flasks each containing 300 mL of culture broth) at 25 °C without shaking for 4 weeks.

3.3. Extraction and Isolation

The fungal cultures were filtered through cheesecloth, and the filtrate was extracted with EtOAc (3 × 20 L, 24 h each). The organic extracts were concentrated in vacuo to yield an oily residue (23.6 g), which was subjected to silica gel column chromatography (CC) (petroleum ether, EtOAc v/v, gradient 100:0–0:100) to generate five fractions (Fr. 1–Fr. 5). Fraction Fr. 4 (5 g) was separated by silica gel CC and eluted with petroleum ether-EtOAc (from 3:1 to 0:1) to afford four subfractions (4a4d). Subfraction 4b were further separated by semi-preparative HPLC (MeOH–H2O, 45:55, v/v) to obtain 1 (6.0 mg), 4 (4.5 mg), and 5 (5.0 mg), and subfraction 4c was further separated by semi-preparative HPLC with MeOH–H2O (35:65 v/v) to give 2 (3.3 mg) and 3 (3.5 mg).
Penicixanthene A (1): white amorphous powder. [α] D 24 − 10.7 (c = 0.02, CHCl3). CD (c 2 × 10-4 mol/L, MeOH) λmaxε) 211 (+11.8), 242 (−0.1), 269 (+1.1) nm; IR (KBr) νmax 3405, 2923, 1747, 1740, 1230, 1064, 756 cm-1; 1H and 13C NMR see Table 1 and Table 2; HRESIMS m/z 255.1012 [M + H]+ (calcd. For C16H15O3, 255.1016).
Penicixanthene B (2): white amorphous powder. [α] D 24 – 9.5 (c = 0.02, CHCl3). CD (c 2 × 10-4 mol/L, MeOH) λmaxε) 207 (−19.6), 232 (+4.2), 252 (+0.8), 286 (+0.8) nm; IR (KBr) νmax 3402, 2944, 1747, 1410, 1250, 1064, 818 cm-1; 1H and 13C NMR see Table 1 and Table 2; HRESIMS m/z 343.1537 [M + H]+ (calcd. for C20H23O5, 343.1540).
Penicixanthene C (3): white amorphous powder. [α] D 24 + 10.7 (c = 0.10, CHCl3). CD (c 2 × 10-4 mol/L, MeOH) λmaxε) 207 (+8.3), 232 (−2.1), 252 (−0.7), 286 (−0.7) nm; IR (KBr) νmax 3405, 2943, 1740, 1413, 1255, 1061, 820 cm-1; 1H and 13C NMR see Table 1 and Table 2; HRESIMS m/z 343.1537 [M + H]+ (calcd. for C20H23O5, 343.1540).
Penicixanthene D (4): white amorphous powder. [α] D 24 + 30.7 (c = 0.10, CHCl3). CD (c 2 × 10-4 mol/L, MeOH) λmaxε) 196 (+25.1), 215 (−5.2), 231 (+1.9), 276 (+0.5) nm; IR (KBr) νmax 3265, 1631, 1237 cm-1; 1H and 13C NMR see Table 1 and Table 2; HRESIMS m/z 357.1700 [M + H]+ (calcd. for C21H25O5 357.1697)

3.4. Preparation of (S)- and (R)-MTPA Ester Derivatives of Compounds 2 and 3

Preparations of (S)- and (R)-MTPA ester derivatives of 2 and 3 were performed as described previously [6].
S-MTPA ester of 2 (2a): 1H NMR (CDCl3, 400 MHz) δH: 7.23 (1H, d, J = 8.4 Hz, H-8), 7.15 (1H, t, J = 6.8 Hz, H-3), 6.96 (1H, dd, J = 8.8, 0.8 Hz, H-2), 6.81 (1H, d, J = 8.4 Hz, H-7), 6.46 (1H, dd, J = 8.8, 0.8 Hz, H-4), 5.34 (1H, m, H-11), 4.12 (1H, m, H-9), 3.59 (2H, m, H-14), 2.53 (1H, m, H-10b), 2.01 (2H, m, H-15), 1.74 (1H, m, H-10a), 1.17 (3H, d, J = 6.0 Hz, H-12), 0.77 (3H, t, J = 7.2 Hz, H-16). ESI-MS m/z 773.1 [M − H].
R-MTPA ester of 2 (2b): 1H NMR (CDCl3, 400 MHz) δH: 7.29 (1H, m, H-8), 7.10 (1H, m, H-3), 6.99 (1H, m, H-2), 6.46 (1H, m, H-7), 6.23 (1H, m, H-4), 5.32 (1H, m, H-11), 4.08 (1H, m, H-9), 3.58 (2H, m, H-14), 2.33 (1H, m, H-10b), 1.99 (2H, m, H-15), 1.69 (1H, m, H-10a), 1.20 (3H, m, H-12), 0.72 (3H, m, H-16). ESI-MS m/z 773.1 [M − H].
S-MTPA ester of 3 (3a): 1H NMR (CDCl3, 400 MHz) δH: 7.35 (1H, m, H-8), 7.30 (1H, m, H-3), 6.81 (1H, m, H-2), 6.69 (1H, m, H-7), 6.49 (1H, m, H-4), 5.34 (1H, m, H-11), 4.52 (1H, m, H-9), 3.58 (2H, m, H-14), 2.61 (1H, m, H-10b), 2.11 (2H, m, H-15), 2.01 (1H, m, H-10a), 1.30 (3H, d, J = 4.8 Hz, H-12), 0.74 (3H, t, J = 7.2 Hz, H-16). ESI-MS m/z 773.1 [M − H].
R-MTPA ester of 3 (3b): 1H NMR (CDCl3, 400 MHz) δH: 7.30 (1H, m, H-8), 7.07 (1H, m, H-3), 6.96 (1H, m, H-2), 6.56 (1H, m, H-7), 6.36 (1H, m, H-4), 5.35 (1H, m, H-11), 4.59 (1H, m, H-9), 3.55 (2H, m, H-14), 2.56 (1H, m, H-10b), 2.10 (2H, m, H-15), 1.89 (1H, m, H-10a), 1.31 (3H, d, J = 6.0 Hz, H-12), 0.74 (3H, t, J = 6.4 Hz, H-16). ESI-MS m/z 773.1 [M − H].

3.5. Computational Section

As previously reported [6,15,16].

3.6. Insecticidal Activities against Newly Hatched Larvae of Helicoverpa armigera Hubner

Insecticidal activity against newly hatched larvae of H. armigera Hubner was evaluated according to the previously reported methods [6]. Newly hatched larvae were raised under 25 ± 1 °C and a relative humidity of 80%. DMSO was used as the negative control, azadirachtin was used as the positive control, and artificial diet was used as the blank control. The number of dead larvae was recorded on the 2nd, 4th, 6th, and 8th day after treatment, respectively [17].

3.7. Insecticidal Activities against Newly Hatched Larvae of Culex quinquefasciatus

Insecticidal activity against newly hatched larvae of C. quinquefasciatus was evaluated according to the previously reported methods [17,18]. DMSO was used as the negative control, azadirachtin was used as the positive control, and 10 mL dechlorinated water was used as the blank control. The number of dead larvae was recorded on the 1st, 2nd, 3rd, and 4th day after treatment, respectively.

4. Conclusions

In summary, four new xanthene derivatives, penicixanthenes A–D (14), along with one known compound 5, were obtained from the mangrove-derived fungus Penicillium sp. JY246. Furthermore, the absolute configurations of 2 and 3 were determined by modified Mosher’s method and calculated electronic circular dichroism data. Compounds 2 and 3 showed growth inhibition activity against newly hatched larvae of H. armigera Hubner with the IC50 values 100 and 200 μg/mL, respectively. Compounds 1, 3, and 4 showed insecticidal activity against newly hatched larvae of C. quinquefasciatus with LC50 values of 38.5 (±1.16), 11.6 (±0.58), and 20.5 (±1) μg/mL, respectively.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/1660-3397/17/12/649/s1, 1H NMR, 13C NMR, DEPT, HMQC, COSY, HMBC, NOESY, and HRESIMS spectra of compounds 14. Figure S1: 1H NMR (DMSO-d6, 400 MHz) spectrum of 1. Figure S2: 13C NMR (DMSO-d6, 100 MHz) spectrum of 1. Figure S3: DEPT (DMSO-d6, 100 MHz) spectrum of 1. Figure S4: HMQC spectrum of 1. Figure S5: HMBC spectrum of 1. Figure S6: 1H-1H COSY spectrum of 1. Figure S7: NOESY spectrum of 1. Figure S8: HRESIMS spectrum of 1. Figure S9: 1H NMR (CD3OD, 400 MHz) spectrum of 2. Figure S10: 13C NMR (CD3OD, 100 MHz) spectrum of 2. Figure S11: DEPT (CD3OD, 100 MHz) spectrum of 2. Figure S12: HMQC spectrum of 2. Figure S13: HMBC spectrum of 2. Figure S14: 1H-1H COSY spectrum of 2. Figure S15: NOESY spectrum of 2. Figure S16: HRESIMS spectrum of 2. Figure S17: 1H NMR (CD3OD, 400 MHz) spectrum of 3. Figure S18: 13C NMR (CD3OD, 100 MHz) spectrum of 3. Figure S19: DEPT (CD3OD, 100 MHz) spectrum of 3. Figure S20: HMQC spectrum of 3. Figure S21: HMBC spectrum of 3. Figure S22: 1H-1H COSY spectrum of 3. Figure S23: NOESY spectrum of 3. Figure S24: HRESIMS spectrum of 3. Figure S25: 1H NMR (CD3OD, 400 MHz) spectrum of 4. Figure S26: 13C NMR (CD3OD, 100 MHz) spectrum of 4. Figure S27: DEPT (CD3OD, 100 MHz) spectrum of 4. Figure S28: HMQC spectrum of 4. Figure S29: HMBC spectrum of 4. Figure S30: COSY spectrum of 4. Figure S31: NOESY spectrum of 4. Figure S32: HRESIMS spectrum of 4. Figure S33: 1H NMR (CDCl3, 400 MHz) of S-MTPA ester of 2a. Figure S34: ESIMS spectrum of 2a. Figure S35: 1H NMR (CDCl3, 400 MHz) of R-MTPA ester of 2b. Figure S36: ESIMS spectrum of 2b. Figure S37: 1H NMR (CDCl3, 400 MHz) of S-MTPA ester of 3a. Figure S38: ESIMS spectrum of 3a. Figure S39: 1H NMR (CDCl3, 400 MHz) of R-MTPA ester of 3b. Figure S40: ESIMS spectrum of 3b.

Author Contributions

M.B. performed the experiments for the isolation, structure elucidation, insecticidal activity, and antimicrobial evaluation, and prepared the manuscript; Y.-P.L. performed the 1D and 2D NMR experiments; X.-H.N. contributed to part of the structure determination; X.-M.Z. contributed to the optimization of fermentation; C.-J.Z. contributed to part of the structure determination and jointly supervised the research; G.-Y.C. supervised the research work and revised the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (Nos. 21662012 and 41866005), Program for Innovative Research Team in University (No. IRT-16R19), and the Key Research & Development Program of Hainan (ZDYF2019116).

Conflicts of Interest

The authors declare no competing financial interest.

References

  1. Tang, J.W.; Kong, L.M.; Zu, W.Y.; Hu, K.; Li, X.N.; Yang, B.C.; Wang, W.G.; Sun, H.D.; Li, Y.; Puno, P.T. Isopenicins A-C: Two types of antitumor meroterpenoids from the plant endophytic fungus Penicillium sp. sh18. Org. Lett. 2019, 21, 711–775. [Google Scholar] [CrossRef] [PubMed]
  2. Liu, S.; Su, M.Z.; Song, S.J.; Hong, J.K.; Chung, H.Y.; Jung, J.H. An anti-inflammatory PPAR-γ agonist from the jellyfish-derived fungus Penicillium chrysogenum J08NF-4. J. Nat. Prod. 2018, 81, 356–363. [Google Scholar] [CrossRef] [PubMed]
  3. Liu, S.; Su, M.Z.; Song, S.J.; Jung, J.H. Marine-derived Penicillium species as producers of cytotoxic metabolites. Mar. Drugs 2017, 15, 329. [Google Scholar] [CrossRef] [PubMed]
  4. Ariantari, N.P.; Ancheeva, E.; Wang, C.Y.; Mándi, A.; Knedel, T.O.; Kurtán, T.; Chaidir, C.; Müller, W.E.G.; Kassack, M.U.; Janiak, C.; et al. Indole diterpenoids from an endophytic Penicillium sp. J. Nat. Prod. 2019, 82, 1412–1423. [Google Scholar] [CrossRef] [PubMed]
  5. Kong, F.D.; Fan, P.; Zhou, L.M.; Ma, Q.Y.; Xie, Q.Y.; Zheng, H.Z.; Zheng, Z.H.; Zhang, R.S.; Yuan, J.Z.; Dai, H.F.; et al. Penerpenes A-D, Four indole terpenoids with potent protein tyrosine phosphatase inhibitory activity from the marine-derived fungus Penicillium sp. KFD28. Org. Lett. 2019, 21, 4864–4867. [Google Scholar] [CrossRef] [PubMed]
  6. Bai, M.; Zheng, C.J.; Huang, G.L.; Mei, R.Q.; Wang, B.; Luo, Y.P.; Zheng, C.; Niu, Z.G.; Chen, G.Y. Bioactive meroterpenoids and isocoumarins from the mangrove-derived fungus Penicillium sp. TGM112. J. Nat. Prod. 2019, 82, 1155–1164. [Google Scholar] [CrossRef] [PubMed]
  7. Bie, Q.; Chen, C.M.; Yu, M.Y.; Guo, J.R.; Wang, J.P.; Liu, J.J.; Zhou, Y.; Zhu, H.C.; Zhang, Y.H. Dongtingnoids A-G: fusicoccane diterpenoids from a Penicillium species. J. Nat. Prod. 2019, 82, 80–86. [Google Scholar] [CrossRef] [PubMed]
  8. Huang, G.L.; Zhou, X.M.; Bai, M.; Liu, Y.X.; Zhao, Y.L.; Luo, Y.P.; Niu, Y.Y.; Zheng, C.J.; Chen, G.Y. Dihydroisocoumarins from the mangrove-derived fungus Penicillium citrinum. Mar. Drugs 2016, 14, 177. [Google Scholar] [CrossRef] [PubMed]
  9. Bai, M.; Huang, G.L.; Mei, R.Q.; Wang, B.; Luo, Y.P.; Nong, X.H.; Chen, G.Y.; Zheng, C.J. Bioactive lactones from the mangrove-derived fungus Penicillium sp. TGM112. Mar. Drugs 2019, 17, 433. [Google Scholar] [CrossRef] [PubMed]
  10. Cao, J.Q.; Yao, Y.; Chen, H.; Qiao, L.; Zhou, Y.Z.; Pei, Y.H. A New xanthene from Blumea riparia. Chin. Chem. Lett. 2007, 18, 303–305. [Google Scholar] [CrossRef]
  11. Choukchou-Braham, N.; Asakawa, Y.; Lepoittevin, J.P. Isolation, structure determination and synthesis of new dihydroisocoumarins from Ginkgo biloba L. Tetrahedron Lett. 1994, 35, 3949–3952. [Google Scholar] [CrossRef]
  12. Kusumi, T.; Fujita, Y.; Ohtani, I.; Kakisawa, H. Anomaly in the modified Mosher’s method: absolute configurations of some marine cembranolides. Tetrahedron Lett. 1991, 32, 2923–2926. [Google Scholar] [CrossRef]
  13. Dai, J.Q.; Krohn, K.; Draeger, S.; Schulz, B. New naphthalene-chroman coupling products from the endophytic fungus, Nodulisporium sp. from Erica arborea. Eur. J. Org. Chem. 2009, 10, 1564–1569. [Google Scholar] [CrossRef]
  14. Yang, L.J.; Liao, H.X.; Bai, M.; Huang, G.L.; Luo, Y.P.; Niu, Y.Y.; Zheng, C.J.; Chen, G.Y. One new cytochalasin metabolite isolated from a mangrovederived fungus Daldinia eschscholtzii HJ001. Nat. Prod. Res. 2018, 32, 208–213. [Google Scholar] [CrossRef] [PubMed]
  15. Elnaggar, M.S.; Ebrahim, W.; Mándi, A.; Kurtán, T.; Müller, W.E.G.; Kalscheuer, R.; Singab, A.; Lin, W.H.; Liu, Z.; Proksch, P. Hydroquinone derivatives from the marine-derived fungus Gliomastix sp. RSC Adv. 2017, 7, 30640–30649. [Google Scholar] [CrossRef]
  16. Ren, J.; Ding, S.S.; Zhu, A.; Cao, F.; Zhu, H.J. Bioactive azaphilone derivatives from the fungus Talaromyces aculeatus. J. Nat. Prod. 2017, 80, 2199–2203. [Google Scholar] [CrossRef] [PubMed]
  17. Guo, Z.K.; Gai, C.J.; Cai, C.H.; Chen, L.L.; Liu, S.B.; Zeng, Y.B.; Yuan, J.Z.; Mei, W.L.; Dai, H.F. Metabolites with insecticidal activity from Aspergillus fumigatus JRJ111048 isolated from mangrove plant Acrostichum specioum endemic to hainan island. Mar. Drugs 2017, 15, 381. [Google Scholar] [CrossRef] [PubMed]
  18. Zhang, W.F.; Crickmore, N.; George, Z.; Xie, L.; He, Y.Q.; Li, Y.Z.; Tang, J.L.; Tian, L.; Wang, X.; Fang, X.J. Characterization of new highly mosquitocidal isolate of Bacillus thuringiensis—an alternative to Bti? J. Invertebr. Pathol. 2012, 109, 217–222. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Chemical structures of compounds 15.
Figure 1. Chemical structures of compounds 15.
Marinedrugs 17 00649 g001
Figure 2. 1H-1H COSY correlations and key HMBC correlations for compounds 14.
Figure 2. 1H-1H COSY correlations and key HMBC correlations for compounds 14.
Marinedrugs 17 00649 g002
Figure 3. Energy-minimized 3D models of 1 and 4 with selected NOESY correlations.
Figure 3. Energy-minimized 3D models of 1 and 4 with selected NOESY correlations.
Marinedrugs 17 00649 g003
Figure 4. Experimental and calculated electronic circular dichroism (ECD) spectra of compounds 14.
Figure 4. Experimental and calculated electronic circular dichroism (ECD) spectra of compounds 14.
Marinedrugs 17 00649 g004
Figure 5. Reactions of compounds 2 and 3 with Mosher esters.
Figure 5. Reactions of compounds 2 and 3 with Mosher esters.
Marinedrugs 17 00649 g005
Table 1. 1H NMR data (400 MHz, δ in ppm, J in Hz) for 14.
Table 1. 1H NMR data (400 MHz, δ in ppm, J in Hz) for 14.
Position1 a2 b3 b4 b
26.55, dd (8.0, 0.8)6.35, dd (8.0, 0.8)6.35, dd (8.0, 0.8)6.81, d (8.0)
37.04, dd (8.0, 8.0)6.90, dd (8.0, 8.0)6.94, dd (8.0, 8.0)7.24, dd (8.0, 8.0)
46.58, dd (8.0, 0.8)6.26, dd (8.0, 0.8)6.29, dd (8.0, 0.8)7.06, d (8.0)
56.52, dd (8.4, 1.2)
67.05, dd (8.4, 8.4)
76.50, dd (8.4, 1.2)6.18, d (8.4)6.16, d (8.4)6.05, d (8.4)
8 6.69, d (8.4)6.57, d (8.4)6.22, d (8.4)
93.99, dd (12.4, 4.0)4.52, dd (10.4, 8.0)4.50, dd (7.6, 1.2)4.63, m
10a1.79, td (12.4, 6.4)1.49, td (13.6, 10.4)1.76, ddd (13.6, 7.6, 4.4)1.81, m
10b3.17, ddd (13.2, 4.0, 1.6)2.44, ddd (13.6, 8.0, 1.6)2.06, td (13.6, 1.2)2.22, m
114.64, m4.04, m3.82, m1.74, m
121.39, d (6.4)1.28, d (6.4)1.25, d (6.4)4.72, m
14 3.11, t (7.2)3.13, t (7.2)3.12, t (7.2)
15 1.71, m1.73, m1.72, m
16 0.98, t (7.2)0.99, t (7.2)0.99, t (7.2)
17 3.57, s
a DMSO-d6 b CD3OD.
Table 2. 13C NMR data (100 MHz, δ in ppm) for 14.
Table 2. 13C NMR data (100 MHz, δ in ppm) for 14.
Position1 a2 b3 b4 b
1156.9, C159.7, C159.9, C141.6, C
2107.0, CH109.3, CH108.7, CH110.8, CH
3127.8, CH128.2, CH128.6, CH128.3, CH
4110.4, CH108.7, CH107.4, CH123.3, CH
4a152.3, C157.3, C158.3, C158.3, C
4b150.1, C161,9, C161,9, C162.1, C
5110.2, CH110.7, C110.9, C110.9, C
6127.9, CH157.3, C159.9, C159.7, C
7106.5, CH107.0, CH106.0, CH105.8, CH
8154.3, C134.4, CH136.6, CH135.6, CH
8a109.3, C126.2, C125.3, C125.4, C
8b109.1, C114.3, C111.9, C128.9, C
922.3, CH31.8, CH30.5, CH32.3, CH
1033.0, CH240.4, CH236.1, CH223.9, CH2
1171.7, CH73.4, CH68.7, CH28.2, CH2
1220.7, CH321.6, CH321.7, CH367.9, CH
13 209.7, C209.8, C209.9, C
14 47.6, CH247.6, CH247.2, CH2
15 19.1, CH219.1, CH219.2, CH2
16 14.3, CH314.3, CH314.3, CH3
17 59.9, OCH3
a DMSO-d6 b CD3OD.
Table 3. Biological activities of 15.
Table 3. Biological activities of 15.
CompoundsIC50 (µg/mL)LC50 (µg/mL)
Helicoverpa armigera HubnerCulex quinquefasciatus
1>20038.5 (±1.16)
2100>80
320011.6 (±0.58)
4>20023.5 (±1)
5>200>80
Azadirachtin a258.8 (±0.58)
a Azadirachtin was used as a positive control.

Share and Cite

MDPI and ACS Style

Bai, M.; Zheng, C.-J.; Nong, X.-H.; Zhou, X.-M.; Luo, Y.-P.; Chen, G.-Y. Four New Insecticidal Xanthene Derivatives from the Mangrove-Derived Fungus Penicillium sp. JY246. Mar. Drugs 2019, 17, 649. https://0-doi-org.brum.beds.ac.uk/10.3390/md17120649

AMA Style

Bai M, Zheng C-J, Nong X-H, Zhou X-M, Luo Y-P, Chen G-Y. Four New Insecticidal Xanthene Derivatives from the Mangrove-Derived Fungus Penicillium sp. JY246. Marine Drugs. 2019; 17(12):649. https://0-doi-org.brum.beds.ac.uk/10.3390/md17120649

Chicago/Turabian Style

Bai, Meng, Cai-Juan Zheng, Xu-Hua Nong, Xue-Ming Zhou, You-Ping Luo, and Guang-Ying Chen. 2019. "Four New Insecticidal Xanthene Derivatives from the Mangrove-Derived Fungus Penicillium sp. JY246" Marine Drugs 17, no. 12: 649. https://0-doi-org.brum.beds.ac.uk/10.3390/md17120649

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