Next Article in Journal
Contribution of Glycation and Oxidative Stress to Thyroid Gland Pathology—A Pilot Study
Previous Article in Journal
Short- and Long-Term Social Recognition Memory Are Differentially Modulated by Neuronal Histamine
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Chemical Constituents of the Egg Cases of Tenodera angustipennis (Mantidis ootheca) with Intracellular Reactive Oxygen Species Scavenging Activity

1
Herbal Medicine Resources Research Center, Korea Institute of Oriental Medicine (KIOM), Naju 58245, Korea
2
Department of Plant Biotechnology, College of Life Sciences and Biotechnology, Korea University, Seoul 02841, Korea
*
Author to whom correspondence should be addressed.
Biomolecules 2021, 11(4), 556; https://doi.org/10.3390/biom11040556
Submission received: 30 March 2021 / Revised: 7 April 2021 / Accepted: 8 April 2021 / Published: 10 April 2021
(This article belongs to the Section Natural and Bio-inspired Molecules)

Abstract

:
As a traditional medicine with potential antioxidant effects, Tenodera angustipennis egg cases (Mantidis ootheca) are a potential source of new bioactive substances. Herein, three new N-acetyldopamine derivatives, namely, (+)-tenoderin A (1a), (−)-tenoderin A (1b), and tenoderin B (2), along with thirteen known compounds (315), were isolated from a 70% EtOH extract of T. angustipennis egg cases. Compound 1 was isolated as a racemic mixture, and two enantiomers (1a and 1b) were successfully separated by chiral-phase preparative HPLC. The chemical structures of the new compounds were established by NMR spectroscopy and high-resolution electrospray ionization mass spectrometry, and the absolute configurations of enantiomers 1a and 1b were determined by electronic circular dichroism spectroscopy. All the new compounds exhibited antioxidant activities with IC50 values of 19.45–81.98 μM, as evaluated using free-radical scavenging assays, with the highest activity observed for compound 2. In addition, compounds 1a, 1b, and 2 exhibited inhibitory activities on intracellular reactive oxygen species generation.

1. Introduction

Mantidis ootheca, which refers to egg cases of the Mantidae family, such as Tenodera angustipennis Saussure, Hierodula patellifera Serville, Statilia maculate Thunberg, and Tenodera sinensis Saussure, has been used as a source of traditional medicines in East Asia, including Korea and China [1]. Mantidis ootheca has been used in traditional medicines to treat incontinence, lumbago, spermatorrhea, acidosis, renal failure, and leukorrhea [2], and has been reported to have various biological activities, including anti-inflammatory, antidiuretic, anticancer, and antioxidant activities, as well as vascular relaxant effects [3,4,5,6]. Although various pharmacological studies on Mantidis ootheca have been reported, only a few studies have investigated its chemical constituents. The chemical constituents of Mantidis ootheca have been revealed to include fatty acids and phenols that exhibit pharmacological properties, such as antibacterial, antibiofilm, antioxidant, and anti-atherosclerotic effects [7,8,9]. Notably, previous studies have indicated that commercially available Mantidis ootheca products contain the egg cases of several mantis species within one package [10]. As such mixtures may be used for research, the exact identification of research samples is of particular importance.
As part of our continuing research on the discovery of new bioactive substances in Mantidis ootheca [10,11], the egg cases of T. angustipennis, which constitute a major raw material within Mantidis ootheca product mixtures, were selected for this study through sample identification. Using a 70% EtOH extract of T. angustipennis egg cases, which was selected as an antioxidant resource through our screening system, three new N-acetyldopamine derivatives (1a, 1b, and 2) and thirteen known compounds (3–15) were isolated. Compound 1 was isolated as an enantiomeric mixture, and two stereoisomers (1a and 1b) were successfully separated by chiral-phase preparative HPLC. Their chemical structures were identified using spectroscopic/spectrometric techniques including NMR spectroscopy, electronic circular dichroism (ECD) spectroscopy, and high-resolution electrospray ionization mass spectrometry (HRESIMS). To assess the antioxidant activities of the isolated compounds, their free-radical 1,1-diphenyl-2-picrylhydrazyl (DPPH) and 2,2′-azino-bis-3-ethylbenzthiazoline-6-sulfonic acid (ABTS) scavenging activities and their inhibitory activity on intracellular reactive oxygen species (ROS) generation were evaluated.

2. Materials and Methods

2.1. General Experimental Procedures

UV, optical rotation, and ECD spectra were recorded using a Pop UV-vis spectrophotometer (Mecasys, Daejeon, Korea), a P-2000 polarimeter (Jasco, Tokyo, Japan), and a J-1100 spectrometer (Jasco), respectively. LC/MS and HRESIMS data were acquired using a UPLC Q-TOF MS spectrometer (Waters, Milford, MA, USA). GC/MS data were recorded using an Agilent 6890/JMS-700 system (Agilent, Palo Alto, CA, USA). NMR data were acquired using a 500 MHz NMR spectrometer (Bruker, Karlsruhe, Germany). MPLC was performed using a Biotage Selekt system (Biotage AB, Uppsala, Sweden). Preparative HPLC was performed using a Waters system with an YMC-Pack ODS-A column (5 µm, 250 × 20 mm I.D.) and a Chiralpak IH column (5 µm, 250 × 4.6 mm I.D.).

2.2. Insect Material

The insect material (Mantidis ootheca) was purchased from a medicinal herb company (Gwangmyeongdang Co., Ulsan, Korea) in July 2019, authenticated, and deposited in the Korean Herbarium of Standard Herbal Resources (Index Herbarium code KIOM, specimen no. 2-20-0223) at the Korea Institute of Oriental Medicine. To prepare the sample for this study, the egg cases of T. angustipennis were identified and selected based on key morphological characteristics, such as shape, color, texture, and angle of the distal end [10]. All materials had a fusiform shape, a tapered distal end with angles of 22°–50°, and a lusterless and brownish textured external wall (Figure S16).

2.3. Extraction and Isolation

Dried T. angustipennis egg cases (900.0 g) were ground and extracted with 70% EtOH (3 × 10.0 L) at room temperature. After the solvent was evaporated, the residue was suspended in distilled water (4.0 L) and extracted using EtOAc (3 × 4.0 L) to obtain the EtOAc-soluble extract (23.6 g). The EtOAc-soluble extract (20.0 g) was fractionated by MPLC using Diaion HP20 (Mitsubishi Chemical, Tokyo, Japan) cartridges (400 g, H2O–MeOH, 100:0 to 0:100 in 18 CV, 40 mL/min) to produce six fractions (Fr 1–6). Fr 2 (1368.3 mg) was fractionated by MPLC using Sfär Silica HC (Biotage AB, Uppsala, Sweden) cartridges (200 g, n-hexane–EtOAc, 80:20 to 0:100 in 30 CV, 30 mL/min) to produce nine sub-fractions (Fr 2.1–2.9). Fr 2.2 (3.7 mg) was purified using preparative HPLC (H2O–ACN, 80:20 to 40:60 in 50 min, 6 mL/min) to yield 5 (1.2 mg). Fr 2.4 (44.3 mg) was purified using preparative HPLC (H2O–ACN, 85:15 to 70:30 in 50 min, 6 mL/min) to yield 6 (23.4 mg). Fr 2.5 (43.6 mg) was purified using preparative HPLC (H2O–ACN, 85:15 to 70:30 in 50 min, 6 mL/min) to yield 10 (14.4 mg). Fr 2.6 (34.9 mg) was purified using preparative HPLC (H2O–ACN, 85:15 to 70:30 in 50 min, 6 mL/min) to yield 9 (15.5 mg), 11 (3.2 mg), 12 (1.8 mg), and 13 (2.4 mg). Fr 2.9 (1648.3 mg) was fractionated by MPLC using Sephadex LH20 (Pharmacia Fine Chemicals Inc., Piscataway, NJ, USA) cartridges (120 g, H2O–MeOH, 100:0 to 0:100 in 16 CV, 20 mL/min) to produce nine sub-fractions (Fr 2.9.1–2.9.9). Fr 2.9.3 (114.6 mg) was purified using preparative HPLC (H2O–ACN, 90:10 to 80:20 in 50 min, 6 mL/min) to yield 3 (39.8 mg) and 4 (40.1 mg). Fr 2.9.5 (55.1 mg) was purified using preparative HPLC (H2O–ACN, 85:15 to 75:25 in 50 min, 6 mL/min) to yield 1 (6.3 mg). Racemic mixture 1 was separated by preparative HPLC (n-hexane–EtOH–MeOH, 70:20:10 isocratic in 15 min, 1 mL/min) with a chiral-phase stationary column (Chiralpak IH) to yield 1a (2.0 mg, tR 4.8 min) and 1b (2.1 mg, tR 11.7 min). Fr 4 (863.9 mg) was fractionated by MPLC using Sfär Silica HC cartridges (100 g, CHCl3–MeOH, 100:0 to 50:50 in 40 CV, 70 mL/min) to produce seven sub-fractions (Fr 4.1–4.7). Fr 4.1 (53.5 mg) was purified using preparative HPLC (H2O–ACN, 80:20 to 60:40 in 50 min, 6 mL/min) to yield 7 (6.9 mg) and 8 (11.2 mg). Fr 4.4 (198.5 mg) was purified using preparative HPLC (H2O–ACN, 80:20 to 70:30 in 50 min, 6 mL/min) to yield 2 (14.6 mg). Fr 5 (2176.2 mg) was fractionated by MPLC using Sfär Silica HC cartridges (200 g, CHCl3–MeOH, 100:0 to 50:50 in 21 CV, 80 mL/min) to produce five sub-fractions (Fr 5.1–5.5). Fr 5.3 (473.7 mg) was fractionated by MPLC using Sephadex LH20 cartridges (120 g, H2O–MeOH, 100:0 to 0:100 in 17 CV, 30 mL/min) to produce nine sub-fractions (Fr 5.3.1–5.3.7). Fr 5.3.2 (32.7 mg) was purified using preparative HPLC (H2O–ACN, 75:25 to 65:35 in 50 min, 6 mL/min) to yield 14 (1.6 mg). Fr 6 (9810.0 mg) was purified by MPLC using Sfär Silica HC cartridges (350 g, n-hexane–CHCl3–MeOH, 100:0:0 to 50:50:0, 0:100:0 to 0:0:100 in 20 CV, 50 mL/min) to yield 15 (2012.2 mg).

2.3.1. Tenoderin A (1)

Brown gum; 1H and 13C NMR (500 and 125 MHz, CD3OD), see Table 1; ESIMS (positive) m/z 403 [M + H]+; ESIMS (negative) m/z 401 [M − H]; HRESIMS m/z 401.1349 [M − H] (calcd for C20H21N2O7, 401.1349).

(+)-tenoderin A (1a)

Brown gum; [α]26D +21.2 (c 0.01, MeOH); UV (MeOH) λmax (log ε) 215 (3.91), 285 (3.37), 332 (3.23) nm; ECD (c 0.5 mM, MeOH) Δε +33.9 (212), −19.5 (248), +20.1 (282), +7.6 (302), +11.3 (317).

(−)-tenoderin A (1b)

Brown gum; [α]26D −21.0 (c 0.01, MeOH); UV (MeOH) λmax (log ε) 215 (3.91), 285 (3.36), 332 (3.25) nm; ECD (c 0.5 mM, MeOH) Δε −36.6 (212), +8.1 (247), −25.1 (284), −12.6 (302), −13.1 (317).

2.3.2. Tenoderin B (2)

Brown gum; UV (MeOH) λmax (log ε) 215 (3.95), 285 (3.91) nm; 1H and 13C NMR (500 and 125 MHz, CD3OD), see Table 1; ESIMS (positive) m/z 316 [M + H]+; ESIMS (negative) m/z 314 [M − H]; HRESIMS m/z 314.1042 [M − H] (calcd for C17H16NO5, 314.1028).

2.4. Computational Methods

ECD calculations were performed for 1a and 1b. A conformer distribution was constructed using the Spartan’14 software (Wavefunction, Inc., Irvine, CA, USA) with a Merck molecular force field. The conformers were optimized at the DFT [B3LYP/6-31+G(d,p)] level, and ECD calculations were performed at the TDDFT (CAM-B3LYP/SVP) level with a CPCM solvent model in MeOH using the Gaussian 09 software (Gaussian, Inc., Wallingford, CT, USA).

2.5. DPPH and ABTS Radical Scavenging Activities

The DPPH (Sigma-Aldrich, St. Louis, MO, USA) and ABTS (Sigma-Aldrich) free-radical scavenging activities were measured according to previous studies [12,13]. All measurements were independently repeated three times. The DPPH and ABTS scavenging activities were calculated as follows: DPPH or ABTS scavenging activity (%) = [A(free radical) − A(standard)]/A(free radical), where A is the absorbance value of DPPH at 517 nm or ABTS at 734 nm.

2.6. Detection of Intracellular ROS

The HUVECs used in the ROS scavenging experiments were purchased from the American Type Culture Collection (ATCC, Manassas, VA, USA). The cells were cultured in α-MEM medium, and 20% FBS in a humid atmosphere of 5% CO2 at 37 °C.
To confirm the antioxidant effect of each compound against oxidative stress, the cells were treated with each sample at a concentration of 5, 10, 50 or 100 µM and hydrogen peroxide (1 mM), cultured for 24 h, washed with PBS, and treated with 100 μM DCFDA (ThermoFisher, Waltham, MA, USA) for 30 min at 37 °C. Reacted. After DCFDA was removed, and the amount of DCFDA remaining in the cells was observed using a fluorescence microscope. The fluorescence intensity at an excitation wavelength of 485 nm and an emission wavelength of 520 nm was measured using a microplate reader.

3. Results and Discussion

Dried T. angustipennis egg cases selected from a commercial Mantidis ootheca material were ground and extracted with 70% EtOH. Following fractionation, fifteen compounds were isolated from the extract and structural determination was performed (Figure 1).
Compound 1 was obtained as a brown gum, and HRESIMS analysis revealed its elemental formula to be C20H22N2O7, suggesting eleven degrees of unsaturation. The 1H NMR data indicated the presence of two methyl groups (δH 1.80 (3H, s, CH3-10) and 2.02 (3H, s, CH3-10′)), two methylene groups (δH 2.55 (2H, t, J = 7.2 Hz, H-7) and 3.24 (2H, t, J = 7.2 Hz, H-8)), a methine group (δH 6.66 (1H, s, H-8′)), two meta-coupled aromatic methine groups (δH 6.61 (1H, d, J = 1.8 Hz, H-2) and 6.39 (1H, d, J = 1.9 Hz, H-6)), and three aromatic methine groups in an ABX spin system (δH 6.72 (1H, d, J = 8.4 Hz, H-5′), 7.44 (1H, d, J = 2.0 Hz, H-2′), and 7.47 (1H, dd, J = 8.4, 2.1 Hz, H-6′)) (Table 1). The 13C NMR data corresponded to 20 carbons, including two methyl carbons, two methylene carbons, six methine carbons, three carbonyl carbons, and seven additional quaternary carbons (Table 1). The 1H and 13C NMR data for compound 1 were similar to those for the dimeric N-acetyldopamine derivative polyrhadopamine B [14], except that the 2-oxo-N-acetyldopamine (4) unit linked at C-6 in polyrhadopamine B was linked at C-5 in compound 1. This result was supported by the presence of two doublet proton signals with meta coupling at H-2 (δH 6.61) and H-6 (δH 6.39), as well as the HMBC correlations between H-6/C-8′ (δC 54.4), C-7 (δC 36.0), C-2 (δC 116.8), and C-4 (δC 142.8) and those between H-8′/C-6 (δC 120.3), C-4 (δC 142.8), C-5 (δC 125.3), C-9′ (δC 173.2), and C-7′ (δC 196.3) (Figure 2).
Compound 1 is a racemic mixture, which was suggested by the lack of an optical rotation value. Therefore, chiral separation was performed on a Chiralpak IH column to yield optically pure enantiomers 1a (+21.2 (c 0.01, MeOH)) and 1b (−21.0 (c 0.01, MeOH)) (Figure 3). The absolute configurations of 1a and 1b were determined by comparing the calculated and experimental ECD spectra. The calculated ECD spectra of the 8′R and 8′S stereoisomers were in good agreement with the experimental spectra of 1a and 1b, respectively (Figure 4). Therefore, the structures of new compounds 1a and 1b were assigned as (+)-tenoderin A and (−)-tenoderin A, respectively.
Compound 2 was obtained as a brown gum, and its elemental formula was determined to be C17H17NO5 using HRESIMS analysis. The 1D NMR data revealed the presence of a N-acetyldopamine (3) moiety as well as a 1,4-disubstituted benzene ring (δH 7.68 (2H, d, J = 8.2 Hz, CH3-2′ and 6′) and 6.84 (2H, d, J = 8.3 Hz, CH3-3′ and 5′); δC 134.3 (C-2′ and 6′) and 116.2 (C-3′ and 5′)) and a ketone group (δC 199.1 (C-7′)) (Table 1). A detailed analysis of the NMR data showed that compound 2 was similar to periplanetol A [15], except for the absence of a hydroxy group at C-5′ and the presence of an aromatic methine signal (δH 6.84, H-5′) for compound 2. These results are supported by the COSY correlations between H-5′/H-6′ and the HMBC correlations between H-5′/ C-4′ (δC 164.0), C-3′ (δC 116.2), and C-1′ (δC 131.3) (Figure 2). Accordingly, the structure of new compound 2 was elucidated, and this compound was given the trivial name tenoderin B.
In addition to the above-described new compounds, thirteen known compounds (3–15) were isolated and identified as N-acetyldopamine (3), 2-oxo-N-acetyldopamine (4) [16], 4-hydroxybenzaldehyde (5) [17], 4-hydroxybenzoic acid (6) [18], apocynin (7) [19], benzoic acid (8) [20], protocatechuic acid (9) [21], 4-hydroxyphenylacetic acid (10) [22], (S)-1-phenylenthane-1,2-diol (11) [23], 4-hydroxyphenylglyoxylic acid amide (12) [24], (±)-hydroxybutenolide (13) [25], scoparone (14) [26], and oleic acid (15) (Figure 1). Notably, many of the known compounds (4–14) were first isolated from Mantidis ootheca.
The antioxidant activity of the extract, as measured using DPPH and ABTS radical scavenging assays, was significant (81.99 ± 1.98% and 99.74% ± 0.13%, respectively, at 100 μg/mL) (Table S1). Furthermore, all the isolated compounds were screened for antioxidant activity at a concentration of 100 μM using DPPH and ABTS radical scavenging assays. DPPH and ABTS radical scavenging activities were observed for compounds 1a, 1b, 2, 3, 4, and 9 (Table S1). In particular, new compounds 1a, 1b, and 2 displayed antioxidant activities with IC50 values between 19.45 and 81.98 μM. Among the new compounds, the antioxidant activity of compound 2 was the highest. In addition, although compounds 1a and 1b were enantiomers, there was a difference in their antioxidant activities (Table 2). Compound 1b exhibited stronger antioxidant effects than 1a, and the ratios of their IC50 values were 1:1.8 (DPPH) and 1:1.3 (ABTS). These results suggest that the difference in antioxidant activity between the two enantiomers is due to the chirality of C-8′.
In addition, to determine whether the antioxidant activities were associated with protective effects in H2O2-treated human umbilical vein/vascular endothelium cells (HUVECs), intracellular 2′,7′-dichlorofluorescin diacetate (DCFDA) levels were measured. Oxidative-stress-induced cell damage has been implicated in various types of disease. H2O2 is a major ROS produced intracellularly during pathological processes and causes oxidative injury. H2O2 has been extensively used as an inducer of oxidative stress for in vitro models [27]. Therefore, H2O2 was selected to promote oxidative stress in the current investigation. In this study, as shown in Figure 5, H2O2 significantly increased the intracellular DCFDA level. In this assay system, compound 2 showed a potent antioxidant activity in a dose-dependent manner and compounds 1a and 1b also showed significant antioxidant activity, which presented the identical activity pattern supporting the results in the free-radical scavenging assay.

4. Conclusions

In conclusion, this study on the chemical constituents of T. angustipennis egg cases (Mantidis ootheca) revealed sixteen compounds, including three new N-acetyldopamine derivatives (1a, 1b, and 2) and thirteen known compounds (3–15). Two enantiomers (1a and 1b) were successfully separated by chiral-phase preparative HPLC, and the absolute configurations were determined by ECD spectroscopy. All the isolated compounds were evaluated for antioxidant activity, and the new compounds appeared to show antioxidant effects in HUVECs. These findings not only reveal various chemical constituents in the egg cases of T. angustipennis, they also provide guidance for clarifying the pharmacodynamic basis of the antioxidant effects of Mantidis ootheca.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/biom11040556/s1, Figure S1: Structures of known compounds (315), Figure S2: Spectroscopic data of known compounds (314), Figure S3: GC-MS spectrum of compound 15, Figure S4: 1H NMR spectrum of tenoderin A (1) (CD3OD, 500 MHz), Figure S5: 13C NMR spectrum of tenoderin A (1) (CD3OD, 125 MHz), Figure S6: HSQC NMR spectrum of tenoderin A (1) (CD3OD), Figure S7: HMBC NMR spectrum of tenoderin A (1) (CD3OD), Figure S8: COSY NMR spectrum of tenoderin A (1) (CD3OD), Figure S9: HRESIMS spectrum of tenoderin A (1) (CD3OD), Figure S10: 1H NMR spectrum of tenoderin B (2) (CD3OH, 500 MHz), Figure S11: 13C NMR spectrum of tenoderin B (2) (CD3OH, 125 MHz), Figure S12: HSQC NMR spectrum of tenoderin B (2) (CD3OH), Figure S13: HMBC NMR spectrum of tenoderin B (2) (CD3OH), Figure S14: COSY NMR spectrum of tenoderin B (2) (CD3OH), Figure S15: HRESIMS spectrum of tenoderin B (2), Figure S16: Stereomicroscope micrographs showing the ootheca morphology of Tenodera angustipennis. (A) Dorsal view; (B) Lateral view; (C) Surface pattern on lateral view; Scale bars = 1 cm (A, B). 1 mm (C), Table S1: Screening of the antioxidant activity of extract (100 μg/mL) and isolated compounds (100 μM).

Author Contributions

S.M.R. and J.L. conceived and designed the experiments; S.M.R., H.-h.N., J.S.K., and J.-h.S. performed the experiments; S.M.R., Y.H.S., H.S.K., A.Y.L., W.J.K., and D.L. analyzed the data; B.C.M. and J.L. reviewed and edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by a grant on the Development of Sustainable Application for Standard Herbal Resources (KSN2012320) from the Korea Institute of Oriental Medicine, Korea.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

In this section, you can acknowledge any support given which is not covered by the author contribution or funding sections. This may include administrative and technical support, or donations in kind (e.g., materials used for experiments).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Defining Dictionary for Medicinal Herbs. Korea Institute of Oriental Medicine. Available online: http://boncho.kiom.re.kr/codex (accessed on 4 April 2019).
  2. Kim, C.M.; Shin, M.K.; Ahn, D.K.; Lee, K.S. Chinese Medicine Dictionary; Jungdam: Seoul, Korea, 2006; Volume 5, pp. 2184–2186. [Google Scholar]
  3. Tan, Z.; Lei, Y.; Zhang, B.; Huang, L. Comparison of Pharmacological Studies on Ootheca Mantidis. China J. Chin. Mater. Med. 1997, 22, 496–499. [Google Scholar]
  4. Kim, H.Y.; Lee, Y.J.; Han, B.H.; Yoon, J.J.; Ahn, Y.M.; Hong, M.H.; Tan, R.; Kang, D.G.; Lee, H.S. Mantidis Ootheca Induces Vascular Relaxation through Pi3k/Akt-Mediated Nitric Oxid-Cyclic Gmp-Protein Kinase G Signaling in Endothelial Cells. J. Physiol. Pharmacol. 2017, 68, 215–221. [Google Scholar] [PubMed]
  5. Ahn, M.Y.; Ryu, K.S.; Lee, Y.W.; Kim, Y.S. Cytotoxicity and L-Amino Acid Oxidase Activity of Crude Insect Drugs. Arch. Pharmacal Res. 2000, 23, 477–481. [Google Scholar] [CrossRef] [PubMed]
  6. Park, G.; Moon, B.C.; Lim, H. Effects of 14 Chung-Bu Medicinal Materials Described in the Dongui Bogam on Inflammatory Cytokines Production in Hacat Keratinocytes. J. Soc. Cosmet. Sci. Korea 2020, 46, 195–204. [Google Scholar]
  7. Wang, W.; Zhang, N.; Chanda, W.; Liu, M.; UD Din, S.R.; Diao, Y.; Liu, L.; Cao, J.; Wang, X.; Li, X. Antibacterial and Anti-Biofilm Activity of the Lipid Extract from Mantidis Ootheca on Pseudomonas aeruginosa. J. Zhejiang Univ. Sci. B 2018, 19, 364–371. [Google Scholar] [CrossRef]
  8. Xu, M. Study of the Anti-Dpph Free Radical Component from Mantidis Ootheca Extracts. J. Anhui Agric. Univ. 2014, 2014, 11619–11620. [Google Scholar]
  9. Xu, M. Anti-Atherosclerotic Activities of Two Compounds from Mantidis Ootheca. J. Anhui Agric. Univ. 2012, 2012, 15722–15723. [Google Scholar]
  10. Song, J.; Cha, J.; Moon, B.C.; Kim, W.J.; Yang, S.; Choi, G. Mantidis Oötheca (Mantis Egg Case) Original Species Identification Via Morphological Analysis and DNA Barcoding. J. Ethnopharmacol. 2020, 252, 112574–112584. [Google Scholar] [CrossRef]
  11. Lim, H.; Seo, Y.S.; Ryu, S.M.; Moon, B.C.; Choi, G.; Kim, J. Two-Week Repeated Oral Dose Toxicity Study of Mantidis Ootheca Water Extract in C57bl/6 Mice. J. Evid. Based Complementary Altern. Med. 2019, 2019, 6180236. [Google Scholar] [CrossRef]
  12. Kedare, S.B.; Singh, R.P. Genesis and Development of Dpph Method of Antioxidant Assay. J. Food Sci. Technol. 2011, 48, 412–422. [Google Scholar] [CrossRef] [Green Version]
  13. Rajurkar, N.S.; Hande, S.M. Estimation of Phytochemical Content and Antioxidant Activity of Some Selected Traditional Indian Medicinal Plants. Indian J. Pharm. Sci. 2011, 73, 146–151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Tang, J.; Zhang, L.; Jiang, L.; Di, L.; Yan, Y.; Tu, Z.; Yang, C.; Zuo, Z.; Hou, B.; Xia, H.; et al. Dopamine Derivatives from the Insect Polyrhachis Dives as Inhibitors of Rock1/2 and Stimulators of Neural Stem Cell Proliferation. Tetrahedron 2014, 70, 8852–8857. [Google Scholar] [CrossRef]
  15. Bai, H.; Li, Y.; Qin, F.; Yan, Y.; Wang, S.; Zhang, H.; Cheng, Y. Periplanetols A–F, Phenolic Compounds from Periplaneta Americana with Potent Cox-2 Inhibitory Activity. Fitoterapia 2020, 143, 104589. [Google Scholar] [CrossRef]
  16. Noda, N.; Kubota, S.; Miyata, Y.; Miyahara, K. Optically Active N-Acetyldopamine Dimer of the Crude Drug “Zentai,” the Cast-Off Shell of the Cicada Cryptotympana sp. Chem. Pharm. Bull. 2000, 48, 1749–1752. [Google Scholar] [CrossRef] [Green Version]
  17. Kim, H.; Ralph, J.; Lu, F.; Ralph, S.A.; Boudet, A.M.; MacKay, J.J.; Sederoff, R.R.; Ito, T.; Kawai, S.; Ohashi, H. Nmr Analysis of Lignins in Cad-Deficient Plants. Part 1. Incorporation of Hydroxycinnamaldehydes and Hydroxybenzaldehydes into Lignins. Org. Biomol. Chem. 2003, 1, 268–281. [Google Scholar] [CrossRef] [PubMed]
  18. Jia, J.; Jiang, C.; Zhang, X.; Jiang, Y.; Ma, D. Cui-Catalyzed Hydroxylation of Aryl Bromides under the Assistance of 5-Bromo-2-(1h-Imidazol-2-Yl) Pyridine and Related Ligands. Tetrahedron Lett. 2011, 52, 5593–5595. [Google Scholar] [CrossRef]
  19. Luo, J.; Ma, Q.; Zhao, Y.; Yi, T.; Li, C.; Zhou, J. Palaeophytochemical Components from the Miocene-Fossil Wood of Pinus Griffithii. J. Chin. Chem. Soc. 2009, 56, 600–605. [Google Scholar] [CrossRef]
  20. Bonaparte, A.C.; Betush, M.P.; Panseri, B.M.; Mastarone, D.J.; Murphy, R.K.; Murphree, S.S. Novel Aerobic Oxidation of Primary Sulfones to Carboxylic Acids. Org. Lett. 2011, 13, 1447–1449. [Google Scholar] [CrossRef] [PubMed]
  21. Matsui, T.; Kudo, A.; Tokuda, S.; Matsumoto, K.; Hosoyama, H. Identification of a New Natural Vasorelaxatant Compound,(+)-Osbeckic Acid, from Rutin-Free Tartary Buckwheat Extract. J. Agric. Food Chem. 2010, 58, 10876–10879. [Google Scholar] [CrossRef] [PubMed]
  22. Milne, J.E.; Storz, T.; Colyer, J.T.; Thiel, O.R.; Dilmeghani Seran, M.; Larsen, R.D.; Murry, J.A. Iodide-Catalyzed Reductions: Development of a Synthesis of Phenylacetic Acids. J. Org. Chem. 2011, 76, 9519–9524. [Google Scholar] [CrossRef]
  23. Chang, L.; Ouyang, L.M.; Xu, Y.; Pan, J.; Xu, J.H. Highly Enantioselective Hydrolysis of Phenyl-1, 2-Ethanediol Cyclic Carbonates by Newly Isolated Bacillus Sp. Ecu0015. J. Mol. Catal. B Enzym. 2010, 66, 95–100. [Google Scholar] [CrossRef]
  24. Matsuki, Y.; Ito, T.; Komatsu, S.; Nambara, T. Studies on the Metabolism of Atenolol. Characterization and Determination of a New Urinary Metabolite in the Rat. Chem. Pharm. Bull. 1982, 30, 196–201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Rasmussen, S.; Wolff, C.; Rudolph, H. Compartmentalization of Phenolic Constituents in Sphagnum. Phytochemistry 1995, 38, 35–39. [Google Scholar] [CrossRef]
  26. Fillion, E.; Dumas, A.M.; Kuropatwa, B.A.; Malhotra, N.R.; Sitler, T.C. Yb (Otf) 3-Catalyzed Reactions of 5-Alkylidene Meldrum’s Acids with Phenols: One-Pot Assembly of 3, 4-Dihydrocoumarins, 4-Chromanones, Coumarins, and Chromones. J. Org. Chem. 2006, 71, 409–412. [Google Scholar] [CrossRef] [PubMed]
  27. Coyle, C.H.; Martinez, L.J.; Coleman, M.C.; Spitz, D.R.; Weintraub, N.L.; Kader, K.N. Mechanisms of H2o2-Induced Oxidative Stress in Endothelial Cells. Free Radic. Biol. Med. 2006, 40, 2206–2213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. All isolated compounds from a 70% EtOH extract of T. angustipennis egg cases.
Figure 1. All isolated compounds from a 70% EtOH extract of T. angustipennis egg cases.
Biomolecules 11 00556 g001
Figure 2. Key HMBC (arrow) and COSY (bold) correlations (1 and 2).
Figure 2. Key HMBC (arrow) and COSY (bold) correlations (1 and 2).
Biomolecules 11 00556 g002
Figure 3. (A) UPLC chromatogram of compound 1 (stationary phase: ACQUITY UPLC BEH C18 column (2.1 mm × 100 mm, 1.7 μm); mobile phase: H2O–ACN, 95:5 to 50:50 in 5 min, 0.3 mL/min; UV 280 nm). (B) HPLC chromatogram of the enantiomeric mixture of compound 1 (stationary phase: Chiralpak IH column (4.6 mm × 250 mm, 5 μm); mobile phase: n-hexane–EtOH–MeOH, 70:20:10 isocratic in 15 min, 1 mL/min; UV 280 nm).
Figure 3. (A) UPLC chromatogram of compound 1 (stationary phase: ACQUITY UPLC BEH C18 column (2.1 mm × 100 mm, 1.7 μm); mobile phase: H2O–ACN, 95:5 to 50:50 in 5 min, 0.3 mL/min; UV 280 nm). (B) HPLC chromatogram of the enantiomeric mixture of compound 1 (stationary phase: Chiralpak IH column (4.6 mm × 250 mm, 5 μm); mobile phase: n-hexane–EtOH–MeOH, 70:20:10 isocratic in 15 min, 1 mL/min; UV 280 nm).
Biomolecules 11 00556 g003
Figure 4. Calculated and experimental electronic circular dichroism (ECD) spectra of 1a and 1b.
Figure 4. Calculated and experimental electronic circular dichroism (ECD) spectra of 1a and 1b.
Biomolecules 11 00556 g004
Figure 5. (A) Reactive oxygen species (ROS) scavenging activities as revealed by images of human umbilical vein/vascular endothelium cells (HUVECs) with 2′,7′-dichlorofluorescin diacetate (DCFDA) treated with 5–100 μM of compound 1a, 1b, or 2. (B) Scavenging percentages of the tested compounds. * p < 0.05, ** p < 0.01.
Figure 5. (A) Reactive oxygen species (ROS) scavenging activities as revealed by images of human umbilical vein/vascular endothelium cells (HUVECs) with 2′,7′-dichlorofluorescin diacetate (DCFDA) treated with 5–100 μM of compound 1a, 1b, or 2. (B) Scavenging percentages of the tested compounds. * p < 0.05, ** p < 0.01.
Biomolecules 11 00556 g005
Table 1. 1D NMR (1H and 13C) and 2D NMR (HMBC and COSY) data for compounds 1 and 2 in CD3OD.
Table 1. 1D NMR (1H and 13C) and 2D NMR (HMBC and COSY) data for compounds 1 and 2 in CD3OD.
Position12
δC, Type 2δH, Multi
(J in Hz) 1
COSYHMBCδC, TypeδH, Multi
(J in Hz)
COSYHMBC
1132.4, C 132.5, C
2116.8, CH6.61, d (1.8) 3, 4, 6, 7118.7, CH6.78, s 3, 4, 6, 7
3147.4, C 149.1, C
4142.8, C 144.1, C
5125.3, C 118.1, CH6.75, s 1, 3, 4, 7′
6120.3, CH6.39, d (1.9) 2, 4, 7, 8′131.7, C
736.0, CH22.55, t (7.2)81, 2, 6, 833.1, CH22.73, t (7.0)81, 2, 6, 8
842.2, CH23.24, t (7.2)71, 7, 942.8, CH23.30, overlap 371, 7, 9
9173.4, CO 173.4, CO
1022.6, CH31.80, s 922.7, CH31.83, s 9
1′128.5, C 131.3, C
2′116.7, CH7.44, d (2.0) 3′, 4′, 6′, 7′134.3, CH7.68, d (8.2)3′3′, 4′, 6′, 7′
3′146.5, C 116.2, CH6.84, d (8.3)2′1′, 4′, 5′
4′152.6, C 164.0, C
5′115.9, CH6.72, d (8.4)6′1′, 3′, 4′116.2, CH6.84, d (8.3)6′1′, 3′, 4′
6′123.9, CH7.47, dd (8.4, 2.1)5′2′, 4′, 7′134.3, CH7.68, d (8.2)5′2′, 4′, 5′, 7′
7′196.3, CO 199.1, CO
8′54.4, CH6.66, s 4, 5, 6, 7′, 9′
9′173.2, CO
10′22.4, CH32.02, s 9′
1 Measured at 500 MHz. 2 Measured at 125 MHz. 3 Overlap with NMR solvent (CD3OD).
Table 2. IC50 values for the antioxidant activities of new compounds 1a, 1b, and 2.
Table 2. IC50 values for the antioxidant activities of new compounds 1a, 1b, and 2.
SamplesIC50 (μM)
DPPHABTS
1a81.50 ± 0.77 181.98 ± 0.48
1b46.54 ± 0.5662.74 ± 0.69
219.45 ± 0.4237.23 ± 0.26
Gallic acid8.95 ± 0.2010.82 ± 0.97
1 Values are reported as mean ± SD (n = 3).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ryu, S.M.; Nam, H.-h.; Kim, J.S.; Song, J.-h.; Seo, Y.H.; Kim, H.S.; Lee, A.Y.; Kim, W.J.; Lee, D.; Moon, B.C.; et al. Chemical Constituents of the Egg Cases of Tenodera angustipennis (Mantidis ootheca) with Intracellular Reactive Oxygen Species Scavenging Activity. Biomolecules 2021, 11, 556. https://doi.org/10.3390/biom11040556

AMA Style

Ryu SM, Nam H-h, Kim JS, Song J-h, Seo YH, Kim HS, Lee AY, Kim WJ, Lee D, Moon BC, et al. Chemical Constituents of the Egg Cases of Tenodera angustipennis (Mantidis ootheca) with Intracellular Reactive Oxygen Species Scavenging Activity. Biomolecules. 2021; 11(4):556. https://doi.org/10.3390/biom11040556

Chicago/Turabian Style

Ryu, Seung Mok, Hyeon-hwa Nam, Joong Sun Kim, Jun-ho Song, Young Hye Seo, Hyo Seon Kim, A Yeong Lee, Wook Jin Kim, Dongho Lee, Byeong Cheol Moon, and et al. 2021. "Chemical Constituents of the Egg Cases of Tenodera angustipennis (Mantidis ootheca) with Intracellular Reactive Oxygen Species Scavenging Activity" Biomolecules 11, no. 4: 556. https://doi.org/10.3390/biom11040556

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