DOI QR코드

DOI QR Code

Two new triterpenoid saponins derived from the leaves of Panax ginseng and their antiinflammatory activity

  • Li, Fu (Key Laboratory of Mountain Ecological Restoration and Bioresource Utilization and Ecological Restoration Biodiversity Conservation Key Laboratory of Sichuan Province, Chengdu Institute of Biology, Chinese Academy of Sciences) ;
  • Cao, Yufeng (Jiangsu Collaborative Innovation Center of Regional Modern Agriculture & Environmental Protection/Jiangsu Key Laboratory for Eco-Agricultural Biotechnology Around Hongze Lake, Huaiyin Normal University) ;
  • Luo, Yanyan (Jiangsu Collaborative Innovation Center of Regional Modern Agriculture & Environmental Protection/Jiangsu Key Laboratory for Eco-Agricultural Biotechnology Around Hongze Lake, Huaiyin Normal University) ;
  • Liu, Tingwu (Jiangsu Collaborative Innovation Center of Regional Modern Agriculture & Environmental Protection/Jiangsu Key Laboratory for Eco-Agricultural Biotechnology Around Hongze Lake, Huaiyin Normal University) ;
  • Yan, Guilong (Jiangsu Collaborative Innovation Center of Regional Modern Agriculture & Environmental Protection/Jiangsu Key Laboratory for Eco-Agricultural Biotechnology Around Hongze Lake, Huaiyin Normal University) ;
  • Chen, Liang (Jiangsu Collaborative Innovation Center of Regional Modern Agriculture & Environmental Protection/Jiangsu Key Laboratory for Eco-Agricultural Biotechnology Around Hongze Lake, Huaiyin Normal University) ;
  • Ji, Lilian (Jiangsu Collaborative Innovation Center of Regional Modern Agriculture & Environmental Protection/Jiangsu Key Laboratory for Eco-Agricultural Biotechnology Around Hongze Lake, Huaiyin Normal University) ;
  • Wang, Lun (Key Laboratory of Mountain Ecological Restoration and Bioresource Utilization and Ecological Restoration Biodiversity Conservation Key Laboratory of Sichuan Province, Chengdu Institute of Biology, Chinese Academy of Sciences) ;
  • Chen, Bin (Key Laboratory of Mountain Ecological Restoration and Bioresource Utilization and Ecological Restoration Biodiversity Conservation Key Laboratory of Sichuan Province, Chengdu Institute of Biology, Chinese Academy of Sciences) ;
  • Yaseen, Aftab (Key Laboratory of Mountain Ecological Restoration and Bioresource Utilization and Ecological Restoration Biodiversity Conservation Key Laboratory of Sichuan Province, Chengdu Institute of Biology, Chinese Academy of Sciences) ;
  • Khan, Ashfaq A. (Department of Chemistry, Women University of Azad Jammu and Kashmir) ;
  • Zhang, Guolin (Key Laboratory of Mountain Ecological Restoration and Bioresource Utilization and Ecological Restoration Biodiversity Conservation Key Laboratory of Sichuan Province, Chengdu Institute of Biology, Chinese Academy of Sciences) ;
  • Jiang, Yunyao (Xiyuan Hospital, China Academy of Chinese Medical Sciences) ;
  • Liu, Jianxun (Xiyuan Hospital, China Academy of Chinese Medical Sciences) ;
  • Wang, Gongcheng (Department of Urology, The Affiliated Huaian No. 1 People's Hospital of Nanjing Medical University) ;
  • Wang, Ming-Kui (Key Laboratory of Mountain Ecological Restoration and Bioresource Utilization and Ecological Restoration Biodiversity Conservation Key Laboratory of Sichuan Province, Chengdu Institute of Biology, Chinese Academy of Sciences) ;
  • Hu, Weicheng (Jiangsu Collaborative Innovation Center of Regional Modern Agriculture & Environmental Protection/Jiangsu Key Laboratory for Eco-Agricultural Biotechnology Around Hongze Lake, Huaiyin Normal University)
  • Received : 2018.07.26
  • Accepted : 2018.09.18
  • Published : 2019.10.15

Abstract

Background: The leaves and roots of Panax ginseng are rich in ginsenosides. However, the chemical compositions of the leaves and roots of P. ginseng differ, resulting in different medicinal functions. In recent years, the aerial parts of members of the Panax genus have received great attention from natural product chemists as producers of bioactive ginsenosides. The aim of this study was the isolation and structural elucidation of novel, minor ginsenosides in the leaves of P. ginseng and evaluation of their antiinflammatory activity in vitro. Methods: Various chromatographic techniques were applied to obtain pure individual compounds, and their structures were determined by nuclear magnetic resonance and high-resolution mass spectrometry, as well as chemical methods. The antiinflammatory effect of the new compounds was evaluated on lipopolysaccharide-stimulated RAW 264.7 cells. Results and conclusions: Two novel, minor triterpenoid saponins, ginsenoside $LS_1$ (1) and 5,6-didehydroginsenoside $Rg_3$ (2), were isolated from the leaves of P. ginseng. The isolated compounds 1 and 2 were assayed for their inhibitory effect on nitric oxide production in LPS-stimulated RAW 264.7 cells, and Compound 2 showed a significant inhibitory effect with $IC_{50}$ of $37.38{\mu}M$ compared with that of NG-monomethyl-L-arginine ($IC_{50}=90.76{\mu}M$). Moreover, Compound 2 significantly decreased secretion of cytokines such as prostaglandin $E_2$ and tumor necrosis factor-${\alpha}$. In addition, Compound 2 significantly suppressed protein expression of inducible nitric oxide synthase and cyclooxygenase-2. These results suggested that Compound 2 could be used as a valuable candidate for medicinal use or functional food, and the mechanism is warranted for further exploration.

Keywords

1. Introduction

Ginseng (Panax ginseng Meyer) has been used as an herbal medicine or a functional food in East Asian countries because of its healthenhancing effects. The principle bioactive ingredients, termed ginsenosides, can be secreted from almost every part of this medicinal plant [1]. However, the chemical compositions of the aerial and underground parts of P. ginseng differ significantly. The roots are rich in protopanaxdiol saponins (PPDs; e.g., the ginsenosides Rb1, Rb2, and Rc), whereas the leaves contain high levels of protopanaxtriol saponins (PPTs; e.g., the ginsenosides Rg1 and Re) [2-4]. Pharmacological studies have demonstrated the different biological activities of these two categories of ginsenosides. For example, the ginsenoside Rg1 has been shown to induce angiogenesis, whereas Rb1 exerted an opposite effect [5,6]. Thus, the roots and leaves of ginseng can be used for different medicinal purposes. Ginseng roots, the most commonly used part of the P. ginseng plant, have been extensively investigated. The leaves, which contain a higher level of total ginsenosides, have been relatively less studied [7].

In recent years, many ginsenosides with unique C-17 side chains have been isolated from the aerial components of ginseng [8-11]. In particular, the ginsenosides Rg6, F4, Rh4, Rk1, and Rg5 were previously isolated from red ginseng in which the ginsenoside profile of ginseng products was altered by steaming [12-14]. These minor ginsenosides showed greater biological activities than the main constituents, such as ginsenosides Rg1, Rc, and Rd [15,16]. In terms of cost and availability, ginseng leaves are more advantageous than the roots.

Inflammation is a clinically common and significant basic pathological process. Chronic inflammation may lead to many serious diseases. Nowadays, scientists are paying more and more attention to develop antiinflammatory agents from natural resources [17-19]. Some ginsenosides from the genus Panax also exhibited good antiinflammatory activities [20-22]. To adequately understand and explore the easily obtained ginseng leaves, this study focused on the isolation and antiinflammatory evaluation of previously undiscovered ginsenosides from this valuable source.

2. Materials and methods

2.1. Plant material and chemicals

Panax ginseng leaves were purchased in August 2013 from Jingyu County, Jilin Province, China. A voucher specimen (no. 20130820) of this plant was deposited in our laboratory. 1-(4,5-Dimethylthiazol-2- yl)-3,5-diphenylformazan and lipopolysaccharide (LPS) were obtained from Sigma-Aldrich (St. Louis, MO, USA). RPMI 1640 medium and antibiotics were obtained from Gibco BRL (Life Technologies, Shanghai, China). Fetal bovine serum was obtained from Corning (Mediatech, Inc., Manassas, VA, USA). NG-monomethyl-l-arginine was purchased from Beyotime (Haimen, China). Immun-Blot PVDF membrane for protein blotting (0.2 μm, #1620177) was acquired from Bio-Rad (Hercules, CA, USA). Primary antibodies including inducible nitric oxide synthase (iNOS; D6B6S, #13120), cyclooxygenase-2 (COX-2; D5H5, #12282), and β-actin (8H10D10, #3700) were acquired from Cell Signaling Technology (Beverly, MA, USA). Goat anti-rabbit IgG H&L (ab6721) and the enzyme-linked immunosorbent assay kits for prostaglandin E2 (PGE2) and tumor necrosis factor-α (TNF-α) were purchased from Abcam (Cambridge, MA, USA). The enhanced chemiluminescence Western blot kit, radioimmunoprecipitation assay (RIPA) lysis buffer, and bicinchoninic acid (BCA) protein assay kit were obtained from CWBIO (Taizhou, China). Silica gel (100-200 mesh) for open column chromatography (CC) was purchased from Qingdao Haiyang Chemical Group Co., Ltd. (Qingdao, China). Methanol (MeOH), acetonitrile (CH3CN), n-butanol (n-BuOH), and dichloromethane (CH2Cl2) were purchased from Chengdu Kelong Chemical Co., Ltd. (Chengdu, China).

2.2. General experimental procedures

Optical rotations were measured using a PerkinElmer 341 polarimeter (PerkinElmer, MA, US). Infrared (IR) spectra were obtained using a PerkinElmer 1725X-FT spectrometer with potassium bromide disks. The high resolution electrospray ionization mass spectroscopy (HRESIMS) spectra were acquired on a Vion IMS QTof (Waters Corp., Milford, Massachusetts, USA) in positive ion mode. One-dimensional (1D) and two-dimensional (2D) NMR data were obtained using a Bruker Avance-600 spectrometer in C5D5N. Analysis of the sugar residues by HPLC was carried out on an Alltech series III apparatus with an evaporative light-scattering detector with a COSMOSIL Sugar-D packed column (Nacalai Tesque, Inc., Tokyo, Japan) using a mobile phase of 80% acetonitrile (CH3CN). Preparative-scale HPLC was implemented on a CXTH system, equipped with a C18 column (50 × 250 μm i.d., 10 mm, Daiso SP-100-10-ODS-P) from Daiso Co., Ltd. (Osaka, Japan) at a flow rate of 90 mL/min.

2.3. Extraction and isolation

The leaves (100.0 kg) of P. ginseng plants were powdered and extracted with 80% MeOH (3 × 250 L, each 24 h) at 60°C. The concentration of the extract was implemented under a rotary evaporator (R-114, Buchi, Switzerland) to yield a dark residue, which was suspended in H2O and partitioned successively with CH2Cl2 (3 × 15 L) and n-BuOH (3 × 15 L) to generate CH2Cl2 (1.1 kg) and n-BuOH (4.2 kg) soluble fractions.

A portion of the n-BuOH (4.0 kg) extract was applied to silica gel CC for separation using an H2O-saturated CH2Cl2/MeOH solvent system (gradient: 15:1 to 1:1) to generate 12 fractions (Fr.1 to Fr.12). Fr.2 (32.6 g) was purified by preparative HPLC and eluted with 58% and 80% MeOH, producing five subfractions (Fr.2-1 to Fr.2-5). Fr.2-1 (5.6 g) was further purified by preparative HPLC using 35% CH3CN to yield Compound 1 (tR = 33.8 min, 26 mg). Fr.4 (40 g) was fractionated by preparative HPLC using 68%, 80%, and 90% MeOH as eluent to generate nine subfractions (Fr.4-1 to Fr.4-9). Further separation of Fr.4-6 (2.3 g) by preparative HPLC with 65% CH3CN yielded Compound 2 (tR = 27.4 min, 43 mg).

2.4. Spectroscopic data

Ginsenoside LS1 (1): amorphous white power; \([\alpha]^{20}_D\)+20.7° (c 0.11, MeOH); IRνmax 3393, 2950, 2351, 1386, 1305, 1078, 1036, 931, 894, and 655 cm-1 ; 1H-NMR (600 MHz, C5D5N) and 13C-NMR (150 MHz, C5D5N) data are shown in Table 1; m/z 659.4144 [M+Na]+ (calculated for C36H60O9) from HRESIMS.

5,6-Didehydroginsenoside Rg3 (2): amorphous white power; \([\alpha]^{20}_D\)+7.1° (c 0.15, MeOH); IRνmax 3362, 2925, 2351, 1447, 1376, 1078, 1028, 896, and 655 cm-1 ; 1H-NMR (600 MHz, C5D5N) and 13C-NMR (150 MHz, C5D5N) data are shown in Table 1; m/z 805.4741 [M+Na]+ (calculated for C42H70O13) from HRESIMS.

Table 1 1H and13C NMR (600 MHz, 150 MHz in C5D5N) data for Compounds 1 and 2

2.5. Determination of the absolute configurations of sugar moieties

The previously described method was used to determine the absolute configurations of the sugar residues [23]. Compounds 1 and 2 (5 mg each) were mixed and heated with 5% sulfuric acid (2 mL) at 105°C for 8 h. The resultant reaction mixture was exhaustedly leached with CH2Cl2. Then, the water part was neutralized with barium hydroxide, filtered, and subjected to HPLC analysis compared to a standard glucose sample. The determined optical rotation value of \([\alpha]^{20}_D\)+48.8° (c 0.05, H2O) suggested the D-form for the glucose moieties in the identified compounds.

2.6. Cell lines and cell culture

RAW 264.7 cells were maintained in RPMI 1640 medium containing 10% fetal bovine serum and 1% antibiotics at 37°C under 5% CO2/95% air.

2.7. Cell viability assay

RAW 264.7 macrophage cells (1 × 105 cells/well) were seeded into 96-well plates and cultured for 18 h. Then, the culture medium was replaced with fresh medium along with the test compounds at various concentrations for 24 h. Cytotoxicity was investigated by 1- (4,5-dimethylthiazol-2-yl)-3,5-diphenylformazan assay as previously described [24]. Cell viability was expressed as a percentage of control cells.

2.8. Determination of NO, PGE2, and TNF-α production

RAW 264.7 macrophage cells (1 × 105 cells/well) were cultured for 18 h in a 96-well plate; the cells were pretreated with different concentrations of test samples for 30 min before treatment with LPS for 24 h. Nitrite contents were determined by the Griess reaction. The contents of TNF-α and PGE2 were evaluated according to the instruction on the kits.

2.9. Western blot analysis

RAW 264.7 cells were seeded onto 60-mm plates at 5 × 106 cells/ well. After treatment, the cells were washed with cold phosphate buffered saline (PBS), and the whole proteinwas extracted with RIPA lysis buffer containing protease inhibitor cocktail tablets (Complete ultra, Roche, Germany) under a FastPrep-24 homogenizer (MP, Solon, OH) with glass beads. Equal amounts of proteins were separated using 10% sodium dodecyl sulfate polyacrylamide gel electrophoresis and transferred onto a polyvinylidenefluoride (PVDF) membrane. The transferred protein was detected by iNOS (1:1000 dilution), COX-2 (1:1000 dilution), and β-actin (1:1000 dilution) using an enhanced chemiluminescence Western blot kit. The bands were visualized and photographed using the Tanon-5200 multiimaging system (Tanon Science & Technology, Shanghai, China).

2.10. Data analysis

The values are presented as the mean ± standard derivation. Differences between various experimental and control groups were compared using one-way analysis of variance followed by unpaired Studentttest, and p-values less than 0.05 were considered significant.

3. Results and discussion

Separation of the n-BuOH fraction of P. ginseng leaves by open silica gel CC and preparative reversed HPLC yielded Compounds 1 and 2 (Fig. 1). Compound 1, an amorphous white power, had a molecular formula of C36H60O9, which was based on the m/z of 659.4144 [M+Na]+ determined in the HRESIMS analysis. Thin layer chromatography visualization by heating at 105°C after spraying with 5% H2SO4 ethanol solution and the 13C-NMR spectrum of Compound 1 suggested that it was a triterpene glycoside. The 1HNMR spectrum of 1 revealed one anomeric proton signal at δH 5.20 (d, 7.7 Hz), whereas the 13C-NMR spectrum displayed one anomeric carbon signal at δC 98.4. The sugar residue was determined as Dglucose by acid hydrolysis experiment. The β configuration for the glucose residue was verified according to the coupling constant of the anomeric proton. The signals at δH 4.98 (s), 5.04 (s), 6.08 (m), and 6.42 (d, 15.6 Hz) in the 1H-NMR spectrum, along with the signals at δC 127.5 (C-23), 136.0 (C-24), 142.6 (C-25), and 115.0 (C26) in the 13C-NMR spectrum suggested the presence of two double bonds in 1. Compound 1 showed a maximum absorbance at 230 nm in its UV spectrum, indicating that the two double bonds were conjugated. Comparison of the 13C-NMR data of 1 with those of quinquenoside L1 demonstrated that they shared similar aglycones [25]. The appearance of carbon signals at δC 61.8 (C-5) and 67.8 (C6) in 1 rather than at around δC 56.1 (C-5) and 18.2 (C-6) suggested the existence of a hydroxyl group at C-6 in 1 [8]. The 1H- and 13CNMR signals of compound 1 were further assigned by extensive two-dimensional NMR spectra. The heteronuclear multiple bond correlation (HMBC) correlation between H-1 (δH 5.20) of the glucose residue and C-20 (δC 83.3) of the aglycone suggested the location of the sugar residue (Fig. 2). The positions of the two double bonds in the side chain were confirmed by correlations between H1-23 (δ 6.08) and C-22 and C-25 (δ 40.2 and 142.6); H1- 24 (δ 6.42) and C-22, C-25, and C-27 (δ 40.2, 142.6, and 19.0); H2-26 (δ 4.98 and 5.04) and C-24 and C-27 (δ 136.0 and 19.0); and H3-27 (δ 1.93) and C-24, C-25, and C-26 (δ 136.0, 142.6, and 115.0). Accordingly, Compound 1 was identified and named ginsenoside LS1.

Compound 2, an amorphous white power, possessed a molecular formula of C42H70O13, which was based on the m/z of 805.4741 [M+Na]+ determined in the HRESIMS analysis. Thin layer chromatography and 13C-NMR analysis demonstrated that Compound 2 was also a triterpene glycoside. The 1H- and 13C-NMR spectra of 2 exhibited signals that could be assigned to two sugar moieties at δH 4.89 (d, 7.4 Hz), 5.36 (d, 7.6 Hz) and δC 105.0, 106.2, respectively. Acid hydrolysis of 2 indicated that Dglucose was the only sugar component. The β configuration for both of the glucose residues was determined as described previously. The 1H-NMR spectrum of 2 showed signals due to eight tertiary methyl residues at δH 0.95, 1.04, 1.12, 1.45, 1.46, 1.52, 1.65, and 1.67 and two olefinic protons at δH 5.33 and 5.63 for the aglycone moiety. The existence of two double bonds in 2 was suggested from corresponding signals at δC 119.9, 126.4, 130.9, and 147.2 in the 13C-NMR spectrum. Comparison of the NMR data of 2 with those of 5,6-didehydroginsenoside Rd indicated they had comparable chemical structures except for the absence of the glucose residue at C-20 in 2 [26]. Sugar sequence was indicated to be a 1→2 linkage type attaching to C-3 of the aglycone from the HMBC correlations between δC 88.0 (C-3) and δH 4.89 (H-1 of the glc) and δ83.6 (C-2 of the glc) and δH 5.36 (H1 of the glc’) (Fig. 2). The chemical shifts of C-17, C-21, and C-22 observed at δC 54.8, 27.3, and 36.2, respectively, suggested that the configuration of C-20 was S [27]. As a result, Compound 2 was identified and named 5,6-didehydroginsenoside Rg3.

Fig. 1. Chemical structures of Compounds 1 and 2 isolated from the leaves of Panax ginseng

Fig. 2. Key HMBC (arrow) and 1H-1COSY (bold) correlations of Compounds 1 and 2. COSY, correlation spectroscopy

Macrophages play a key role in the process of inflammation on account of their functions in innate immunity and adaptive immunity. The activated macrophages by bacterial LPS initiate defensive reactions and release inflammatory mediators including NO and PGE2 and proinflammatory cytokines such as TNF-α, interleukin 1β (IL-1β), and interleukin 6 (IL-6) to improve the defensive ability [28]. Overproduction of NO can be harmful and mediate a broad spectrum of inflammatory disorders. Therefore, the LPS triggered the release of NO in RAW 264.7 cells is an excellent model to assess the effects of drugs on the inflammatory process [29].

As shown in Fig. 3A, the nitrite concentration in unstimulated RAW 264.7 cells was almost undetectable; however, the nitrite concentration was increased significantly upon the treatment of LPS. Among the two compounds, Compound 2 significantly reduced NO production in LPS-stimulated RAW 264.7 with IC50 value of 37.38 μM, which was more potent than the positive control (L-NMA, 90.76 μM), whereas Compound 1 did not exert an inhibitory effect.

To exclude the possibility that the cytotoxicity of Compound 2 might contribute to its antiinflammatory effects, we incubated RAW 264.7 cells with varying concentrations of Compound 2 for 24 h. As shown in Fig. 3B, Compound 2 treatment did not induce cell death in RAW 264.7 cells at concentrations up to 100 mM. Furthermore, the levels of PGE2 and TNF-α were determined using an enzyme-linked immunosorbent assay kit. As shown in Fig. 3C and D, Compound 2 inhibited TNF-α and PGE2 production in a concentration-dependent manner.

Fig. 3. Inhibitory effect of Compounds 1 and 2 on NO, PGE2, and TNF-α production. Cells were cultured for 18 h in a 96-well plate; the cells were pretreated with various concentrations of test samples for 30 min before treatment with LPS for 24 h. (A) Nitrite contents were determined by the Griess reaction. (B) Cell viability was investigated by MTT assay. (C and D) TNF-α and PGEwere determined using commercial kits according to the manufacturers’ instructions. The data are presented as means ± SD (n = 3). *p < 0.05 vs LPS+ group; #p < 0.05 vs control group.

Several ginsenosides, such as ginsenosides Rg1, Rf, Rh1, Rh2, Rb1, Rb2, Rc, Rd, Rg3, and Rg5, were previously reported to exert good antiinflammatory activities [30-35]. Ginsenosides Rg1, Rf, and Rh1 (PPT) shared one or two glucose residues at C-6. Ginsenosides Rh2, Rb1, Rb2, Rc, Rd, Rg3, and Rg5 (PPD) had one or two glucose moieties at C-3 position. This conclusion also well fit the results of the present article. Compound 1 (PPT) had no glucose residues at C-6 and did not exhibit antiinflammatory activity, whereas Compound 2 (PPD) bore two glucose moieties at C-3 and showed excellent antiinflammatory activity.

iNOS is a product of transcription expression of activated macrophages and is the reason for the prolonged and profound production of NO [36]. COX-2 is an inducible enzyme, and overexpression of it has been deemed to involve in the pathogenesis of various inflammatory diseases, angiogenesis, and cancer cell invasion [37]. The expression of iNOS and COX-2 induces the generation of excessive amounts of NO and PGE2 [38]. In the present study, protein expression of iNOS and COX-2 was attenuated by Compound 2 treatment, which might indicate that Compound 2 inhibited the production of NO and PGE2 by the downregulating protein expression level of iNOS and COX-2 (Fig. 4).

Fig. 4. Effects of Compound 2 on the expression of iNOS and COX-2. (A) Cells were plated at a density of 5 × 106 cells/dish. Cells were pretreated with various concentrations of Compound 2 for 30 min before treatment with LPS for 6 h. After preparation of the whole protein, the protein expression levels of iNOS and COX-2 were measured by Western blot. Results are representative of three experiments

In conclusion, we reported the isolation of two new and minor ginsenosides from the leaves of P. ginseng and found that Compound 2 showed excellent antiinflammatory activity. This study might provide further insights into the chemical compositions and functions of the leaves of P. ginseng and help to explore this invaluable source.

Conflicts of interest

The authors declare no conflicts of interest.

Acknowledgments

This work was financially supported by the West Light Foundation of Chinese Academy of Sciences (Y7C1031100), Huai’an 2016 Annual Promotion Project for Science and Technology International Cooperation (HAC201620 and HAC201622), National Natural Science Foundation of China (21561142003 and 41501262), Natural Science Foundation of Jiangsu Province (BK20171269 and BK20140455), National Basic Research Program of China (973 Program, 2015CB554400), Natural Science Foundation of the Higher Education Institutions of Jiangsu Province (17KJA550001 and 18KJA550001), and Qing Lan Project of Jiangsu Province.

References

  1. Oh JY, Kim YJ, Jang MG, Joo SC, Kwon WS, Kim SY, Jung SK, Yang DC. Investigation of ginsenosides in different tissues after elicitor treatment in Panax ginseng. J Ginseng Res 2014;38:270-7. https://doi.org/10.1016/j.jgr.2014.04.004
  2. Cheng YJ, Zhang M, Liang QL, Hu P, Wang YM, Jun FW, Luo GA. Two-step preparation of ginsenoside Re, $Rb_{1}$, Rc and $Rb_{2}$ from the root of Panax ginseng by high-performance counter-current chromatography. Sep Purif Technol 2011;77:347-54. https://doi.org/10.1016/j.seppur.2011.01.003
  3. Xie JT, Mehendale SR, Wang A, Aung HH, Wu J, Osinski J, Yuan CS. American ginseng leaf: ginsenoside analysis and hypoglycemic activity. Pharmacol Res 2004;49:113-7. https://doi.org/10.1016/j.phrs.2003.07.015
  4. Li TSC, Mazza G, Cottrell AC, Gao L. Ginsenosides in roots and leaves of American ginseng. J Agric Food Chem 1996;44:717-20. https://doi.org/10.1021/jf950309f
  5. Leung KW, Pon YL, Wong RNS, Wong AST. Ginsenoside-$Rg_{1}$ induces vascular endothelial growth factor expression through the glucocorticoid receptorrelated phosphatidylinositol 3-kinase/akt and ${\beta}$-catenin/T-cell factordependent pathway in human endothelial cells. J Biol Chem 2006;281:36280-8. https://doi.org/10.1074/jbc.M606698200
  6. Leung KW, Cheung LWT, Pon YL, Wong RNS, Mak NK, Fan TPD. Au SCI, Tombran-Tink J, Wong AST. Ginsenoside $Rb_{1}$ inhibits tube-like structure formation of endothelial cells by regulating pigment epithelium-derived factor through the oestrogen ${\beta}$ receptor. Brit J Pharmacol 2007;152:207-15. https://doi.org/10.1038/sj.bjp.0707359
  7. Wang HW, Peng DC, Xie JT. Ginseng leaf-stem: bioactive constituents and pharmacological functions. Chin Med UK 2009;4:20. https://doi.org/10.1186/1749-8546-4-20
  8. Liu GY, Li XW, Wang NB, Zhou HY, Wei W, Gui MY, Yang B, Jin YR. Three new dammarane-type triterpene saponins from the leaves of Panax ginseng C.A. Meyer. J Asian Nat Prod Res 2010;12:865-73. https://doi.org/10.1080/10286020.2010.508035
  9. Zhu TT, Li F, Chen B, Deng Y, Wang MK, Li LH. Studies on the saponins from the leaves of Panax ginseng. Chin J Appl Environ Biol 2016;22:70-4.
  10. Yang XW, Li LY, Tian JM, Zhang ZW, Ye JM, Gu WF. Ginsenoside-$Rg_{6}$, a novel triterpenoid saponin from the stem-leaves of Panax ginseng C.A. Meyer. Chin Chem Lett 2000;11:909-12.
  11. Cong DL, Song CC, Xu JD. Isolation and identification of 20 (s)-ginsenoside-Rh1, -Rh2 and ginsenoside-Rh3 from the leaves of Panax quinquefolium. Chin Pharm J 2000;35:82-4.
  12. Ryu JH, Park JH, Eun JH, Jung JH, Sohn DH. A dammrane glycoside from Korean red ginseng. Phytochemistry 1997;44:931-3. https://doi.org/10.1016/S0031-9422(96)00661-9
  13. Park IH, Kim NY, Han SB, Kim JM, Kwon SW, Kim HJ, Park MK, Park JH. Three new dammarane glycosides from heat processed ginseng. Arch Pharm Res 2002;25:428-32. https://doi.org/10.1007/BF02976595
  14. Ryu JH, Park JH, Kim TH, Sohn DH, Kim JM, Park JH. A genuine dammarane glycoside, (20E)-ginsenoside $F_{4}$ from Korean red ginseng. Arch Pharm Res 1996;19:335-6. https://doi.org/10.1007/BF02976251
  15. Ying A, Yu QT, Guo L, Zhang WS, Liu JF, Li Y, Song H, Li P, Qi LW, Ge YZ, et al. Structuraleactivity relationship of ginsenosides from steamed ginseng in the treatment of erectile dysfunction. Am J Chin Med 2018:1-19.
  16. Baek SH, Shin BK, Kim NJ, Chang SY, Park JH. Protective effect of ginsenosides $Rk_{3}$ and $Rh_{4}$ on cisplatin-induced acute kidney injury in vitro and in vivo. J Ginseng Res 2017;41:233-9. https://doi.org/10.1016/j.jgr.2016.03.008
  17. Wen R, Lv HN, Jiang Y, Tu PF. Anti-inflammatory flavone and chalcone derivatives from the roots of Pongamia pinnata (L.) Pierre. Phytochemistry 2018;149:56-63. https://doi.org/10.1016/j.phytochem.2018.02.005
  18. Ahmad TB, Liu L, Kokiw M, Benkendorff K. Review of anti-inflammatory, immune-modulatory and wound healing properties of molluscs. J Ethnopharmacol 2018;210:156-78. https://doi.org/10.1016/j.jep.2017.08.008
  19. Liu MT, Zhou Q, Wang JP, Liu JJ, Qi CX, Lai YJ, Zhu HC, Xue YB, Hu ZX, Zhang YH. Anti-inflammatory butenolide derivatives from the coral-derived fungus Aspergillus terreus and structure revisions of aspernolides D and G, butyrolactone VI and 40,80 0-diacetoxy butyrolactone VI. RSC Adv 2018;8:13040-7. https://doi.org/10.1039/C8RA01840E
  20. Gao Y, Chu SF, Li JW, Li JP, Zhang Z, Xia CY, Heng Y, Zhang MJ, Hu JF, Wei GN, et al. Anti-inflammatory function of ginsenoside Rg1 on alcoholic hepatitis through glucocorticoid receptor related nuclear factor-kappa B pathway. J Ethnopharmacol 2015;173:231-40. https://doi.org/10.1016/j.jep.2015.07.020
  21. Kang S, Park SJ, Lee AY, Huang J, Chung HY, Im DS. Ginsenoside Rg3 promotes inflammation resolution through M2 macrophage polarization. J Ginseng Res 2018;42:68-74. https://doi.org/10.1016/j.jgr.2016.12.012
  22. Tam DNH, Truong DH, Nguyen TTH, Quynh LN, Tran L, Nguyen HD, Shamandy B, Le TMH, Tran DK, Sayed D, et al. Ginsenoside Rh1: a systematic review of its pharmacological properties. Planta Med 2018;84:139-52. https://doi.org/10.1055/s-0043-124087
  23. Li F, Yang FM, Liu X, Wang L, Chen B, Li LH, Wang MK. Cucurbitane glycosides from the fruit of Siraitia grosvenori and their effects on glucose uptake in human HepG2 cells in vitro. Food Chem 2017;228:567-73. https://doi.org/10.1016/j.foodchem.2017.02.018
  24. Baek KS, Hong YD, Kim Y, Sung NY, Yang S, Lee KM, Park JY, Park JS, Rho HS, Shin SS, et al. Anti-inflammatory activity of AP-SF, a ginsenoside-enriched fraction, from Korean ginseng. J Ginseng Res 2015;39:155-61. https://doi.org/10.1016/j.jgr.2014.10.004
  25. Wang JH, Li X, Wang YJ. A new triterpenoid in the leaves and stems of Panax Quinquefolium L. from Canada. J Shenyang Pharm Univ 1997;14:135-6.
  26. Wan JB, Zhang QW, Hong SJ, Guan J, Ye WC, Li SP, Simon Lee MY, Wang YT. 5,6-didehydroginsenosides from the roots of Panax notoginseng. Molecules 2010;15:8169-76. https://doi.org/10.3390/molecules15118169
  27. Yang H, Kim JY, Kim SO, Yoo YH, Sung SH. Complete $^{1}H$-NMR and $^{13}C$-NMR spectral analysis of the pairs of 20(S) and 20(R) ginsenosides. J Ginseng Res 2014;38:194-202. https://doi.org/10.1016/j.jgr.2014.05.002
  28. Kmiec Z, Cyman M, Slebioda TJ. Cells of the innate and adaptive immunity and their interactions in inflammatory bowel disease. Adv Med Sci 2017;62(1):1-16. https://doi.org/10.1016/j.advms.2016.09.001
  29. Danie LO, Rosina LF. Immunomodulating peptides for food allergy prevention and treatment. Crit Rev Food Sci Nutr 2018;58:1629-49. https://doi.org/10.1080/10408398.2016.1275519
  30. Zhu G, Wang H, Wang T, Shi F. Ginsenoside Rg1 attenuates the inflammatory response in DSS-induced mice colitis. Int Immunopharmacol 2017;50:1-5. https://doi.org/10.1016/j.intimp.2017.06.002
  31. Zhang YX, Wang L, Xiao EL, Li SJ, Chen JJ, Gao B, Min GN, Wang ZP, Wu YJ. Ginsenoside-Rd exhibits anti-inflammatory activities through elevation of antioxidant enzyme activities and inhibition of JNK and ERK activation in vivo. Int Immunopharmacol 2013;17:1094-100. https://doi.org/10.1016/j.intimp.2013.10.013
  32. Choi WY, Lim HW, Lim CJ. Anti-inflammatory, antioxidative and matrix metalloproteinase inhibitory properties of 20(R)-ginsenoside Rh2 in cultured macrophages and keratinocytes. J Pharm Pharmacol 2012;65:310-6. https://doi.org/10.1111/j.2042-7158.2012.01598.x
  33. Li J, Du J, Liu D, Cheng BB, Fang FF, Weng L, Wang C, Ling CQ. Ginsenoside Rh1 potentiates dexamethasone0s anti-inflammatory effects for chronic inflammatory disease by reversing dexamethasone-induced resistance. Arthritis Res Ther 2014;16:R106. https://doi.org/10.1186/ar4556
  34. Yu T, Yang YY, Kwak YS, Song GG, Kim MY, Rhee MH, Cho JY. Ginsenoside Rc from Panax ginseng exerts anti-inflammatory activity by targeting TANKbinding kinase 1/interferon regulatory factor-3 and p38/ATF-2. J Ginseng Res 2017;41:127-33. https://doi.org/10.1016/j.jgr.2016.02.001
  35. Lee YY, Park JS, Jung JS, Kim DH, Kim HS. Anti-inflammatory effect of ginsenoside Rg5 in lipopolysaccharide-stimulated BV2 microglial cells. Int J Mol Sci 2013;14:9820-33. https://doi.org/10.3390/ijms14059820
  36. Bueno-Silva B, Kawamoto D, Ando-Suguimoto ES, Casarin RCV, Alencar SM, Rosalen PL, Mayer MPA. Brazilian red propolis effects on peritoneal macrophage activity: nitric oxide, cell viability, pro-inflammatory cytokines and gene expression. J Ethnopharmacol 2017;207:100-7. https://doi.org/10.1016/j.jep.2017.06.015
  37. Shehzad A, Parveen S, Qureshi M, Subhan F, Lee YS. Decursin and decursinol angelate: molecular mechanism and therapeutic potential in inflammatory diseases. Inflamm Res 2018;67:209-18. https://doi.org/10.1007/s00011-017-1114-7
  38. Kim E, Yi YS, Son YJ, Han SY, Kim DH, Nam G, Hossain MA, Kim JH, Park J, Chol JY. BIOGF1K, a compound K-rich fraction of ginseng, plays an antiinflammatory role by targeting an activator protein-1 signaling pathway in RAW 264.7 macrophage-like cells. J Ginseng Res 2018;42(2):233. https://doi.org/10.1016/j.jgr.2018.02.001

Cited by

  1. Panax ginseng callus, suspension, and root cultures: extraction and qualitative analysis vol.8, pp.2, 2019, https://doi.org/10.21603/2308-4057-2020-2-369-376
  2. Anti-inflammatory effect of Barringtonia angusta methanol extract is mediated by targeting of Src in the NF-κB signalling pathway vol.59, pp.1, 2019, https://doi.org/10.1080/13880209.2021.1938613
  3. Anti-neuroinflammatory effect of agaves and cantalasaponin-1 in a model of LPS-induced damage vol.35, pp.5, 2019, https://doi.org/10.1080/14786419.2019.1608537
  4. Identification of Bioactive Natural Product from the Stems and Stem Barks of Cornus walteri: Benzyl Salicylate Shows Potential Anti-Inflammatory Activity in Lipopolysaccharide-Stimulated RAW 264.7 Mac vol.13, pp.4, 2019, https://doi.org/10.3390/pharmaceutics13040443