Next Article in Journal
The Interplay of RNA Binding Proteins, Oxidative Stress and Mitochondrial Dysfunction in ALS
Next Article in Special Issue
The Effects of Prunus spinosa L. Flower Extracts, Model Polyphenols and Phenolic Metabolites on Oxidative/Nitrative Modifications of Human Plasma Components with Particular Emphasis on Fibrinogen In Vitro
Previous Article in Journal
Antioxidative Molecules in Human Milk and Environmental Contaminants
Previous Article in Special Issue
Composition and Antioxidant Activity of Phenolic Compounds in Fruit of the Genus Rosa L.
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Phenolic Constituents of Chinese Quince Twigs (Chaenomeles sinensis Koehne) and Their Anti-Neuroinflammatory, Neurotrophic, and Cytotoxic Activities

1
School of Pharmacy, Sungkyunkwan University, Suwon 16419, Korea
2
Gachon Institute of Pharmaceutical Science, Gachon University, Incheon 21936, Korea
3
Korea Research Institute of Chemical Technology, Daejeon 34114, Korea
4
College of Pharmacy, Gachon University, #191, Hambakmoero, Yeonsu-gu, Incheon 21936, Korea
5
Department of Biopharmaceutical Convergence, Sungkyunkwan University, Suwon 16419, Korea
*
Author to whom correspondence should be addressed.
Antioxidants 2021, 10(4), 551; https://doi.org/10.3390/antiox10040551
Submission received: 8 March 2021 / Revised: 26 March 2021 / Accepted: 30 March 2021 / Published: 2 April 2021
(This article belongs to the Special Issue Phenolics as Antioxidant Agents)

Abstract

:
Chaenomeles sinensis has been used as a food and traditional medicines. However, most of research on discovering bioactive constituents from this plant have been focused on its yellow fruit, Chinese quince, due to its wide usage. Here, we isolated and characterized three new phenolic compounds (1, 9, and 11) and 21 known compounds (28, 10, and 1224) from the twigs of C. sinensis. Their chemical structures were established by spectroscopic and spectrometric data analysis including 1D and 2D NMR, high-resolution mass spectrometry (HRMS), electronic circular dichroism (ECD), and LC-MS analysis. Some of the isolated compounds (124) showed anti-neuroinflammatory effects on nitric oxide (NO) production in lipopolysaccharide (LPS)-activated BV-2 cells, neurotrophic activity in C6 cells through the secretion of nerve growth factor (NGF) and/or cytotoxicity against four human cancer cell lines (A549, SK-OV-3, SK-MEL-2, MKN-1).

1. Introduction

Phenolic compounds are one of the largest natural products groups, especially in the plant kingdom, and are characterized by the presence of at least a benzene ring with hydroxy group(s) attached to it [1]. They can be classified into several subgroups according to the number of carbon atoms and their arrangement as follows: phenolic acids (C6–C1), acetophenones and phenylacetic acids (C6–C2), phenylpropanoids (C6–C3), naphthoquinones (C6–C4), xanthones (C6–C1–C6), stilbenoids (C6–C2–C6), flavonoids (C6–C3–C6), and lignans (C6–C3–C3–C6) [1,2]. Many of these well-known natural products in the phenolic compounds class have been reported to possess diverse pharmacological activities including anti-inflammatory (e.g., salicylic acid), neurotrophic (e.g., 6-shogaol) [2,3], antioxidant (e.g., resveratrol, quercetin) [2], and anticancer (e.g., podophyllotoxin) [2] effects.
Neurodegenerative diseases such as dementia and Parkinson’s disease are exploding due to rapid aging populations [4]. Despite various efforts to overcome these neurodegenerative diseases, there are not sufficient reports on drugs that modify the disease [5]. Therefore, it is very important to draw meaningful leading compounds from natural resources in order to regulate neurodegenerative diseases. As neurodegenerative diseases are highly complex and incurable things, there are some aspects that make it difficult to overcome diseases with a single target such as amyloid beta [5]. Thus, we should think about the study of exploring phytochemicals that overcome neurodegenerative diseases by regulating multiple targets. In this study, the effects of phytochemicals with anti-neuroinflammation and neurotrophic factor potentiation will be shown. They are important multi-targets that cause the induction and progression of neurodegenerative diseases.
Chaenomeles sinensis Koehne (or Pseudocydonia sinensis, Chinese Quince, Rosaceae) is a semi-evergreen tree in the family Rosaceae and is widely distributed in East Asia, including Korea, mainland China, and Japan. The yellow fruit of this plant has been consumed as a tea, in a candied form, or in liquor [6], and also used as a Korean traditional medicine over 400 years to treat myalgia, beriberi, vomiting, and diarrhea [7]. Due to the wide range of usages of C. sinensis fruit, the phytochemical and biological research on this plant has been mainly focused on the fruit [8,9,10,11,12,13] rather than the other parts such as twigs [14,15] or leaves. Based on these previous studies, it was found that the major compound classes of C. sinensis were triterpenoids [9,14,15,16] and flavonoids [8,13,15]. However, minor components such as non-flavonoid phenolic compounds in the C. sinensis twigs with anti-neurodegerative effects remain largely unknown. In this study, 24 phenolic compounds (1–24) including three new compounds (1, 9, and 11) were isolated and characterized from the twigs of C. sinensis (Figure 1). Their structures were elucidated by conventional NMR techniques coupled with simulation of 1H NMR peak, analysis of HRMS and ECD data, and hydrolysis followed by LC-MS analysis. The isolates (124) were tested for their anti-neuroinflammatory, neurotrophic, and cytotoxic activities, and herein, we report the isolation and structure elucidation of phenolic phytochemicals from the twigs of C. sinensis as well as their biological activities.

2. Materials and Methods

2.1. General Experimental Procedures

JASCO P-1020 polarimeter equipped with the sodium D line (590 nm) (JASCO, Easton, MD, USA) was performed to measure specific rotations. Bruker AVANCE III 700 NMR spectrometer at 700 MHz (1H) and 175 MHz (13C) was carried out to measure the NMR spectra (1H, 13C, COSY, HSQC, and HMBC) at 700 MHz (1H) and 175 MHz (13C) with chemical shifts given in ppm (δ) (Bruker, Karlsruhe, Germany). High-resolution fast atom bombardment mass spectroscopy (HRFABMS) were measured on either a Waters SYNAPT G2 (Milford, MA, USA) or a JEOL JMS700 mass spectrometer (Tokyo, Japan). LC-MS data were measured using an Agilent 1290 Infinity II HPLC instrument (Foster City, CA, USA) coupled to a G6545B quadrupole time-of-flight (Q-TOF) mass spectrometer (Agilent Technologies) with a Kinetex C18 5 µm column (250 mm length × 4.6 mm i.d.; Phenomenex, Torrance, CA, USA). The semipreparative high-performance liquid chromatography (HPLC) furnished with a Gilson 306 pump (Middleton, WI, USA) and a Shodex refractive index detector (New York, NY, USA) was used. Column chromatography was performed employing either silica gel 60 (70−230 and 230−400 mesh; Merck, Darmstadt, Germany) or RP-C18 silica gel (Merck, 230−400 mesh). Sephadex LH-20 (Pharmacia, Uppsala, Sweden) was utilized for molecular sieve column chromatography. Merck precoated silica gel F254 plates and RP-C18 F254s plates (Merck, Darmstadt, Germany) were used for thin-layer chromatography (TLC) analysis. Spots were detected on TLC under UV light or by heating after spraying with anisaldehyde−sulfuric acid.

2.2. Plant Material

Twigs of two-year-old C. sinensis (7.0 kg) were purchased at Yangjae Flower Market in Seoul, Korea, in January 2012. A voucher specimen of the plants (SKKU-NPL 1206) was authenticated by Prof. Kang Ro Lee and deposited in the herbarium of the School of Pharmacy, Sungkyunkwan University, Suwon, Korea.

2.3. Extraction and Isolation

Twigs of C. sinensis (7.0 kg) were extracted with 80% aqueous MeOH (each 10 L × 1 day, 3 times) under reflux and filtered. The filtrate was evaporated in vacuo to yield a MeOH extract (320 g). The 80% MeOH extract was suspended in distilled water and successively partitioned with n-hexane (3 g), CHCl3 (15 g), EtOAc (6 g), and n-BuOH (30 g). Repeated column chromatographical fractionation and isolation of CHCl3, EtOAc, and n-BuOH soluble layers afforded the compounds 124. (See Supplementary Materials for the detailed isolation process).
(7S,8R)-4,7,9,9′-tetrahydroxy-3,5,3′,5′-tetramethoxy-7′-oxo-8-O-4′-neolignan (1). Colorless gum; [α] D 25 −12 (c 0.1, MeOH); ECD (MeOH) λmaxε) 244 (−2.3) nm; 1H and 13C NMR data, see Table 1; HRFABMS m/z 451.1600 [M − H] (calcd for C22H27O10, 451.1599) (Figures S1–S7).
(7R,8S)-4,7,9,9′-tetrahydroxy-3,3′,5′-trimethoxy-8-O-4′-neolignan (2). Colorless gum; ECD (MeOH) λmaxε) 241 (+2.1) nm; 1H and 13C NMR data, see Table 1.
1:1 mixture of (7S,8R)- and (7R,8S)-4,7,9,9-tetrahydroxy-3,5,3′,5′-tetramethoxy-8-O-4-neolignan (3). Colorless gum; 1H and 13C NMR data, see Table 1.
7R,7′R,8S,8′S-icariol A2 9′-O-α-l-rhamnopyranoside (9). Colorless gum; [α] D 25 +30 (c 0.1, MeOH); ECD (MeOH) λmaxε) 246 (+5.3), 208 (−10.1) nm; 1H and 13C NMR data, see Table 1; HRFABMS m/z 581.2231 [M − H] (calcd for C28H37O13, 581.2229) (Figures S8–S15).
β-hydroxypropiosyringone 4′-O-β-d-glucopyranoside (11). Colorless gum; [α] D 25 +24 (c 0.25, MeOH); 1H and 13C NMR data, see Table 2; HRFABMS m/z 411.1263 [M + Na]+ (calcd for C17H24O10Na, 411.1262) (Figures S16–S22).

2.4. 1H NMR Simulation for H-7 of Compound 1

MestReNova Version 14.1.2-25024 was used to simulate the H-7 peak. The “Spin Simulation” under the “Prediction” tab was clicked to open the dialog box. The “Spin Groups” was set to be “2” (default value), the “Shifts (ppm)” values of “A” and “B” were given “4.987” and “4.232”, respectively, with the “Line Width (Hz)” of both “3.0”. After the coupling constant between “A” and “B” was set as “3.5” or “6.7”, the icon with “New Simulation” at the top left was clicked to generate the simulated 1H NMR peaks of H-7 and H-8 of 1.

2.5. Acid Hydrolysis of Compounds 6 and 11 and Sugar Analysis

The sugar analysis experiment was performed as described in previous communication by Tanaka et al. with some modifications [17]. Compounds 9 and 11 (each of 0.5 mg) were individually hydrolyzed with 1 N HCl (1.0 mL) for 2 h at 100 °C. Each hydrolysate was extracted using CHCl3, and then the aqueous layers were evaporated to remove the HCl and to afford sugar from each hydrolysate. Each H2O-soluble fraction was added to anhydrous pyridine (400 µL) containing l-cysteine methyl ester hydrochloride (1.1 mg), which were heated at 60 °C for 1 h. Then, O-tolyl isothiocyanate (30 µL) was added to the reaction mixture and heated at 60 °C for another 1 h. After the reaction, each reaction mixture was analyzed by LC-MS (0.7 mL/min; 25% aqueous CH3CN with 0.1% formic acid for 40 min). The authentic samples of l-rhamnopyranose and d-glucopyranose were derivatized and analyzed by the same method mentioned above, and these monosaccharides were also derivatized with d-cysteine methyl ester hydrochloride instead of its l-form to prepare their enantiomeric derivatives. The hydrolysate derivatives of 9 and 11 were 25.4 min (l-rhamnopyranose) and 14.9 min (d-glucopyranose) which matched well with those of authentic l-rhamnopyranose (25.3 min) and d-glucopyranose (14.9 min), respectively, rather than their enantiomeric derivatives (13.2 min (d-rhamnopyranose) and 13.7 min (l-glucopyranose)) (Figures S15 and S22).

2.6. NO Production Assays

Analogous as described in [18]. The BV-2 cells, developed by Dr. V. Bocchini at the University of Perugia (Perugia, Italy), were used for this study [19,20]. The cells were seeded in a 96-well plate (4 × 104 cells/well) and incubated in the presence or absence of various doses of tested compounds (124). Lipopolysaccharide (LPS) (100 ng/mL) was added to all the pre-treated wells except the control one and grown for 1 d. The produced levels of nitrite (NO2), a soluble oxidized product of NO, was evaluated with 0.1% N-1-napthylethylenediamine dihydrochloride and 1% sulfanilamide in 5% phosphoric acid, aka the Griess reagent. The supernatant (50 μL) was mixed with the Griess reagent (50 μL). After 10 min the absorbance was gauged at 570 nm. For a positive control, the reported nitric oxide synthase (NOS) inhibitor NG-monomethyl-l-arginine (l-NMMA) was employed. Graded sodium nitrite solution was utilized to determine nitrite concentrations. A 3-[4,5-dimethylthiazol-2-yl]-2,5-diphenyltetrazolium bromide (MTT) assay was used for the cell viability assay.

2.7. Assays for Nerve Growth Factor NGF Release from C6 Cells

Analogous as described in [21]. C6 glioma cell lines were used to measure the nerve growth factor (NGF) of the culture medium, which were fixed with 10% fetal bovine serum (FBS) and 1% penicillin-streptomycin (PS) in an incubator filled with 5% CO2. The cells were seeded in a 24-well culture plate (1 × 105 cells/well) and incubated for 24 h. The cells were treated with or without 20 µM of the compounds (124), together with serum-free Dulbecco’s modified Eagle’s medium (DMEM) for another 24 h. Released NGF levels from the supernatants from each cell were measured using an ELISA development kit (R&D System, Minneapolis, MN, USA). Besides, the cell viability was evaluated by 3-[4,5-dimethylthiazol-2-yl]-2,5-diphenyltetrazolium bromide (MTT) assay by comparison with 6-shogaol as a positive control and the results are expressed as percentage of the control group.

2.8. Cytotoxicity Assessment

The cytotoxicity of purified compounds (124) were tested against the A-549 (non-small cell lung adenocarcinoma), SK-OV-3 (ovary malignant ascites), SK-MEL-2 (skin melanoma), and MKN-1 (adenosquamous carcinoma), utilizing the sulforhodamine B colorimetric (SRB) method [22]. Cisplatin (≥98%; Sigma–Aldrich) served as a positive control.

3. Results and Discussion

3.1. Structure Elucidation of Compounds 124

Compound 1 was purified as a colorless gum and its molecular formula was determined as C22H28O10 according to the deprotonated molecular ion at m/z at 451.1600 [M – H] in the HRFABMS data (calcd. for C22H27O10, 451.1599). Analysis of the 1H NMR spectrum of 1 indicated the presence of two 1,3,4,5-tetrasubstituted benzene rings (δH 7.29 (2H, s, H-2′,6′) and 6.58 (2H, s, H-2,6)), two oxygenated methines (δH 4.99 (1H, brs, H-7) and 4.23 (1H, ddd, J = 6.6, 3.5, 3.0 Hz, H-8)), two oxygenated methylenes (δH 4.06 (2H, t, J = 5.3 Hz, H-9′), 3.93 (1H, overlap, H-9a), and 3.54 (1H, m, H-9b)), a methylene (δH 3.23 (2H, t, J = 5.4 Hz, H-8′)), and four methoxy groups (δH 3.96 (6H, s, 3′,5′-OCH3) and 3.88 (6H, s, 3,5-OCH3)). The 13C spectrum of 1 showed 22 carbon signals for two 1,3,4,5-tetrasubstituted benzene rings (δC 153.6 (×2), 147.2 (×2), 139.9, 134.2, 133.1, 130.3, 105.6 (×2), and 102.5 (×2)), a carbonyl carbon (δC 199.1), four methoxy carbons (δC 56.6 (×2) and 56.5 (×2)), four oxygenated carbons (δC 87.5, 73.1, 60.7, and 58.3), and a methylene carbon (δC 40.4). These NMR spectra of 1 were similar to those of 3 (Table 1) [23], except for the presence of the carbonyl carbon signal at δC 199.1 (C-7′) in 1 rather than the methylene carbon signal at δC 34.4 (C-7′) in 3. The HMBC correlations from H-2′,6′/H-8′/H-9′ to C-7′ were observed, suggesting that carbonyl carbon is located at C-7′. The planar structure of 1 was established through 2D NMR spectra analysis including COSY, HSQC, and HMBC spectra (Figure 2A). In 8-O-4′-oxyneolignan derivatives, the relatively small (~4 Hz) and large (~8 Hz) coupling constants between H-7 and H-8 indicate erythro- or threo-configuration, respectively, in chloroform-d [24,25,26]. When the H-7 peak was analyzed to measure the 3JH-7,H-8 value, however, its shape was found to be a broad singlet. Alternatively, the H-8 peak showed the splitting pattern of doublet of doublet of doublet (ddd) with coupling constants of 6.7, 3.5, and 3.0 Hz. In order to deduce the 3JH-7,H-8 value among them, the H-7 peak was simulated with coupling constants of 3.5 and 6.7 Hz. As shown in Figure 2B (middle) the simulated peak with the 3.5 Hz coupling constant was more similar to that of experimental peak rather than that with 6.7 Hz coupling constant. Therefore, the 3JH-7,H-8 value was assigned as 3.0 or 3.5 Hz which suggested erythro-form. The difference of 13C NMR chemical shift between C-7 and C-8 can be also used for its relative configuration assignment. The ΔδC-8,C-7 value was ~14.6 ppm in the erythro-form whereas ~15.5 ppm was observed in the threo-form [24]. By observing 14.4 ppm of ΔδC-8,C-7 value in the 13C NMR data of 1 (Figure 2C), the relative configuration between C-7 and C-8 was deduced as erythro-form. The 8R configuration of 1 was confirmed by means of analysis of its ECD spectrum displaying a characteristic negative Cotton effect at 244 nm (Figure 2D) [26,27]. Thus, the structure of 1 was established as (7S,8R)-7,9,9′-trihydroxy-3,5,3′,5′-tetramethoxy-7′-oxo-8-O-4′-neolignan.
Compounds 2 and 3 were previously reported from Bursera tonkinensis [23], and Epimedium sagittatum [28] and Fraxinus mandshurica [29], respectively, without determination of the absolute configuration of C-7 and C-8 (Figure S23–S26). The relative configuration between C-7 and C-8 were assigned as erythro by the same empirical rule applied for 1; the relatively small 3JH-7,H-8 value was observed from both 2 (3.5 Hz) and 3 (3.2 Hz) (Table 1). A positive Cotton effect at 241 nm from ECD spectrum of 2 indicated an 8S configuration of 2 and the absence of any characteristic Cotton effect from the ECD spectrum of 3, suggesting 3 as a racemic mixture (Figure 2D).
Compound 9 was isolated as a colorless gum with a molecular formula of C28H38O13. The 1H NMR spectrum of 9 showed two 1,3,4,5-tetrasubstituted benzene rings (δH 6.76 (2H, s, H-2′,6′) and 6.75 (2H, s, H-2,6)), two oxygenated methines (δH 5.06 (1H, d, J = 8.4 Hz, H-7′) and 5.03 (1H, d, J = 8.4 Hz, H-7)), two oxygenated methylenes (δH 3.82 (1H, m, H-9a), 3.78 (1H, dd, J = 11.2, 5.2 Hz, H-9′a), 3.69 (1H, dd, J = 11.2, 4.4 Hz, H-9′b), and 3.57 (1H, dd, J = 9.9, 4.9 Hz, H-9′b)), two methines (2.51 (1H, m, H-8) and 2.38 (1H, m, H-8′)), four methoxy groups (δH 3.90 (12H, s, 3,5,3′,5′-OCH3)), and an α-rhamnopyranosyl unit (δH 4.70 (1H, d, J = 1.3 Hz, H-1′′), 3.81 (1H, m, H-2′′), 3.65 (1H, dd, J = 4.5, 3.2 Hz, H-3′′), 3.61 (1H, dq, J = 9.5, 6.5 Hz, H-5′′), 3.40 (1H, m, H-4′′), and 1.27 (3H, d, J = 6.5 Hz, H-6′′)). The 13C NMR spectrum of 2 revealed 28 resonances, comprising 12 benzene ring carbons [δC 149.5 (×4), 136.4, 136.3, 134.7, 134.3, 105.1 (×2), and 104.8 (×2)], four methoxy carbons (δC 57.0 (×4)), four oxygenated carbons (δC 85.2, 84.5, 67.8, and 61.6), two methine carbons (δC 54.6 and 51.8), and a rhamnopyranosyl unit (δC 102.3, 74.1, 72.7, 72.3, 70.3, and 18.2). These spectroscopic data were almost identical with those of 7S,7′S,8R,8′R-icariol A2 9′-O-α-l-rhamnopyranoside (25) [30]. Intensive 2D NMR data analysis (Figure 3A) confirmed that 9 and 25 have the same the planar structures. The absolute configuration of the known compound 25 was determined as 7S,7′S,8R,8′R by observing the negative Cotton effect around 240–250 nm in the ECD spectrum (Figure S9) [30], however, the new compound 9 showed the positive Cotton effect around that area (246 nm, Figure 3B), indicating 7R,7′R,8S,8′S-configuration. In addition, the opposite sign of the specific rotation values of 9, [α] D 25 + 30 (c 0.1, MeOH), and 25, [α] D 25 − 51.4 (c 0.1, MeOH) [30], supported the initial assignment. Finally, the l-form of rhamnopyranosyl moiety was assigned by LC-MS analysis on the chiral derivatized monosaccharide obtained by hydrolysis of 9 and following chemical reaction [17]. The retention time of derivatized rhamnopyranose (25.4 min) was matched with that of standard l-rhamnopyranose (25.3 min) rather than d-rhamnopyranose (13.2 min) (Figure 2C). Thus, the structure of 9 was determined as 7R,7′R,8S,8′S-icariol A2 9′-O-α-l-rhamnopyranoside.
Compound 11 was obtained as a colorless gum, and its molecular formula was deduced as C17H24O10 based upon the HRFABMS data. The 1H and 13C NMR spectra of 11 were comparable to those of 3-hydroxyl-1-(4-hydroxy-3,5-dimethoxyphenyl)-1-propanone (or β-hydroxypropiosyringone, 12) [31], except for the presence of resonances for a β-glucopyranosyl unit (δH 5.00 (1H, d, J = 7.5 Hz, H-1′), 3.67 (1H, dd, J = 11.9, 2.3 Hz, H-6′a), 3.55 (1H, dd, J = 11.9, 5.3 Hz, H-6′b), 3.40 (1H, dd, J = 8.8, 7.5 Hz, H-2′), 3.33 (1H, t, J = 8.8 Hz, H-3′), 3.30 (1H, t, J = 8.8 Hz, H-4′), 3.12 (1H, ddd, J = 8.8, 5.3, 2.3 Hz, H-5′); δC 104.6, 78.6, 78.0, 75.9, 71.5, and 62.7). On the basis of the HMBC cross peak of H-1′ with C-4, the location of the β-glucopyranosyl unit was confirmed to be connected at C-4 by an ether bond (Figure 4A). Acid hydrolysis of 11 afforded glucopyranose, which was identified as d-form by the same method as described above (Figure 4B). Thus, the structure of 11 was elucidated as β-hydroxypropiosyringone 4′-O-β-d-glucopyranoside.
The other known compounds were identified as (7S,8R)-1-[4-(O-β-d-glucopyranosyl)-3-methoxyphenyl]-2-[4-(3-hydroxypropyl)-2,6-dimethoxyphenoxyl]-1,3-propanediol (4) [32], 3-[4-[(1R,2S)-2-hydroxy-2-(4-hydroxy-3-methoxyphenyl)-1-(hydroxymethyl)ethoxy]-3-methoxyphenyl]propyl β-d-glucopyranoside (5) [27], syringatesinol (6) [33], pinoresinol-4-O-glucoside (7) [34], 8-hydroxypinoresinol-4′-O-β-d-glucose (8) [35], (–)-lyoniresinol 3α-O-β-d-glucopyranoside (10) [36], 3-hydroxyl-1-(4-hydroxy-3,5-dimethoxyphenyl)-1-propanone (12) [31], (2R)-O-[4′-(3″-hydroxypropyl)-2′,6′-dimethoxyphenyl]-3-O-β-d-glucopyranosyl-sn-glycerol (13) [37], 4-allyl-2,6-dimethoxyphenyl 1-O-β-d-xylopyranosyl-(1→6)-β-d-glucopyranoside (14) [38], 4-allyl-2-hydroxyphenyl 1-O-β-d-apiosyl-(1→6)-β-d-glucoside (15) [39], 4-allyl-2-methoxyphenyl 6-O-β-d-apiosyl-(1→6)-β-d-glucoside (16) [40], 3,5-dimethoxy-1-hydroxy-4-O-β-d-glucoside (17) [41], 3,4,5-trimethoxyphenol 1-O-β-d-glucopyranoside (18) [42], 3,4,5-trimethoxyphenyl 1-O-β-d-apiofuranosyl-(1′′→6′)-β-d-glucoside (19) [43], tachioside (20), isotachioside (21) [44], vanillyl alcohol-4-O-β-d-glucopyranoside (22) [45], benzyl β-d-xylopyranosyl-(1→6)-β-d-glucopyranoside (23) [46], and 4-hydroxy-2,6-dimethoxyphenyl 6′-O-vanilloyl-β-d-glucopyranoside (24) [47] by comparing their spectroscopic data with reported data.

3.2. Anti-Neuroinflammatory Activity of the Isolated Compounds (124)

Alzheimer’s disease, a neurodegenerative disease, is an unsolved social and medical problem and is strongly associated with neurotrophins (e.g., NGF, brain-derived neurotrophic factor (BDNF), neurotrohpin-3 (NT-3), and NT-4/5) and neuroinflammatory mediators (e.g., nitric oxide (NO), tumor necrosis factor alpha (TNF-α), interleukin (IL)-1β, nuclear factor kappa B (NF-κB), and prostaglandin (PG)E2) [48]. Therefore, many phytochemicals have been tested to discover novel anti-neurodegenerative agents, which revealed that phenolic constituents are one of the major neurotrophic and anti-neuroinflammatory compounds [48,49]. In line with this finding, the isolated phenolic compounds in this study (124) were assessed for their anti-neuroinflammatory activities by measuring NO levels in LPS-activated BV-2 cells (Table 3) and for their neurotrophic effects evaluated by assessing the secretion of NGF from C6 cells (Table 4). Compounds 13, 8–O–4′ type neolignans, significantly reduced NO levels with IC50 values of 18.63–19.75 μM, which showed better activity than a well-known inhibitor of NO synthase (NOS), NG-monomethyl-l-arginine (l-NMMA, IC50 21.35 μM) [50]. These data were consistent with the previous studies in which diverse 8–O–4′ type neolignans exhibited potent NO inhibitory activity [49]. However, compounds 4 and 5, the other 8–O–4′ type neolignans, were not active (>100 μM) in this study, which might be attributed to the different configuration at C-7 and/or the glucopyranosyl unit attached to the molecules. The most potent compound in this study was 6, a furanofuran-type lignan, with IC50 value of 12.69 μM, whereas the other glucosylated furanofuran-type lignans 7 and 8 did not display NO inhibitory activity. The simple phenolic glucoside 21 was a potent NO inhibitor (IC50 15.00 μM) but transfer of the methoxy group at C-3 to C-2 decreased the potency (IC50 41.67 μM, 20). The other compounds showed weak or no activity.

3.3. Neurotrophic Activity of the Isolated Compounds (124)

Similar to the NO inhibitory activity profile of the 8–O–4′ type neolignans, only 13 displayed effect on NGF release with stimulation levels of 110.27 ± 2.18%, 127.98 ± 3.95%, and 149.89 ± 1.97%, respectively, while the latter was comparable to that of positive control, 6-shogaol (149.53 ± 5.36%) [51]. The addition of a ketone group at C-7′ (1) and removal of the methoxy group at C-5 (2) in 3 decreased NGF secretion activity. Compound 20 showed powerful NGF release effect with stimulation levels of 157.73 ± 1.50% but this compound was cytotoxic to the tested C6 cells (cell viability: 11.83 ± 8.24%). Compounds 6, 12, and 21 were mild neurotrophic agents (120.09 ± 2.72%, 135.10 ± 5.59%, and 125.00 ± 1.50%, respectively) and the other compounds were inactive (<100%).

3.4. Cytotoxicity of the Isolated Compounds (124)

Podophyllotoxin is a representative anticancer natural product with a lignan scaffold and its synthetic derivative etoposide (VePesid®) is chemotherapy medication used for the treatments of diverse types of cancer [2]. Since most of the isolated phytochemicals (1–24) are derivatives of lignan or its monomer, phenylpropanoid, their cytotoxic activity against four human cancer cell lines, A549 (non-small-cell lung adenocarcinoma), SK-OV-3 (ovary malignant ascites), SK-MEL-2 (skin melanoma), and MKN-1 (adenosquamous carcinoma) were evaluated via the SRB bioassay method with etoposide as a positive control substance (IC50 0.98, 2.15, 1.80, and 3.47 μM, respectively). As a result, four 8–O–4′ type neolignans, 14, and a phenolic glucoside 24 showed an inhibitory effect on MKN-1 cancer cell with IC50 values of 11.15, 14.03, 14.16, 7.96, and 19.04 μM, respectively (Table 5).

4. Conclusions

A total of 24 phenolic compounds (124) including three new compounds (1, 9, and 11) were isolated and characterized from the twigs of C. sinensis. All isolates were evaluated for their anti-neuroinflammatory, neurotrophic, and cytotoxic activities. As a result, compounds 13, 8–O–4′ type neolignans, displayed significant results in all activity evaluations. Compounds 6, 12 and 21 showed anti-neuroinflammatory effects and/or neurotrophic activity. Compounds 24 displayed inhibitory effect on MKN-1 cancer cell. Through these results, it was suggested that the phenolic compounds in C. sinensis would be responsible for one of the various activities of this plant which could be the basis for the traditional use of C. sinensis for inflammatory diseases. Especially, 8–O–4′ type neolignans showed promising multi-target potentials as a drug against neurodegenerative diseases.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/antiox10040551/s1, General Experimental Procedures: Detailed isolation process. Scheme S1. Fractionation and isolation process of compounds 124 from C. sinensis twigs. Figures S1–S7: HRMS, NMR, and ECD spectra of 1, Figures S8–S14: HRMS, NMR, and ECD spectra of 9. Figure S15: Extracted ion chromatograms (EICs, m/z 431.1311) of chiral derivatized l- and d-rhamnopyranose purchased or obtained by hydrolysis of 9. Figures S16–S21: HRMS and NMR spectra of 11. Figure S22: Extracted ion chromatograms (EICs, m/z 447.1260) of chiral derivatized d- and l-glucopyranose purchased or obtained by hydrolysis of 11. Figures S23–S26: 1H and 13C NMR spectra of 2 and 3.

Author Contributions

D.H.K. performed the experiments, analyzed the data, and wrote the manuscript. L.S., and H.R.K. performed the experiments and analyzed the data. S.Y.K. and S.U.C. oversaw the biological assays. C.S.K. conceived the study, oversaw experiments, and wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Research Foundation (NRF) of Korea grant funded by the government of Korea (MSIT) (No. 2012M3A9C4048775) and by the BK21 FOUR Project.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

We are thankful to the Korea Basic Science Institute (KBSI) for the measurements on mass spectra.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Crozier, A.; Jaganath, I.B.; Clifford, M.N. Dietary phenolics: Chemistry, bioavailability and effects on health. Nat. Prod. Rep. 2009, 26, 1001–1043. [Google Scholar] [CrossRef] [PubMed]
  2. Walsh, C.T.; Tang, Y. Natural Product Biosynthesis; Royal Society of Chemistry: London, UK, 2017; pp. 357–410. [Google Scholar]
  3. Moon, M.; Kim, H.G.; Choi, J.G.; Oh, H.; Lee, P.K.; Ha, S.K.; Kim, S.Y.; Park, Y.; Huh, Y.; Oh, M.S. 6-Shogaol, an active constituent of ginger, attenuates neuroinflammation and cognitive deficits in animal models of dementia. Biochem. Biophys. Res. Commun. 2014, 449, 8–13. [Google Scholar] [CrossRef] [PubMed]
  4. Wyss-Coray, T. Ageing, neurodegeneration and brain rejuvenation. Nature 2016, 539, 180–186. [Google Scholar] [CrossRef] [PubMed]
  5. Liu, Z.; Zhang, A.; Sun, H.; Han, Y.; Kong, L.; Wang, X. Two decades of new drug discovery and development for Alzheimer’s disease. RSC Adv. 2017, 7, 6046–6058. [Google Scholar] [CrossRef] [Green Version]
  6. Mihara, S.; Tateba, H.; Nishimura, O.; Machii, Y.; Kishino, K. Volatile components of Chinese quince (Pseudocydonia sinensis Schneid). J. Agric. Food. Chem. 1987, 35, 532–537. [Google Scholar] [CrossRef]
  7. Korea Institute of Oriental Medicine. MEDICLASSICS. Available online: https://mediclassics.kr/books/8/volume/6#content_1437; https://mediclassics.kr/books/8/volume/8#content_396; https://mediclassics.kr/books/8/volume/13#content_64 (accessed on 15 January 2021).
  8. Lee, M.H.; Son, Y.K.; Han, Y.N. Tissue factor inhibitory flavonoids from the fruits ofChaenomeles sinensis. Arch. Pharm. Res. 2002, 25, 842–850. [Google Scholar] [CrossRef] [PubMed]
  9. Lee, M.H.; Han, Y.N. A new in vitro tissue factor inhibitory triterpene from the fruits of Chaenomeles sinensis. Planta Med. 2003, 69, 327–331. [Google Scholar] [CrossRef] [PubMed]
  10. Oku, H.; Ueda, Y.; Ishiguro, K. Antipruritic effects of the fruits of Chaenomeles sinensis. Biol. Pharm. Bull. 2003, 26, 1031–1034. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Osawa, K.; Arakawa, T.; Shimura, S.; Takeya, K. New quinic acid derivatives from the fruits of Chaenomeles sinensis (Chinese quince). Nat. Med. 2001, 55, 255–257. [Google Scholar]
  12. Lianna, S.; Yongfu, H. Chemical constituents of Chaenomeles sinensis (Thouin.) Koehne. J. Chin. Pharm. Sci. 2000, 9, 6–9. [Google Scholar]
  13. Kim, H.-K.; Jeon, W.-K.; Ko, B.-S. Flavanone glycoside from the fruits of Chaenomeles sinensis. Nat. Prod. Sci. 2000, 6, 79–81. [Google Scholar]
  14. Shin, J.-E.; Jin, Q.-L.; Jin, H.-G.; Woo, E.-R. Phytochemical constituents Isolated from the stems of Chaenomeles sinensis Koehne. Kor. J. Pharmacogn. 2011, 42, 223–228. [Google Scholar]
  15. Gao, H.; Wu, L.; Kuroyanagi, M.; Harada, K.; Kawahara, N.; Nakane, T.; Umehara, K.; Hirasawa, A.; Nakamura, Y. Antitumor-promoting constituents from Chaenomeles sinensis KOEHNE and their activities in JB6 mouse epidermal cells. Chem. Pharm. Bull. 2003, 51, 1318–1321. [Google Scholar] [CrossRef] [Green Version]
  16. Osawa, K.; Yasuda, H.; Morita, H.; Takeya, K.; Itokawa, H. Antibacterial and antihemolytic activity of triterpenes and β-sitosterol isolated from Chinese Quince (Chaenomeles sinensis). Nat. Med. 1997, 51, 365–367. [Google Scholar]
  17. Tanaka, T.; Nakashima, T.; Ueda, T.; Tomii, K.; Kouno, I. Facile discrimination of aldose enantiomers by reversed-phase HPLC. Chem. Pharm. Bull. 2007, 55, 899–901. [Google Scholar] [CrossRef] [Green Version]
  18. Kim, C.S.; Oh, J.; Suh, W.S.; Jang, S.W.; Subedi, L.; Kim, S.Y.; Choi, S.U.; Lee, K.R. Investigation of chemical constituents from Spiraea prunifolia var. simpliciflora and their biological activities. Phytochem. Lett. 2017, 22, 255–260. [Google Scholar] [CrossRef]
  19. Blasi, E.; Barluzzi, R.; Bocchini, V.; Mazzolla, R.; Bistoni, F. Immortalization of murine microglial cells by a v-raf/v-myc carrying retrovirus. J. Neuroimmunol. 1990, 27, 229–237. [Google Scholar] [CrossRef]
  20. Choi, Y.; Lee, M.K.; Lim, S.Y.; Sung, S.H.; Kim, Y.C. Inhibition of inducible NO synthase, cyclooxygenase-2 and interleukin-1β by torilin is mediated by mitogen-activated protein kinases in microglial BV2 cells. Br. J. Pharmacol. 2009, 156, 933–940. [Google Scholar] [CrossRef] [Green Version]
  21. Kim, C.S.; Subedi, L.; Oh, J.; Kim, S.Y.; Choi, S.U.; Lee, K.R. Bioactive Triterpenoids from the Twigs of Chaenomeles sinensis. J. Nat. Prod. 2017, 80, 1134–1140. [Google Scholar] [CrossRef]
  22. Skehan, P.; Storeng, R.; Scudiero, D.; Monks, A.; McMahon, J.; Vistica, D.; Warren, J.T.; Bokesch, H.; Kenney, S.; Boyd, M.R. New colorimetric cytotoxicity assay for anticancer-drug screening. JNCI J. Natl. Cancer Inst. 1990, 82, 1107–1112. [Google Scholar] [CrossRef]
  23. Jutiviboonsuk, A.; Zhang, H.; Tan, G.T.; Ma, C.; Van Hung, N.; Cuong, N.M.; Bunyapraphatsara, N.; Soejarto, D.D.; Fong, H.H. Bioactive constituents from roots of Bursera tonkinensis. Phytochemistry 2005, 66, 2745–2751. [Google Scholar] [CrossRef] [PubMed]
  24. Gan, M.; Zhang, Y.; Lin, S.; Liu, M.; Song, W.; Zi, J.; Yang, Y.; Fan, X.; Shi, J.; Hu, J. Glycosides from the root of Iodes cirrhosa. J. Nat. Prod. 2008, 71, 647–654. [Google Scholar] [CrossRef] [PubMed]
  25. Huo, C.; Liang, H.; Zhao, Y.; Wang, B.; Zhang, Q. Neolignan glycosides from Symplocos caudata. Phytochemistry 2008, 69, 788–795. [Google Scholar] [CrossRef] [PubMed]
  26. Cha, J.M.; Lee, T.H.; Subedi, L.; Ha, Y.J.; Kim, H.R.; Kim, S.Y.; Choi, S.U.; Kim, C.S. Isolation and structural characterization of four diastereomeric lignan glycosides from Abies holophylla and their neuroprotective activity. Tetrahedron 2020, 131735. [Google Scholar]
  27. Matsuda, N.; Kikuchi, M. Studies on the constituents of Lonicera species. X. Neolignan glycosides from the leaves of Lonicera gracilipes var. glandulosa MAXIM. Chem. Pharm. Bull. 1996, 44, 1676–1679. [Google Scholar] [CrossRef] [Green Version]
  28. Matsushita, H.; Miyase, T.; Ueno, A. Lignan and terpene glycosides from Epimedium sagittatum. Phytochemistry 1991, 30, 2025–2027. [Google Scholar] [CrossRef]
  29. Sakakibara, A.; Sudo, K.; Kishi, M.; Aoyama, M.; Hwang, B. Hydrogenolysis of protolignin. XVI. Isolation of β-O-4 and β-β type compounds. Mokuzai Gakkaishi 1980, 26, 628–632. [Google Scholar]
  30. Lee, I.-S.; Jung, S.-H.; Lee, Y.M.; Choi, S.-J.; Sun, H.; Kim, J.S. Phenolic compounds from the leaves and twigs of Osteomeles schwerinae that inhibit rat lens aldose reductase and vessel dilation in zebrafish larvae. J. Nat. Prod. 2015, 78, 2249–2254. [Google Scholar] [CrossRef]
  31. Nakasone, Y.; Takara, K.; Wada, K.; Tanaka, J.; Yogi, S.; Nakatani, N. Antioxidative compounds isolated from Kokuto, non-centrifugal cane sugar. Biosci. Biotechnol. Biochem. 1996, 60, 1714–1716. [Google Scholar] [CrossRef] [Green Version]
  32. Tang, W.-Z.; Liu, Y.-B.; Yu, S.-S.; Qu, J.; Su, D.-M. New sesquiterpene lactone and neolignan glycosides with antioxidant and anti-inflammatory activities from the fruits of Illicium oligandrum. Planta Med. 2007, 53, 484–490. [Google Scholar] [CrossRef]
  33. Son, Y.K.; Lee, M.H.; Han, Y.N. A new antipsychotic effective neolignan from firmiana simplex. Arch. Pharm. Res. 2005, 28, 34–38. [Google Scholar] [CrossRef]
  34. Kim, D.K.; Lim, J.P.; Kim, J.W.; Park, H.W.; Eun, J.S. Antitumor and antiinflammatory constituents Fromceltis sinensis. Arch. Pharm. Res. 2005, 28, 39–43. [Google Scholar] [CrossRef]
  35. Schumacher, B.; Scholle, S.; Hölzl, J.; Khudeir, N.; Hess, S.; Müller, C.E. Lignans isolated from valerian: Identification and characterization of a new olivil derivative with partial agonistic activity at A1 adenosine receptors. J. Nat. Prod. 2002, 65, 1479–1485. [Google Scholar] [CrossRef]
  36. Ohashi, K.; Watanabe, H.; Okumura, Y.; UJI, T.; Kitagawa, I. Indonesian medicinal plants. XII. Four isomaric lignan-glucosides from the bark of Aegle marmelos (Rutaceae). Chem. Pharm. Bull. 1994, 42, 1924–1926. [Google Scholar] [CrossRef] [Green Version]
  37. Zhou, Y.; Li, L.-Y.; Ren, H.-C.; Qin, R.-D.; Li, Q.; Tu, P.-F.; Dou, G.-F.; Zhang, Q.-Y.; Liang, H. Chemical constituents from the whole plants of Pilea cavaleriei Levl subsp. cavaleriei. Fitoterapia 2017, 119, 100–107. [Google Scholar] [CrossRef]
  38. Liu, Y.-F.; Shi, G.-R.; Wang, X.; Zhang, C.-L.; Wang, Y.; Chen, R.-Y.; Yu, D.-Q. Nine new compounds from the whole plants of Rehmannia chingii. J. Asian Nat. Prod. Res. 2016, 18, 509–519. [Google Scholar] [CrossRef]
  39. Yamauchi, H.; Kakuda, R.; Yaoita, Y.; Machida, K.; Kikuchi, M. Two new glycosides from the whole plants of Glechoma hederacea L. Chem. Pharm. Bull. 2007, 55, 346–347. [Google Scholar] [CrossRef] [Green Version]
  40. Machida, K.; Nakano, Y.; Kikuchi, M. Phenolic glycosides from Viburnum dilatatum. Phytochemistry 1991, 30, 2013–2014. [Google Scholar] [CrossRef]
  41. Chung, M.-I.; Lai, M.-H.; Yen, M.-H.; Wu, R.-R.; Lin, C.-N. Phenolics from Hypericum geminiflorum. Phytochemistry 1997, 44, 943–947. [Google Scholar] [CrossRef]
  42. Verotta, L.; Dell’Agli, M.; Giolito, A.; Guerrini, M.; Cabalion, P.; Bosisio, E. In vitro antiplasmodial activity of extracts of Tristaniopsis species and identification of the active constituents: Ellagic acid and 3, 4, 5-trimethoxyphenyl-(6‘-O-galloyl)-O-β-D-glucopyranoside. J. Nat. Prod. 2001, 64, 603–607. [Google Scholar] [CrossRef]
  43. Kanchanapoom, T.; Kasai, R.; Yamasaki, K. Iridoid and phenolic glycosides from Morinda coreia. Phytochemistry 2002, 59, 551–556. [Google Scholar] [CrossRef]
  44. Zhong, X.-N.; Otsuka, H.; Ide, T.; Hirata, E.; Takeda, Y. Hydroquinone diglycoside acyl esters from the leaves of Myrsine seguinii. Phytochemistry 1999, 52, 923–927. [Google Scholar] [CrossRef]
  45. Kanho, H.; Yaoya, S.; Kawahara, N.; Nakane, T.; Takase, Y.; Masuda, K.; Kuroyanagi, M. Biotransformation of benzaldehyde-type and acetophenone-type derivatives by Pharbitis nil hairy roots. Chem. Pharm. Bull. 2005, 53, 361–365. [Google Scholar] [CrossRef] [Green Version]
  46. Kawahara, E.; Fujii, M.; Kato, K.; Ida, Y.; Akita, H. Chemoenzymatic synthesis of naturally occurring benzyl 6-O-glycosyl-β-D-glucopyranosides. Chem. Pharm. Bull. 2005, 53, 1058–1061. [Google Scholar] [CrossRef] [Green Version]
  47. Lee, M.A.; Lee, H.K.; Kim, S.H.; Kim, Y.C.; Sung, S.H. Chemical constituents of Alnus firma and their inhibitory activity on lipopolysaccharide-induced nitric oxide production in BV2 microglia. Planta Med. 2010, 76, 1007–1010. [Google Scholar] [CrossRef] [Green Version]
  48. Venkatesan, R.; Ji, E.; Kim, S.Y. Phytochemicals that regulate neurodegenerative disease by targeting neurotrophins: A comprehensive review. BioMed Res. Int. 2015, 2015, 814068. [Google Scholar] [CrossRef]
  49. Teponno, R.B.; Kusari, S.; Spiteller, M. Recent advances in research on lignans and neolignans. Nat. Prod. Rep. 2016, 33, 1044–1092. [Google Scholar] [CrossRef] [Green Version]
  50. Ha, S.K.; Moon, E.; Ju, M.S.; Kim, D.H.; Ryu, J.H.; Oh, M.S.; Kim, S.Y. 6-Shogaol, a ginger product, modulates neuroinflammation: A new approach to neuroprotection. Neuropharmacology 2012, 63, 211–223. [Google Scholar] [CrossRef] [PubMed]
  51. Park, G.; Kim, H.G.; Ju, M.S.; Ha, S.K.; Park, Y.; Kim, S.Y.; Oh, M.S. 6-Shogaol, an active compound of ginger, protects dopaminergic neurons in Parkinson’s disease models via anti-neuroinflammation. Acta Pharmacol. Sin. 2013, 34, 1131–1139. [Google Scholar] [CrossRef]
Figure 1. Chemical structure of compounds 125.
Figure 1. Chemical structure of compounds 125.
Antioxidants 10 00551 g001
Figure 2. Structure elucidation of 1–3. (A) Key COSY and HMBC correlations of 1. (B) Coupling constants values among H-7, H-8, and H-9 (left) deduced by analysis of 1H NMR peaks of H-7 (middle) and H-8 (right). (C) Zoomed-in 13C NMR spectra of 1 for C-7 and C-8. (D) ECD spectra of 1–3. The negative or positive Cotton effect around 240–245 nm indicated the 8R- or 8S-configuration, respectively.
Figure 2. Structure elucidation of 1–3. (A) Key COSY and HMBC correlations of 1. (B) Coupling constants values among H-7, H-8, and H-9 (left) deduced by analysis of 1H NMR peaks of H-7 (middle) and H-8 (right). (C) Zoomed-in 13C NMR spectra of 1 for C-7 and C-8. (D) ECD spectra of 1–3. The negative or positive Cotton effect around 240–245 nm indicated the 8R- or 8S-configuration, respectively.
Antioxidants 10 00551 g002
Figure 3. Structure elucidation of 9. (A) Key COSY and HMBC correlations of 9. (B) Extracted ion chromatograms (EICs, m/z 431.1311) of chiral derivatized l- and d-rhamnopyranose purchased or obtained by hydrolysis of 9. (C) ECD spectrum of 9. The positive Cotton effect at 246 nm indicated the 7R,7′R,8S,8′S-configuration.
Figure 3. Structure elucidation of 9. (A) Key COSY and HMBC correlations of 9. (B) Extracted ion chromatograms (EICs, m/z 431.1311) of chiral derivatized l- and d-rhamnopyranose purchased or obtained by hydrolysis of 9. (C) ECD spectrum of 9. The positive Cotton effect at 246 nm indicated the 7R,7′R,8S,8′S-configuration.
Antioxidants 10 00551 g003
Figure 4. Structure elucidation of 11. (A) Key COSY and HMBC correlations of 11. (B) Extracted ion chromatograms (EICs, m/z 447.1260) of chiral derivatized d- and l-glucopyranose purchased or obtained by hydrolysis of 11.
Figure 4. Structure elucidation of 11. (A) Key COSY and HMBC correlations of 11. (B) Extracted ion chromatograms (EICs, m/z 447.1260) of chiral derivatized d- and l-glucopyranose purchased or obtained by hydrolysis of 11.
Antioxidants 10 00551 g004
Table 1. 1H (700 MHz) and 13C (175 MHz) NMR data of compounds 13, and 9.
Table 1. 1H (700 MHz) and 13C (175 MHz) NMR data of compounds 13, and 9.
Pos.1 12 13 19 2
δCδH (multi, J in Hz)δCδH (multi, J in Hz)δCδH (multi, J in Hz)δCδH (multi, J in Hz)
1130.3 130.6 134.3
2102.76.58 (s) 6.97 (d, 1.6)102.66.59 (s)105.16.75 (s)
3147.2 147.2 149.5
4134.2 134.0 136.4
5147.2 6.86 (d, 8.1)147.2 149.5
6102.76.58 (s) 6.75 (dd, 8.1, 1.6)102.66.59 (s)105.16.75 (s)
773.14.99 (brs)72.65.00 (d, 3.5)72.74.99 (d, 3.2)85.25.03 (d, 8.4)
887.54.23 (ddd, 6.6, 3.5, 3.0)87.24.11 (dt, 6.7, 2.9)87.24.10 (dt, 6.8, 3.1)51.82.51 (m)
9a60.73.93 (overlap)60.83.87 (overlap)60.83.87 (overlap)67.83.82 (m)
9b 3.54 (m) 3.48 (overlap) 3.46 (m) 3.57 (dd, 9.9, 4.9)
1′133.1 133.1 133.1 134.7
2′/6′105.67.29 (s)105.56.48 (s)105.56.49 (s)104.86.76 (s)
3′/5′153.6 153.3 153.3 149.5
4′139.9 138.8 138.9 136.3
7′199.1 34.42.70 (m)34.42.71 (m)84.55.06 (d, 8.4)
8′40.43.23 (t, 5.4)32.81.91 (m)32.81.92 (m)54.62.38 (m)
9′a58.34.06 (t, 5.3)62.33.71 (t, 6.3)62.33.72 (t, 6.3)61.63.78 (dd, 11.2, 5.2)
9′b 3.69 (dd, 11.2, 4.4)
1′′ 102.34.70 (d, 1.3)
2′′ 72.33.81 (m)
3′′ 72.73.65 (dd, 4.5, 3.2)
4′′ 74.13.40 (m)
5′′ 70.33.61 (dq, 9.5, 6.5)
6′′ 18.21.27 (d, 6.5)
3-OCH356.53.88 (s)56.13.90 (s)56.53.88 (s)57.03.90 (s)
5-OCH356.53.88 (s) 56.53.88 (s)57.03.90 (s)
3′,5′-OCH356.63.96 (s)56.33.87 (s)56.33.88 (s)57.03.90 (s)
4-OH 5.47 (s) 5.55 (s) 5.45 (s)
1 Measured in chloroform-d; 2 Measured in methanol-d4.
Table 2. 1H (700 MHz) and 13C (175 MHz) NMR data of compound 11 in methanol-d4.
Table 2. 1H (700 MHz) and 13C (175 MHz) NMR data of compound 11 in methanol-d4.
Pos.11
δCδH (multi, J in Hz)
1134.6
2/6107.67.23 (s)
3/5154.4
4140.7
7200.0
842.13.11 (t, 6.1)
958.83.86 (t, 6.1)
1′104.65.00 (d, 7.5)
2′75.93.40 (dd, 8.8, 7.5)
3′78.03.33 (t, 8.8)
4′71.53.30 (t, 8.8)
5′78.63.12 (ddd, 8.8, 5.3, 2.3)
6′a62.73.67 (dd, 11.9, 2.3)
6′b 3.55 (dd, 11.9, 5.3)
3,5-OCH357.33.81 (s)
Table 3. Inhibitory effects of selected compounds on nitric oxide (NO) production in lipopolysaccharide (LPS)-activated BV-2 cells.
Table 3. Inhibitory effects of selected compounds on nitric oxide (NO) production in lipopolysaccharide (LPS)-activated BV-2 cells.
CompoundIC50 (μM) 1Cell viability (%) 2
118.66113.97 ± 1.36
219.75105.99 ± 7.78
318.6394.33 ± 3.84
612.69101.33 ± 8.15
997.5097.67 ± 3.20
1272.50114.97 ± 5.17
1768.77106.59 ± 1.27
2041.67103.50 ± 2.62
2115.00103.50 ± 2.62
2470.0894.46 ± 10.89
l-NMMA 321.35104.56 ± 4.20
1 IC50 value of each compound was defined as the concentration (μM) that caused 50% inhibition of NO production in LPS-activated BV-2 cells. 2 Cell viability after treatment with 20 μM of each compound was determined by 3-[4,5-dimethylthiazol-2-yl]-2,5-diphenyltetrazolium bromide (MTT) assay and is expressed in percentage (%). The results are averages of three independent experiments, and the data are expressed as mean ± SD. 3 NG-monomethyl-l-arginine (l-NMMA) as positive control.
Table 4. Effects of selected compounds on nerve growth factor (NGF) secretion in C6 cells.
Table 4. Effects of selected compounds on nerve growth factor (NGF) secretion in C6 cells.
CompoundNGF Secretion 1Cell Viability (%) 2
1110.27 ± 2.1896.59 ± 2.78
2127.98 ± 3.95101.75 ± 2.40
3149.89 ± 1.97105.00 ± 3.21
6120.09 ± 2.72109.25 ± 8.03
12135.10 ± 5.5994.41 ± 1.83
20157.73 ± 1.5011.83 ± 8.24
21125.00 ± 1.50117.97 ± 8.50
6-shogaol 3149.53 ± 5.3697.00 ± 0.17
1 C6 cells were treated with 20 μM of compounds. After 24 h, the content of NGF secretion in C6-conditioned media was measured by ELISA. The level of secreted NGF cells is expressed as percentage of the untreated control. The data shown represent the means ± SD of three independent experiments performed in triplicate. 2 Cell viability after treatment with 20 μM of each compound was determined by MTT assay and is expressed in percentage (%). The results are averages of three independent experiments, and the data are expressed as mean ± SD. 3 6-shogaol as positive control.
Table 5. Cytotoxicity of selected compounds against four cultured human cell lines.
Table 5. Cytotoxicity of selected compounds against four cultured human cell lines.
CompoundIC50 (μM) 1
A-549SK-OV-3SK-MEL-2MKN-1
1>30.0>30.0>30.011.15
2>30.0>30.0>30.014.03
3>30.0>30.0>30.014.16
4>30.0>30.0>30.07.96
24>30.0>30.0>30.019.04
Etoposide 20.982.151.803.47
1 50% inhibitory concentration; the concentration of compound that caused a 50% inhibition in cell growth. 2 Etoposide as positive control.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kim, D.H.; Subedi, L.; Kim, H.R.; Choi, S.U.; Kim, S.Y.; Kim, C.S. Phenolic Constituents of Chinese Quince Twigs (Chaenomeles sinensis Koehne) and Their Anti-Neuroinflammatory, Neurotrophic, and Cytotoxic Activities. Antioxidants 2021, 10, 551. https://doi.org/10.3390/antiox10040551

AMA Style

Kim DH, Subedi L, Kim HR, Choi SU, Kim SY, Kim CS. Phenolic Constituents of Chinese Quince Twigs (Chaenomeles sinensis Koehne) and Their Anti-Neuroinflammatory, Neurotrophic, and Cytotoxic Activities. Antioxidants. 2021; 10(4):551. https://doi.org/10.3390/antiox10040551

Chicago/Turabian Style

Kim, Dong Hyun, Lalita Subedi, Hye Ryeong Kim, Sang Un Choi, Sun Yeou Kim, and Chung Sub Kim. 2021. "Phenolic Constituents of Chinese Quince Twigs (Chaenomeles sinensis Koehne) and Their Anti-Neuroinflammatory, Neurotrophic, and Cytotoxic Activities" Antioxidants 10, no. 4: 551. https://doi.org/10.3390/antiox10040551

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