Next Article in Journal
Piperlongumine Inhibits Thioredoxin Reductase 1 by Targeting Selenocysteine Residues and Sensitizes Cancer Cells to Erastin
Previous Article in Journal
Effects of Volatile Anaesthetics and Iron Dextran on Chronic Inflammation and Antioxidant Defense System in Rats
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Polyphenolic Compounds from Lespedeza bicolor Protect Neuronal Cells from Oxidative Stress

by
Darya V. Tarbeeva
1,*,
Evgeny A. Pislyagin
1,
Ekaterina S. Menchinskaya
1,
Dmitrii V. Berdyshev
1,
Anatoliy I. Kalinovskiy
1,
Valeria P. Grigorchuk
2,
Natalia P. Mishchenko
1,
Dmitry L. Aminin
1 and
Sergey A. Fedoreyev
1
1
G.B. Elyakov Pacific Institute of Bioorganic Chemistry, Far Eastern Branch, Russian Academy of Science, 690022 Vladivostok, Russia
2
Federal Scientific Center of the East Asia Terrestrial Biodiversity, Far Eastern Branch, Russian Academy of Sciences, 690022 Vladivostok, Russia
*
Author to whom correspondence should be addressed.
Antioxidants 2022, 11(4), 709; https://doi.org/10.3390/antiox11040709
Submission received: 6 March 2022 / Revised: 29 March 2022 / Accepted: 1 April 2022 / Published: 3 April 2022

Abstract

:
Pterocarpans and related polyphenolics are known as promising neuroprotective agents. We used models of rotenone-, paraquat-, and 6-hydroxydopamine-induced neurotoxicity to study the neuroprotective activity of polyphenolic compounds from Lespedeza bicolor and their effects on mitochondrial membrane potential. We isolated 11 polyphenolic compounds: a novel coumestan lespebicoumestan A (10) and a novel stilbenoid 5’-isoprenylbicoloketon (11) as well as three previously known pterocarpans, two pterocarpens, one coumestan, one stilbenoid, and a dimeric flavonoid. Pterocarpans 3 and 6, stilbenoid 5, and dimeric flavonoid 8 significantly increased the percentage of living cells after treatment with paraquat (PQ), but only pterocarpan 6 slightly decreased the ROS level in PQ-treated cells. Pterocarpan 3 and stilbenoid 5 were shown to effectively increase mitochondrial membrane potential in PQ-treated cells. We showed that pterocarpans 2 and 3, containing a 3’-methyl-3’-isohexenylpyran ring; pterocarpens 4 and 9, with a double bond between C-6a and C-11a; and coumestan 10 significantly increased the percentage of living cells by decreasing ROS levels in 6-OHDA-treated cells, which is in accordance with their rather high activity in DPPH and FRAP tests. Compounds 9 and 10 effectively increased the percentage of living cells after treatment with rotenone but did not significantly decrease ROS levels.

1. Introduction

Parkinson’s disease (PD) is a common neurodegenerative disease of older age [1]. The pathogenesis of PD includes the death of neurons from oxidative damage as a result of an increase in the production of intracellular reactive oxygen species (ROS), which causes damage to lipids, proteins, and DNA. A combination of oxidative stress, mitochondrial dysfunction, protein misprocessing, and genetic factors plays a crucial role in the pathogenesis of age-related neurodegeneration [1,2]. The etiology of PD can be associated not only with aging but also with adverse environmental influences and neurotoxins. 6-hydroxydopamine (6-OHDA) and 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) were shown to cause PD symptoms [3]. Several pesticides and herbicides (rotenone, paraquat (PQ), maneb (MB), and mancozeb (MZ)) are also neurotoxins and cause pathologies similar to PD [3]. The neurotoxic properties of these compounds are primarily due to their ability to generate toxic free radicals and reactive oxygen species in cells. For example, PQ affects the redox cycle and activity of the enzyme nitric oxide synthase in neuronal cells, which leads to increased production of ROS and increased levels of α-synuclein and tau protein, α-tubulin hyperacetylation, inhibition of proteasomes, and dysfunction of axonal autophagy. All these factors eventually cause PD symptoms [3]. Being a structural analogue of dopamine, 6-OHDA selectively penetrates dopaminergic and noradrenergic neurons, accumulates in the cytosol, and changes into dihydrophenylacetic acid or oxidizes to form hydrogen peroxide and para-quinone, which leads to ROS formation and oxidative stress in neurons, followed by cell death [3]. Rotenone easily crosses all cellular membranes and inhibits the transition of electrons from the iron-sulfur centers in complex I to ubiquinone. By doing so, rotenone inhibits reduced nicotinamide adenine dinucleotide (NADH)-ubiquinone reductase activity and impairs oxidative phosphorylation [3]. Thus, these neurotoxins can be used to cause neuronal cell damage in models of PD [3,4,5]. Being natural antioxidants, flavonoids and isoflavonoids are promising for the study of their neuroprotective properties [6].
Among isoflavonoids, pterocarpans are the second largest group. They contain a tetracyclic ring system derived from the flavonoid skeleton and have two chiral centers at carbon atoms C-6a and C-11a [7]. Pterocarpans have been reported to exhibit numerous bioactivities, including antioxidant, antineuroinflammatory, antimalarial, anticancer, and antimicrobial effects [7,8]. Pterocarpans have been reported to possess remarkable neuroprotective properties. For example, maackiain, widespread in plants of the Fabaceae family, was shown to significantly reduce dopaminergic neuron damage in 6-OHDA-exposed worms and to diminish the accumulation of α-synuclein [9].
The neuroprotective activity of pterocarpans may be due to their ability to inhibit two isoforms of monoamine oxidase: MAO-A and MAO-B. Maackiain and medicarpin selectively inhibited MAO-B, with one of the lowest IC50 values reported so far [10,11], but they did not effectively inhibit MAO-A. Meanwhile, (–)-4-hydroxy-3-methoxy-8,9-methylenedioxypterocarpan was found to effectively inhibit both the MAO-A and MAO-B isoforms [10]. Maackiain also exhibited inhibitory effects on LPS-induced nitric oxide production in RAW 264.7 macrophages [12] and inhibited 5-lipoxygenase [13], thus preventing damage to dopaminergic neurons [12,13,14].
Naturally occurring coumestans contain an additional carbonyl group compared to pterocarpans. These compounds also showed protective effects against neuronal damage, hypoxia–ischemia-induced cognitive impairment, and neurodegenerative disorders, including PD. Coumestrol attenuated the neurotoxicity induced by LPS and amyloid-beta peptide (Aβ) and inhibited the production of inflammatory cytokines TNF-α, IL-1, and IL-6 [15]. This compound not only provided effective neuroprotection in a cerebral global-ischemia model [16] but also modulated mitochondrial activity and decreased oxidative stress in rats [17,18].
The Far Eastern medicinal plant Lespedeza bicolor and other Lespedeza species produce unusual biologically active prenylated pterocarpans and coumestans, possessing strong antioxidant and neuroprotective activity [19,20,21,22,23,24,25,26,27,28]. Pterocarpan 1-methoxylespeflorin G11 from L. bicolor has recently been reported to exhibit neuroprotective effects against glutamate-induced neurotoxicity in neuronal HT22 cells. This neuroprotective activity was shown to be due to inhibition of oxidative stress and apoptosis [25]. We should note that arylbenzofuran and coumestan compounds did not protect HT22 cells from glutamate-induced cell death, whereas the pterocarpan-type compounds exhibited a significant neuroprotective effect against glutamate neurotoxicity [25].
Here, we studied the neuroprotective effect of polyphenolic compounds from L. bicolor growing in the Russian Far East using models of PQ-, 6-OHDA-, and rotenone-induced neurotoxicity; we also studied their effects on mitochondrial potential.

2. Materials and Methods

2.1. Plant Material

L. bicolor was collected by academician Gorovoy P.G. in Khasansky District (Andreevka village) of the Primorye Territory (The Russian Federation) in August 2016. A voucher specimen (No. 103608) was deposited into the herbarium of the Laboratory of Chemotaxonomy (G.B. Elyakov Pacific Institute of Bioorganic Chemistry, FEB RAS, Vladivostok, Russia).

2.2. Extraction and Isolation

Air-dried root bark of L. bicolor (330 g) was extracted for 3 h at 60 °C twice under reflux with a CHCl3–EtOH mixture (3:1, v/v). The dried extract (3.4 g) was subjected to a polyamide (100 g, 50–160 µm, Sigma-Aldrich, St. Louis, MI, USA) column (15 × 5 cm) and eluted with a hexane–CHCl3 solution system with gradually increasing CHCl3 percentage (hexane/CHCl3, 1:0, 100:1, 50:1, 40:1, v/v) to give fractions 1–16 and then with a CHCl3–EtOH solution system with gradually increasing EtOH amounts (CHCl3/EtOH, 1:0, 100:1, 50:1, 40:1, v/v) to give fractions 17–23.
We subsequently purified the fractions containing polyphenolic compounds according to HPLC data. Fraction 19 (CHCl3–EtOH (50:1), 408 mg) was chromatographed twice over a silica gel (40–63 µm, Sigma-Aldrich, St. Louis, MI, USA) column (11 × 1.4 cm). The column was eluted with a benzene–ethylacetate solution system with gradually increasing ethylacetate percentage (benzene/EtOAc, 1:0, 200:1, 100:1, 50:1, 40:1, v/v) to yield compound 4 (14.4 mg). Fraction 20 (215 mg), washed out with CHCl3–EtOH (40:1), was also chromatographed over a silica gel column using the same solution system to obtain compounds 1 (12.1 mg), 4 (5.5 mg), and 10 (3.7 mg). Fraction 22 (427 mg), eluted with CHCl3−EtOH (20:1), was chromatographed on a silica gel (40–63 µm) column (11×1.4 cm) using the same solution system to obtain compounds 5 (6.9 mg), 7 (2.9 mg), 8 (7.9 mg), and 11 (6.3 mg).
Fraction 15 (CHCl3, 193 mg) was subsequently subjected to a silica gel (40–63 µm) column (11 × 1.4 cm) and eluted with a benzene–ethylacetate solution system with gradually increasing ethylacetate percentage (benzene/EtOAc, 1:0, 200:1, 100:1, 50:1, 40:1, v/v) twice to give compounds 6 (7.2 mg) and 9 (3.6 mg). Fraction 16 (89 mg), washed out with CHCl3, was also further chromatographed over a silica gel (40–63 µm) column twice using the same eluent system to afford compounds 2 (7.2 mg) and 3 (9.0 mg).
A semi-preparative HPLC technique was used as a final step for purification of the isolated compounds.
Lespebicoumestan A (10): White, amorphous powder; UV (MeOH) λmax 211, 257, 358, 371 nm; ECD (3.99 × 10−4 M, CH3CN) λmax (Δε) 196 (+1.33), 216 (−0.77), 219 (−0.76), 280 (+0.17); 1H and 13C NMR data, see Table 1; HR-ESI-MS m/z 417.1337 [M-H] (calculated for [C25H21O6] 417.1344), m/z 419.1479 [M+H]+ (calculated for [C25H23O6]+ 419.1489).
5′-isoprenylbicoloketone A (11): Yellow, amorphous powder; UV (MeOH) λmax 220, 293, 345 nm; 1H and 13C NMR data, see Table 2; HR-ESI-MS m/z 493.2217 [M-H] (calculated for [C29H33O7] 493.2232), m/z 495.2401 [M+H]+ (calculated for [C29H35O7]+ 495.2377).

2.3. General Experimental Procedures

We recorded the UV spectra on a UV-1601 PC spectrophotometer (Shimadzu, Kyoto, Japan). The CD spectra were recorded using a Chirascan-Plus Quick Start CD Spectrometer (Applied Photophysics Limited, Leatherhead, UK) (acetonitrile, 20 °C). The 1H, 13C, and two-dimensional NMR spectra (HSQC, HMBC, COSY, ROESY) were recorded in CDCl3 on a Bruker AVANCE III DRX-700 NMR spectrometer (Bruker, Karlsruhe, Germany). The chemical shift values (δ) and the coupling constants (J) are given in parts per million and Hz, respectively.

2.4. Analytical and Semi-Preparative HPLC

We carried out analytical HPLC using an Agilent Technologies 1260 Infinity II HPLC system (Agilent Technologies, Waldbronn, Germany) equipped with a VWD detector (λ = 280 nm). The extracts and fractions were analyzed at a flowrate of 0.8 mL/min using an HS-C18 column (3 μm, 4.6, 75 mm, Supelco Analytical, Bellefonte, PA, USA). The column was thermostated at 30 °C. We used a mixture of 1% aqueous acetic acid (A) and acetonitrile containing 1% acetic acid (B) as mobile phase. We programmed the following gradient steps for elution: 0–2 min–5% B, 2–4 min–5–20% B, 5–17 min–20–50% B, 18–23 min–50–90% B, 24–25 min–90–100% B, 16–27 min–100% B, and 28–33 min–100–5% B. The data were analyzed using OpenLab CDC software v2.4 (Agilent Technologies, Waldbronn, Germany).
Semi-reparative HPLC was performed using a Shimadzu HPLC system equipped with an LC-20AT pump and SPD-20A detector (λ = 280 nm) (Shimadzu, Kyoto, Japan). We purified polyphenolic compounds using a silica gel YMC-Pack SIL (Supelco Analytical, Bellefonte, PA, USA) column (5 μm, 10, 250 mm) at a flowrate of 4.5 mL/min. The mobile phase consisted of hexane (97%) and isopropanol (3%). We used Shimadzu LCMS Solution software v5.93 (Shimadzu, Kyoto, Japan) to acquire and process the data.

2.5. HR-ESI-MS

We performed HR-ESI-MS analysis on a Shimadzu hybrid ion-trap–time-of-flight mass spectrometer (Shimadzu, Kyoto, Japan). The following instrument settings were applied for analysis: drying gas (N2) pressure—200 kPa; nebulizer gas (N2) flow—1.5 L/min; electrospray ionization (ESI) source potential—3.8 kV for negative polarity ionization and 4.5 kV for positive polarity ionization; temperature for the curved desolvation line (CDL) and heat block—200 °C; detector voltage—1.5 kV, detection range—100–900 m/z. The mass accuracy was below 4 ppm. We acquired and processed the data using Shimadzu LCMS Solution software v3.60.361 (Shimadzu, Kyoto, Japan).

2.6. Antiradical Activity

The 2,2-diphenyl-1-picrylhydrazyl (DPPH) radical-scavenging effect of polyphenolic compounds 711 was determined as described in [24]. Polyphenolic compounds 7, 8, 9, 10, and 11 were added to DPPH solution in MeOH (10−4 M) at concentrations from 6 to 34 µM. The reacting mixture was kept in the dark for 30 min at room temperature. Then, we measured the absorbance at 517 nm using a Shimadzu UV 1240 spectrophotometer (Shimadzu, Kyoto, Japan). Equation (1) was used to calculate the DPPH radical-scavenging effect (%):
DPPH   scavenging   effect , % = A 0 A x A 0
where A0 is the absorbance of DPPH solution of a blank sample (without polyphenolic compounds); AX is the absorbance of DPPH solution in the presence of different concentrations of polyphenolic compounds.
Quercetin was used as a reference compound. All experiments were performed in triplicate. We calculated the half-maximal scavenging concentration (SC50) for polyphenolic compounds by plotting the DPPH scavenging effect (%) against the concentrations of polyphenolic compounds. SC50 values are given as the mean value ± SEM.

2.7. Ferric Reducing Antioxidant Power (FRAP) Assay

We performed the FRAP assay as described in [29]. We prepared the FRAP reagent by mixing 2.5 mL of TPTZ (2,4,6-tris(2-pyridyl)-s-triazine) solution (10 mM) in 40 mM HCl and 25 mL of FeCl3 solution (20 mM) in acetate buffer solution (300 mM, pH 3.6). Polyphenolics 711 were added to 3 mL of FRAP reagent at concentrations from 6 to 34 µM. The mixture was kept in the dark at room temperature for 4 min. Then, we measured the absorbance at 595 nm using a Shimadzu UV 1240 spectrophotometer. Equation (2) was used to calculate the FRAP values for polyphenolic compounds 711:
FRAP = C F e C x
where CFe is the concentration of Fe2+ (µM) formed in the reaction; CX is the concentration of polyphenolic compounds in the reacting mixture.
The concentration of Fe2+ (µM) formed in the reaction was determined using the calibration curve obtained for different concentrations of FeSO4·7H2O.

2.8. Quantum-Chemical Modeling

We applied density functional theory (DFT) with the nonlocal exchange-correlation functional B3LYP [30], the polarization continuum model (PCM) [31], and split-valence basis set 6-311G(d), implemented in the Gaussian 16 package of programs [32] to perform the quantum-chemical calculations for compounds 9 and 10 in acetonitrile solvent. The molecular cavity was modeled according to unified force field (radii = UFF). The detailed conformational analysis of compounds 9 and 10 preceded the following calculations of their chiroptical properties.
The statistical weights (gim) of different conformations were calculated according to Equation (3):
g i m = e Δ G i m R T i e Δ G i m R T
where the summation was performed over all found stable conformations of the stereoisomer under study, and ΔGim = GiGm; G = Eel + Gtr,T + Grot,T + Gvib,T is the sum of electronic, translational, rotational, and vibrational contributions to the Gibbs free energy, respectively, calculated at temperature T = 298.15 K; the subscript “m” denotes the conformation, for which G is minimal.
The excitation energies and the rotatory strengths were calculated using time-dependent density functional theory (TDDFT), cam-B3LYP functional theory [33], and a PCM model and basis set, used previously for conformational analysis. Each individual transition from the electronic ground state to the i-th calculated excited electronic state (1 ≤ i ≤ 115) was simulated as a Gaussian-type function. The bandwidths, taken at 1/e peak heights, were chosen to be σ = 0.24 eV.
The total theoretical ECD spectrum was obtained after statistical averaging over all selected conformations using Equation (4):
Δ ε c a l c ( λ ) = i g i × Δ ε i ,     c a l c ( λ )
where i denotes different conformations of the stereoisomer under study. Conformations with Gibbs free energies in the region of ΔGim ≤ 5 kcal/mol were accounted for.

2.9. Cell Line and Culture Conditions

The murine neuroblastoma cell line Neuro-2a (CCL-131) was purchased from American Type Culture Collection (ATCC®) (Manassas, VA, USA). Cells were cultured in Dulbecco’s Modified Eagle Medium (Biolot, St. Petersburg, Russia). The medium contained 10% fetal bovine serum (Biolot, St. Petersburg, Russia) and 1% penicillin/streptomycin (Biolot, St. Petersburg, Russia). Cells were incubated at 37 °C in a humidified atmosphere containing 5% (v/v) CO2.

2.10. Cell Viability Assay (MTT Method)

Stock solutions of polyphenolic compounds were prepared in DMSO at a concentration of 10 mM. We diluted the solutions of all tested compounds in PBS to a volume of 20 μL and added them to the wells of the plates at the following final concentrations: 0.01, 0.1, 1.0, and 10.0 µM (final concentration of DMSO <1%).
We incubated Neuro-2a cells (1 × 104 cells/well) at 37 °C in a CO2 incubator for 24 h until they formed an adherent monolayer. Then, 20 μL of the tested solution was loaded onto the cells and incubated for 24 h. After incubation, the medium containing the polyphenolic test compounds was replaced by 100 μL of fresh medium. Then, we added 10 μL of MTT (3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide) (Sigma-Aldrich, St. Louis, MO, USA) stock solution (5 mg/mL) to each well and incubated the microplate for 4 h. Then, 100 μL of SDS-HCl solution (1 g SDS/10 mL dH2O/17 μL 6 M HCl) was added to each well, followed by incubation for 18 h. We measured the absorbance of the converted dye formazan on a Multiskan FC microplate photometer (Thermo Scientific, Waltham, MA, USA) at a wavelength of 570 nm [34]. We performed all experiments in triplicate and expressed the cytotoxic activity as percent of living cells.

2.11. In Vitro Model of PQ-, Rotenone-, and 6-OHDA-Induced Neurotoxicity

After 24 h of adhesion, Neuro-2a cells (1 × 104 cells/well) were treated with polyphenolic compounds at concentrations of 0.01–10 μM for 1 h. Subsequently, 1 mM of PQ, 10 μM rotenone or 80 μM of 6-OHDA (Sigma-Aldrich, St. Louis, MO, USA) were added. We used MTT assay to determine the percentage of living cells (cell viability) after 24 h of incubation. Cells incubated without inductors or with inductors were used as positive and negative control, respectively. The results are presented as percentages of the positive control value.

2.12. Reactive Oxygen Species (ROS) Level Evaluation in PQ-Treated Cells

After 24 h of adhesion, Neuro-2a cells (1 × 104 cells/well) were incubated with compounds (0.01–10 µM) for 1 h. Then, we added rotenone (10 μM), 6-OHDA (80 μM), or PQ (1 mM) to each well and incubated cells for 1 h, 1 h, or 3 h, respectively. To evaluate the level of intracellular ROS, we added 20 µL of 2,7-dichlorodihydrofluorescein diacetate solution (10 µM, H2DCF-DA, Molecular Probes, Eugene, OR, USA) to each well, so that the final concentration was 10 µM, and incubated the microplate at 37 °C for an additional 30 min. Quercetin was used as a reference compound.

2.13. Mitochondrial Membrane Potential (MMP) Detection

The cells were incubated for 1 h in a 96-well plate (1 × 104 cells/well) with polyphenolic compounds (1 and 10 µM). Then, PQ (500 µM) was added, and the cell suspension was incubated for 1 h. Cells incubated without PQ and compounds were used as positive control, and cells with PQ alone were used as negative control. We added the tetramethylrhodamine methyl (TMRM) (Sigma-Aldrich, St. Louis, MO, USA) solution (500 nM) to each well, and incubated cells for 30 min at 37 °C. After the incubation, we measured the intensity of fluorescence with a PHERAstar FSplate reader (BMG Labtech, Ortenberg, Germany) at λex = 540 nm and λem = 590 nm. We processed the data using MARS Data Analysis v3.01R2 (BMG Labtech, Ortenberg, Germany) and presented results as percentages of the positive control value.

2.14. Statistical Analysis

We carried out all the experiments in triplicate and performed Student’s t-test using SigmaPlot 14.0 (Systat Software Inc., San Jose, CA, USA) to determine statistical significance.

3. Results

3.1. Isolation and Structure Elucidation of Compounds 10 and 11

We managed to isolate a new coumestan 10 and stilbenoid 11 as well as previously known pterocarpans: (6aR,11aR)-6a,11a-dihydrolespedezol A2 (1), (6aR,11aR,3′S)-6a,11a-dihydrolespedezol A3 (2), 6aR,11aR,3′R-6a,11a-dihydrolespedezol A3 (3), (6aR,11aR)-2-isoprenyl-6a,11a-dihydrolespedezol A2 (6), pterocarpens: lespedezol A2 (4), lespedezol A3 (9), coumestan lespedezol A6 (7), stilbenoid bicoloketone (5), and dimeric flavonoid lespebicolin B (8) from L. bicolor root bark (Figure 1). Compounds 19 were identified by comparison of their HPLC-PDA-MS and NMR spectra with previously published data [21,22,24,26].
Compound 10 was obtained as a white amorphous powder. The molecular formulae of 10 was determined to be C25H22O6 based on the presence of [M-H]- and [M+H]+ ions at m/z 417.1337 (calculated for C25H21O6- 417.1344) and 419.1479 (calculated for C25H23O6+ 419.1489), respectively, in the HR-MS-ESI spectrum of 10. The 13C spectrum of 10 contained 25 signals of carbon atoms (Table 1). Fifteen carbon atoms belonged to the coumestan skeleton (rings A–D), and 10 atoms formed a 3′-methyl-3′-isohexenylpyran ring (E). An ester carbonyl carbon signal was observed in the 13C NMR spectrum of 10 at δC 159.0 and was assigned to C-6 of the coumestan skeleton [22]. The 1H NMR spectra of 10 showed the presence of an ABX spin system: the signals at δH 7.85 (d, J = 8.5), 6.95 (d, J = 8.5), and 7.13 (s) were attributed to protons H-1, H-2, and H-4 of ring A, respectively (Table 1). A singlet signal at δH 7.44 was ascribed to the H-7 proton (ring D of the coumestan skeleton). We also observed two broad singlet signals at δH 6.54 and 5.50 due to OH-3 and OH-8 protons, respectively, in the 1H NMR spectra of 10 (Table 1). The 1H NMR spectra of 10 revealed signals at δH 6.89 (1H, d, J = 9.9 Hz) and 5.78 (1H, d, J = 9.9 Hz) due to the H-1′ and H-2′ protons of the AX-type olefinic proton system, suggesting that 10 had an oxidatively cyclized geranyl side chain similar to the structures of compounds 2 and 9 (Table 1) [24].
We completely assigned all signals in the 1H and 13C NMR spectra of 10 on the basis of COSY, HMBC, and ROESY spectral data. Thus, compound 10 was determined to be a derivative of lespedezol A6 (7), previously isolated from L. homoloba but containing a cyclized geranyl side chain [22]. Compound 10 was named lespebicoumestan A.
Compounds 9 and 10 had a 3′-methyl-3′-isohexenylpyran ring (E) and, hence, an asymmetric center at C-3′. Although lespedezol A3 (9) had previously been isolated from L. homoloba [21], the absolute configuration of the asymmetric center at C-3′ had not yet been determined. In order to determine the absolute configuration of the asymmetric center at C-3′ in compounds 9 and 10, we compared their calculated theoretical ECD spectra with corresponding experimental ECD data. We used the cam-B3LYP exchange-correlation functional set [33] along with the 6-311G(d) basis set and polarization continuum model (PCM) [31], implemented in the Gaussian 16 suit of programs [32], to calculate the energies and rotatory strengths of vertical electronic transitions. First, conformational analysis was performed for each compound at the B3LYP/6-311G(d)_PCM level of theoretical modeling; optimized geometries and relative Gibbs free energies as well as statistical weights were thus obtained.
The choice of the cam-B3LYP functional model for excited states is approved because it accounts for long-range interactions, to some extent better than the B3LYP functional model. Finally, statistically averaged ECD spectra were obtained as a weighted superposition of Gaussian-type functions and chosen for simulation of the Δεi(λ) function shapes for individual electronic transitions. The same value of σ = 0.24 eV for the bandwidths at 1/e peak heights was used. One hundred fifteen excited states were calculated for each conformation.
A comparison of the experimental and theoretical ECD spectra obtained for 9 and 10 is presented in Figure 2a,b. The positions and relative intensities of individual bands in the characteristic region 200 ≤ λ ≤ 300 nm are well-reproduced. The discrepancies in properties of Δεcalc (λ) and Δεexp (λ) occur in the long-wave region λ ≥ 300 nm, caused to some extent by underestimation of the contribution to the Δεcalc (λ) from the E+ conformations (Schemes S60–S61, Figure S62, Supplementary data). Thus, we performed several simulations of the averaged ECD spectrum for 9, in which relative amounts of the E and E+ conformations were varied manually. We found that the shapes, positions, and relative intensities of individual bands in the characteristic region 200 ≤ λ ≤ 300 nm are refractory to these variations, while they change dramatically in the λ ≥ 300 nm region Schemes S60–S61, Figure S62 (S61–S62, Supplementary data). We previously observed analogous behavior for compound 2 [24].
The good qualitative coincidence between theoretical and experimental ECD spectra allowed us to determine the absolute configuration of the asymmetric center at C-3′ in compounds 9 and 10 as C3′-S.
We obtained compound 11 as a yellow, amorphous powder. We elucidated the structure of the new compound using extensive spectroscopic analyses. The molecular formula C29H34O7 was confirmed by the presence of the [M-H] molecular ion at m/z 493.2217 (negative ion mode, calculated for C29H33O7 493.2232) and the [M+H]+ molecular ion at m/z 495.2401 (positive ion mode, calculated for C29H35O7+ 495.2377) in the HR-ESI-MS spectrum of 11. The 13C NMR spectrum of 11 showed the presence of 12 carbon atoms of two aromatic rings, 10 carbon atoms of a geranyl side chain, 5 carbon atoms of an isoprenyl side chain, and 2 carbon atoms of two carbonyl groups. The 1H NMR spectrum of 11 exhibited resonances at δH 6.44 (1H, s) and 7.18 (1H, s) of protons H-3′ and H-6′, respectively (ring B), and a singlet at δH 6.84 (1H) was attributed to H-8 (ring A) (Table 2). The downfield-shifted chemical shift values of hydroxyl groups OH-4 and OH-2′ at δH 11.92 (1H, s) and 11.59 (1H, s) confirmed that they formed hydrogen bonds with carbonyl groups at C-2 and C-1, respectively. The signals of the geranyl side chain protons were observed at δH 3.51 (2H, d, J = 7.3 Hz, H-1″), 5.33 (1H, t, J = 7.3 Hz, H-2″), 2.09 (2H, m, H-4″), 2.13 (2H, m, H-5″), 5.07 (1H, m, H-6″), 1.68 (3H, s, H-8″), 1.84 (3H, s, H-9″), and 1.60 (3H, s, H-10″) (Table 2). The geranyl substituent was determined to be located at C-5 because in the HMBC spectrum of 11 we observed cross-peaks between the proton signal at δH 3.51 (1H, d, J = 7.3 Hz) of H-1″ and the C-4, C-5, and C-6 carbon signals at δC 158.9, 115.3, and 152.5, respectively (Table 2). The signals of the isoprenyl side chain protons were observed at δH 3.23 (2H, d, J = 7.2 Hz, H-1‴), 5.20 (1H, t, J = 7.2 Hz, H-2‴), 1.73 (3H, s, H-4‴), and 1.71 (3H, s, H-5‴) (Table 2). The isoprenyl substituent was determined to be located at C-5′ because in the HMBC spectrum of 11 we observed cross-peaks between the proton signal at δH 3.23 (1H, d, J = 7.2 Hz) of H-1‴ and the C-4′, C-5′, and C-6′ carbon signals at δC 163.7, 119.9, and 133.9, respectively (Table 2). Thus, compound 11 was shown to be a stilbenoid, and the structure of 11 differed from that of bicoloketone (5) only by the presence of an additional isoprenyl side chain at C-5′. Compound 11 was named 5′-isoprenylbicoloketone.

3.2. Antioxidant Activity of Polyphenolic Compounds

The data on antioxidant activity (DPPH-scavenging effect and FRAP assay) of compounds 16 have been previously published in [24]. Here, we evaluated the DPPH-scavenging effect and FRAP of polyphenolic compounds 711 isolated from L. bicolor. In the FRAP assay, polyphenolic antioxidants reduced the light blue Fe3+–TPTZ complex to the dark blue Fe2+–TPTZ complex. The change in color resulted in an increase in absorbance at 595 nm. The results of both tests are shown in Table 3.
Compounds 111 exhibited a moderate DPPH-scavenging effect, which was comparable to the effect of ascorbic acid but smaller than that of quercetin (Table 3). Lespedezol A2 (4) and lespedezol A3 (9) possessed the most-significant DPPH-scavenging effect and FRAP among compounds 1-11 (Table 3) [24]. In the FRAP assay, all tested polyphenolics also showed moderate effects but were less active than quercetin and ascorbic acid. Lespedezol A2 (4) was the most effective in the FRAP assay.

3.3. Cytotoxicity of Polyphenolic Compounds from L. bicolor against Neuro-2a Cells

The cytotoxic activity of 11 polyphenolic compounds from L. bicolor root bark was evaluated using the MTT assay (Table 4). The cytotoxicity of the isolated compounds was determined on neuroblastoma Neuro-2a cells. Compounds 3 (EC50 = 44.0 µM), 6 (EC50 = 40.6 µM), and 11 (EC50 = 44.0 µM) possessed moderate toxicity. Polyphenolics 1, 2, 4, 9, and 11 showed low toxicity (EC50 = 72–87 µM). It should be noted that compound 4 has previously been shown to be rather cytotoxic against three cancer cell lines (human triple-negative breast cancer HTB-19, esophageal squamous cell carcinoma Kyse-30, and hepatocellular carcinoma HEPG-2) and two normal cell lines (retina pigmented epithelium cells RPE-1 and human embryonic kidney cells HEK-293) [11]. The other compounds did not demonstrate cytotoxic effects at concentrations up to 100 µM.

3.4. Influence on Viability and ROS Level in PQ-Treated Neuro-2a Cells

The percentage of living cells after treatment with neurotoxin was assessed by the MTT test. The percentage of living cells after treatment with PQ significantly increased when polyphenolic compounds 3, 5, 6, and 8 were added (Figure 3a–c). The neuroprotective effect of compound 6, measured by the MTT assay, was detected in the concentration range of 0.01–1 μM, and at a concentration of 1 μM this compound increased the percentage of living cells by 17%. Compound 3 increased the percentage of living cells after treatment with PQ by 8% at a concentration of 10 μM in the MTT assay. Compounds 5 and 8 increased the percentage of living cells after PQ treatment by 8–10% at concentrations of 1 and 10 μM in the MTT assay. Compounds 1, 2, 4, and 911 did not exhibit any effect on PQ-treated Neuro-2a cell viability in this test. Polyphenolics from L. bicolor did not significantly reduce ROS levels in PQ-treated cells (Figure 3d).

3.5. Mitochondrial Membrane Potential (MMP) Detection

We studied the effect of polyphenolic compounds from L. bicolor on PQ-induced mitochondrial dysfunction. The 23% decrease in tetramethylrhodamine methyl (TMRM) fluorescence after a 1 h exposure of Neuro-2a cells with PQ indicates that PQ causes depolarization of the mitochondrial membrane (Figure 3e). Among the tested compounds, pterocarpan 3 and stilbenoid 5 at a concentration of 10 µM were the most effective in this assay and increased the value of mitochondrial membrane potential by 16% and 23%, respectively.

3.6. Influence on Viability and ROS Levels in 6-OHDA-Treated Neuro-2a Cells

We evaluated the effects of the polyphenolic-compound set on cell viability in a 6-OHDA-induced neurotoxicity model. The percentage of living Neuro-2a cells after 6-OHDA treatment increased from 45% to 65% in the presence of polyphenolic compounds, compared with the control (Figure 4a–c). Pre-treatment of cells with the test compounds for 1 h before 6-OHDA addition provided an increase in the percentage of living cells with varying levels of statistical confidence.
Compounds 9 and 10 increased the percentage of living Neuro-2a cells after 6-OHDA treatment by 31.4% and 10.9% versus 6-OHDA-treated cells, respectively (p < 0.05). Compounds 2, 3, and 4 increased the viability of 6-OHDA-treated cells by 8.7%, 12.9%, and 7.0%, respectively (p < 0.05). The other compounds did not show significant improvement in this assay. Polyphenolic compounds 24, 9, and 10 significantly decreased ROS levels in 6-OHDA-treated cells (Figure 4d). Pterocarpans 2 and 3 were the most-active in this test and decreased the ROS level 4.5 times compared to that of 6-OHDA-treated cells. All tested polyphenolic compounds decreased the intracellular ROS level much more effectively than quercetin (Figure 4d).

3.7. Influence on Viability and ROS Levels in Rotenone-Treated Neuro-2a Cells

We examined the effect of polyphenolic compounds from L. bicolor on cell viability in a rotenone-induced neurotoxicity model. The percentage of living Nuero-2a cells treated with rotenone was 68% compared to the control (Figure 5a–c). Pre-treatment of cells with polyphenolic compounds from L. bicolor for 1 h before rotenone addition provided an increase in the percentage of living Neuro-2a cells with different levels of statistical confidence.
Compounds 9 and 10 increased the percentage of living Neuro-2a cells after treatment with rotenone by 10.4% and 13.2%, respectively, compared to rotenone-treated cells (p < 0.05). The other compounds did not show significant improvement in this assay (Figure 3a–d).

4. Discussion

We continued to study the chemical composition of polyphenolic compounds from L. bicolor root bark and isolated a new coumestan, lespebicoumestan A (10), and a stilbenoid, 5′-isoprenylbicoloketon (11), as well as previously known pterocarpans 13, and 6; pterocarpens 4 and 9; coumestan 7; stilbenoid 5; and dimeric flavonoid 8 (Figure 1). There was a considerable difference between the chemical compositions of polyphenolic compounds of L. bicolor growing in the Primorskiy region (Russian Far East) and in the Republic of Korea. Lee P.J. and coworkers isolated from L. bicolor 11 pterocarpans, 2 coumestans, and 2 arylbenzofurans with isoprenyl and geranyl substituents in their structures [25,27,28]. These compounds contained a methoxy group at C-1, whereas the C-1 position in pterocarpans and coumestans from stem bark and root bark of L. bicolor collected in the Primorskiy region was unsubstituted [24,26]. Besides, some pterorarpans and coumestans from L. bicolor growing in the Russian Far East had a 3′-methyl-3′-isohexenylpyran ring (E), presumably formed by oxidative cyclization of a geranyl side chain. However, pterocarpans and coumestans from L. bicolor growing in the Republic of Korea contained a 3′,3′-dimethylpyran ring (E) produced by oxidative cyclization of an isoprenyl side chain. In contrast to Far-Eastern L. bicolor, Korean L. bicolor did not contain stilbenoids and dimeric flavonoids.
Being natural antioxidants, pterocarpans and coumestans are promising candidates for the study of neuroprotective properties [10,11,25]. Nerve cells treated with various inducers of oxidative stress (PQ, 6-OHDA, rotenone) are often used as one of the generally accepted models for studying neurotoxic disorders, including PD [35,36].
In our study, pterocarpans 3 and 6, stilbenoid 5, and dimeric flavonoid 8 significantly increased the percentage of living Neuro-2a cells after treatment with PQ, but only pterocarpan 6 slightly decreased the ROS level in PQ-treated cells, which is in accordance with its rather high activity in the DPPH and FRAP tests [24]. It is known that the effect of PQ on neurons is accompanied by impaired functioning of mitochondria due to changes in mitochondrial membrane permeability, membrane potential, and depolarization of mitochondrial membranes [37]. We presumed that these compounds could also increase cell viability by preventing depolarization of the mitochondrial membrane. In fact, pterocarpan 3 and stilbenoid 5 were shown to effectively increase the mitochondrial membrane potential of PQ-treated neuronal cells.
We also examined the effects of polyphenolic compounds on cell viability in a 6-OHDA-induced neurotoxicity model. We showed that pterocarpans 2 and 3, containing a 3′-methyl-3′-isohexenylpyran ring (E); pterocarpens 4 and 9, with a double bond between C-6a and C-11a; and lespebicoumestan A (10) significantly increased the percentage of living Neuro-2a cells and decreased ROS levels after treatment with 6-OHDA. Notably, compounds 9 and 10 both contain a 3′-methyl-3′-isohexenylpyran ring (E) and a double bond between C-6a and C-11a. Pterocarpans 2 and 3 were the most active in this assay and decreased the ROS level 4.5 times compared to 6-OHDA-treated cells. Notably, polyphenolic compounds from L. bicolor decreased the level of intracellular ROS much more effectively than quercetin (Figure 4d). Thus, pterocarpans and coumestans with an additional 3′-methyl-3′-isohexenylpyran ring demonstrated the most-significant activity.
We also studied the effect of polyphenolic compounds from L. bicolor on cell viability in a rotenone-induced neurotoxicity model. Pre-treatment of cells with polyphenolic compounds from L. bicolor before rotenone addition resulted in an increase in the percentage of living Neuro-2a cells, with compounds 9 and 10 being the most active. The presence of a 3′-methyl-3′-isohexenyl pyran ring (E) and a double bond between C-6a and C-11a in 9 and 10 may be responsible for the increase in cell viability. The other compounds did not show significant improvement of cell viability in this assay.
Previously, researchers from the College of Pharmacy, the Research Institute of Pharmaceutical Science and Technology (The Republic of Korea), and the Korea Research Institute of Standards and Science demonstrated that pterocarpan-type compounds (1-methoxylespeflorin G11, bicolosin A, 1-methoxyerythrabyssin II, 8-methoxybicolosin C, 2-geranyl-1-methoxylespeflorin G11, and 2-geranylbicolosin A) exhibited significant neuroprotective effects against glutamate neurotoxicity in neuronal HT22 hippocampal cells [25]. The compound 2-geranyl-1-methoxylespeflorin G11—which has a geranyl group at C-2, a prenyl group at C-10, and a methyl group at C-8—was shown to attenuate apoptosis in HT22 cells by inhibiting intracellular ROS generation and mitochondrial dysfunction. Although in our study coumestan 10 effectively increased the percentage of living Neuro-2a cells after treatment with 6-OHDA- and rotenone, arylbenzofurans and coumestan isolated from Korean L. bicolor exhibited no protective effects [25].
Considering that isoflavonoids can quickly penetrate the blood–brain barrier [38], such compounds may be prospective agents in the treatment of PD.

5. Conclusions

Thus, pterocarpans 2, 3, and 6 and pterocarpens 4 and 9, as well as coumestan 10 from L. bicolor effectively protected PQ- and 6-OHDA-treated Neuro-2a cells from oxidative stress. The effect of polyphenolic compounds 3 and 5 from L. bicolor is mainly due to their ability to impair PQ-induced depolarization of the mitochondrial membrane, whereas compounds 24, 9, and 10 decreased ROS levels in 6-OHDA-treated Neuro-2a cells more effectively than quercetin. The effect of polyphenolic compounds on the viability of rotenone-treated cells was less dramatic.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/antiox11040709/s1, Figures S1–S2: Mass-spectra of compound 10; Figures S3–S25: NMR spectra of compound 10; Figures S26–S27: Mass-spectra of compound 11; Figures S28–S57: NMR spectra of compound 11; Figure S58: The geometrical structure and atom numeration for compounds 9 and 10; Scheme S59: Abbreviations for conformations of compounds 9 and 10; Scheme S60: The influence of inversion of ring B on the geometrical structure of compound 9; Scheme S61: Equation for recalculation of the relative contributions of the E and E+ conformations to the ECD spectra of compounds 9 and 10; Figure S62: The influence of variation in the relative amounts of E and E+ conformations on the shape of the statistically averaged scaled ECD spectrum of compound 9.

Author Contributions

Conceptualization, D.V.T., S.A.F., E.A.P., N.P.M. and D.L.A.; methodology, D.V.T., E.A.P., E.S.M., D.V.B., A.I.K. and V.P.G.; investigation, D.V.T., S.A.F., D.V.B., E.A.P., E.S.M. and D.L.A.; writing—original draft preparation, D.V.T.; writing—review and editing, D.V.T., S.A.F., E.A.P., E.S.M., N.P.M. and D.L.A.; data analysis and interpretation, D.V.T., S.A.F., E.A.P., E.S.M. and D.L.A.; supervision, S.A.F. and D.L.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data are contained within the article and Supplementary Materials.

Acknowledgments

The authors are grateful to academician P.G. Gorovoy for collecting the samples of L. bicolor for this study.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Calabrese, V.; Santoro, A.; Monti, D.; Crupi, R.; Di Paola, R.; Latteri, S.; Cuzzocrea, S.; Zappia, M.; Giordano, J.; Calabrese, E.J.; et al. Aging and Parkinson’s disease: Inflammaging, neuroinflammation and biological remodeling as key factors in pathogenesis. Free Radic. Biol. Med. 2018, 115, 80–91. [Google Scholar] [CrossRef] [PubMed]
  2. Borrageiro, G.; Haylett, W.; Seedat, S.; Kuivaniemi, H.; Bardien, S. A review of genome-wide transcriptomics studies in Parkinson’s disease. Eur. J. Neurosci. 2018, 47, 1–16. [Google Scholar] [CrossRef] [PubMed]
  3. Bove, J.; Prou, D.; Perier, C.; Przedborski, S. Toxin-induced models of Parkinson’s disease. NeuroRX 2005, 2, 484–494. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Bastias-Candia, S.; Zolezzi, J.M.; Inestrosa, N.C. Revisiting the paraquat-induced sporadic Parkinson’s disease-like model. Mol. Neurobiol. 2019, 56, 1044–1055. [Google Scholar] [CrossRef] [PubMed]
  5. Menchinskaya, E.; Chingizova, E.; Pislyagin, E.; Likhatskaya, G.; Sabutski, Y.; Pelageev, D.; Polonik, S.; Aminin, D. Neuroprotective effect of 1,4-naphthoquinones in an in vitro model of paraquat and 6-OHDA-induced neurotoxicity. Int. J. Mol. Sci. 2021, 22, 9933. [Google Scholar] [CrossRef] [PubMed]
  6. Hussain, G.; Zhang, L.; Rasul, A.; Anwar, H.; Sohail, M.U.; Razzaq, A.; Aziz, N.; Shabbir, A.; Ali, M.; Sun, T. Role of plant-derived flavonoids and their mechanism in attenuation of Alzheimer’s and Parkinson’s diseases: An update of recent data. Molecules 2018, 23, 814. [Google Scholar] [CrossRef] [Green Version]
  7. Selvam, C.; Jordan, B.C.; Prakash, S.; Mutisya, D.; Thilagavathi, R. Pterocarpan scaffold: A natural lead molecule with diverse pharmacological properties. Eur. J. Med. Chem. 2017, 128, 219–236. [Google Scholar] [CrossRef]
  8. Xia, W.; Luo, P.; Hua, P.; Ding, P.; Li, C.; Xu, J.; Zhou, H.; Gu, Q. Discovery of a new pterocarpan-type antineuroinflammatory compound from Sophora tonkinensis through suppression of the TLR4/NFκB/MAPK signaling pathway with PU 1 as a potential target. ACS Chem. Neurosci. 2018, 10, 295–303. [Google Scholar] [CrossRef]
  9. Tsai, R.T.; Tsai, C.W.; Liu, S.P.; Gao, J.X.; Kuo, Y.H.; Chao, P.M.; Hung, H.S.; Shyu, W.C.; Lin, S.Z.; Fu, R.H. Maackiain ameliorates 6-hydroxydopamine and SNCA pathologies by modulating the PINK1/Parkin pathway in models of Parkinson’s Disease in Caenorhabditis elegans and the SH-SY5Y cell line. Int. J. Mol. Sci. 2020, 21, 4455. [Google Scholar] [CrossRef]
  10. Lee, H.W.; Ryu, H.W.; Kang, M.G.; Park, D.; Oh, S.R.; Kim, H. Potent selective monoamine oxidase B inhibition by maackiain, a pterocarpan from the roots of Sophora flavescens. Bioorg. Med. Chem. Lett. 2016, 26, 4714–4719. [Google Scholar] [CrossRef]
  11. Oh, J.M.; Jang, H.J.; Kim, W.J.; Kang, M.G.; Baek, S.C.; Lee, J.P.; Park, D.; Oh, S.R.; Kim, H. Calycosin and 8-O-methylretusin isolated from Maackia amurensis as potent and selective reversible inhibitors of human monoamine oxidase-B. Int. J. Biol. Macromol. 2020, 151, 441–448. [Google Scholar] [CrossRef] [PubMed]
  12. Lee, J.W.; Lee, J.H.; Lee, C.; Jin, Q.; Lee, D.; Kim, Y.; Hong, J.T.; Lee, M.K.; Hwang, B.Y. Inhibitory constituents of Sophora tonkinensis on nitric oxide production in RAW264.7 macrophages. Bioorg. Med. Chem. Lett. 2015, 25, 960–962. [Google Scholar] [CrossRef] [PubMed]
  13. Yoo, H.; Kang, M.; Pyo, S.; Chae, H.S.; Ryu, K.H.; Kim, J.; Chin, Y.W. SKI3301, a purified herbal extract from Sophora tonkinensis, inhibited airway inflammation and bronchospasm in allergic asthma animal models in vivo. J. Ethnopharmacol. 2017, 206, 298–305. [Google Scholar] [CrossRef] [PubMed]
  14. Joshi, N.; Singh, S. Updates on immunity and inflammation in Parkinson disease pathology. J. Neurosci. Res. 2018, 96, 379–390. [Google Scholar] [CrossRef]
  15. Liu, M.H.; Tsuang, F.Y.; Sheu, S.Y.; Sun, J.S.; Shih, C.M. The protective effects of coumestrol against amyloid-beta peptide-and lipopolysaccharide-induced toxicity on mice astrocytes. Neurol. Res. 2011, 33, 663–672. [Google Scholar] [CrossRef]
  16. Castro, C.C.; Pagnussat, A.S.; Orlandi, L.; Worm, P.; Moura, N.; Etgen, A.M.; Netto, C.A. Coumestrol has neuroprotective effects before and after global cerebral ischemia in female rats. Brain Res. 2012, 1474, 82–90. [Google Scholar] [CrossRef] [Green Version]
  17. Anastacio, J.B.R.; Sanches, E.F.; Nicola, F.; Odorcyk, F.; Fabres, R.B.; Netto, C.A. Phytoestrogen coumestrol attenuates brain mitochondrial dysfunction and longterm cognitive deficits following neonatal hypoxia–ischemia. Int. J. Dev. Neurosci. 2019, 79, 86–95. [Google Scholar] [CrossRef]
  18. Moreira, A.C.; Silva, A.M.; Branco, A.F.; Baldeiras, I.; Pereira, G.C.; Seiça, R.; Santos, M.S.; Sardao, V.A. Phytoestrogen coumestrol improves mitochondrial activity and decreases oxidative stress in the brain of ovariectomized Wistar-Han rats. J. Funct. Foods. 2017, 34, 329–339. [Google Scholar] [CrossRef]
  19. Mori-Hongo, M.; Yamaguchi, H.; Warashina, T.; Miyase, T. Melanin synthesis inhibitors from Lespedeza cyrtobotrya. J. Nat. Prod. 2009, 72, 63–71. [Google Scholar] [CrossRef]
  20. Mori-Hongo, M.; Takimoto, H.; Katagiri, T.; Kimura, M.; Ikeda, Y.; Miyase, T. Melanin synthesis inhibitors from Lespedeza floribunda. J. Nat. Prod. 2009, 72, 194–203. [Google Scholar] [CrossRef]
  21. Miyase, T.; Sano, M.; Nakai, H.; Muraoka, M.; Nakazawa, M.; Suzuki, M.; Yoshino, K.; Nishihara, Y.; Tanai, J. Antioxidants from Lespedeza homoloba (I). Phytochemistry 1999, 52, 303–310. [Google Scholar] [CrossRef]
  22. Miyase, T.; Sano, M.; Yoshino, K.; Nonaka, K. Antioxidants from Lespedeza homoloba (II). Phytochemistry 1999, 52, 311–319. [Google Scholar] [CrossRef]
  23. Dyshlovoy, S.A.; Tarbeeva, D.V.; Fedoreyev, S.A.; Busenbender, T.; Kaune, M.; Veselova, M.V.; Kalinovskiy, A.I.; Hauschild, J.; Grigorchuk, V.P.; Kim, N.Y.; et al. Polyphenolic compounds from Lespedeza bicolor root bark inhibit progression of human prostate cancer cells via induction of apoptosis and cell cycle arrest. Biomolecules 2020, 10, 451. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Tarbeeva, D.V.; Fedoreyev, S.A.; Veselova, M.V.; Blagodatski, A.S.; Klimenko, A.M.; Kalinovskiy, A.I.; Grigorchuk, V.P.; Berdyshev, D.V.; Gorovoy, P.G. Cytotoxic polyphenolic compounds from Lespedeza bicolor stem bark. Fitoterapia 2019, 135, 64–72. [Google Scholar] [CrossRef]
  25. Lee, P.J.; Pham, C.H.; Thuy, N.T.T.; Park, H.J.; Lee, S.H.; Yoo, H.M.; Cho, N. 1-Methoxylespeflorin G11 protects HT22 cells from glutamate-induced cell death through inhibition of ROS production and apoptosis. J. Microbiol. Biotechnol. 2021, 31, 217–225. [Google Scholar] [CrossRef]
  26. Tarbeeva, D.V.; Krylova, N.V.; Iunikhina, O.V.; Likhatskaya, G.N.; Kalinovskiy, A.I.; Grigorchuk, V.P.; Shchelkanov, M.Y.; Fedoreyev, S.A. Biologically active polyphenolic compounds from Lespedeza bicolor. Fitoterapia 2022, 157, 105121. [Google Scholar] [CrossRef]
  27. Thuy, N.T.T.; Lee, J.E.; Yoo, H.M.; Cho, N. Antiproliferative pterocarpans and coumestans from Lespedeza bicolor. J. Nat. Prod. 2019, 82, 3025–3032. [Google Scholar] [CrossRef]
  28. Woo, H.S.; Kim, D.W.; Curtis-Long, M.J.; Lee, B.W.; Lee, J.H.; Kim, J.Y.; Kang, J.E.; Park, K.H. Potent inhibition of bacterial neuraminidase activity by pterocarpans isolated from the roots of Lespedeza bicolor. Bioorg. Med. Chem. Lett. 2011, 21, 6100–6103. [Google Scholar] [CrossRef]
  29. Dehghan, G.; Khoshkam, Z. Tin(II)-quercetin complex: Synthesis, spectral characterization and antioxidant activity. Food Chem. 2012, 131, 422–426. [Google Scholar] [CrossRef]
  30. Stephens, P.J.; Devlin, F.J.; Chabalowski, C.F.; Frisch, M.J. Ab Initio calculation of vibrational absorption and circular dichroism spectra using density functional force fields. J. Phys. Chem. 1994, 98, 11623–11627. [Google Scholar] [CrossRef]
  31. Miertus, S.; Scrocco, E.; Tomasi, J. Electrostatic interaction of a solute with a continuum. A direct utilization of AB initio molecular potentials for the prevision of solvent effects. Chem. Phys. 1981, 55, 117–129. [Google Scholar] [CrossRef]
  32. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16, Revision, A.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  33. Yanai, T.; Tew, D.; Handy, N. A new hybrid exchange-correlation functional using the Coulomb-attenuating method (CAM-B3LYP). Chem. Phys. Lett. 2004, 393, 51–57. [Google Scholar] [CrossRef] [Green Version]
  34. Mosmann, T. Rapid colorimetric assay for cellular growth and survival: Application to proliferation and cytotoxicity assays. J. Immunol. Methods 1983, 65, 55–63. [Google Scholar] [CrossRef]
  35. Pislyagin, E.; Kozlovskiy, S.; Menchinskaya, E.; Chingizova, E.; Likhatskaya, G.; Gorpenchenko, T.; Sabutski, Y.; Polonik, S.; Aminin, D. Synthetic 1,4-naphthoquinones inhibit P2X7 receptors in murine neuroblastoma cells. Bioorg. Med. Chem. 2021, 31, 115975. [Google Scholar] [CrossRef] [PubMed]
  36. Chia, S.J.; Tan, E.K.; Chao, Y.X. Historical perspective: Models of Parkinson’s disease. Int. J. Mol. Sci. 2020, 21, 2464. [Google Scholar] [CrossRef] [Green Version]
  37. Huang, C.-L.; Chao, C.-C.; Lee, Y.-C.; Lu, M.-K.; Cheng, J.-J.; Yang, Y.-C.; Huang, N.-K. Paraquat induces cell death through impairing mitochondrial membrane permeability. Mol. Neurobiol. 2015, 53, 2169–2188. [Google Scholar] [CrossRef]
  38. Xiao, B.; Sun, Z.; Cao, F.; Wang, L.; Liao, Y.; Liu, X.; Pan, R.; Chang, Q. Brain pharmacokinetics and the pharmacological effects on striatal neurotransmitter levels of Pueraria lobata isoflavonoids in rat. Front. Pharmacol. 2017, 8, 599. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Structures of polyphenolic compounds isolated from L. bicolor root bark.
Figure 1. Structures of polyphenolic compounds isolated from L. bicolor root bark.
Antioxidants 11 00709 g001aAntioxidants 11 00709 g001b
Figure 2. Experimental and calculated ECD spectra for compounds 9 (a) and 10 (b).
Figure 2. Experimental and calculated ECD spectra for compounds 9 (a) and 10 (b).
Antioxidants 11 00709 g002
Figure 3. The influence of polyphenolic compounds from L. bicolor on cell viability (ac), ROS levels (d), and mitochondrial membrane potential (e) in Neuro-2a cells treated with PQ (1 mM). The percentage of living cells treated with various compounds and PQ was measured by MTT assay. Each bar represents the mean ± SEM of three independent replicates. (*), (**), and (***) indicate, respectively, p < 0.05, p < 0.005, and p < 0.001 versus PQ-treated cells. The difference between control and PQ-treated cells was considered significant.
Figure 3. The influence of polyphenolic compounds from L. bicolor on cell viability (ac), ROS levels (d), and mitochondrial membrane potential (e) in Neuro-2a cells treated with PQ (1 mM). The percentage of living cells treated with various compounds and PQ was measured by MTT assay. Each bar represents the mean ± SEM of three independent replicates. (*), (**), and (***) indicate, respectively, p < 0.05, p < 0.005, and p < 0.001 versus PQ-treated cells. The difference between control and PQ-treated cells was considered significant.
Antioxidants 11 00709 g003aAntioxidants 11 00709 g003b
Figure 4. The influence of polyphenolic compounds from L. bicolor on cell viability (ac) and ROS levels (d) in Neuro-2a cells treated with 6-OHDA (80 µM). The percentage of living cells treated with compounds and 6-OHDA was measured by MTT assay. Each bar represents the mean ± SEM of three independent replicates. (*), (**), and (***) indicate, respectively, p < 0.05, p < 0.005, and p < 0.001 versus 6-OHDA-treated cells. The difference between control and 6-OHDA-treated cells was considered significant.
Figure 4. The influence of polyphenolic compounds from L. bicolor on cell viability (ac) and ROS levels (d) in Neuro-2a cells treated with 6-OHDA (80 µM). The percentage of living cells treated with compounds and 6-OHDA was measured by MTT assay. Each bar represents the mean ± SEM of three independent replicates. (*), (**), and (***) indicate, respectively, p < 0.05, p < 0.005, and p < 0.001 versus 6-OHDA-treated cells. The difference between control and 6-OHDA-treated cells was considered significant.
Antioxidants 11 00709 g004
Figure 5. The influence of polyphenolic compounds from L. bicolor on cell viability (ac) and ROS levels (d) in Neuro-2a cells treated with rotenone (10 µM). The percentage of living cells treated with compounds and rotenone was measured by the MTT assay. Each bar represents the mean ± SEM of three independent replicates. (*), (**), and (***) indicate, respectively, p < 0.05, p < 0.005, and p < 0.001 versus rotenone-treated cells. The difference between control and rotenone-treated cells was considered significant.
Figure 5. The influence of polyphenolic compounds from L. bicolor on cell viability (ac) and ROS levels (d) in Neuro-2a cells treated with rotenone (10 µM). The percentage of living cells treated with compounds and rotenone was measured by the MTT assay. Each bar represents the mean ± SEM of three independent replicates. (*), (**), and (***) indicate, respectively, p < 0.05, p < 0.005, and p < 0.001 versus rotenone-treated cells. The difference between control and rotenone-treated cells was considered significant.
Antioxidants 11 00709 g005
Table 1. 1H (700 MHz), 13C (175 MHz), HMBC, COSY, and ROESY NMR data for compound 10 (δ in ppm, J in Hz, CDCL3).
Table 1. 1H (700 MHz), 13C (175 MHz), HMBC, COSY, and ROESY NMR data for compound 10 (δ in ppm, J in Hz, CDCL3).
Position13C1HHMBCCOSYROESY
1122.77.85, d, J = 8.5, 1HC-3, 4, 4aH-2H-2
2113.66.95, d, J = 8.5, 1HC-3, 4, 11bH-1H-1
3159.2
4104.07.13, s, 1HC-2, 3, 4a, 11b
4a154.8
6159.0
6a103.8
6b116.0
7105.47.44, s, 1HC-6a, 6b, 8, 9, 10, 10a OH-8
8143.1
9138.7
10106.4
10a145.2
11a160.1
11b106.3
1′115.96.89, d, J = 9.9, 1HC-8, 9, 10, 10a, 3′, 9′H-2′H-2′
2′130.65.78, d, J = 9.9, 1HC-9, 10, 3′, 4′, 9′H-1′H-1′, 4′, 5′, 9′
3′80.6
4′40.91.78, m, 1HC-2′, 3′, 5′, 6′, 9′H-5′H-2′, 5′, 9′
1.84, m, 1HC-2′, 3′, 5′, 6′, 9′H-5′H-2′, 5′, 9′
5′22.82.13, m, 1HC-4′, 6′, 7′H-4′, 6′H-2′, 4′, 9′
2.15, m, 1HC-4′, 6′, 7′H-4′, 6′H-2′, 4′, 9′
6′123.65.10, t, J = 6.9, 1HC-4′, 6′, 8′, 10′H-5′, 8′, 10′H-5′, 8′
7′132.2
8′25.61.67, s, 3HC-4′ (weak), 6′, 7′, 10′H-6′H-6′
9′26.31.50, s, 3HC-2′, 3′, 4′, 5′ (weak) H-2′, 4′, 5′
10′17.61.57, s, 3HC-4′ (weak), 6′, 7′, 8′H-6′
OH-3 6.54, bs, 1H
OH-8 5.50, bs, 1HC-7 H-7
Table 2. 1H (700 MHz), 13C (175 MHz), HMBC, COSY, and ROESY NMR data for compound 11 (δ in ppm, J in Hz, CDCL3).
Table 2. 1H (700 MHz), 13C (175 MHz), HMBC, COSY, and ROESY NMR data for compound 11 (δ in ppm, J in Hz, CDCL3).
Position13C1HHMBCCOSYROESY
1194.7
2194.6
3109.1
4158.9
5115.3
6152.5
7136.8
8113.76.84, s, 1HC-2, 4, 6, 7 OH-7
1′111.1
2′165.2
3′104.36.44, s, 1HC-1′, 2′, 4′, 5′
4′163.7
5′119.9
6′133.97.18, s, 1HC-1, 1′, 2′, 5′, 1‴ H-1‴
1″22.03.51, d, J = 7.3, 2HC-4, 5, 6, 2″, 3″H-2″, 9″H-2″, 9″, OH-6
2″120.55.33, t, J = 7.3, 1HC-1″H-1″, 9″H-1″, 4″, OH-6
3″139.7
4″39.72.09, m, 2HC-2″, 3″, 5″, 6″ H-2″, 6″
5″26.42.13, m, 2HC-3″, 6″, 7″H-6″H-6″
6″123.85.07, m, 1H H-5″, 8″, 10″H-4″, 5″
7″132.1
8″25.71.68, s, 3HC-6″, 7″, 10″H-6″
9″16.31.84, s, 3HC-2″, 3″, 4″H-1″, 2″H-1″
10″17.71.60, s, 3HC-6″, 7″, 8″H-6″
1‴29.13.23, d, J = 7.2, 2HC-4′, 5′, 6′, 2‴, 3‴H-2‴, 4‴, 5‴H-2‴, 4‴, 5‴, 6′, OH-4′
2‴120.85.20, t, J = 7.2, 2H H-1‴, 4‴, 5‴H-1‴, 5‴
3‴136.4
4‴25.71.73, s, 3HC-2‴, 3‴, 5‴H-1‴, 2‴H-1‴
5‴17.91.71, s, 3HC-2‴, 3‴, 4‴H-1‴, 2‴H-1‴, 2‴
OH-4 11.92, s, 1HC-3, 4, 5
OH-6 6.68, s, 1HC-5, 6, 7 H-2″, 1″, OH-7
OH-7 5.05, bs, 1H H-8, OH-6
OH-2′ 11.59, s, 1HC1′, 2′, 3′, 4′
OH-4′ 6.11, s, 1HC-3′, 4′, 5′ H-1‴
Table 3. DPPH-scavenging activity and FRAP of compounds 711.
Table 3. DPPH-scavenging activity and FRAP of compounds 711.
CompoundDPPH-Scavenging EffectFRAP
SC50 µM, 30 minmol (Fe2+)/mol (polyphenolic compound X)
Quercetin8.9 ± 0.25.11 ± 0.31
Ascorbic acid30.5 ± 2.13.43 ± 0.33
124.3 ± 3.7 *1.78 ± 0.16 ***
225.0 ± 2.2 **1.31 ± 0.02 **
323.7 ± 2.1 **1.60 ± 0.08 **
418.0 ± 0.2 ***2.78 ± 0.18 **
526.1 ± 3.0 *0.86 ± 0.04 *
621.8 ± 2.0 *1.43 ± 0.06 **
725.7 ± 3.4 *1.83 ± 0.16 *
826.7 ± 3.8 *1.95 ± 0.19 *
913.5 ± 1.2 **0.75 ± 0.05 **
1022.7 ± 3.5 *1.75 ± 0.20 *
1119.4 ± 2.2 **1.45 ± 0.08 **
1 Data are presented as the mean ± SEM, n = 3. *** p < 0.001, ** p < 0.005, and * p < 0.05 compared to quercetin. 2 Data for compounds 16 have been previously published in [24].
Table 4. Cytotoxic activity of polyphenolic compounds 111 from L. bicolor.
Table 4. Cytotoxic activity of polyphenolic compounds 111 from L. bicolor.
CompoundEC50, µM
172.0
275.0
344.0
476.0
5100.0
640.6
7>100
8>100
975.0
1044.0
1187.0
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tarbeeva, D.V.; Pislyagin, E.A.; Menchinskaya, E.S.; Berdyshev, D.V.; Kalinovskiy, A.I.; Grigorchuk, V.P.; Mishchenko, N.P.; Aminin, D.L.; Fedoreyev, S.A. Polyphenolic Compounds from Lespedeza bicolor Protect Neuronal Cells from Oxidative Stress. Antioxidants 2022, 11, 709. https://doi.org/10.3390/antiox11040709

AMA Style

Tarbeeva DV, Pislyagin EA, Menchinskaya ES, Berdyshev DV, Kalinovskiy AI, Grigorchuk VP, Mishchenko NP, Aminin DL, Fedoreyev SA. Polyphenolic Compounds from Lespedeza bicolor Protect Neuronal Cells from Oxidative Stress. Antioxidants. 2022; 11(4):709. https://doi.org/10.3390/antiox11040709

Chicago/Turabian Style

Tarbeeva, Darya V., Evgeny A. Pislyagin, Ekaterina S. Menchinskaya, Dmitrii V. Berdyshev, Anatoliy I. Kalinovskiy, Valeria P. Grigorchuk, Natalia P. Mishchenko, Dmitry L. Aminin, and Sergey A. Fedoreyev. 2022. "Polyphenolic Compounds from Lespedeza bicolor Protect Neuronal Cells from Oxidative Stress" Antioxidants 11, no. 4: 709. https://doi.org/10.3390/antiox11040709

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