Next Article in Journal
Effect of Steam Blanching and Drying on Phenolic Compounds of Litchi Pericarp
Next Article in Special Issue
Probing Steroidal Substrate Specificity of Cytochrome P450 BM3 Variants
Previous Article in Journal
Aza-Henry Reactions on C-Alkyl Substituted Aldimines
Previous Article in Special Issue
Analysis of Potential Amino Acid Biomarkers in Brain Tissue and the Effect of Galangin on Cerebral Ischemia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Polyphenols from Erythrina crista-galli: Structures, Molecular Docking and Phytoestrogenic Activity

1
Department of Pharmacognosy, Faculty of Pharmacy, Ain Shams University, Cairo 11566, Egypt
2
Department of Biology, Institute of Pharmacy and Molecular Biotechnology, Heidelberg University, Heidelberg 69120, Germany
3
Graduate Institute of Natural Products, Kaohsiung Medical University, Kaohsiung 807, Taiwan
4
Department of Pharmacognosy, Faculty of Pharmacy, British University of Egypt (BUE), Cairo 11837, Egypt
5
Department of Pharmacology and Toxicology, Faculty of Pharmacy, Ain Shams University, Cairo 11566, Egypt
6
Department of Pharmacology, Faculty of Medicine, Umm Al Qura University, Makkah 21955, Saudi Arabia
*
Author to whom correspondence should be addressed.
Molecules 2016, 21(6), 726; https://doi.org/10.3390/molecules21060726
Submission received: 30 March 2016 / Revised: 21 May 2016 / Accepted: 27 May 2016 / Published: 3 June 2016

Abstract

:
Objectives: The current study aimed at exploring the secondary metabolites content of Erythrina crista-galli aqueous methanol extract and assessing its phytoestrogenic and cytoprotective activities. Methods: Isolation of the compounds was carried out using conventional chromatographic techniques. The structures of the isolated compounds were elucidated based on the UV, NMR spectral data along with their mass-spectrometric analyses. The phytoestrogenic activity was evaluated in-silico and in vitro using the Arabidopsis thaliana pER8: GUS reporter assay and the proliferation-enhancing activity of MCF-7 cells. Key findings: Phytochemical investigation of E. crista-galli aqueous methanol extract resulted in the isolation and identification of five flavonoids. The plant extract and its fractions showed significant estrogenic activities compared to controls. Conclusion: Five flavonoids were identified from E. crista-galli aqueous methanol extract. To the best of our knowledge, among these flavonoids, apigenin-7-O-rhamnosyl-6-C-glucoside was isolated for the first time from nature. Moreover, luteolin-6-C-glucoside was isolated for the first time from this plant. The plant revealed promising phytoestrogenic activities. This gives rationale to some of its pharmacological properties and suggests additional phytoestrogenic effects, which have not been reported yet.

1. Introduction

Phenolic compounds are the most widely distributed secondary metabolites in terrestrial plants. They are typical multi-target compounds exerting multiple biological effects [1]. These effects, such as antibacterial [2], antioxidant [3], phytoestrogenic [4,5], cytotoxic [6] and hepatoprotective activities, have been thoroughly investigated [7,8].
The family Fabaceae (the legume family) includes about 730 genera with more than 19,500 species. Plants belonging to this family biosynthesize and accumulate large amounts of phenolics, particularly flavonoids, phenolic acids and tannins [9]. Red clover (Trifolium pratense), rich in phytoestrogenic isoflavones, such as genistein and daidzein, and soybean (Glycine max), with the methylated isoflavones, biochanin A and formononetin [10], are known examples of phenolic-rich and bioactive legumes. The heart wood of Acacia [11] and the roots of Glycyrrhiza glabra are important sources of phenolic compounds [12]. Some of these phenolics exhibit a potent hepatoprotective effect [13].
The genus Erythrina (Fabaceae) contains more than 100 species distributed in the tropics and subtropics of America, Africa and Australasia [14]. The origin of the name Erythrina comes from the Greek word “erythros” which means red, referring to the bright red flowers of the genus trees. Many Erythrina species, with their attractive flowers, are cultivated as ornamental plants in many tropical and subtropical areas of the world. These species have been widely used in traditional medicine for the treatment of insomnia, malaria fever, venereal disease, asthma and toothache [15]. Most of the phytochemical investigations on Erythrina species were directed towards their neurotoxic alkaloids, which are a special class of isoquinoline alkaloids [16,17,18]. Additionally, some Erythrina species were reported to contain a novel class of phytoestrogenic alkaloids, named as erythroidine alkaloids [19]. Furthermore, studies on E. poeppigiana indicated the presence of phenolic constituents with inhibitory potential against human glyoxalase I [20]. Other secondary metabolites such as flavonoids [21], isoflavonoids [22,23], pterocarpanes [24,25], flavanones [26,27], isoflavans [28], chalcones [29,30] and cinnamoyl phenols [31,32] were also reported.
Erythrina extracts exhibit a wide range of pharmacological properties, including antibacterial [22,33], antimalarial [34], antioxidant [17,35], anti-inflammatory [36], cytotoxic [37,38,39], tyrosine-protein phosphatase inhibitory [40], phytoestrogenic [19,41,42], anti-osteoporotic [43] and neurotoxic activities [44]. E. crista-galli is widely distributed in subtropical and tropical regions of South America. It is known as Cockspur Coral Tree and commonly called “Corticeira” in Brazil. Its bark is used for the treatment of many ailments related to rheumatism, and for hepatitis [45]. Phytochemical studies on the underground parts revealed several Erythrina alkaloids, isoflavonoids and pterocarpans [46]. The present study focused on the identification of the major phenolic secondary metabolites from the aerial parts (leaves) of Erythrina crista-galli. Potential phytoestrogenic activities of E. crista-galli leaves extracts and fractions were also investigated.

2. Results and Discussion

Apigenin-7-O-rhamnosyl-6-C-glucoside (1), a minor compound, was obtained as an amorphous yellow powder, which appeared as a dark purple spot on the PC and turned yellow upon exposure to ammonia vapors, under UV light. Rf = 0.34 (6% AcOH) and 0.27 (BAW). The UV spectral data of the compound revealed two distinct maxima, one at 266 nm for band II and one at 336 nm for band I, indicating a flavone skeleton. A bathochromic shift (50 nm) was observed in band I upon addition of sodium methoxide, indicating the probable presence of a free hydroxyl group at position 4′. No bathochromic shift was observed in band I on the addition of aluminium chloride, indicating the absence of ortho-dihydroxyl groups. The absence of ortho-dihydroxyl groups was further confirmed by the absence of the bathochromic shift in band I in the sodium acetate/boric acid spectrum. No bathochromic shift was observed in band II in the presence of sodium acetate, indicating the absence of a free hydroxyl group at position 7. Negative ESI-mass spectral analysis ([M − 1] at m/z = 577, corresponding to a molecular mass of 578), together with the above given analytical data, suggested an apigenin molecule containing a hexoside and rhamnoside moieties; at least one of these moieties occupies position 7.
The proposed structure of this minor constituent was then proved by 1H-NMR and COSY analysis (DMSO-d6, room temperature). The presence of 7-O-substitution of apigenin was concluded from the present downfield shifts of the aromatic protons on ring A, which showed a signal at δ 6.72 (1H, s, H-8). The identity of apigenin was confirmed by the presence of signals at δ 6.91 (2H, d, J = 9 Hz, H-3′ and H-5′), 7.92 (2H, d, J = 9 Hz, H-2′ and H-6′) and 6.91 (1H, s, H-3). The absence of the H-6 signal indicated C-substitution at the position of this carbon. Moreover, the anomeric proton of the hexosyl moiety displayed a signal at δ 4.57 (1H, d, J = 9 Hz, Hglu-1′′). The appearance of the H-1′′ proton of this moiety at δ 4.57 suggests a C-hexosyl linkage, because in case of O-glycosidic linkage, the same proton should appear at a downfield location. That the hexoosyl moiety in compound 1 exists in β-configuration, was evidenced by the splitting pattern of its anomeric proton resonance into doublet at δ 4.57 (1H, d, J = 7 Hz, H-1′′), which indicated an axial-axial coupling between H-1′′ and H-2′′ sugar protons. The anomeric proton of rhamnose (identified as rhamnose from the methyl substituent) displayed a signal at δ 5.47 (1H, d, J = 2 Hz, Hrha-1′′′), a location which is downfield enough to suggest an O-type of glycosidation. A C-rhamnosyl substitution will bring that anomeric proton comparatively upfield (e.g., at δ = 4.8 ppm). Other signals of sugar protons appeared between 3.1–4.2 and 1.2 (3H, d, J = 6.0 Hz, CH3-6′′′). Finally, enzymatic hydrolysis of compound 1 by α rhamnase enzyme yielded compound 2.
The above given data suggested, therefore, an apigenin-7-O-rhamnosyl-6-C-glucoside structure for the minor constituent (1), a new natural flavone glycoside (Figure 1) [47].
Apigenin-6-C-glucoside (2) (Figure 1) was obtained as an amorphous yellow powder, which appeared as a dark purple spot on the PC, which turned yellow upon exposure to ammonia vapors, as seen under UV light. Rf = 29 (6% AcOH) and 48 (BAW). 1H-NMR (500 MHz, DMSO-d6, room temp.) δ: 7.80 (2H, d, J = 8.5 Hz, 2′, 6′-H), 6.89 (2H, d, J = 8.4 Hz, 3′, 5′-H), 6.53 (1H, s, 3-H), 6.18 (1H, s, 8-H), 4.55 (1H, d, J = 9.8 Hz, 1′′-H), 4.03–3.11 (6H, m, sugar). 13C-NMR (500 MHz, DMSO-d6) δ: 163.32 (C-2), 102.60 (C-3), 181.73 (C-4), 160.64 (C-5), 109.77 (C-6), 163.32 (C-7), 95 (C-8), 157.38 (C-9), 102.95 (C-10), 121.46 (C-1′), 128.35 (C-2′, 6′), 116.53 (C-3′, 5′), 161.32 (C-4′), 81.41 (C-1′′), 70.44 (C-2′′), 74.25 (C-3′′), 70.20 (C-4′′), 79.54 (C-5′′), 61.37 (C-6′′). The UV (neat and after addition of shifting reagents) and NMR spectral data were identical to those reported for apigenin-6-C-glucoside (isovitexin) [48].
Luteolin-6-C-glucoside (3) (Figure 1) was obtained as an amorphous yellow powder, which appeared as a dark purple spot on the PC under UV light and turned yellowish green upon exposure to ammonia vapors and then turned dirty green after spraying with 1% methanol ferric chloride solution. Rf = 37 (6% AcOH) and 44 (BAW). 1H-NMR (500 MHz, DMSO-d6) δ: 7.38 (1H, dd, J = 2.5 Hz, 9.0 Hz, 6′-H), 7.38 (1H, d, J = 2.5 Hz, 2′-H), 6.89 (1H, d, J = 9.0 Hz, 5′-H), 6.52 (1H, S, 3-H), 4.45 (1H, d, J = 10.0 Hz, 1′′-H). 13C-NMR (500 MHz, DMSO-d6) δ: 163.45 (C-2), 103.75 (C-3), 182.54 (C-4), 160.57 (C-5), 107.72 (C-6), 164.79 (C-7), 93.76 (C-8), 157.24 (C-9), 102.46 (C-10), 121.56 (C-1′), 112.92 (C-2′), 145.50 (C-3′), 149.50 (C-4′), 115.35 (C-5′), 118.88 (C-6′), 81.18 (C-1′′), 70.36 (C-2′′), 73.86 (C-3′′), 71.16 (C-4′′), 78.69 (C-5′′), 61.44 (C-6′′).The UV (neat and after addition of shifting reagents) and NMR spectral data were similar to those reported for luteolin-6-C-glucoside (Isoorientin) [49], which is reported here for the first time in Erythrina crista-galli [48].
Compound 4 5,7,4′-trihydroxyflavone (apigenin) (Figure 1), was obtained as a yellow crystalline material, Rf = 49 (6% acetic acid) and 86 (BAW). Compound 4 appeared as a dark purple spot on the PC under UV. The 1H-NMR spectral data showed the following resonance peaks at δ (ppm) 6.03 (d, J = 2.5 Hz, H-6), 6.31 (d, J = 2.5 Hz, H-8), 6.64 (s, H-3), 6.88 (d, J = 8 Hz, H-3′, 5′), 7.85 (d, J = 8 Hz, H-2′, 6′). Comparison of the obtained spectral data of compound 4 with those reported for apigenin confirmed its identity [49]. Apigenin is reported here for the first time from Erythrina crista-galli.
Compound 5 5,7,3′,4′-tetrahydroxyflavone (luteolin) (Figure 1), was obtained as a yellow crystalline material, Rf = 66 (6% acetic acid) and 78 (BAW). Compound 5 appeared as a dark purple spot on the PC under UV. The 1H-NMR spectral data showed the following resonance peaks at δ (ppm) 6.19 (d, J = 2.5 Hz, H-6), 6.46 (d, J = 2.5 Hz, H-8), 6.85 (s, H-3), 7.5 (d, J = 2.5 Hz, H-2′), 6.89 (d, J = 8 Hz, H-5′), 7.45 (dd, J = 2.5 Hz and J = 8 Hz, H-6′). Comparison of the obtained spectral data of compound 5 with those reported for luteolin confirmed its identity [50]. Luteolin is reported here for the first time from Erythrina crista-galli.

2.1. Docking Study

The binding mode of compound E2 revealed formation of three hydrogen bonds with the residues Arg 394, Glu 353 and His 524 of the of human estrogen receptor R (ERR) (PDB ID 1A52) [51]. The isolated compounds were able to interact with the active site in the same pattern as the lead E2. The results showed that among the isolated compounds, apigenin-7-O-rhamnosyl-6-C-glucoside, apigenin-6-C-glucoside, luteolin-6-C-glucoside, apigenin and luteolin exhibit the same binding mode of the lead compound E2, with higher fitting scores (Table 1). The apigenin-7-O-rhamnosyl-6-C-glucoside binding mode is illustrated in (Figure 2A–C), with a fitting score of 44.7, which revealed an extra hydrogen bonding with more amino acids in the active pocket, in addition to the three key amino acids Arg 394, Glu 353 and His 524, indicating a firmer binding. However, the three compounds of interest share the same H-bonding of the lead compound with the same key amino acids, which suggests an identical binding mode with the active site with lead and, hence, probably the same biological activity. Interestingly, all the docked compounds are characterized by possessing one or more phenolic groups, which dissociate under physiological conditions resulting in O ions [1]. The polyphenols can form ionic bonds with positively charged side chains of the aspartate aminoacids (Asp 351).These bonds are quite weak, but because several of them are formed concomitantly, the effect is much stronger, leading to more fitting in the binding pocket. In addition, the docking study also revealed that simple phenyl-substituted coumarin derivatives are more likely to have the same binding mode as lead and hence the same biological activity. Accordingly, the molecular modeling studies that were performed indicated that the proposed molecules are promising candidates for phytoestrogenic activity, as compared to 17β-estradiol.

2.2. Phytoestrogenic Activity (pER8: GUS Reporter Assay)

The phytoestrogenic activity of the extract and the fractions as well as the reference compound 17β-estradiol were evaluated utilizing the Arabidopsis thaliana pER8: GUS reporter assay system [52]. The results of the minimal active concentration (MAC) of the 17β-estradiol (E2) and samples are presented in Figure 3. Fractions I and II exhibited the most obvious phytoestrogenic activity with MAC values lower than 6.25 µg/mL. Aqueous and aqueous methanol extracts showed less activity with MAC values of 25 and 12.5 µg/mL, respectively. The observed phytoestrogenic activity of Erythrina crista-galli and its fractions is supported by several in vitro and in vivo studies reporting the estrogenic activity of another Erythrina sp., E. poeppigiana [19,42]. In E. lysistemon, the phytoestrogenic activity is related to the prenylated flavonoid abyssinone V-4′-methyl-ether [53]. This finding is in agreement with our previous work showing the binding effect of prenylated flavonoids to the uterine estrogen receptor of rats [54]; the presence of a 5-hydroxy group and an 8-prenyl group were essential for such binding. In our previous papers, the reliability of the method was validated through comparing the phytoestrogenic results obtained by the pER8: GUS reporter assay with the results obtained by the standard in vitro model of the SEAP reporter assay system in the MCF-7 breast cancer cell-line [55,56]. The results of the pER8: GUS reporter assay were in agreement with the results of the SEAP reporter assay system in the MCF-7 breast cancer cell-line. However, the major limitation for this assay is its relative lower sensitivity (minimum active concentration of E2 at 1.25–0.63 nM) compared to the in vitro SEAP reporter assay (minimum active concentration 0.01 nM) [55].

2.3. Proliferation-Enhancing Activity in MCF-7 Cells

The phytoestrogenic activity of Erythrina crista-galli aqueous and aqueous methanol extracts and its active fractions (I and II) was further substantiated through assessing their effect on the proliferation of the estrogen-responsive cell line MCF-7. Initially, the cytotoxicity of the extracts and its active fractions (I and II) was determined using the sulphorhodamine B (SRB) assay. The aqueous extract showed an IC50 value of 187 ± 11.2 µg/mL, while that of the aqueous methanol extract was above 1000 µg/mL. IC50 values of fractions I and II were 23.3 ± 1.9 and 51.5 ± 3.5 µg/mL, respectively (Table 2). The aqueous extract of Erythrina was previously reported to possess cytotoxic activity against breast and lung cancer cell lines [39]. The observed superior activity of the aqueous extract over the aqueous methanol extract is referred to the presence of many other secondary metabolites in the multi-component plant extract. The presence of low amount of erythrinan (with their unique skeleton of a tetracyclic spiroamine) and benzylisoquinoline alkaloids may be regarded as the main reason for the relatively higher cytotoxic activity. These alkaloids showed IC50 values higher than 25 µg/mL when they were assessed against the murine macrophages RAW264.7 [46]. However, more potent activity of the total alkaloids fraction was noticed against HepG-2, HEP-2, HCT116, MCF-7 and HFB4, with IC50 values in the range of 2.97‒21.1 µg/mL [56].
In this context, isolation of these alkaloids was not the scope of this work. The main objective was to evaluate the phytoestrogenic and cytoprotective activity of the plant extract. The aqueous methanol extract showed nontoxic effect on the tested cells, even in a high concentration. This can be attributed to the presence of only a low amount of the alkaloids for a solubility reason. However, the relatively higher cytotoxicity of the fractions may be attributed to the presence of other secondary metabolites, especially pterocarpan. These compounds showed potent inhibitory activity on protein-tyrosine phosphatase-1B, which is involved in many phosphorylation processes associated with cell growth and mitotic division [40]. Therefore, we selected the aqueous methanol extract together with fractions I and II for MCF-7 proliferation studies. The aqueous methanol extract (10 µg/mL) resulted in enhanced proliferation for MCF-7 cells. Sub-cytotoxic concentrations of fractions I and II (1 and 10 µg/mL) showed significant increases in the proliferation rate of MCF-7 cells at day 7 and day 10 (Table 3 and Figure 4). These results further support the phytoestrogenic activity evidenced by the Arabidopsis thaliana pER8: GUS reporter assay elaborated above.

3. Materials and Methods

3.1. Plant Material

Fresh leaves of Erythrina crista-galli L. were collected in April 2014 from plants grown in El-Giza Zoo garden, Giza. They were kindly authenticated morphologically by Mrs. Treese Labib, agricultural expert, El-Orman Botanical Garden, Giza, Egypt. Voucher specimens of the plant material were deposited at the Department of Pharmacognosy, Faculty of Pharmacy, Ain Shams University, Cairo, Egypt (P-NA-302).

3.2. Cell Lines and Chemicals

The human breast adenocarcinoma (MCF-7) cell line was obtained from the Egyptian Holding Company for Biological Products and Vaccines (VACSERA, Giza, Egypt). Ammonium nitrate, magnesium sulfate, potassium nitrate, boric acid, cobalt chloride, cupric sulfate, manganese sulfate, potassium iodine, sodium molybdate, zinc sulfate, citric acid and sodium citrate were purchased from Wako Pure Chemical Industries® (Tokyo, Japan). Calcium chloride, dextrose, potassium hexacyanoferrate II, potassium hexacyanoferrate III and sodium carbonate were obtained from Merck® (Darmstadt, Germany), while potassium phosphate, sodium chloride, ferrous sulfate, Na2-EDTA2H2O, nicotinic acid, pyridoxineHCl, thiamineHCl, Fe-EDTA, 4-morpholineethanesulfonic acid, τ-inositol, sucrose, phytoagar, sodium phosphate, silymarin, triton X-100, 5-bromo-4-chloro-3-indoly-β-d-glucuronide and α-Rhamnase were purchased from Sigma-Aldrich Corp. (St. Louis, MO, USA). 17β-estradiol was purchased from Takeda Chemical Industries Ltd. (Osaka, Japan). Solvents used were of analytical grade unless mentioned and were purchased from Merck, J.T. Backer® (Deventer, The Netherlands) and Theo Seulberger® (Karlsruhe, Germany). Media and supplements for cell cultures were obtained from Gibco®/Invitrogen (Karlsruhe, Germany) and Greiner Labortechnik® (Frickenhausen, Germany).

3.3. Phytochemistry

1H- and 13C-NMR were measured on a Varian AC 500 MHz NMR spectrometer (Varian®, Santa Clara, CA, USA) at the operating frequencies of 500.13 and 125.67 MHz, respectively. Samples were dissolved in dimethyl sulphoxide (Deutero®, Kastellaun, Germany). The chemical shift (δ) values are reported in ppm in DMSO, and coupling constants (J-value) are recorded in Hz. UV recording was done on a Shimadzu UV Visible-1601 spectrophotometer, Shimadzu® (Kyoto, Japan). The molecular weights were recorded using a Bruker micrOTOF-QII (Bruker Daltonics, Bremen, Germany) mass spectrometer operating at 70 eV with an ion source temperature 200 °C over the range 200–900 m/z. Chromatograms were processed using Bruker Compass Data Analysis 4.0 software.
Column chromatography (CC) was carried out on either silica gel 60 (0.063–0.2 mm) (Merck®, Darmstadt, Germany), polyamide S6, (Riedel-De Haen®, Hannover, Germany) and a Sephadex LH-20 (25–100 μm) (GE Healthcare Bio-Sciences®, Uppsala, Sweden). Fractions were monitored by thin layer chromatography (TLC) on either pre-coated silica gel 60 F254 (0.25 mm) (Merck®) using p-anisaldehyde/sulphuric acid reagent and heating at 100 °C for 7 min [57]. Paper chromatography (PC) was carried out on Whatmann No. 1 and No. 3 papers (Whatmann™, Kent, UK); localization of the spots on PC was performed by exposure to ammonia vapour.

3.4. Extraction and Isolation

E. crista-galli fresh leaves (1 kg) were chopped in double-distilled water (4 L) for 6 h three times. The aqueous extract was concentrated under reduced pressure followed by lyophilization. The dry lyophilized extract was then further extracted with methanol (HPLC grade) for 30 min at 40 °C to ensure complete extraction of the phenolic components. The extract was completely evaporated in vacuo (45 °C) until constant weight, to yield a solid residue (30 g).
Fractionation of the extract on a cellulose column using water, followed by water–methanol mixtures of decreasing polarities yielded two main fractions (I–II), which were individually subjected to further purification using sub-columns and preparative paper chromatography (PPC). Re-fractionation of fraction I over a polyamide 6S column and elution with H2O, followed by H2O/MeOH mixtures of decreasing polarities, yielded 6 main sub-fractions (I-A-IF). Compound 1 was obtained from sub-fraction I-C (eluted with 40% methanol) through subsequent fractionation over column chromatography using Sephadex LH-20 with H2O, followed by H2O/MeOH mixtures of decreasing polarities as the solvent system. Further purification was done using PPC eluted with n-butanol:acetic acid:water (BAW). Compounds 2 and 3 were obtained from sub-fractions I-D (eluted from 60% MeOH) through subsequent fractionation over a Sephadex LH-20 column, with n-butanol saturated with water as the solvent system. Further purification was done using PPC with BAW as the solvent system separating 2 from 3. Compounds 4 and 5 were obtained from fraction II (eluted with neat MeOH) and purified through subsequent PPC using BAW as the solvent system [58].

3.5. Molecular Docking Study

The molecular docking study of the isolated compounds from the plant on the human estrogen receptor was carried out using Discovery Studio 2.5 software (Accelrys Inc., San Diego, CA, USA) Gold protocol. The X-ray crystal structure of human estrogen receptor R (ERR) (PDB ID 1A52) was downloaded from the Protein Data Bank (www.pdb.org). The structure of the receptor was established using the default protein preparation protocol of Accelry’s discovery studio 2.5 [59]. The study was started by determining the binding mode of bioactive conformation of the reported lead compound E2 crystallized with ER. The structure was obtained from the Protein Data Bank without change in its conformation to investigate the detailed intermolecular interactions between the ligand and the target receptor. A validation for the ideal pose was also performed by alignment of the X-ray bioactive conformer, with the best fitted pose of the same compound. The alignment showed good coincidence between them (RMSD = 0.21 Å), indicating the validity of the selected pose. Interactive docking using Gold protocol was carried out between the test set of molecules and the binding site of the prepared ER. Each test set molecule gave 10 possible docked poses. The ideal pose of each molecule was selected according to the similarity of its binding mode in the binding site to that of the lead compound. The binding pattern of the ideal pose for each of the test set molecules and the corresponding Gold fitness score were considered in our study to prioritize their virtual affinity to the binding site, in comparison to the ideal pose of the lead molecule E2. The predicted binding energies of the compounds are listed in Table 1.

3.6. Biological Activity

3.6.1. Estrogenic Activity: pER8: GUS Reporter System

The Arabidopsis pER8: GUS reporter assay system was originally developed by Brand et al. [60]. The seeds of Arabidopsis pER8: GUS were sown and grown on MS solid medium (3% sucrose, 0.9% agar) in the dark for 24–36 h at 4 °C for vernalization, and then left at 24 °C for three days under continuous light. The seedlings were transferred into 24-well microtiter plates containing MS liquid medium with or without the tested compounds in each well, and incubated at 24 °C for 48 h. The culture medium was removed, then the staining buffer (50 mM Na3PO4 buffer (pH 7.0), 10 mM EDTA (pH 8.0), 0.2 mM X-Gluc, 0.5 mM K3Fe(CN)6, 0.5 mM K4Fe(CN)6 and 0.1% Triton X-100) was added for GUS staining at 37 °C for three hours [61]. The reference compound, 17β-estradiol, was used as a positive control (0.31–10 nM). A ZEISS Axiovert 200 inverted microscope (Carl Zeiss, Oberkochen, Germany) was used to examine GUS staining and images were captured with a digital camera.

3.6.2. Cell Culture, Cytotoxicity and Cell Proliferation Assay

MCF-7 cells were maintained in Dulbecco’s minimal essential medium (DMEM), free of serum and phenol red and supplemented with 10% charcoal-stripped fetal bovine serum and 1% antibiotic-antimycotic solution. Cells were grown at 37 °C in a humidified atmosphere of 5% CO2. All experiments were performed with cells in the logarithmic growth phase.
Sensitivity to drugs was determined in triplicate using the SRB (sulforhodamine B) cell viability assay [62]. Exponentially growing MCF-7 cells were collected using 0.25% Trypsin-EDTA and plated in 96-well plates from Greiner Labortechnik® at 1000 and 2000 cells/well, respectively. The cells were cultured for 24 h, then incubated with various concentrations (0.1–1000 µg/mL) of the tested samples (stock solution 1 mg/mL) at 37 °C for 72 h, and subsequently fixed with 10% trichloroacetic acid (TCA) for 1 h at 4 °C. After several washing steps, cells were exposed to 0.4% SRB for 10 min in the dark and subsequently washed with 1% glacial acetic acid. After drying overnight, Tris-HCl was used to dissolve the SRB-stained cells for 5 min on a shaker at 1600 rpm and optical density intensity was measured photometrically at 545 nm, using a microplate reader, ChroMate® 4300, (Awareness Technology Inc., Palm City, FL, USA). The cell viability (%) of the three independent experiments was calculated using the Emax model as follows:
%  Cell viability = ( 100 R ) × ( 1 [ D ] m K d m + [ D ] m ) + R
where R is the residual unaffected fraction (the resistance fraction), [D] is the drug concentration used, Kd is the drug concentration that produces a 50% reduction of the maximum inhibition rate and m is a Hill-type coefficient [47]. The assay was repeated with minor modification in order to evaluate the effect of sub-lethal concentrations (1 and 10 µg/mL) of fractions I and II and the aqueous methanol extract for up to 10 days. Cells were fixed at 1, 3, 7 and 10 days after exposure to different treatments.

3.6.3. Statistical Analysis

All experiments were carried out three times unless mentioned otherwise. Continuous variables were presented as mean ± S.E. The IC50 was determined as the concentration of a sample, which resulted in a 50% reduction in cell viability or inhibition of the biological activity. IC50 values were calculated using a four-parameter logistic curve (SigmaPlot 11.0, Systat Software, Inc., Richmond, CA, USA), and all the data were sketched and statistically evaluated using one-way ANOVA followed by Tukey–Kramer multiple comparisons tests (GraphPad Prism 6.01, GraphPad Software, Inc., San Diego, CA, USA) when the significance value is <0.05 using the same significance level. The criterion for statistical significance was taken as p < 0.05.

4. Conclusions

This phytochemical study indicates the isolation of apigenin-7-O-rhamnosyl-6-C-glucoside as a new flavonoid glycoside and the isolation of luteolin-6-C-glucoside for the first time from Erythrina crista-galli. The plant exhibits interesting phytoestrogenic activities. Epidemiologal evidence indicates that the incidence of breast cancer is lower in Asian populations who consume high dietary concentrations of soy products which have a high phytoestogens content [63,64]. The current data give support to the traditional use of the herb in some ailments and warrant further investigations.

Acknowledgments

The authors would like to thank Rabah Taha (Department of Pharmaceutical Chemistry, Ain Shams University) for her valuable help in carrying out the molecular modeling experiment, T. Timmermann (Department of Pharmaceutical Chemistry, Heidelberg University) for his technical assistance in recording NMR spectra, Yu-Chi Tsai (Graduate Institute of Natural Products, Kaohsiung Medical University) for helping in assessing the estrogenic activity and Ahmed M. El-Abd, Department of Pharmacology, Medical Division, National Research Centre, Dokki, Giza, Egypt, for his help in the proliferation assays.

Author Contributions

N.S.A., M.L.A., F.-R.C., A.B.A.-N., and N.A. conceived and designed the experiments; N.S.A., M.L.A., M.E.-S. and N.S. performed the experiments; A.B.A.-N., M.W. and N.S. analyzed the data. M.W., M.E.-S. and N.A. contributed some reagents, materials and analysis tools. N.S.A., M.L.A., N.S., A.B.A.-N. and N.A.A. wrote the paper.

Conflicts of Interest

The authors declare that they have no competing interests.

References

  1. Wink, M. Evolutionary advantage and molecular modes of action of multi-component mixtures used in phytomedicine. Curr. Drug Metab. 2008, 9, 996–1009. [Google Scholar] [CrossRef] [PubMed]
  2. Psotova, J.; Kolář, M.; Soušek, J.; Švagera, Z.; Vičar, J.; Ulrichová, J. Biological activities of Prunella vulgaris extract. Phytother. Res. 2003, 17, 1082–1087. [Google Scholar] [CrossRef] [PubMed]
  3. Fukumoto, L.; Mazza, G. Assessing antioxidant and prooxidant activities of phenolic compounds. J. Agric. Food Chem. 2000, 48, 3597–3604. [Google Scholar] [CrossRef] [PubMed]
  4. Nomura, T.; Fukai, T.; Akiyama, T. Chemistry of phenolic compounds of licorice (Glycyrrhiza species) and their estrogenic and cytotoxic activities. Pure Appl. Chem. 2002, 74, 1199–1206. [Google Scholar] [CrossRef]
  5. Miksicek, R.J. Commonly occurring plant flavonoids have estrogenic activity. Mol. Pharmacol. 1993, 44, 37–43. [Google Scholar] [PubMed]
  6. Rao, Y.K.; Geethangili, M.; Fang, S.-H.; Tzeng, Y.-M. Antioxidant and cytotoxic activities of naturally occurring phenolic and related compounds: A comparative study. Food Chem. Toxicol. 2007, 45, 1770–1776. [Google Scholar] [CrossRef] [PubMed]
  7. Yoshikawa, M.; Ninomiya, K.; Shimoda, H.; Nishida, N.; Matsuda, H. Hepatoprotective and antioxidative properties of Salacia reticulata: Preventive effects of phenolic constituents on CCl4-induced liver injury in mice. Biol. Pharm. Bull. 2002, 25, 72–76. [Google Scholar] [CrossRef] [PubMed]
  8. Jain, A.; Soni, M.; Deb, L.; Jain, A.; Rout, S.; Gupta, V.; Krishna, K. Antioxidant and hepatoprotective activity of ethanolic and aqueous extracts of Momordica dioica Roxb. Leaves. J. Ethnopharmacol. 2008, 115, 61–66. [Google Scholar] [CrossRef] [PubMed]
  9. Hegnauer, R.; Renpe, J.; Gpayer-Barkmeijer, R.J. Relevance of seed polysaccharides and flavonoids for the classification of the leguminosae: A chemotaxonomic approach. Phytochemistry 1993, 34, 3–16. [Google Scholar] [CrossRef]
  10. Nestel, P.J.; Pomeroy, S.; Kay, S.; Komesaroff, P.; Behrsing, J.; Cameron, J.D.; West, L. Isoflavones from red clover improve systemic arterial compliance but not plasma lipids in menopausal women. J. Clin. Endocrinol. Metab. 1999, 84, 895–898. [Google Scholar] [CrossRef] [PubMed]
  11. Tindale, M.D.; Roux, D. A phytochemical survey of the Australian species of Acacia. Phytochemistry 1969, 8, 1713–1727. [Google Scholar] [CrossRef]
  12. Li, W.; Asada, Y.; Yoshikawa, T. Flavonoid constituents from Glycyrrhiza glabra hairy root cultures. Phytochemistry 2000, 55, 447–456. [Google Scholar] [CrossRef]
  13. Wang, G.-S.; Han, Z.-W. Effects of flavonoids of glycyrrhiza on ethanol-induced liver injury in mice. Chin. Pharmacol. Bull. 1993, 9, 271–273. [Google Scholar]
  14. Kone, W.M.; Solange, K.-N.E.; Dosso, M. Assessing sub-Saharan Erythrina for efficacy: Traditional uses, biological activities and phytochemistry. Pak. J. Biol. Sci. 2011, 14, 560–571. [Google Scholar] [PubMed]
  15. De Lima, M.R.F.; de Souza Luna, J.; dos Santos, A.F.; de Andrade, M.C.C.; Sant’Ana, A.E.G.; Genet, J.-P.; Marquez, B.; Neuville, L.; Moreau, N. Anti-bacterial activity of some Brazilian medicinal plants. J. Ethnopharmacol. 2006, 105, 137–147. [Google Scholar] [CrossRef] [PubMed]
  16. Flausino, O.A., Jr.; Pereira, A.M.; da Silva Bolzani, V.; Nunes-de-Souza, R.L. Effects of erythrinian alkaloids isolated from Erythrina mulungu (Papilionaceae) in mice submitted to animal models of anxiety. Biol. Pharm. Bull. 2007, 30, 375–378. [Google Scholar] [CrossRef] [PubMed]
  17. Juma, B.F.; Majinda, R.R.T. Erythrinaline alkaloids from the flowers and pods of Erythrina lysistemon and their DPPH radical scavenging properties. Phytochemistry 2004, 65, 1397–1404. [Google Scholar] [CrossRef] [PubMed]
  18. Wanjala, C.C.; Juma, B.F.; Bojase, G.; Gashe, B.A.; Majinda, R.R. Erythrinaline alkaloids and antimicrobial flavonoids from Erythrina latissima. Planta Med. 2002, 68, 640–642. [Google Scholar] [CrossRef] [PubMed]
  19. Djiogue, S.; Halabalaki, M.; Njamen, D.; Kretzschmar, G.; Lambrinidis, G.; Hoepping, J.; Raffaelli, F.M.; Mikros, E.; Skaltsounis, A.L.; Vollmer, G. Erythroidine alkaloids: A novel class of phytoestrogens. Planta Med. 2014, 80, 861–869. [Google Scholar] [CrossRef] [PubMed]
  20. Hikita, K.; Tanaka, H.; Murata, T.; Kato, K.; Hirata, M.; Sakai, T.; Kaneda, N. Phenolic constituents from stem bark of Erythrina poeppigiana and their inhibitory activity on human glyoxalase I. J. Nat. Med. 2014, 68, 636–642. [Google Scholar] [CrossRef] [PubMed]
  21. KamdemWaffo, A.F.; Coombes, P.H.; Mulholland, D.A.; Nkengfack, A.E.; Fomum, Z.T. Flavones and isoflavones from the west African Fabaceae Erythrina vogelii. Phytochemistry 2006, 67, 459–463. [Google Scholar] [CrossRef] [PubMed]
  22. Tanaka, H.; Atsumi, I.; Shirota, O.; Sekita, S.; Sakai, E.; Sato, M.; Murata, J.; Murata, H.; Darnaedi, D.; Chen, I.S. Three new constituents from the roots of Erythrina variegata and their antibacterial activity against methicillin-resistant Staphylococcus aureus. Chem. Biodivers. 2011, 8, 476–482. [Google Scholar] [CrossRef] [PubMed]
  23. Chukwujekwu, J.C.; van Heerden, F.R.; van Staden, J. Antibacterial activity of flavonoids from the stem bark of Erythrina caffrathunb. Phytother. Res. 2011, 25, 46–48. [Google Scholar] [CrossRef] [PubMed]
  24. Innok, P.; Rukachaisirikul, T.; Phongpaichit, S.; Suksamrarn, A. Fuscacarpans A–C, new pterocarpans from the stems of Erythrina fusca. Fitoterapia 2010, 81, 518–523. [Google Scholar] [CrossRef] [PubMed]
  25. Dao, T.T.; Nguyen, P.H.; Thuong, P.T.; Kang, K.W.; Na, M.K.; Ndinteh, D.T.; Mbafor, J.T.; Oh, W.K. Pterocarpans with inhibitory effects on protein tyrosine phosphatase 1B from Erythrina lysistemon Hutch. Phytochemistry 2009, 70, 2053–2057. [Google Scholar] [CrossRef] [PubMed]
  26. Innok, P.; Rukachaisirikul, T.; Suksamrarn, A. Flavanoids and pterocarpans from the bark of Erythrinafusca. Chem. Pharm. Bull. 2009, 57, 993–996. [Google Scholar] [CrossRef] [PubMed]
  27. Watjen, W.; Suckow-Schnitker, A.K.; Rohrig, R.; Kulawik, A.; Addae-Kyereme, J.; Wright, C.W.; Passreiter, C.M. Prenylated flavonoid derivatives from the bark of Erythrina addisoniae. J. Nat. Prod. 2008, 71, 735–738. [Google Scholar] [CrossRef] [PubMed]
  28. Jang, J.; Na, M.; Thuong, P.T.; Njamen, D.; Mbafor, J.T.; Fomum, Z.T.; Woo, E.R.; Oh, W.K. Prenylated flavonoids with PTP1B inhibitory activity from the root bark of Erythrina mildbraedii. Chem. Pharm. Bull. 2008, 56, 85–88. [Google Scholar] [CrossRef] [PubMed]
  29. Cui, L.; Thuong, P.T.; Lee, H.S.; Njamen, D.; Mbafor, J.T.; Fomum, Z.T.; Lee, J.; Kim, Y.H.; Oh, W.K. Four new chalcones from Erythrina abyssinica. Planta Med. 2008, 74, 422–426. [Google Scholar] [CrossRef] [PubMed]
  30. Na, M.; Jang, J.; Njamen, D.; Mbafor, J.T.; Fomum, Z.T.; Kim, B.Y.; Oh, W.K.; Ahn, J.S. Protein tyrosine phosphatase-1B inhibitory activity of isoprenylated flavonoids isolated from Erythrina mildbraedii. J. Nat. Prod. 2006, 69, 1572–1576. [Google Scholar] [CrossRef] [PubMed]
  31. Sato, M.; Tanaka, H.; Yamaguchi, R.; Oh-Uchi, T.; Etoh, H. Erythrinapoeppigiana-derived phytochemical exhibiting antimicrobial activity against Candida albicans and methicillin-resistant Staphylococcus aureus. Lett. Appl. Microbiol. 2003, 37, 81–85. [Google Scholar] [CrossRef] [PubMed]
  32. Iinuma, M.; Okawa, Y.; Tanaka, T. Three new cinnamylphenols in heartwood of Erythrina crista-galli. Phytochemistry 1994, 37, 1153–1155. [Google Scholar] [CrossRef]
  33. Tanaka, H.; Sudo, M.; Kawamura, T.; Sato, M.; Yamaguchi, R.; Fukai, T.; Sakai, E.; Tanaka, N. Antibacterial constituents from the roots of Erythrina herbacea against methicillin-resistant Staphylococcus aureus. Planta Med. 2010, 76, 916–919. [Google Scholar] [CrossRef] [PubMed]
  34. Andayi, A.W.; Yenesew, A.; Derese, S.; Midiwo, J.O.; Gitu, P.M.; Jondiko, O.J.I.; Akala, H.; Liyala, P.; Wangui, J.; Waters, N.C. Antiplasmodial flavonoids from Erythrina sacleuxii. Planta Med. 2006, 72, 187–188. [Google Scholar] [CrossRef] [PubMed]
  35. Sakat, S.; Juvekar, A. Comparative study of Erythrina indica Lam. (Febaceae) leaves extracts for antioxidant activity. J. Young Pharm. 2010, 2, 63–67. [Google Scholar] [CrossRef] [PubMed]
  36. Njamen, D.; Mbafor, J.T.; Fomum, Z.T.; Kamanyi, A.; Mbanya, J.C.; Recio, M.C.; Giner, R.M.; Manez, S.; Rios, J.L. Anti-inflammatory activities of two flavanones, sigmoidin A and sigmoidin B, from Erythrina sigmoidea. Planta Med. 2004, 70, 104–107. [Google Scholar] [PubMed]
  37. Agrawal, S.K.; Agrawal, M.; Sharma, P.R.; Gupta, B.D.; Arora, S.; Saxena, A.K. Induction of apoptosis in human promyelocytic leukemia HL60 cells by an extract from Erythrina suberosa stem bark. Nutr. Cancer 2011, 63, 802–813. [Google Scholar] [CrossRef] [PubMed]
  38. Kuete, V.; Sandjo, L.P.; Djeussi, D.E.; Zeino, M.; Kwamou, G.M.; Ngadjui, B.; Efferth, T. Cytotoxic flavonoids and isoflavonoids from Erythrina sigmoidea towards multi-factorial drug resistant cancer cells. Investig. New Drugs 2014, 32, 1053–1062. [Google Scholar] [CrossRef] [PubMed]
  39. RathiSre, P.R.; Reka, M.; Poovazhagi, R.; Arul Kumar, M.; Murugesan, K. Antibacterial and cytotoxic effect of biologically synthesized silver nanoparticles using aqueous root extract of Erythrina indica Lam. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2015, 135, 1137–1144. [Google Scholar] [CrossRef] [PubMed]
  40. Nguyen, P.H.; Le, T.V.; Thuong, P.T.; Dao, T.T.; Ndinteh, D.T.; Mbafor, J.T.; Kang, K.W.; Oh, W.K. Cytotoxic and PTP1B inhibitory activities from Erythrina abyssinica. Bioorg. Med. Chem. Lett. 2009, 19, 6745–6749. [Google Scholar] [CrossRef] [PubMed]
  41. Djiogue, S.; Njamen, D.; Halabalaki, M.; Kretzschmar, G.; Beyer, A.; Mbanya, J.C.; Skaltsounis, A.L.; Vollmer, G. Estrogenic properties of naturally occurring prenylated isoflavones in U2OS human osteosarcoma cells: Structure-activity relationships. J. Steroid. Biochem. Mol. Biol. 2010, 120, 184–191. [Google Scholar] [CrossRef] [PubMed]
  42. Njamen, D.; Djiogue, S.; Zingue, S.; Mvondo, M.A.; Nkeh-Chungag, B. In vivo and in vitro estrogenic activity of extracts from Erythrina poeppigiana (Fabaceae). J. Complement. Integr. Med. 2013, 10, 63–73. [Google Scholar] [CrossRef] [PubMed]
  43. Zhang, Y.; Li, Q.; Li, X.; Wan, H.Y.; Wong, M.S. Erythrina variegata extract exerts osteoprotective effects by suppression of the process of bone resorption. Br. J. Nutr. 2010, 104, 965–971. [Google Scholar] [CrossRef] [PubMed]
  44. Vasconcelos, S.M.; Macedo, D.S.; de Melo, C.T.; Paiva Monteiro, A.; Rodrigues, A.C.; Silveira, E.R.; Cunha, G.M.; Sousa, F.C.; Viana, G.S. Central activity of hydro alcoholic extracts from Erythrina velutina and Erythrin amulungu in mice. J. Pharm. Pharmacol. 2004, 56, 389–393. [Google Scholar] [CrossRef] [PubMed]
  45. Hashimoto, G. Illustrated Cyclopedia of Brazilian Medicinal Plants; ABOC-SHA: Kamakura, Japan, 1996; pp. 678–679. [Google Scholar]
  46. Ozawa, M.; Kawamata, S.; Etoh, T.; Hayashi, M.; Komiyama, K.; Kishida, A.; Kuroda, C.; Ohsaki, A. Structures of new erythrinan alkaloids and nitric oxide production inhibitors from Erythrina crista-galli. Chem. Pharm. Bull. 2010, 58, 1119–1122. [Google Scholar] [CrossRef] [PubMed]
  47. Ramarathnam, N.; Osawa, T.; Namiki, M.; Kawakishi, S. Chemical studies on novel rice hull antioxidants. 2. Identification of isovitexin, a C-glycosyl flavonoid. J. Agric. Food Chem. 1989, 37, 316–319. [Google Scholar] [CrossRef]
  48. Koeppen, B.; Roux, D. C-Glycosylflavonoids. The chemistry of orientin and iso-orientin. Biochem. J. 1965, 97, 444–448. [Google Scholar] [CrossRef] [PubMed]
  49. Peng, J.; Fan, G.; Hong, Z.; Chai, Y.; Wu, Y. Preparative separation of isovitexin and isoorientin from Patrinia villosa Juss by high-speed counter-current chromatography. J. Chromatogr. A 2005, 1074, 111–115. [Google Scholar] [CrossRef] [PubMed]
  50. Mabry, T.J.; Markham, K.R.; Thomas, M.B. The Systematic Identification of Flavonoids; Springer Verlag: New York, NY, USA; Heidelberg/Berlin, Germany, 1970; pp. 253–273. [Google Scholar]
  51. Celik, L.; Lund, J.D.; Schiøtt, B. Conformational dynamics of the estrogen receptor α: Molecular dynamics simulations of the influence of binding site structure on protein dynamics. Biochem. J. 2007, 46, 1743–1758. [Google Scholar] [CrossRef] [PubMed]
  52. Chang, F.R.; Hayashi, K.; Chua, N.H.; Kamio, S.; Huang, Z.Y.; Nozaki, H.; Wu, Y.C. The transgenic Arabidopsis plant system, pER8-GFP, as a powerful tool in searching for natural product estrogen-agonists/antagonists. J. Nat. Prod. 2005, 68, 971–973. [Google Scholar] [CrossRef] [PubMed]
  53. MagneNde, C.B.; Njamen, D.; TaneeFomum, S.; Wandji, J.; Simpson, E.; Clyne, C.; Vollmer, G. In vitro estrogenic activity of two major compounds from the stem bark of Erythrina lysistemon (Fabaceae). Eur. J. Pharmacol. 2012, 674, 87–94. [Google Scholar] [CrossRef] [PubMed]
  54. Hillerns, P.I.; Wink, M. Binding of flavonoids from Sophora flavescens to the rat uterine estrogen receptor. Planta Med. 2005, 71, 1065–1068. [Google Scholar] [CrossRef] [PubMed]
  55. Tsai, Y.C.; Chiang, S.Y.; El-Shazly, M.; Wu, C.C.; Beerhues, L.; Lai, W.C.; Wu, S.F.; Yen, M.H.; Wu, Y.C.; Chang, F.R. The oestrogenic and anti-platelet activities of dihydrobenzofuroisocoumarins and homoisoflavonoids from Liriope platyphylla roots. Food Chem. 2013, 140, 305–314. [Google Scholar] [CrossRef] [PubMed]
  56. Mohammed, M.M.; Ibrahim, N.A.; Awad, N.E.; Matloub, A.A.; Mohamed-Ali, A.G.; Barakat, E.E.; Mohamed, A.E.; Colla, P.L. Anti-HIV-1 and cytotoxicity of the alkaloids of Erythrina abyssinica Lam. growing in Sudan. Nat. Prod. Res. 2012, 26, 1565–1575. [Google Scholar] [CrossRef] [PubMed]
  57. Wagner, H.; Bladt, S. Plant Drug Analysis, A Thin Layer Chromatography Atlas, 2nd ed.; Springer-Verlag: Berlin/Heidelberg, Germany, 2009; pp. 3–51. [Google Scholar]
  58. Hassaan, Y.; Handoussa, H.; El-Khatib, A.H.; Linscheid, M.W.; El Sayed, N.; Ayoub, N. Evaluation of plant phenolic metabolites as a source of Alzheimer’s drug leads. Biomed. Res. Int. 2014, 2014, 843263. [Google Scholar] [CrossRef] [PubMed]
  59. Tanenbaum, D.M.; Wang, Y.; Williams, S.P.; Sigler, P.B. Crystallographic comparison of the estrogen and progesterone receptor’s ligand binding domains. Proc. Natl. Acad. Sci. USA 1998, 95, 5998–6003. [Google Scholar] [CrossRef] [PubMed]
  60. Brand, L.; Hörler, M.; Nüesch, E.; Vassalli, S.; Barrell, P.; Yang, W.; Jefferson, R.A.; Grossniklaus, U.; Curtis, M.D. A versatile and reliable two-component system for tissue-specific gene induction in Arabidopsis. Plant Physiol. 2006, 141, 1194–1204. [Google Scholar] [CrossRef] [PubMed]
  61. Lai, W.C.; Wang, H.C.; Chen, G.Y.; Yang, J.C.; Korinek, M.; Hsieh, C.J.; Nozaki, H.; Hayashi, K.I.; Wu, C.C.; Wu, Y.C.; et al. Using the pER8: GUS reporter system to screen for phytoestrogens from Caesalpinia sappan. J. Nat. Prod. 2011, 74, 1698–1706. [Google Scholar] [CrossRef] [PubMed]
  62. 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. J. Natl. Cancer Inst. 1990, 82, 1107–1112. [Google Scholar] [CrossRef] [PubMed]
  63. Mense, S.M.; Hei, T.K.; Ganju, R.K.; Bhat, H.K. Phytoestrogens and breast cancer prevention: Possible mechanisms of action. Environ. Health Perspect. 2008, 116, 426–433. [Google Scholar] [CrossRef] [PubMed]
  64. Miller, P.E.; Snyder, D.C. Phytochemicals and cancer risk: A review of the epidemiological evidence. Nutr. Clin. Pract. 2012, 27, 599–612. [Google Scholar] [CrossRef] [PubMed]
  • Sample Availability: Samples of the compounds 25 are available from the authors.
Figure 1. Chemical structures of the isolated compounds from Erythrina crista-galli. Compound 1: apigenin-7-O-rhamnosyl-6-C-glucoside.
Figure 1. Chemical structures of the isolated compounds from Erythrina crista-galli. Compound 1: apigenin-7-O-rhamnosyl-6-C-glucoside.
Molecules 21 00726 g001

CompoundR1R2R3
Apigenin-7-O-rhamnosyl-6-C-glucoside (1)Hglucoserhamnose
Apigenin-6-C-glucoside (isovitexin) (2)HglucoseH
Luteolin-6-C-glucoside (Isoorientin) (3)OHglucoseH
Apigenin (4)HHH
Luteolin (5)OHHH
Figure 2. Binding mode for apigenin-7-O-rhamnosyl-6-C-glucoside on human estrogen receptor R. (A) 2D binding mode for apigenin-7-O-rhamnosyl-6-C-glucoside on human estrogen receptor R; (B) 3D binding mode for apigenin-7-O-rhamnosyl-6-C-glucoside on human estrogen receptor R; (C) apigenin-7-O-rhamnosyl-6-C-glucoside into the binding site surface of human estrogen receptor R.
Figure 2. Binding mode for apigenin-7-O-rhamnosyl-6-C-glucoside on human estrogen receptor R. (A) 2D binding mode for apigenin-7-O-rhamnosyl-6-C-glucoside on human estrogen receptor R; (B) 3D binding mode for apigenin-7-O-rhamnosyl-6-C-glucoside on human estrogen receptor R; (C) apigenin-7-O-rhamnosyl-6-C-glucoside into the binding site surface of human estrogen receptor R.
Molecules 21 00726 g002
Figure 3. Phytoestrogenic effects of (a) 17β-estradiol (E2) and (b) Fractions I and II from Erythrina crista-galli in the Arabidopsis thaliana pER: GUS system. (a) Estrogenic activity of the reference compound 17β-estradiol was evaluated utilizing the pER8: GUS reporter assay system. MAC of 17β-estradiol was 2.5 nM; (b) Fractions I and II exhibited the most obvious estrogenic activity with the MAC values lower than 6.25 µg/mL. Aqueous and aqueous methanol extracts showed less activity with the MAC values of 25 and 12.5 µg/mL, respectively.
Figure 3. Phytoestrogenic effects of (a) 17β-estradiol (E2) and (b) Fractions I and II from Erythrina crista-galli in the Arabidopsis thaliana pER: GUS system. (a) Estrogenic activity of the reference compound 17β-estradiol was evaluated utilizing the pER8: GUS reporter assay system. MAC of 17β-estradiol was 2.5 nM; (b) Fractions I and II exhibited the most obvious estrogenic activity with the MAC values lower than 6.25 µg/mL. Aqueous and aqueous methanol extracts showed less activity with the MAC values of 25 and 12.5 µg/mL, respectively.

Molecules 21 00726 i001 Molecules 21 00726 i002 Molecules 21 00726 i003 Molecules 21 00726 i004 Molecules 21 00726 i005 Molecules 21 00726 i006 Molecules 21 00726 i007
Blank0.3125 nM0.625 nM1.25 nM2.5 nM5 nM10 nM
(a)
6.25 µg/mL12.5 µg/mL25 µg/mL50 µg/mL100 µg/mL200 µg/mL
Aqueous extract Molecules 21 00726 i008 Molecules 21 00726 i009 Molecules 21 00726 i010 Molecules 21 00726 i011 Molecules 21 00726 i012 Molecules 21 00726 i013
Aqueous methanol extract Molecules 21 00726 i014 Molecules 21 00726 i015 Molecules 21 00726 i016 Molecules 21 00726 i017 Molecules 21 00726 i018 Molecules 21 00726 i019
Fraction I Molecules 21 00726 i020 Molecules 21 00726 i021 Molecules 21 00726 i022 Molecules 21 00726 i023 Molecules 21 00726 i024 Molecules 21 00726 i025
Fraction II Molecules 21 00726 i026 Molecules 21 00726 i027 Molecules 21 00726 i028 Molecules 21 00726 i029 Molecules 21 00726 i030 Molecules 21 00726 i031
(b)
Figure 4. Effect of Erythrina crista-galli extract and fractions on the rate of proliferation of the MCF-7 cell line. * Significantly different from corresponding control at p < 0.05.
Figure 4. Effect of Erythrina crista-galli extract and fractions on the rate of proliferation of the MCF-7 cell line. * Significantly different from corresponding control at p < 0.05.
Molecules 21 00726 g004
Table 1. The computed binding energy values (ΔGbinding) for the molecular docking study for the binding of the isolated components in Erythrina crista-galli extract with human Estrogen Receptor R (ERR).
Table 1. The computed binding energy values (ΔGbinding) for the molecular docking study for the binding of the isolated components in Erythrina crista-galli extract with human Estrogen Receptor R (ERR).
Compound NameBinding Energy (ΔGbinding)
Apigenin-7-O-rhamnosyl-6-C-glucoside−44.7
Apigenin-6-C-glucoside−39.11
Luteolin-6-C-glucoside−41.91
Apigenin−43.33
Luteolin−40.97
17β-estradiol−36.38
The ligand-enzyme interaction energy value (∆Gbinding) was calculated using the following equation: ∆Gbinding = Ecomplex − (EERRα + Eligand), where Ecomplex was the potential energy for the complex of human Estrogen Receptor R (ERR) bound with the ligand, EERRα was the potential energy of the receptor alone, and Eligand was the potential energy for the ligand alone.
Table 2. IC50 values of extracts and fractions from Erythrina crista-galli on the growth of MCF-7 cells.
Table 2. IC50 values of extracts and fractions from Erythrina crista-galli on the growth of MCF-7 cells.
AdditionIC50 (µg/mL)
Aqueous extract187.0 ± 11.2
Aqueous methanol extract>1000
Fraction I23.3 ± 1.9
Fraction II51.5 ± 3.5
Table 3. The proliferation-enhancing activity of Erythrina crista-galli in MCF-7 cells using the SRB assay relative to the control.
Table 3. The proliferation-enhancing activity of Erythrina crista-galli in MCF-7 cells using the SRB assay relative to the control.
DrugViability %
Day 1Day 3Day 7Day 10
Control100.00 ± 9.2200.00± 22.92210.97 ± 22.36210.97 ± 25.36
Aq.alc. extract (10 µg/mL)88.37 ± 9.63204.87 ± 24.14234.14 ± 29.02234.14 ± 29.02
Fraction I (10 µg/mL)91.46 ± 8.41198.78 ± 18.7290.24 ± 27.22 *290.24 ± 21.22 *
(1 µg/mL)86.58 ± 7.19193.90 ± 20.26261.95 ± 32.12279.83 ± 26.34 *
Fraction II (10 µg/mL)93.41 ± 5.97198.78 ± 21.66290.24 ± 31.22 *290.24 ± 21.22 *
(1 µg/mL)92.68 ± 10.85203.66 ± 24.14262.73 ± 30.12276.83 ± 26.34 *
Data are represented in terms of mean ± SEM (n = 3). * Significantly different from control at p < 0.05.

Share and Cite

MDPI and ACS Style

Ashmawy, N.S.; Ashour, M.L.; Wink, M.; El-Shazly, M.; Chang, F.-R.; Swilam, N.; Abdel-Naim, A.B.; Ayoub, N. Polyphenols from Erythrina crista-galli: Structures, Molecular Docking and Phytoestrogenic Activity. Molecules 2016, 21, 726. https://doi.org/10.3390/molecules21060726

AMA Style

Ashmawy NS, Ashour ML, Wink M, El-Shazly M, Chang F-R, Swilam N, Abdel-Naim AB, Ayoub N. Polyphenols from Erythrina crista-galli: Structures, Molecular Docking and Phytoestrogenic Activity. Molecules. 2016; 21(6):726. https://doi.org/10.3390/molecules21060726

Chicago/Turabian Style

Ashmawy, Naglaa S., Mohamed L. Ashour, Michael Wink, Mohamed El-Shazly, Fang-Rong Chang, Noha Swilam, Ashraf B. Abdel-Naim, and Nahla Ayoub. 2016. "Polyphenols from Erythrina crista-galli: Structures, Molecular Docking and Phytoestrogenic Activity" Molecules 21, no. 6: 726. https://doi.org/10.3390/molecules21060726

Article Metrics

Back to TopTop