2,017
Views
11
CrossRef citations to date
0
Altmetric
Research Article

Phenolic metabolites, biological activities, and isolated compounds of Terminalia muelleri extract

, , , , &
Pages 2277-2284 | Received 02 Aug 2017, Accepted 14 Nov 2017, Published online: 27 Nov 2017

Abstract

Context: Terminalia muelleri Benth. (Combretaceae), is rich with phenolics that have antioxidant and cytotoxic activities. No screening studies were published before on T. muelleri.

Objective: The study focused on isolation and identification of secondary metabolites from aqueous methanol leaf extract of T. muelleri and evaluation of its biological activities.

Materials and methods: The n-butanol extract was chromatographed on polyamide 6, and eluted with H2O/MeOH mixtures of decreasing polarity, then separated by different chromatographic tools that yielded 10 phenolic compounds. The antioxidant activity of the extract was evaluated by investigating its total phenolic and flavonoid content and DPPH scavenging effectiveness. The extract and the two acylated flavones were evaluated for their anticancer activity towards MCF-7 and PC3 cancer cell lines. Molecular docking study of the acylated flavones was performed against topoisomerase enzyme.

Results and discussion: Two acylated flavonoids, apigenin-8-C-(2″-O-galloyl) glucoside 1 and luteolin-8-C-(2″-O-galloyl) glucoside 2, were isolated and identified for the second time in nature, with eight tannins (310), from the leaves of T. muelleri. The extract and compound 10 showed the most significant antioxidant activity (IC50 = 3.55 and 6.34 μg/mL), respectively. The total extract and compound 2 demonstrated cytotoxic effect against MCF-7 with IC50 = 29.7 and 45.2 μg/mL respectively, while compound 1 showed cytotoxic effect against PC3 (IC50 = 40.8 μg/mL). The docking study of compounds 1 and 2 confirmed unique binding mode in the active site of human DNA topoisomerase enzyme.

Conclusions: Terminalia muelleri is a promising medicinal plant as it possesses high antioxidant activity and moderate cytotoxic activity against MCF-7.

Introduction

For centuries to date, herbal medicines have been treating both human and animal diseases worldwide while various pharmacological studies continue to validate their uses (Fahmy et al. Citation2015). Those studies focus on a variety of biological activities that include antioxidant, antibacterial, anti-inflammatory and cytotoxic activities. Phenolics are safe natural antioxidants (Farnsworth and Bingel Citation1977), as they retard the progress of many diseases by protecting the body from free radicals. Although many studies have already been undertaken on a number of medicinal plants for their antioxidant activity (Kumaran and Karunakaran Citation2007), great interest in searching for new plant sources of antioxidant activity continues (Göktürk Baydar et al. Citation2007; Prakash et al. Citation2007). The continual use of the Egyptian plants in folk medicine (Abdel-Azim et al. Citation2011; Omara et al. Citation2012), for treatment of various disorders, prompted the researchers to carry out intensive studies on these plants. One of these Egyptian plants is Terminalia (Combretaceae), which comprises 200 species in tropical and subtropical regions around the world (de Morais Lima et al. Citation2012). Terminalia genus ranks amongst the most widely used plants for folk medicine (Cock Citation2015). As such Terminalia species have wide pharmacological actions such as are antioxidant (Lin et al. Citation1997; Cheng et al. Citation2003; Lee et al. Citation2005; Pfundstein et al. Citation2010), antiviral (El Mekkawy et al. Citation1995; Valsaraj et al. Citation1997), antimicrobial (Elizabeth Citation2005), hepatoprotective (Marzouk et al. Citation2002; Eesha et al. Citation2011), antihyperlipidemic, antidiabetic (Latha and Daisy Citation2011), anti-inflammatory (Nair et al. Citation2010) and cytotoxic activities (Ponou et al. Citation2010). Some Terminalia species also possess wound healing properties (Dougnon et al. Citation2014). Different secondary metabolites such as triterpenes, tannins, flavonoids, phenolic acids, lignan and lignan derivatives were reported from Terminalia species (Lee et al. Citation2007; Eloff et al. Citation2008; Pfundstein et al. Citation2010; Fahmy et al. Citation2015). Originated in Australia, T. muelleri Benth. possesses some good antimicrobial activity against Staphylococcus aureus, methicillin-resistant S. aureus, Escherichia coli and Candida albicans (Anan et al. Citation2010; Silva and Serramo Citation2015). The presence of a strong antibacterial activity of T. muelleri is attributed to gallic acid while its antioxidant activity is due to a high phenolic content (Fahmy et al. Citation2015). The main purpose of this study is to investigate, for the first time, the chemical and biological constituents of Terminalia muelleri as there are no screening studies on this plant species. Molecular docking has been carried out to assist with understanding the anticancer activity of the isolated compounds of T. muelleri and explore the binding mode of flavonoids to topoisomerase enzyme.

Materials and methods

General methods

Nuclear magnetic resonance (NMR) were performed on a Bruker AMX 500 instrument (Billerica, MA) with standard pulse sequences operating at 500 MHz in 1H NMR and 125 MHz in 13C NMR, chemical shifts are given in δ values (ppm) using tetramethylsilane as the internal standard and DMSO-d6 as solvent at room temperature and coupling constants were reported in Hz. Infra-red (IR) spectra were recorded with Nexus 670 FT-IR FT-Ramen spectrometer (Gaithersburg, MD) as potassium bromide discs. UV spectra were recorded with Shimadzu UV-1601 (Kyoto, Japan). HRESI-MS was taken on a Micromass Autospec (70 eV) spectrometer. Column chromatography was performed on polyamide 6 and Sephadex LH-20.

Plant material

The leaves of flowering T. muelleri were collected in 2013 from the Zoo garden (Giza, Egypt). The plant species was identified by Dr. M. El-Gibaly, Lecturer of Taxonomy and Consultant for Central Administration of Plantation and Environment. A voucher sample (No.: TM # 0610) was deposited at Department of Pharmacognosy, Faculty of Pharmacy, October-6-University, Egypt.

Extraction, isolation and purification

Powder of the air-dried leaves of T. muelleri (1 kg) was defatted with CHCl3 (3 × 1 L) and extracted with CH3OH:H2O [7:3 (5 × 3 L)] at room temperature. The combined extracts were filtered, evaporated under reduced pressure and lyophilized (150 g). The dry extract (100 g) was redissolved in 2 L with distilled H2O and extracted with n-butanol (5 × 2 L). After evaporation of solvents, the n-butanol extract and the remaining H2O phase gave dark brown solids 30 and 50 g, respectively. The n-butanol extract was loaded on a polyamide 6 column chromatography (50 × 3 cm). The column was eluted with H2O, and then H2O–MeOH mixtures of decreasing polarity and 10 fractions (1 L, each) were collected. The major phenolic fractions obtained were combined into five fractions after chromatographic analysis using paper chromatography. Fraction I (3.5 g) was fractionated by column chromatography on Sephadex LH-20 with aqueous EtOH (0–70%) for elution to give compounds 1 (35 mg) and 7 (25 mg). Fraction II (2.5 g) was subjected to column chromatography on cellulose and n-BuOH saturated with H2O as an eluent to give two major subfractions, then each of them was separately fractionated on a Sephadex LH-20 to yield pure samples 2 (30 mg) and 5 (18 mg). Using the same procedure fraction III (2.2 g) gave chromatographically pure samples 3 (15 mg) and 4 (38 mg). Fraction IV (1.5 g) was chromatographed on Sephadex LH-20 using aqueous acetone (0–25%) for elution to give pure sample 6 (20 mg). Fraction V (1.4 mg) has been separated on Sephadex LH-20 CC using n-butanol-water saturated to give three compounds which were further purified on Sephadex LH-20 CC using MeOH:H2O (1:1) to give pure samples of 8 (18 mg), 9 (20 mg) and 10 (25 mg).

Apigenin 8-C-(2″-O-galloyl) glucoside 1

Yellow amorphous powder; UV λmax (MeOH) nm: 267, 332; IR (KBr) υmax cm−1: 1612, 1654, 3409; 1H NMR (500 MHz, DMSO-d6) δ (ppm): 8.12 (2H, d, J = 8.7 Hz, H-2′,6′), 6.94 (2H, J = 8.7 Hz, H-3′,5′), 6.82 (1H, s, H-3), 6.75 (2H, s, H-2‴,6‴), 6.10 (1H, s, H-6), 4.95 (1H, d, J = 10.1, Hz, H-1″), 5.49 (1H, t, J = 9.80, 9.35 Hz, H-2″), 3.83 (1H, dd, Ha-6″), 3.62 (1H, m, Hb-6″), 3.59 (1H, m, H-3″), 3.54 (1H, t, J = 9.30, 8.95 Hz, H-4″), 3.38 (1H, m, H-5″).13C NMR (125 MHz, DMSO-d6) (see ).

Luteolin 8-C-(2″-O-galloyl) glucoside 2

Yellow amorphous powder; UV λmax (MeOH) nm: 271, 345; IR(KBr) υmax cm−1: 1614, 1651, 3422; 1H NMR δ (ppm): (500 MHz, DMSO-d6) 7.64 (1H, d, J = 8.3, 2 Hz, H-6′), 7.56 (1H, d, J = 2 Hz, H-2′), 6.91 (2H, J = 8.3 Hz, H-5′), 6.75 (2H, s, H-2‴,6‴), 6.68 (1H, s, H-3), 6.10 (1H, s, H-6), 4.95 (1H, d, J = 10.10 Hz, H-1″), 5.49 (1H, t, J = 9.80, 9.40 Hz, H-2″), 3.86 (1H, dd, Ha-6″), 3.64 (1H, dd, Hb-6″), 3.60 (1H, m, H-3″), 3.52 (1H, dd, J = 9.45, 8.95 Hz, H-4″), 3.40 (1H, m, H-5″). 13C NMR (125 MHz, DMSO-d6) (see ).

Table 1. 13C NMR data of compounds 1 and 2 in DMSO-d6 at 125 MHz.

Determination of total phenolic and flavonoids contents

The concentration of total phenolics of the plant extract and fractions was determined according to the method described by Kumar et al. (Citation2008), gallic acid was used as standard. Briefly, a mixture of 100 µL of plant extract (100 µg/mL), 500 µL of Folin–Ciocalteu reagent and 1.5 mL of Na2CO3 (20%) was shaken and diluted up to 10 mL with water. After 2 h, the absorbance was measured at 765 nm using a spectrophotometer. All determinations were carried out in triplicate. The total phenolic concentration was expressed as gallic acid equivalents (GAE). Total flavonoid concentration of plant extract and fractions was determined according to the reported procedure (Kumaran and Karunakaran Citation2007). One hundred microlitres of plant extract (10 mg/mL) in methanol was mixed with 100 µL of 20% AlCl3 in methanol and a drop of acetic acid, and then diluted to 5 mL with methanol. The absorbance was measured at 415 nm after 40 min against the blank. The blank consisted of all reagents and solvent without AlCl3. All determinations were carried out in triplicate; the total flavonoid concentration was expressed as rutin equivalents (RE).

Radical scavenging activity towards DPPH

The 1,1-diphenyl-2-picrylhydrazyl is a stable deep violet radical due to its unpaired electron. In the presence of an antioxidant radical scavenger, which can donate an electron to DPPH, the deep violet colour decolourize to the pale yellow non-radical form (Ratty et al. Citation1988). The change in colourization and the subsequent fall in absorbance are monitored spectrophotometrically at 520 nm. In a flat bottom 96-well microplate, a total test volume of 200 µL was used. In each well, 20 µL of different concentrations (6.25–100 µg/mL final concentration) of tested samples were mixed with 180 µL of ethanolic DPPH and incubated for 30 min at 37 °C. Triplicate wells were prepared for each concentration and the average was calculated. Then photometric determination of absorbance at 520 nm was performed by microplate ELISA reader. The half maximal scavenging capacity (SC50) values for each tested sample and ascorbic acid was estimated via dose curve and was calculated using the curve equation.

Measurement of potential cytotoxicity by SRB assay

Potential cytotoxicity was tested using the method of Skehan et al. (Citation1990). Cells were plated in 96-multiwell plate (104 cells/well) for 24 h before treatment with the tested samples to allow the attachment of cells to the wall of the plate. Different concentrations of the tested samples (0, 1, 2.5, 5 and 10) were added to the cell monolayer, triplicate wells were prepared for each individual dose. Monolayer cells were incubated with the tested samples for 48 h at 37 °C and in atmosphere of 5% CO2. After 48 h cells were fixed, washed and stained with Sulfo-Rhodamine-B stain. Excess stain was washed with acetic acid and attached stain was recovered with Tris EDTA buffer. Finally, the colour intensity was measured in an ELISA reader. The relation between surviving fraction and drug concentration is plotted to get the survival curve of each tumour cell line after the specified compounds. IC50 was calculated as the mean of triplicate determinations ± standard deviation, concentrations in µg/mL.

Docking

Docking simulation study has been performed in the CADD Lab. (Dept. of Medicinal Chemistry; Faculty of Pharmacy; October 6 University) using Molecular Operating Environment (MOE®) version 2014.09, Chemical Computing Group Inc. (Montreal, Canada). The computational software operated under ‘Windows XP’ installed on an Intel Pentium IV PC with a 1.6 GHz processor and 512 MB memory.

Target compounds optimization

The target compounds were constructed into a 3D model using the builder interface of the MOE program. After checking their structures and the formal charges on atoms by 2D depiction, the target compounds were subjected to a conformational search. All conformers were subjected to energy minimization, all the minimizations were performed with MOE until a RMSD gradient of 0.7823 kcal/mol and RMS distance of 0.1Å´ with MMFF94X force-field and the partial charges were automatically calculated. The obtained database was then saved as MDB (Molecular Data Base) file to be used in the docking calculations.

Optimization of the enzymes active site

The X-ray crystallographic structures of human DNA topoisomerase I (70 kDa) complexed with camptothecin and covalent complex with a 22 base pair DNA duplex (Code: 1T8I) were obtained from RCSB-Protein data bank. The enzyme was prepared for docking studies by: (i) adding hydrogen atoms to the system with their standard geometry, (ii) checking the atoms connection and type for any errors with automatic correction, (iii) fixation of the selection of the receptor and its atoms potential and (iv) using MOE Alpha Site Finder for the active site search in the enzyme structure using all default items. Dummy atoms were created from the obtained alpha spheres.

Docking of the target molecules to human DNA topoisomerase I (code: 1T8I) active sites

Docking of the selected conformer in database of the target compounds was done using MOE-Dock software. The enzyme active site file was loaded and the Dock tool was initiated. The program specifications were adjusted to dummy atoms as the docking site, triangle matcher as the placement methodology and London dG adjusted to its default values as the scoring methodology to be used and was adjusted to its default values. The MDB file of the ligand to be docked was loaded and Dock calculations were run automatically, finally the obtained poses were studied and the poses showed best ligand–enzyme interactions were selected and stored for energy calculations.

Results and discussion

Structural elucidation

Two galloylated flavonoids and eight tannins were separated from the alcoholic extract of T. muelleri leaves for the first time using different chromatographic tools (). Compound 1 was isolated as yellow amorphous powder that displayed a dark purple spot on paper chromatogram under UV light, which turned to yellow color when fumed with ammonia vapor, as well as, giving a characteristic color reaction (lemon yellow) with Naturestoff reagent. UV spectral data of 1 was measured and showed that it has a flavone nature (Mabry et al. Citation1970; Harborne et al. Citation1975). The molecular formula of compound 1 was estimated as C28H24O14 with a pseudo molecular ion at m/z 607.1057 [M + Na]. The IR spectrum showed carbonyl, aromatic and hydroxyl bands. 1H NMR spectrum of compound 1 showed an AA′ BB′ system at δ 8.12 and 6.94 ppm for p-disubstituted benzene ring, in addition to, two singlets at δ 6.82 for H-3 and 6.10 for H-6 indicated the presence of apigenin structure, which is substituted at position 8. The appearance of one sharp singlet at δ 6.75 ppm is for the two magnetically equivalent protons of the galloyl group. In the aliphatic region, the appearance of a doublet at δ ppm 4.95 is attributed to the anomeric proton of the sugar with large J value ≥9 Hz of β-C-glycoside. The conspicuous downfield of the triplet signal of proton 2″ of the glucose moiety at δ 5.49 is attributed to the attachment of the gallic acid (Liu et al. Citation1997; Rayyan et al. Citation2010) and the downfield of the anomeric proton confirmed that that the substitution is at 2″. 13C NMR spectrum of compound 1 was very close to those for apigenin-8-C-glycoside as shown in (Markham and Chari Citation1982), the spectrum showed a peak at δ 103.82 due to the glycosylation of the aglycone at C-8 (Agrawal and Bansal Citation1989), in addition to a singlet at δ 181.96 assigned to the carbonyl carbon. Twenty-eight carbon signals, one methylene, 13 methine and 14 quaternary carbons are shown in 13C DEPT experiments, and further confirmation is achieved by HSQC experiment for all carbons carrying protons. The site of attachment of the sugar is confirmed by the observed cross peak between the anomeric proton appeared at δ 4.95 and carbon 8 at δ 103.82 in the HMBC spectrum, indicating that attachment is at C-8, while the attachment of gallic acid to the sugar moiety is confirmed by a remarkable cross peak between proton 2″ at δ 5.49 and the carbonyl carbon of gallic acid at δ 164.79. Depending upon all mentioned data, 1 was elucidated to be apigenin 8-C-(2″-O-galloyl) glucoside. Compound 2 has the same physical properties as 1 except in the color reaction as it gave an orange color with Naturestoff reagent as it has two hydroxyl groups in B ring. The molecular formula of compound 2 was estimated as C28H24O15 from the HRESI-MS with a pseudo molecular ion at m/z 623.1019 [M + Na]. The chemical structure of compound 2 was very similar to that of 1; in 1H NMR the most significant difference between them was the ABX system for B substituted benzene ring that appeared at δ ppm 7.64, 7.56 and 6.91 for protons 6′, 2′ and 5′, respectively, indicating the presence of luteolin moiety. 13C NMR spectrum of compound 2 is illustrated in which was very close to those for luteolin-8-C-glycoside (Agrawal and Bansal Citation1989).13C DEPT experiments revealed the presence of one methylene, 12 methine and 15 quaternary. Based upon spectral data, compound 2 was confirmed to be luteolin 8-C-(2″-O-galloyl) glucoside. HMBC and COSY correlations of compounds 1 and 2 are shown in . Compounds 1 and 2 are isolated for the second time from natural source (Latté et al. Citation2002). The known compounds were identified as 1-O-galloyl-2,3,4,6-dihexahydroxydiphenoyl-β-d-glucopyranoside (3), 1,4,6-tri-O-galloyl-2,3-hex ahydroxydiphenoyl-β-D-glucopyranoside 4, 1,2-di-O-galloyl-4,6-hexahydroxydiphenoyl-β-d-glucopyranoside (5), 1-O-galloyl-2,3,hexahydroxydiphenoyl-β-d-glucopyranoside (isostrictinin) (6), 1-O-galloyl-β-d-glucopyranoside (7), 3,3′,4,tri-O-methylellagic acid (8), ellagic acid (9) and gallic acid (10), by using different spectroscopic analysis and comparing with previously reported data (Nonaka et al. Citation1989; Abdulladzhanova et al. Citation2003; Grace et al. Citation2014).

Figure 1. Chemical structures of isolated compounds of Terminalia muelleri.

Figure 1. Chemical structures of isolated compounds of Terminalia muelleri.

Figure 2. [GRAPHIC] HMBC correlations for 1 and 2. [GRAPHIC] Cosy correlations for H-2″ and H-1″ & 3″.

Figure 2. [GRAPHIC] HMBC correlations for 1 and 2. [GRAPHIC] Cosy correlations for H-2″ and H-1″ & 3″.

Evaluation of antioxidant activity

The free radical scavenging activity of the alcoholic extract and the isolated compounds were evaluated on DPPH (2,2-diphenyl-1-(2,4,6-trinitrophenyl)hydrazinyl). The total extract and compounds 2, 9 and 10 exhibited potent antioxidant scavenging activity towards DPPH, with IC50 values of 3.55, 56.32, 67.54 and 6.34 µg/mL, respectively, while the other compounds exhibited moderate activity (), using ascorbic acid as a positive control. The DPPH radical scavenging activity is known to be related to the nature of phenolics contributing to their electron transfer/hydrogen donating ability (Brand-Williams et al. Citation1995). As previously reported, the antioxidant activity of phenolic compounds depends on their skeleton and the number of hydroxyl groups and their arrangement (Bouchet et al. Citation1998), where the ortho dihydroxy systems of B ring increase radical scavenging activity (Rice-Evans et al. Citation1996). The higher H-donating ability of the free hydroxyl groups, explained the higher radical scavenging activity of the isolated compounds (Natella et al. Citation1999). The total flavonoidal and phenolic content of the extract of T. muelleri leaves was found to be 276.48 ± 28.07 and 162.09 ± 16.67 mg/mL, respectively.

Table 2. DPPH radical scavenging activity of the total extract and the isolated compounds.

Anticancer activity of the acylated compounds and the total extract

The anticancer effect of the total extract and two acylated flavonoids against MCF-7 and PC3 cancer cell lines was evaluated and expressed by median inhibitory growth concentration (IC50) using doxorubicin as a standard (IC50 = 5.10 µg/mL for PC3 and IC50 = 3.83 µg/mL for MCF-7). The extract showed the most significant effect against MCF-7 displaying IC50 value of 29.7 ± 1.54 µg/mL. Compound 2 showed moderate anticancer activity towards MCF-7 (IC50 = 45.2 ± 1.61 µg/mL), while the activity was diminished for 1 as IC50 > 50 µg/mL. On the other hand, compound 1 demonstrated antiprostate cancer activity against PC3 with IC50 value of 40.8 ± 1.75 µg/mL, while the extract and compound 2 showed lower activity towards PC3 with IC50>50 µg/mL ().

Figure 3. The correlations between different concentrations of the total extract of T. muelleri and the surviving fraction of MCF-7 and PC3 cancer cells.

Figure 3. The correlations between different concentrations of the total extract of T. muelleri and the surviving fraction of MCF-7 and PC3 cancer cells.

Figure 4. The correlations between different concentrations of compound 1 of T. muelleri and the surviving fraction of MCF-7 and PC3 cancer cells.

Figure 4. The correlations between different concentrations of compound 1 of T. muelleri and the surviving fraction of MCF-7 and PC3 cancer cells.

Figure 5. The correlations between different concentrations of compound 1 of T. muelleri and the surviving fraction of MCF-7 and PC3 cancer cells.

Figure 5. The correlations between different concentrations of compound 1 of T. muelleri and the surviving fraction of MCF-7 and PC3 cancer cells.

Molecular docking of compounds 1 and 2

The polyphenolic compounds present richly in human diet introduce several chemo-preventative agents (Fresco et al. Citation2006). Many flavonoids undergo redox chemistry including oxidative reactions, but their mechanism of action on topoisomerase is still not well defined (Bandele and Osheroff Citation2007), despite many performed studies. It has been proposed that genistein acts as traditional poison and that (–)-epigallocatechin-3-gallate acts as a redox-dependent poison (Bandele et al. Citation2008). The isolated compounds 1 and 2 are related to this class of polyphenolic compounds. In vitro anticancer activity of the two compounds has been determined on MCF-7 and PC3 cell lines using SRB assay. The results illustrate good activity of these compounds. Moreover, the antioxidant activity of the compounds using DPPH method illustrates pronounced effect. Based on the aforementioned observation, a molecular docking experiment has been carried out to assist understanding the anticancer activity of the isolated compounds. Molecular docking of compounds 1 and 2 was performed on human DNA topoisomerase I (PDB: 1T8I) enzyme and their interactions are illustrated in . Using London dG force, docking of camptothecin, as the co-crystallized ligand, was performed to study the scoring energy (S), amino acid interactions and root mean standard deviation (RMSD), which enables to carry out precise validation of the docking protocol. Force field energy was used to elaborate the outcome. Camptothecin binds in the active site pocket with S = –20.0331 kcal/mol and RMSD = 0.7823. Camptothecin interacts with the amino acid ASP 533 by hydrogen bonds, the hydrogen bond is 3.27Å´. In addition, several pi–pi interactions between the aromatic rings of the ligand and the DNA residues of the target are displayed in . The isolated compounds 1 and 2 were docked on the enzyme/DNA complex active site and the docking scores, amino acids interactions are shown in . The compounds bound to the active site of human DNA topoisomerase I enzyme (PDB: 1T8I) with energy scores –28.64 and –30.81 kcal/mol, respectively, exceeding the co-crystallized ligand camptothecin. The results indicate that there is correlation between the number of phenolic (OH) groups and the binding score. Ligand interactions and 3D representations of compounds 1 and 2 are illustrated in .

Figure 6. Ligand interaction representation of compounds 1 and 2 of T. muelleri with human topo I–DNA complex.

Figure 6. Ligand interaction representation of compounds 1 and 2 of T. muelleri with human topo I–DNA complex.

Figure 7. Ligand interaction representation (left) of camptothecin with human topo I–DNA complex and the corresponding 3D form (right).

Figure 7. Ligand interaction representation (left) of camptothecin with human topo I–DNA complex and the corresponding 3D form (right).

Figure 8. 3D representation of compounds 1 and 2 of T. muelleri with human topo I–DNA complex.

Figure 8. 3D representation of compounds 1 and 2 of T. muelleri with human topo I–DNA complex.

Table 3. Binding scores and amino acid interactions of the docked compounds on the active site of human DNA topoisomerase I enzyme (PDB: 1T8I).

Conclusions

Terminalia muelleri is a promising medicinal plant as it exhibited potent antioxidant scavenging activity towards DPPH. Our study tends to support the therapeutic value of this plant as antioxidant drug. It demonstrated moderate cytotoxicity against MCF-7 cell line. The acylated compounds confirmed unique binding mode in the active site of human DNA topoisomerase enzyme.

Disclosure statement

The authors declare no conflict of interest.

References

  • Abdel-Azim NS, Shams KA, Shahat AAA, El Missi MM, Ismail SI, Hammouda FM. 2011. Egyptian herbal drug industry: challenges and future prospects. Res J Med Plant. 5:136–144.
  • Abdulladzhanova NG, Mavlyanov SM, Dalimov DN. 2003. Polyphenols of certain plants of the euphorbiaceae family. Chem Nat Compd. 39:399–400.
  • Agrawal PK, Bansal MC. 1989. Flavonoid glycosides. In: Agrawal PK, editor. Carbon-13 NMR of flavonoids. New York (NY): Elsevier; p. 283–364.
  • Anan K, Suganda AG, Sukandar EY, Kardono LBS. 2010. Antibacterial effect of component of Terminalia muelleri Benth. against Staphylococcus aureus. Int J Pharmacol. 6:407–412.
  • Bandele OJ, Clawson SJ, Osheroff N. 2008. Dietary polyphenols as topoisomerase II poisons: B ring and C ring substituents determine the mechanism of enzyme-mediated DNA cleavage enhancement. Chem Res Toxicol. 21:1253–1260.
  • Bandele OJ, Osheroff N. 2007. Bioflavonoids as poisons of human topoisomerase II alpha and II beta. Biochemistry. 46:6097–6108.
  • Bouchet N, Laurence B, Fauconneau B. 1998. Radical scavenging activity and antioxidant properties of tannins from Guiera senegalensis (Combretaceae). Phytother Res. 12:159–162.
  • Brand-Williams W, Cuvelier ME, Berset C. 1995. Use of a free radical method to evaluate antioxidant activity. Lebenson Wiss Technol. 28:25–30.
  • Cheng HY, Lin TC, Yu KH, Yang CM, Lin CC. 2003. Antioxidant and free radical scavenging activities of Terminalia chebula. Biol Pharm Bull. 26:1331–1335.
  • Cock E. 2015. The medicinal properties and phytochemistry of plants of genus Terminalia (Combretaceae). Inflammopharmacology. 23:203–229.
  • de Morais Lima GR, de Sales IR, Caldas Filho MR, de Jesus NZ, de Sousa Falcão H, Barbosa-Filho JM, Cabral AG, Souto AL, Tavares JF, Batista LM. 2012. Bioactivities of the genus Combretum (Combretaceae): a review. Molecules. 17:9142–9206.
  • Dougnon TV, Klotoe JR, Bankole HS, Yaya Nadjo S, Fanou B, Loko F. 2014. Antibacterial and wound healing properties of Terminalia superba Engl. and Diels (Combretaceae) in Albino Wistar Rats. J Bacteriol Parasitol. 5:206.
  • Eesha BR, Mohanbabu AV, Meena KK, Babu S, Vijay M, Lalit M, Rajput R. 2011. Hepatoprotective activity of Terminalia paniculata against paracetamol induced hepatocellular damage in Wistar albino rats. Asian Pac J Trop Med. 4:466–469.
  • El Mekkawy S, Meselhy MR, Kusumoto IT, Kadota S, Hattori M, Namba T. 1995. Inhibitory effects of Egyptian folk medicines on human immunodeficiency virus (HIV) reverse transcriptase. Chem Pharm Bull. 43:641–648.
  • Elizabeth KM. 2005. Antimicrobial activity of Terminalia bellerica. Indian J Clin Biochem. 20:150–153.
  • Eloff JN, Katerere DR, McGaw LJ. 2008. The biological activity and chemistry of the southern African combretaceae. J Ethnopharmacol. 119:686–699.
  • Fahmy NM, Al-Sayed E, Singab AN. 2015. Genus Terminalia: a phytochemical and biological review. Med Aromat Plants. 4:1–21.
  • Farnsworth NR, Bingel AS. 1977. Problems and prospects of discovering drugs from higher plants by pharmacological screening. In: Wagner H, Wolff P, editors. New natural products and plant drugs with pharmacological, biological or therapeutical activity. Berlin (Germany): Springer-Verlag; p. 1–22.
  • Fresco P, Borges F, Diniz C, Marques MPM. 2006. New insights on the anticancer properties of dietary polyphenols. Med Res Rev. 26:747–766.
  • Göktürk Baydar N, Özkan G, Yaşar S. 2007. Evaluation of the antiradical and antioxidant potential of grape extracts. Food Cont. 18:1131–1136.
  • Grace MH, Warlick CW, Neff SA, Lila MA. 2014. Efficient preparative isolation and identification of Walnut bioactive components using high-speed counter-current chromatography and LC-ESI-IT-TOF-MS. Food Chem. 158:229–238.
  • Harborne JB, Mabry TJ, Mabry H. 1975. The flavonoids. London (UK): Chapman and Hall.
  • Kumar KS, Ganesan K, Rao PV. 2008. Antioxidant potential of solvent extracts of Kappaphycus alverezii (Doty) Doty – an edible seaweed. Food Chem. 107:289–295.
  • Kumaran A, Karunakaran J. 2007. In vitro antioxidant activities of methanol extracts of five Phyllanthus species from India. Food Sci Technol. 40:344–352.
  • Latha RC, Daisy P. 2011. Insulin-secretagogue, antihyperlipidemic and other protective effects of gallic acid isolated from Terminalia bellerica Roxb. in streptozotocin-induced diabetic rats. Chem Biol Interact. 189:112–118.
  • Latté KP, Ferreira D, Venkatraman MS, Kolodziej H. 2002. O-Galloyl-C-glycosylflavones from Pelargonium reniforme. Phytochemistry. 59:419–424.
  • Lee HS, Jung SH, Yun BS, Lee KW. 2007. Isolation of chebulic acid from Terminalia chebula Retz. and its antioxidant effect in isolated rat hepatocytes. Arch Toxicol. 81:211–218.
  • Lee HS, Won NH, Kim KH, Lee H, Jun W, Lee KW. 2005. Antioxidant effects of aqueous extract of Terminalia chebula in vivo and in vitro. Biol Pharm Bull. 28:1639–1644.
  • Lin CC, Chen YL, Lin JM, Ujiie T. 1997. Evaluation of the antioxidant and hepatoprotective activity of Terminalia catappa. Am J Chin Med. 25:153–161.
  • Liu Y, Wu Y, Yuan K, Ji C, Hou A, Yoshida T, Okuda T. 1997. Astragalin 2″, 6″- di-O-gallate from Loropetalum chinense. Phytochemistry. 46:389–391.
  • Mabry TJ, Markham KR, Thomas MB. 1970. The ultraviolet spectra of flavones and flavonols. Berlin Heidelberg (Germany): Springer; p. 41–164.
  • Markham KR, Chari VM. 1982. Carbon-13 NMR spectroscopy of flavonoids. In: Harborne JB, Mabry TJ, editors. The flavonoids: advances in research. London (UK): Chapman and Hall; p. 19–132.
  • Marzouk MS, El-Toumy SA, Moharram FA, Shalaby N, Ahmed AE. 2002. Pharmacologically active ellagitannins from Terminalia myriocarpa. Planta Med. 68:523–527.
  • Nair V, Singh S, Gupta YK. 2010. Anti-arthritic and disease modifying activity of Terminalia chebula Retz. in experimental models. J Pharm Pharmacol. 62:1801–1806.
  • Natella F, Nardini M, Di Felice M, Scaccini C. 1999. Benzoic and cinnamic acid derivatives as antioxidants: structure–activity relation. J Agric Food Chem. 47:1453–1459.
  • Nonaka G, Ishimatsu M, Ageta M, Nishioka I. 1989. Tannins and related compounds LXXVI. Isolation and characterization of cercidnins A and B and cuspinin, unusual 2,3-(R)-hexahydroxydiphenoyl glucose from Cercidiphyllum japonicum and Castanopsis cuspidata var. sieboldii. Chem Pharm Bull. 37:50–53.
  • Omara EA, Nada SA, Farrag AR, Sharaf WM, El-Toumy SA. 2012. Therapeutic effect of Acacia nilotica pods extract on streptozotocin induced diabetic nephropathy in rat. Phytomedicine. 19:1059–1067.
  • Pfundstein B, El Desouky SK, Hull WE, Haubner R, Erben G, Owen RW. 2010. Polyphenolic compounds in the fruits of Egyptian medicinal plants (Terminalia bellerica, Terminalia chebula and Terminalia horrida): characterization, quantitation and determination of antioxidant capacities. Phytochemistry. 71:1132–1148.
  • Ponou BK, Teponno RB, Ricciutelli M, Quassinti L, Bramucci M, Lupidi G, Barboni L, Tapondjou LA. 2010. Dimeric antioxidant and cytotoxic triterpenoid saponins from Terminalia ivorensis A. Chev. Phytochemistry. 71:2108–2115.
  • Prakash D, Suri S, Upadhyay G, Singh BN. 2007. Total phenol, antioxidant and free radical scavenging activities of some medicinal plants. Int J Food Sci Nutr. 58:18–28.
  • Ratty AK, Sunamoto J, Das NP. 1988. Interaction of flavonoids with 1,1-diphenyl-2-picrylhydrazyl free radical, liposomal membranes and soybean lipoxygenase-1. Biochem Pharmacol. 37:989–995.
  • Rayyan S, Fossen T, Andersen ØM. 2010. Flavone C-glycosides from seeds of fenugreek, Trigonella foenum-graecum L. J Agric Food Chem. 58:7211–7217.
  • Rice-Evans CA, Miller NJ, Paganga G. 1996. Structure-antioxidant activity relationships of flavonoids and phenolic acids. Free Radic Biol Med. 20:933–956.
  • Silva O, Serramo R. 2015. Terminalia genus as source of antimicrobial agents. In: Mendez-Vilas A, editor. The battle against microbial pathogens: basic science, technological advances and educational programs. Spain: Formatex Research Center; p. 236–245.
  • Skehan P, Storeng R, Scudiero D, Monks A, McMahon J, Vistica D, Warren JT, Bokesch H, Kenney S, Boyd MR. 1990. New colorimetric cytotoxicity assay for anti-cancer drug screening. J Nat Cancer Inst. 82:1107–1112.
  • Valsaraj R, Pushpangadan P, Smitt UW, Adsersen A, Christensen SB, Sittie A, Nyman U, Nielsen C, Olsen CE. 1997. New anti-HIV-1, antimalarial, and antifungal compounds from Terminalia bellerica. J Nat Prod. 60:739–742.