Next Article in Journal
Prediction of Antimalarial Drug-Decorated Nanoparticle Delivery Systems with Random Forest Models
Next Article in Special Issue
Modification and Targeted Design of N-Terminal Truncates Derived from Brevinin with Improved Therapeutic Efficacy
Previous Article in Journal
Cognitive Stimulation Induces Differential Gene Expression in Octopus vulgaris: The Key Role of Protocadherins
Previous Article in Special Issue
Protective Mechanism of Acacia saligna Butanol Extract and Its Nano-Formulations against Ulcerative Colitis in Rats as Revealed via Biochemical and Metabolomic Assays
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Isolation of Antidiabetic Withanolides from Withania coagulans Dunal and Their In Vitro and In Silico Validation

1
Department of Chemistry, Sardar Bahadur Khan Woman University, Quetta 95000, Pakistan
2
H.E.J. Research Institute of Chemistry, International Center for Chemical and Biological Sciences, University of Karachi, Karachi 75270, Pakistan
3
Department of Chemistry, College of Science, King Saud University, Riyadh-11451, Saudi Arabia
4
Faculty of Pharmacy, The University of Lahore, Lahore 54000, Pakistan
5
Akhtar Saeed College of Pharmaceutical Sciences, Bahria Town, Lahore 54000, Pakistan
6
Natural and Medical Sciences Research Center, University of Nizwa, Nizwa 616, Oman
7
Department of Chemistry, Faculty of Sciences, University Malaya, Kuala Lumpur 50603, Malaysia
8
Department of Pharmaceutical and Pharmacological Sciences, Rega Institute for Medical Research, Medicinal Chemistry, University of Leuven, B-3000 Leuven, Belgium
*
Authors to whom correspondence should be addressed.
Biology 2020, 9(8), 197; https://doi.org/10.3390/biology9080197
Submission received: 10 March 2020 / Revised: 6 June 2020 / Accepted: 17 June 2020 / Published: 30 July 2020

Abstract

:
Withania coagulans (W. coagulans) is well-known in herbal medicinal systems for its high biological potential. Different parts of the plant are used against insomnia, liver complications, asthma, and biliousness, as well as it is reported to be sedative, emetic, diuretic, antidiabetic antimicrobial, anti-inflammatory, antitumor, hepatoprotective, antihyperglycemic, cardiovascular, immuno-suppressive and central nervous system depressant. Withanolides present in W. coagulans have attracted an immense interest in the scientific field due to their diverse therapeutic applications. The current study deals with chemical and biological evaluation of chloroform, and n-butanol fractions of W. coagulans. The activity-guided fractionation of both extracts via multiple chromatographic steps and structure elucidation of pure isolates using spectroscopies (NMR, mass spectrometry, FTIR and UV-Vis) led to the identification of a new withanolide glycoside, withacogulanoside-B (1) from n-butanol extract and five known withanolides from chloroform extract [withanolid J (2), coagulin E (3), withaperuvin C (4), 27-hydroxywithanolide I (5), and ajugin E (6)]. Among the tested compounds, compound 5 was the most potent α-glucosidase inhibitor with IC50 = 66.7 ± 3.6 µM, followed by compound 4 (IC50: 407 ± 4.5 µM) and compound 2 (IC50: 683 ± 0.94 µM), while no antiglycation activity was observed with the six isolated compounds. Molecular docking was used to predict the binding potential and binding site interactions of these compounds as α-glucosidase inhibitors. Consequently, this study provides basis to discover specific antidiabetic compounds from W. coagulans.

1. Introduction

Withania is a small genus that belongs to a diverse family of shrubs Solanaceae (comprising approximately 2000–3000 species distributed across 90 genera) [1]. Different species of Withania are widely distributed across the East Mediterranean to South Asian regions. Pakistan is home to its two species Withania coagulans and W. somnifera [2].
The genus Withania, particularly W. coagulans, is well-known for its diverse biological potential. According to the literature, different parts of the plant are used against impotence, wasting diseases, failure to thrive in children, insomnia, and nervous exhaustion. For instance, the active principle of its seeds is used in traditional medicines, and is responsible for milk coagulation [3]. Similarly, the sweet fruit of this plant has been found to be helpful against liver complications, asthma, and biliousness, as well as, it is found to be emetic, diuretic, and sedative. The flowers of W. coagulans are used in the treatment of diabetes. Apart from ethnobotanical applications, several therapeutic applications of this plant have been reported, including antihyperglycemic, anti-inflammatory, antitumor, antimicrobial, hepatoprotective, cardiovascular, immuno-suppressive, free radical scavenging, and central nervous system depressant activities [4,5,6]. Moreover, antimutagenic, antidiabetic, and leishmanicidal activities have also been described for this plant [7,8].
The major bioactive phytoconstituents isolated from W. coagulans are lactone steroids called as withanolides [9,10]. Withanolides have C-28 ergostane steroidal nucleus with γ- or δ-lactone, and a side-chain connected to the C-17 carbon atom [10]. According to the literature, withanolides have only been slightly explored in plants bearing withanolide and related ergostane-type steroids [11,12,13].
α-Glucosidase (EC3.2.1.20) is a small intestinal-membrane bound enzyme which catalyzes the hydrolysis of an oligosaccharide to absorbable monosaccharide, i.e.,, glucose, and thus its inhibition can suppress the postprandial hyperglycemia [14]. Therefore, inhibition of α-glucosidase is a useful intervention to manage type II diabetes [15]. Moreover, α-glucosidase antagonists have also been used as anti-obesity drugs, inhibitors of tumor metastasis, insect’s antifeedants, antiviral, and fungistatic compounds, and immune modulators [16].
Glycation is a chemical process in which biomolecules (such as DNA, lipids, and proteins) are modified by the non-enzymatic linkage of reducing sugars (e.g., glucose), resulting in highly reactive advanced glycation end products (AGEs). This process has been linked to some undesired health effects. e.g., diabetic complications, and aging. Thus, efforts are being made to discover new and safe antiglycation compounds to inhibit the formation of AGEs (e.g., dicarbonyl species, free radicals, protein cross-links, etc.) [17].
Molecular docking studies have discovered several novel scaffolds against specific biological targets in several studies [18,19]. We conducted docking of the newly isolated compounds on Saccharomyces cerevisiae (S. cerevisiae) α-glucosidase enzyme. Because the three-dimensional (3D) structure of S. cerevisiae α-glucosidase is not determined yet, we applied homology modeling to predict the 3D-structure of this protein before docking. Homology modeling has been applied previously to determine the structure and function of several challenging targets which cannot be explored via X-ray crystallography and NMR techniques due to their hydrophobic nature. Several α-glucosidase inhibitors have been reported based on homology model of α-glucosidase [20,21,22,23,24].
Based on antidiabetic and antiglycation potential of W. coagulans [25], current study was aimed to identify active constituents responsible for these activities. Additionally, the binding orientation of investigated compounds with α-glucosidase were investigated through molecular dynamics simulations and binding site residues responsible for inhibitor binding were also elucidated.

2. Experimental

2.1. General Experimental Method

Thin-layer chromatography (TLC) was carried out on pre-coated silica gel plates (Merck, Germany) and the spots were visualized initially under UV light at 254 nm and later sprayed with cerium (IV) sulfate reagent to develop the colors to predict different classes of natural products. The recycling high-performance liquid chromatography (HPLC) was employed for final isolation of pure compounds (JAI, LC-908W, Japan Analytical Industry Co. Ltd., Tokyo, Japan) with L-80 or ODS H-80 columns (YMC, Japan). UV spectra were recorded using a Shimadzu UV240 spectrophotometer in CH3OH for λmax nm (log ε). The FTIR spectra were recorded using KBr disc method on a JASCO A-302 spectrometer and represented in cm−1. Proton (1H, 300 MHz) and carbon (13C, 100 MHz) NMR spectral measurements were performed in CD3OD on a Bruker AV-500 equipment with reference to the residual solvent signals (CD3OD/CDCl3), and the data was presented in δ (ppm). 2D NMR measurements were performed on a Bruker AMX 500 NMR spectrometer. EI-MS were performed at 70 eV on a Finnigan MAT-112 or MAT-312 equipment and main ions are represented as m/z (%). FAB-MS were measured in the glycerol matrix by using a JEOL HX-110 mass spectrometer.

2.2. Plant Material

Plant material was collected from Hub, Sindh, Pakistan and analyzed by Dr. Sher Wali Khan, Taxonomist, Department of Botany, University of Karachi. A voucher specimen was submitted in the same section (Voucher specimen No. 46257).

2.3. Extraction and Isolation

Air-dried whole plant of W. coagulans (3.4 kg) was macerated in methanol for 24 h, filtered and the solvent evaporated on a rotary evaporator under high vacuum. The extract obtained (ca. 120 g) was suspended in water and extracted with hexane to remove nonpolar constituents. The aqueous phase was extracted with chloroform at three pH levels (i.e. 3, 7 and 10) in ascending order [26]. The residual aqueous phase was extracted with butanol and water. The chloroform fraction extracted at pH 3 (19 g) was subjected to medium pressure liquid chromatography (MPLC) using chloroform/hexane gradient elution in the order of increasing polarity with 30, 70, and 100% chloroform resulting in three fractions i.e., MC1, MC2 and MC3, respectively. The fraction MC2 (6.3 g) was then subjected to column chromatography using silica gel and chloroform/hexane solvent system to elute several fractions and the pooling of similar fractions yielded seven fractions (A-G). The fraction A (2.3 g) was exposed to column chromatography with the same solvent system, followed by various chromatographic steps, such as Sephadex LH-20, polyamide and ODS column chromatography and further purified by RP-HPLC. This led to the isolation of compounds 2–6.
The fractionation of n-butanol extract (31.5 g) was carried out on HP-20 column and the methanol/water (1:1) soluble portion was subjected to polyamide column chromatography using chloroform/methanol solvent gradient resulting in four subfractions [6]. The fraction WC2 obtained at 20% CHCl3: MeOH (3.8 gm) was subjected to silica gel column chromatography to afford five main fractions (A-E). Fraction C (0.3 g) was fractionated using silica gel column chromatography, with isocratic elution of 8% CHCl3:MeOH, and purification of RP-HPLC with L-80 column by using MeOH: H2O (1:1) at a flow rate of 3.5 mL/min yielded compound 1 (8 mg), and was characterized as a new compound.
3β-Hydroxy-14R,20R-epoxy-1-oxowitha-5,24-dienolide-27-O-β-D glucopyranoside (1)
White powder; C34H48O11, [α] 25D: -62 (c= 0.2, MeOH); UV (MeOH) λmax nm 223 (unsaturated lactone); IR (KBr) νmax cm−1: 1718 (carbonyl), 1684 (α,β-unsaturated lactone), 3490 (hydroxyl). 1H- and 13C-NMR spectroscopic data: (Table 1); HRFAB-MS m/z (+ve) m/z: [M + H]+ 633.3240 (calcd. for C34H48O11 = 633.3275).

2.4. Acidic Hydrolysis of Compound 1

Compound 1 (2 mg) was solubilized in 5% H2SO4/H2O (2 mL) and refluxed for 12 h. The reaction mixture was cooled to room temperature and extracted with ethyl acetate (1 × 3 mL) to get aglycons. The water layer was neutralized with aqueous sodium hydroxide and the presence of glucose was validated by co-TLC with authentic samples by using reported standard method [27].

2.5. In Vitro α-Glucosidase Inhibition Assay

α-Glucosidase inhibitory potential of test compounds was assayed by using 0.1 M phosphate buffer (pH 6.8) at 37 °C. The enzyme (0.2 U/mL) in phosphate buffer was incubated with selected concentrations of test compounds at 37 °C for 15 min. The substrate, p-nitrophenyl-α-d-glucopyranoside (PNP-G) (0.7 mM) was added in each well of 96-well plate and change in absorbance was monitored at 400 nm for 30 min. DMSO (7.5% final) and 1-deoxynojirimycin were used as negative and positive controls, respectively [28]. The increase in absorption due to the hydrolysis of PNP-G by α-glucosidase was recorded continuously with the spectrophotometer (Spectra Max, MD, USA.) and percent inhibition was calculated by using the following formula:
%   Inhibition   =   100 OD   of   test   sample OD   of   the   control × 100

2.6. In Vitro Antiglycation Assay

Methylglyoxal (14 mM), sodium azide (30 mM) in 0.1 M phosphate buffer, bovine serum albumin (10 mg/mL), and various concentrations of the compounds (prepared in 10% DMSO) were incubated under sterile conditions at 37 °C for 9 days. The samples were observed for the development of fluorescence (excitation, 330 nm; emission, 440 nm) against blank [28]. Standard rutin was used as a positive control and percent inhibition of AGE formation in the test sample versus control was calculated by using the following relation:
%   Inhibition   =   1 Fluorescence   of   test   sample Fluorescence   of   the   control × 100

2.7. Software Used/Statistical Analysis

The experiments were performed by using 96-well microplate reader (SpectraMax M2, Molecular Devices, CA, USA) and the results were analyzed by SoftMaxPro 4.8 and MS-Excel. The experiments were repeated in triplicate, and results were presented as means ± SEM. IC50 values were determined by using EZ-FIT, Enzyme kinetics software by Perrella Scientific, Inc., Amherst, MA, USA.

2.8. Homology Modeling and Molecular Docking

Homology modeling was conducted on the SwissModel server by using an automatic modeling mode. The sequence of S. cerevisiae α-glucosidase enzyme was downloaded from UniprotKB with the accession number P53341 for modeling. SwissModel server searched 3A47 and 3AXH as the best template structures. We used both templates for modeling S. cerevisiae α-glucosidase. Five models were obtained that were evaluated based on their stereochemical and geometrical properties by different servers, including Procheck, ERRAT and verify3D. The best model was selected for further study.

2.9. Molecular Dynamics Simulation and Binding Free Energy Calculations

Molecular docking was performed using AutoDock Vina [29] and best pose was subjected to 20 ns MD simulations. The MD simulations were carried out using AMBER simulation package 18 [30]. The same MD simulation protocol was implemented as described in previous studies [31,32,33]. The topology and coordinate files of complex was generated through tleap program of AMBER, and Antechamber package of AmberTools was utilized and parameters were taken from the GAFF force field (GAFF) [34], After stepwise minimization, the simulation system was heated and equilibrated accordingly and a final production run of 20 ns was performed at standard temperature (300K) and pressure (1 bar). The trajectory analysis was performed using CPPTRAJ program of AMBER [35].
The MM-GBSA approach is well illustrated in binding free energy calculations [36,37,38,39] which was utilized to estimate the total binding free energy ΔGtol of the complex which is the sum of the molecular mechanics binding energy (ΔEMM) and solvation free energy (ΔGsol) as given below:
ΔEMM = ΔEint + ΔEele + ΔEvdw
ΔGsol = ΔGp + ΔGnp
ΔGtol = ΔEMM + ΔGsol

3. Results and Discussion

3.1. Characterization of the Isolated Compounds

Compound 1 was isolated in the form of white amorphous powder. During structural elucidation its IR spectrum depicted absorptions at 3385, 1735, 1660, and 1593 cm−1, demonstrating the presence of hydroxyl group, carbonyl ester, and C = C, respectively. The UV-Vis spectrum showed λmax at 223 nm characteristic of conjugated enone and conjugated δ-lactone chromophores which usually exist in withanolides. The molecular formula of compound 1 was C34H48O11, determined with the help of HR-FABMS (+ve) showed [M + H]+ at 633.3240 (Calc. for C34H48O11 +H = 633.3275).
The 1H-NMR spectrum of compound 1 showed four signals for 3H singlets at δ 1.35, 1.27, 1.59, and 2.01 which were assigned to the C-18, C-19, C-21, and C-28 methyl groups, respectively. The presence of AB doublets at δ 4.30, 4.31 (J27α,27β = 11.4 Hz) implied that the C-27 position is substituted with an oxygenated group. The downfield chemical shift at δ 3.84 br, m, was assigned to C-3, its W1/2 (14.5 Hz) indicated that hydrogen is axially oriented. The characteristic double doublet at δ 4.23 (1H, dd, J = 13.2, 3.4 Hz) was assigned to the C-22 methine proton of the lactone moiety. The multiplicity of the H-22 signal indicated the absence of any proton at the vicinal C-20. A one-proton broad doublet at δ 4.58 (1H, br s) was assigned to the C-6 vinylic proton, while doublet at δ 4.36 (J = 7.6 Hz) was assigned to the anomeric proton of the sugar moiety.
13C-NMR spectra of 1 indicated that there were four methyl, ten methylene, eleven methine, and nine quaternary carbons, respectively. 1H and 13C-NMR spectrum (Table 1) indicated that there is an aglycone part of the compound similar to coagulin J [6] and a sugar moiety. The position of the sugar residue was resolved by using the HMBC correlation technique, in which C-27 (4.30, 4.38) showed correlations with anomeric carbon C-1′ (103.0) Key HMBC correlations in compound 1 is shown in Figure 1B. The stereochemistry at C-17 was inferred to be R with α-oriented ether bridge that occurred between C-14/C-20 [6]. The stereochemistry at other asymmetric centers of 1 was assigned by comparison with coagulin J [6] and NOESY spectrum (Figure 1C). Glucose was identified by hydrolysis followed by co-TLC with an authentic sample. The above analysis depicted the structure of 1 as 3β-hydroxy-14, 20-epoxy-1-oxowitha-5, 24-dienolide-27-O-β-d-glucopyranoside. The known compounds isolated from W. coagulans, were identified as withanolid J (2) [40], coagulin E (3) [6], withaperuvin C (4) [41], 27-hydroxywithanolide I (5) [42], and ajugin E (6) [43] by comparing spectral data to the reported data in the literature. The structures of the compounds 26 are presented in Figure 2.

3.2. Biological Activity

α-Glucosidase inhibitory activity of fractions from W. coagulans and its compounds were evaluated. Among two fractions, i.e., CHCl3 (crude) and BuOH (crude), only CHCl3 was found to be active with an IC50 value of 0.2 ± 0.01 mg/mL. Among the tested hits, compound 5 was the most potent compound against the α-glucosidase enzyme with an IC50 value of 66.7 ± 3.6 µM. The compound 4 and 2 depicted moderate to lower level of activities with IC50 values of 407 ± 4.5, and 683 ± 0.94 µM, respectively, when compared with 1-deoxynojirimycin (i.e., standard inhibitor; IC50 = 440.99 ± 0.01 µM). It was observed that compounds 1, 3 and 6 did not possess α-glucosidase inhibitory activity. The activities of compounds are tabulated in Table 2. Fractions from W. coagulans and pure secondary metabolites were also screened against antiglycation activity. These fractions and compounds showed less than 50% inhibition; therefore, these compounds were considered as inactive.
Based on exhaustive literature search for potential biological activities of compounds closely related to isolates 16, 17β-hydroxywithanolide K and 4β-hydroxywithanolide E were found potent antihyperglycemic and anti-inflammatory in Sprague-Dawley normoglycemic rat model [44].

3.3. Molecular Modeling and Simulations

Molecular docking was carried out to predict the structural requirements of compounds to inhibit the structure and function of α-glucosidase enzyme. For this purpose, we conducted homology modeling of α-glucosidase from S. cerevisiae. Initially, the template structure was searched on NCBI protein BLAST to model the protein of interest, and we obtained a crystal structure of isomaltase enzyme from S. cerevisiae (PDB code: 3A47) with 72% identity and 99% query coverage. The maximum score was 909 with E-value 0. Similarly, another structure of isomaltase enzyme in complex with isomaltose (PDB code: 3AXH) was retrieved from S. cerevisiae with 72% identity, 99% query coverage, 907 scores and E-value 0. Both structures were used to model S. cerevisiae α-glucosidase enzyme. The active site was predicted by superimposing the developed model on 3AXH. The stereochemical aspects of the model were inspected by procheck Ramachandran plot, verify-3D, and ERRAT plots. The Ramachandran plot statistics showed that the protein contains a total of 579 residues, among them 444 (86.7%) residues lie in the most preferred region, while 63 (12.3%) residues were situated in the additional allowed region. The number of residues present in the generously allowed region is 3 (0.6%), however, only two residues (0.4%) are present in disallowed regions. These residues are Ala278 and Thr566, both the residues are not a part of the active site, thus our model is of good quality. The homology model was further refined and optimized by small MD simulations of 20 ns to remove the clashes and to adjust overall geometry of the backbone. We obtained 93.52% quality factors from ERRAT, while 95.5% of residues presented an average 3D-1D score of 0.2 from verify3D, this suggests that the quality of the model is satisfactory and can be used in docking experiments. Homology modeling results are presented in Figure 3.
The overall structure of α-glucosidase was found similar to α-amlylases [45,46] and isomaltase [47]. The catalytic residues are conserved among S. cerevisiae Isomaltase and α-glucosidase. Glu276 and Asp214 of S. cerevisiae α-glucosidase act as proton donor and nucleophile, respectively. While Asp349 act as a transition state stabilizer for substrate molecule. The catalytic residues and the surrounding residues involved in the formation of the active site are tabulated in Table 3.
The generated model was used in docking studies. Molecular docking was carried out by AutoDock Vina. In this study, we only investigated the predictive binding potential of the most active compound 5 which indicated reasonable docking binding energy of −6.68 kcal/mol. As water molecules have prime importance in catalytic activity of this enzyme, therefore MD simulation was performed in explicit solvent environment. Later, the MMGBSA binding free energies were estimated in the presence of solvent effect. The results of MD simulations are interactively displayed in Figure 4. During MD simulation, compound 5 remained stable inside the binding pocket in close proximity of catalytic triad residues as depicted from a final snapshot obtained after 20 ns time scale (Figure 4A,B). Moreover, the stability of 5 was evident from the energy decomposition analysis of residues involved in the formation of the active site followed by H-bond occupancy over a period of 20 ns (Figure 4C,D). The MD simulated binding mode of 5 established four well-populated H-bonds with the sidechains of His239 (average, 1.7 Å; 75.5%) and Glu276 (average, 2.5 Å; 17.6%) and backbone atoms of Phe157 (average 2.08 Å; 62.3%), Arg312 (average 2.53 Å; 45.6%). Other than these, Asn237 and Asn241 also showed favorable H-bond occupancy of 40.9 and 23.8%. These molecular interactions were in-line with the reported experimental studies of f α-glucosidases [20,21,23,48,49,50].
To analyze the residual contribution in overall binding energy of complex, per-residue decomposition analysis was performed which indicated His239 and Arg312 as highly contributing residues in overall binding free energy with a ΔGtotal of −3.12 and −2.89 kcal/mol that mainly interacted electrostatically together with Asn241 (ΔGtotal = −1.39 kcal/mol) and Phe300 (ΔGtotal = −1.42 kcal/mol) which mainly contributed through vdW interactions. All these residues were found highly contributing to stabilize the complex formation during MD simulation. While, the catalytic triad weakly contributed as indicated from ΔGtotal values of Asp214 (−0.15 kcal/mol), Glu276 (−0.89 kcal/mol) and Asp349 (−0.45 kcal/mol). This was due to the contribution of unfavorable solvation energy by blocking out a sufficient amount of solvent that might stabilize these negatively charged catalytic triad residues. The root-mean-square-deviation (RMSD) and root-mean-square-fluctuations (RMSF) was also examined over a period of 20 ns which revealed consistent backbone stability including the active site residues (Figure 4E) which contributed well in stabilizing the complex (Figure 4F,G).

4. Conclusions

In the present study, a phytochemical investigation was carried out on CHCl3 and n-butanol fraction of W. coagulans plant. All the compounds 2–6 isolated from chloroform fraction were known, while a new compound (i.e., compound 1) was isolated from n-butanol fraction. It was observed that n-butanol fraction, and its isolated new compound were inactive in both antiglycation, and α-glucosidase inhibition assay, while CHCl3 fraction and its isolated pure compounds showed activity against α-glucosidase and were inactive in antiglycation assay. The activity was correlated with molecular docking studies that depict that the inactive compounds do not bind appropriately with the α-glucosidase enzyme. This study provides scientific basis to discover specific anti-diabetic compounds from W. coagulans.

Author Contributions

Conceptualization, M.I.C. and S.A.; methodology, S.M. and S.R.; software, S.A.H., M.U.M. and M.F.; validation, F.S., S.A.H. and I.A.; formal analysis, S.M., and I.W.; investigation, M.A.; resources, M.I.C. and M.F.; data curation, S.A. and S.R.; writing—original draft preparation, S.M. and F.S.; writing—review and editing, S.A., I.B.A. and M.I.C.; visualization, F.S.; supervision, M.I.C. and M.F.; project administration, M.I.C. and S.A.; funding acquisition, M.I.C., S.R. and M.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Malaya University, Malaysia, grant number IIRG003A-2019 and in-house resources of H.E.J. Research Institute, Pakistan, and the APC was funded by the both institutions.

Acknowledgments

The Authors are grateful to H.E.J. research institute of chemistry, International Center for Chemical and Biological Sciences and Malaya University for providing research facilities (grant no.: IIRG003A-2019). One of the Authors (M. Iqbal Choudhary) acknowledge the financial support of the Searle Company (Pakistan) for sponsoring the studies through research project entitled, “Discovery of new therapeutic application of existing drugs by repositioning and repurposing approach”.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chadha, Y.R. The wealth of India, New Delhi: Publications and informations directorate. CSIR 1976, X, 582. [Google Scholar]
  2. Yousaf, Z.; Masood, S.; Shinwari, Z.K.; Khan, M.A.; Rabani, A. Evaluation of taxonomic status of medicinal species of the genus Hyoscyamous, Withania, Atropa and Datura based on polyacrylamide gel electrophoresis. Pak. J. Bot. 2008, 40, 2289–2297. [Google Scholar]
  3. Watt, G.A. Dictionary of the Economic Products of India; Cosmo Publications: Delhi, India, 1972; Volume 6, p. 309. [Google Scholar]
  4. Mathur, D.; Agrawal, R. Withania coagulans: A review on the morphological and pharmacological properties of the shrub. Sci. World J. 2011, 1, 30–37. [Google Scholar]
  5. Rahman, A.U.; Abbas, S.; Shahwar, D.E.; Jamal, S.; Choudhary, M. New withanolides from Withania spp. J. Nat. Prod. 1993, 56, 1000–1006. [Google Scholar] [CrossRef]
  6. Atta-ur-Rahman; Choudhary, M.I.; Qureshi, S.; Gul, W.; Yousaf, M. Two new ergostane-type steroidal lactones from Withania coagulans. J. Nat. Prod. 1998, 61, 812–814. [Google Scholar] [CrossRef]
  7. Kuroyanagi, M.; Murata, M.; Nakane, T.; Shirota, O.; Sekita, S.; Fuchino, H.; Shinwari, Z.K. Leishmanicidal active withanolides from a Pakistani medicinal plant, Withania coagulans. Chem. Pharm. Bull. 2012, 60, 892–897. [Google Scholar] [CrossRef] [Green Version]
  8. Verma, P.K.; Rajurkar, S.; Gaikwad, N.; Kamboj, V. The isolation and structure elucidation of new Withanaoloides from Withania cogulance with antidiabetic activity. Acta Pharm. Sci. 2010, 52, 155–157. [Google Scholar]
  9. Ali, A.; Maher, S.; Khan, S.A.; Chaudhary, M.I.; Musharraf, S.G. Sensitive quantification of six steroidal lactones in Withania coagulans extract by UHPLC electrospray tandem mass spectrometry. Steroids 2015, 104, 176–181. [Google Scholar] [CrossRef]
  10. Khan, S.; Khan, S.; Shah, Z.; Asiri, A. Withanolides: Biologically active constituents in the treatment of Alzheimer disease. Med. Chem. 2016, 12, 238–256. [Google Scholar] [CrossRef]
  11. Adam, G.; Chien, N.; Khoi, N. Dunawithanine A and B, first plant withanolide glycosides from Dunalia australis. Naturwissenschaften 1981, 68, 425–426. [Google Scholar] [CrossRef]
  12. Chen, L.-X.; He, H.; Qiu, F. Natural withanolides: An overview. Nat. Prod. Rep. 2011, 28, 705–740. [Google Scholar] [CrossRef] [PubMed]
  13. Glotter, E. Withanolides and related ergostane-type steroids. Nat. Prod. Rep. 1991, 8, 415–440. [Google Scholar] [CrossRef] [PubMed]
  14. Kimura, A.; Takata, M.; Fukushi, Y.; Mori, H.; Matsui, H.; Chiba, S. A catalytic amino acid and primary structure of active site in Aspevgillus niger α-glucosidase. Biosci. Biotechnol. Biochem. 1997, 61, 1091–1098. [Google Scholar] [CrossRef]
  15. Shim, Y.-J.; Doo, H.-K.; Ahn, S.-Y.; Kim, Y.-S.; Seong, J.-K.; Park, I.-S.; Min, B.-H. Inhibitory effect of aqueous extract from the gall of Rhus chinensis on alpha-glucosidase activity and postprandial blood glucose. J. Ethnopharmacol. 2003, 85, 283–287. [Google Scholar] [CrossRef]
  16. Braun, C.; Brayer, G.D.; Withers, S.G. Mechanism-based inhibition of yeast α-glucosidase and human pancreatic α-amylase by a new class of inhibitors 2-deoxy-2,2-difluoro-α-glycosides. J. Biol. Chem. 1995, 270, 26778–26781. [Google Scholar] [CrossRef] [Green Version]
  17. Reddy, V.P.; Beyaz, A. Inhibitors of the Maillard reaction and AGE breakers as therapeutics for multiple diseases. Drug Discov. Today 2006, 11, 646–654. [Google Scholar] [CrossRef]
  18. Talele, T.T.; Khedkar, S.A.; Rigby, A.C. Successful applications of computer aided drug discovery: Moving drugs from concept to the clinic. Curr. Top. Med. Chem. 2010, 10, 127–141. [Google Scholar] [CrossRef]
  19. Kubinyi, H. Success stories of computer-aided design. Comput. Appl. Pharm. Res. Dev. 2006, 2, 377. [Google Scholar]
  20. Ma, H.; Wang, L.; Niesen, D.B.; Cai, A.; Cho, B.P.; Tan, W.; Gu, Q.; Xu, J.; Seeram, N.P. Structure activity related, mechanistic, and modeling studies of gallotannins containing a glucitol-core and α-glucosidase. RSC Adv. 2015, 5, 107904–107915. [Google Scholar] [CrossRef]
  21. Bharatham, K.; Bharatham, N.; Park, K.H.; Lee, K.W. Binding mode analyses and pharmacophore model development for sulfonamide chalcone derivatives, a new class of α-glucosidase inhibitors. J. Mol. Graph. Model. 2008, 26, 1202–1212. [Google Scholar] [CrossRef]
  22. Park, H.; Hwang, K.Y.; Kim, Y.H.; Oh, K.H.; Lee, J.Y.; Kim, K. Discovery and biological evaluation of novel α-glucosidase inhibitors with in vivo antidiabetic effect. Bioorg. Med. Chem. Lett. 2008, 18, 3711–3715. [Google Scholar] [CrossRef] [PubMed]
  23. Park, H.; Hwang, K.Y.; Oh, K.H.; Kim, Y.H.; Lee, J.Y.; Kim, K. Discovery of novel α-glucosidase inhibitors based on the virtual screening with the homology-modeled protein structure. Bioorg. Med. Chem. 2008, 16, 284–292. [Google Scholar] [CrossRef] [PubMed]
  24. Park, J.-H.; Ko, S.-M.; Park, H. Toward the virtual screening of α-glucosidase inhibitors with the homology-modeled protein structure. Bull. Korean Chem. Soc. 2008, 29, 921–927. [Google Scholar]
  25. Hemalatha, S.; Wahi, A.K.; Singh, P.N.; Chansouria, J.P.N. Hypoglycemic activity of Withania coagulans Dunal in streptozotocin induced diabetic rats. J. Ethnopharmacol. 2004, 93, 261–264. [Google Scholar] [CrossRef]
  26. Rahman, A.-U.; Choudhary, M.I.; Yousaf, M.; Gul, W.; Qureshi, S. New withanolides from Withania coagulans. Chem. Pharm. Bull. 1998, 46, 1853–1856. [Google Scholar] [CrossRef] [Green Version]
  27. Bedir, E.; Khan, I.A. New steroidal glycosides from the fruits of Tribulus terrestris. J. Nat. Prod. 2000, 63, 1699–1701. [Google Scholar] [CrossRef]
  28. Choudhary, M.I.; Adhikari, A.; Rasheed, S.; Marasini, B.P.; Hussain, N.; Kaleem, W.A. Cyclopeptide alkaloids of Ziziphus oxyphylla Edgw as novel inhibitors of α-glucosidase enzyme and protein glycation. Phytochem. Lett. 2011, 4, 404–406. [Google Scholar] [CrossRef]
  29. Trott, O.; Olson, A.J. AutoDock Vina: Improving the speed and accuracy of docking with a new scoring function, efficient optimization, and multithreading. J. Comput. Chem. 2010, 31, 455–461. [Google Scholar] [CrossRef] [Green Version]
  30. Case, D.A.; Ben-Shalom, I.Y.; Brozell, S.R.; Cerutti, D.S.; Cheatham, T.E., III; Cruzeiro, V.W.D.; Darden, T.A.; Duke, R.E.; Ghoreishi, D.; Gilson, M.K. AMBER 2018; University of California: San Francisco, CA, USA, 2018. [Google Scholar]
  31. Ikram, N.; Mirza, M.U.; Vanmeert, M.; Froeyen, M.; Salo-Ahen, O.M.H.; Tahir, M.; Qazi, A.; Ahmad, S. Inhibition of Oncogenic Kinases: An In Vitro Validated Computational Approach Identified Potential Multi-Target Anticancer Compounds. Biomolecules 2019, 9, 124. [Google Scholar] [CrossRef] [Green Version]
  32. Jabbar, B.; Rafique, S.; Salo-Ahen, O.M.H.; Ali, A.; Munir, M.; Idrees, M.; Mirza, M.U.; Vanmeert, M.; Shah, S.Z.; Jabbar, I. Antigenic Peptide Prediction From E6 and E7 Oncoproteins of HPV Types 16 and 18 for Therapeutic Vaccine Design Using Immunoinformatics and MD Simulation Analysis. Front. Immunol. 2018, 9, 3000. [Google Scholar] [CrossRef] [Green Version]
  33. Mirza, M.U.; Rafique, S.; Ali, A.; Munir, M.; Ikram, N.; Manan, A.; Salo-Ahen, O.M.H.; Idrees, M. Towards peptide vaccines against Zika virus: Immunoinformatics combined with molecular dynamics simulations to predict antigenic epitopes of Zika viral proteins. Sci. Rep. 2016, 6, 37313. [Google Scholar] [CrossRef] [PubMed]
  34. Wang, J.; Wolf, R.M.; Caldwell, J.W.; Kollman, P.A.; Case, D.A. Development and testing of a general amber force field. J. Comput. Chem. 2004, 25, 1157–1174. [Google Scholar] [CrossRef]
  35. Roe, D.R.; Cheatham, T.E., III. PTRAJ and CPPTRAJ: Software for processing and analysis of molecular dynamics trajectory data. J. Chem. Theory Comput. 2013, 9, 3084–3095. [Google Scholar] [CrossRef]
  36. Hou, T.; Wang, J.; Li, Y.; Wang, W. Assessing the performance of the MM/PBSA and MM/GBSA methods. 1. The accuracy of binding free energy calculations based on molecular dynamics simulations. J. Chem. Inf. Model. 2011, 51, 69–82. [Google Scholar] [CrossRef] [PubMed]
  37. Iman, K.; Mirza, M.U.; Mazhar, N.; Vanmeert, M.; Irshad, I.; Kamal, M.A. In silico Structure-based Identification of Novel Acetylcholinesterase Inhibitors Against Alzheimer’s Disease. CNS Neurol. Disord.-Drug Targets (Former. Curr. Drug Targets-CNS Neurol. Disord.) 2018, 17, 54–68. [Google Scholar] [CrossRef] [PubMed]
  38. Chen, F.; Liu, H.; Sun, H.; Pan, P.; Li, Y.; Li, D.; Hou, T. Assessing the performance of the MM/PBSA and MM/GBSA methods. 6. Capability to predict protein–protein binding free energies and re-rank binding poses generated by protein–protein docking. Phys. Chem. Chem. Phys. 2016, 18, 22129–22139. [Google Scholar] [CrossRef]
  39. Sirin, S.; Kumar, R.; Martinez, C.; Karmilowicz, M.J.; Ghosh, P.; Abramov, Y.A.; Martin, V.; Sherman, W. A computational approach to enzyme design: Predicting ω-aminotransferase catalytic activity using docking and MM-GBSA scoring. J. Chem. Inf. Model. 2014, 54, 2334–2346. [Google Scholar] [CrossRef]
  40. Ma, L.; Xie, C.M.; Li, J.; Lou, F.C.; Hu, L.H. Daturametelins H, I, and J: Three new withanolide glycosides from Datura metel L. Chem. Biodivers. 2006, 3, 180–186. [Google Scholar] [CrossRef]
  41. Bagchi, A.; Neogi, P.; Sahai, M.; Ray, A.B.; Oshima, Y.; Hikino, H. Withaperuvin E and nicandrin B, withanolides from Physalis peruviana and Nicandra physaloides. Phytochemistry 1984, 23, 853–855. [Google Scholar] [CrossRef]
  42. Velde, V.V.; Lavie, D. New withanolides of biogenetic interest from Withania somnifera. Phytochemistry 1981, 20, 1359–1364. [Google Scholar] [CrossRef]
  43. Khan, P.M.; Malik, A.; Ahmad, S.; Nawaz, H.R. Withanolides from Ajuga parviflora. J. Nat. Prod. 1999, 62, 1290–1292. [Google Scholar] [CrossRef] [PubMed]
  44. Takimoto, T.; Kanbayashi, Y.; Toyoda, T.; Adachi, Y.; Furuta, C.; Suzuki, K.; Miwa, T.; Bannai, M. 4β-Hydroxywithanolide E isolated from Physalis pruinosa calyx decreases inflammatory responses by inhibiting the NF-κB signaling in diabetic mouse adipose tissue. Int. J. Obes. 2014, 38, 1432–1439. [Google Scholar] [CrossRef]
  45. Brayer, G.D.; Luo, Y.; Withers, S.G. The structure of human pancreatic α-amylase at 1.8 Å resolution and comparisons with related enzymes. Protein Sci. 1995, 4, 1730–1742. [Google Scholar] [CrossRef] [PubMed]
  46. Machius, M.; Wiegand, G.; Huber, R.J.J.o.m.b. Crystal Structure of Calcium-depletedBacillus licheniformisα-amylase at 2.2 Å Resolution. J. Mol. Biol. 1995, 246, 545–559. [Google Scholar] [CrossRef]
  47. Yamamoto, K.; Miyake, H.; Kusunoki, M.; Osaki, S. Crystal structures of isomaltase from Saccharomyces cerevisiae and in complex with its competitive inhibitor maltose. FEBS J. 2010, 277, 4205–4214. [Google Scholar] [CrossRef] [PubMed]
  48. Seo, W.D.; Kim, J.H.; Kang, J.E.; Ryu, H.W.; Curtis-Long, M.J.; Lee, H.S.; Yang, M.S.; Park, K.H. Sulfonamide chalcone as a new class of α-glucosidase inhibitors. Bioorg. Med. Chem. Lett. 2005, 15, 5514–5516. [Google Scholar] [CrossRef] [PubMed]
  49. Yamamoto, K.; Nakayama, A.; Yamamoto, Y.; Tabata, S. Val216 decides the substrate specificity of α-glucosidase in Saccharomyces cerevisiae. Eur. J. Biochem. 2004, 271, 3414–3420. [Google Scholar] [CrossRef] [PubMed]
  50. Saleem, F.; Mehmood, R.; Mehar, S.; Khan, M.T.J.; Khan, Z.-U.-D.; Ashraf, M.; Ali, M.S.; Abdullah, I.; Froeyen, M.; Mirza, M.U.; et al. Bioassay Directed Isolation, Biological Evaluation and in Silico Studies of New Isolates from Pteris cretica L. Antioxidants 2019, 8, 231. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (A) Structure of compound 1, (B) key COSY and HMBC interactions in compound 1, (C) key NOESY interactions in compound 1.
Figure 1. (A) Structure of compound 1, (B) key COSY and HMBC interactions in compound 1, (C) key NOESY interactions in compound 1.
Biology 09 00197 g001
Figure 2. Structures of compounds 2 to 6 isolated from w. coagulans.
Figure 2. Structures of compounds 2 to 6 isolated from w. coagulans.
Biology 09 00197 g002
Figure 3. The superimposed view of model (purple ribbons) and template (3AXH in green ribbons) shows the structural topology of the model is similar to its template and residues of catalytic triad is represented in stick representation.
Figure 3. The superimposed view of model (purple ribbons) and template (3AXH in green ribbons) shows the structural topology of the model is similar to its template and residues of catalytic triad is represented in stick representation.
Biology 09 00197 g003
Figure 4. Molecular surface representation of α-glucosidase. (A) MD simulated binding pose of 5 (orange sticks) inside the binding pocket. Residues are highlighted in stick representation and catalytic triad residues are colored yellow. (B) H-Bonds with distance are displayed beside each bond. (C) Per-residue decomposition analysis of binding pocket residues, while catalytic triad residues are colored yellow. (D) H-Bond occupancy and (E) root-mean-square-deviation (RMSD) of binding pocket residues over a period of 20 ns. (F) RMSD of backbone atoms of α-glucosidase (in black) along with ligand (in red) and root mean square fluctuations and (RMSF) (G) are displayed over a period of 20 ns.
Figure 4. Molecular surface representation of α-glucosidase. (A) MD simulated binding pose of 5 (orange sticks) inside the binding pocket. Residues are highlighted in stick representation and catalytic triad residues are colored yellow. (B) H-Bonds with distance are displayed beside each bond. (C) Per-residue decomposition analysis of binding pocket residues, while catalytic triad residues are colored yellow. (D) H-Bond occupancy and (E) root-mean-square-deviation (RMSD) of binding pocket residues over a period of 20 ns. (F) RMSD of backbone atoms of α-glucosidase (in black) along with ligand (in red) and root mean square fluctuations and (RMSF) (G) are displayed over a period of 20 ns.
Biology 09 00197 g004
Table 1. 1H- and 13C—NMR chemical shift values of compound 1 (ppm, CD3OD, 600 and 150 MHz respectively).
Table 1. 1H- and 13C—NMR chemical shift values of compound 1 (ppm, CD3OD, 600 and 150 MHz respectively).
C. No.δCδH (J, Hz)C. No.δCδH (J, Hz)
1220.4-1820.71.35 s
240.52.31, 2.121918.21.27 s
375.03.84 br m, (W1/2 = 14.5)2078.1-
441.02.04, 2.01 m2121.11.59
5136.3-2283.14.23 dd (13.2, 3.4)
6123.64.58 br s2332.91.60, 1.84 overlap
732.61.77, 1.75 m24160.1-
822.11.63 m25126.4-
921.01.61 m26168.0-
1052.7-2762.74.30 d (11.4), 4.31 d (11.7)
1121.11.27, 1.35 m2818.02.01 s
1220.81.61, 2.01 m1′104.04.36 d (7.6)
1349.0-2′75.03.31 t (8.1)
1485.5-3′74.03.65 overlap
1522.31.63, 1.75 m4′71.63.33 overlap
1632.81.35, 1.09 m5′78.03.33 overlap
1750.52.30 t (9.4)6′63.53.64 dd (11.8, 4.8), 3.84 d (11.8)
Table 2. α-Glucosidase inhibition and antiglycation activities of compounds 16.
Table 2. α-Glucosidase inhibition and antiglycation activities of compounds 16.
Compoundα-Glucosidase InhibitionAntiglycation
IC50 (µM) ± (SEM)IC50 (µM) ± (SEM)
1InactiveInactive
2683 ± 0.94Inactive
3InactiveInactive
4407 ± 4.5Inactive
566.7 ± 3.6Inactive
6InactiveInactive
Standard440.99 ± 0.01 a288.9 ± 1.84 b
a Deoxynojirimycin, b Rutin.
Table 3. Comparison of active site residues of template and model structures.
Table 3. Comparison of active site residues of template and model structures.
Template
(S. cerevisiae Isomaltase)
Model
(S. cerevisiae α-Glucosidase)
Catalytic residues
Asp215Asp214
Glu277Glu276
Asp352Asp349
Extended active site residues
Asp69Asp68
Tyr72Tyr71
Val109Val108
His112His111
Tyr158Phe157
Phe159Phe158
Phe178Phe177
Gln182Gln181
Arg213Arg212
Val216Thr215
Gln279Ala278
Phe303Phe300
Arg315Arg312
His351His348
Gln353Gln350
Glu411Asp408
Arg442Arg439
Arg446Arg443
Interacting water molecules
1021, 1026, 1056, 1058, 1061, 1087, 1102, 1122, 1174, 12281021, 1026, 1056, 1058, 1061, 1087, 1102, 1122, 1174, 1228

Share and Cite

MDPI and ACS Style

Maher, S.; Choudhary, M.I.; Saleem, F.; Rasheed, S.; Waheed, I.; Halim, S.A.; Azeem, M.; Abdullah, I.B.; Froeyen, M.; Mirza, M.U.; et al. Isolation of Antidiabetic Withanolides from Withania coagulans Dunal and Their In Vitro and In Silico Validation. Biology 2020, 9, 197. https://doi.org/10.3390/biology9080197

AMA Style

Maher S, Choudhary MI, Saleem F, Rasheed S, Waheed I, Halim SA, Azeem M, Abdullah IB, Froeyen M, Mirza MU, et al. Isolation of Antidiabetic Withanolides from Withania coagulans Dunal and Their In Vitro and In Silico Validation. Biology. 2020; 9(8):197. https://doi.org/10.3390/biology9080197

Chicago/Turabian Style

Maher, Saima, M. Iqbal Choudhary, Farooq Saleem, Saima Rasheed, Imran Waheed, Sobia Ahsan Halim, Muhammad Azeem, Iskandar Bin Abdullah, Matheus Froeyen, Muhammad Usman Mirza, and et al. 2020. "Isolation of Antidiabetic Withanolides from Withania coagulans Dunal and Their In Vitro and In Silico Validation" Biology 9, no. 8: 197. https://doi.org/10.3390/biology9080197

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