Open Access Article
Hayat Bibia,
Saima Maher
*a,
Noureen Khanb,
Shamim Khanc,
Tahir Ali Chohand,
Hammad Saleemb,
Magda H. Abdellattif
e,
Ajmal Khan*fg and
Ahmed Al-Harrasi
*f
aDepartment of Chemistry, Sardar Bahadur Khan Women University, Quetta, Pakistan. E-mail: saimamaher@yahoo.com
bDepartment of Chemistry, Rawalpindi Women University, 6th Road, Satellite Town, Rawalpindi 46300, Pakistan
cDepartment of Chemistry, Sardar Bahadur Khan Women University, Quetta, Pakistan
dInstitute of Pharmaceutical Sciences (IPS), University of Veterinary & Animal Sciences (UVAS), Lahore, Pakistan
eChemistry Department, College of Sciences, University College of Taraba, Taif University, P.O. Box 11099, Taif 21944, Saudi Arabia
fNatural and Medical Sciences Research Center, University of Nizwa, PO Box 33, 616 Birkat Al Mauz, Nizwa, Oman. E-mail: ajmalkhan@unizwa.edu.om; aharrasi@unizwa.edu.om
gDepartment of Chemical and Biological Engineering, College of Engineering, Korea University, Seoul 02841, Republic of Korea
First published on 9th December 2025
Withania coagulans Dunal is a medicinal plant with potential therapeutic applications in neurodegenerative and metabolic disorders. This study aimed to isolate, structurally characterize, and evaluate the biological potential of withanolide-type compounds from Withania coagulans Dunal, with a focus on their antioxidant and enzyme inhibitory activities relevant to neurodegenerative and metabolic disorders. Two new withanolides (WTH1 and WTH2) and one known compound (WTH3) were isolated from W. coagulans. The methanol extract and its fractions were assessed for total phenolic and flavonoid contents, antioxidant capacity (DPPH, FRAP, phosphomolybdenum assays), and enzyme inhibitory activities against acetylcholinesterase (AChE), butyrylcholinesterase (BChE), lipoxygenase, α-glucosidase, and tyrosinase. The methanol extract exhibited the highest total phenolic (71.53 mg GAE per g) and flavonoid (64.32 mg QE per g) contents, correlating with strong antioxidant activity (DPPH: 91.2%, FRAP: 678.3 µmol Fe2+ per g, phosphomolybdenum: 4.2 mmol TE per g). It showed significant inhibitory effects against AChE (4.10 mg GALAE per g), BChE (3.71 mg GALAE per g), lipoxygenase, α-glucosidase, and tyrosinase, with the ethyl acetate fraction displaying the strongest α-glucosidase inhibition (3.51 mmol ACAE per g). In silico docking revealed strong binding affinities of WTH1 and WTH3 toward AChE (−11.616 and −11.438 kcal mol−1, respectively), while WTH3 also interacted effectively with BChE (−9.30 kcal mol−1), surpassing the standard drug physostigmine (−5.85 kcal mol−1). Pharmacokinetic evaluation of WTH1 predicted high gastrointestinal absorption (97.65%), moderate oral bioavailability (0.55), and absence of hepatotoxicity or AMES toxicity. DFT analysis indicated a stable HOMO–LUMO energy gap (9.923 eV), and binding free energy calculations confirmed strong interaction of WTH1 with AChE using PB (−29.731 kcal mol−1) and GB (−43.54 kcal mol−1) methods, outperforming the reference drug (−15.08 kcal mol−1). The findings demonstrate that W. coagulans methanol extract, particularly the isolated new withanolide WTH1, exhibits potent antioxidant and enzyme inhibitory activities with promising pharmacokinetic properties. These results support further pharmacological and clinical evaluation of W. coagulans as a natural source of therapeutic agents against neurodegenerative and metabolic disorders.
Different parts of the plant have also been employed in folk medicine for the treatment of rheumatism, dropsy, ulcers, and age-related debility.13 The wild fruits of W. coagulans exhibit various nutraceutical and biological activities, making them potentially valuable for promoting health and wellness.14–16 Much of the plant's medicinal value is attributed to its characteristic withanolides a class of steroidal lactones chemically defined by the presence of a γ- or δ-lactone side chain at the C-17 position.17,18 Withanolides have also been identified in other families such as Taccaceae, Leguminosae, and Solanaceae, as well as from marine organisms.19
Previous reports have highlighted the broad pharmacological potential of W. coagulans, including antioxidant, antidiabetic, anti-inflammatory, and neuroprotective effects. However, despite these promising activities, systematic isolation and structural characterization of its active constituents, and their mechanistic evaluation against key biological targets, remain limited. A deeper understanding of its bioactive compounds and their molecular interactions is essential to substantiate its traditional uses and explore its therapeutic potential.
In addition, W. coagulans has shown promising antidiabetic and antiglycation properties, with several bioactive compounds identified as contributors to its antidiabetic activity.23,24 Further studies have demonstrated its antimutagenic, leishmanicidal, and additional antidiabetic effects.21,25,26 Notably, two unique withanolides chantriolides D and E were isolated from Tacca chantrieri and shown to exert selective cytotoxic effects against cancer cell lines.27
Therefore, the present study was designed to isolate and structurally characterize bioactive withanolide compounds from W. coagulans and to evaluate their antioxidant and enzyme inhibitory properties using in vitro and in silico approaches. The investigation particularly focused on enzymes linked to major chronic conditions, including acetylcholinesterase and butyrylcholinesterase (neurodegeneration), α-glucosidase (diabetes), lipoxygenase (inflammation), and tyrosinase (skin disorders). By integrating phytochemical analysis, biological assays, molecular docking, and computational modeling, this study aims to elucidate the pharmacological relevance of W. coagulans and identify potential lead compounds for future drug discovery efforts. The present study aimed to evaluate the enzyme inhibitory activities of whole-plant extracts prepared using solvents of varying polarity. These extracts were assessed for their inhibitory effects against key enzymes linked to major health disorders: acetylcholinesterase and butyrylcholinesterase (neurodegenerative diseases), α-glucosidase (diabetes), lipoxygenase (inflammation), and tyrosinase (hyperpigmentation). Furthermore, the antioxidant potential of the extracts was determined using three standard assays: phosphomolybdenum, DPPH radical scavenging, and ferric reducing antioxidant power (FRAP). In addition, total bioactive contents were quantified via spectrophotometric assays, and major constituents were isolated using chromatographic techniques. Future investigations should focus on the isolation and structural characterization of bioactive compounds from Withania coagulans, particularly withanolides, to better understand their pharmacological mechanisms. In vivo studies are essential to validate its enzyme inhibitory and antioxidant potential. Toxicological assessments will help establish safety for therapeutic use. Formulation of standardized extracts could lead to the development of herbal drugs. Omics-based approaches may reveal biosynthetic pathways for key metabolites. Overall, this plant holds strong potential for future drug discovery and nutraceutical applications.
Despite the extensive traditional uses of W. coagulans, there remains a lack of detailed studies linking its phytochemical constituents to specific biological activities. Therefore, the present study was designed to isolate and structurally characterize bioactive compounds from W. coagulans and to evaluate their antioxidant and enzyme inhibitory properties, targeting key enzymes associated with neurodegenerative, metabolic, and inflammatory disorders. This integrated approach combining phytochemical characterization, bioassays, and computational modeling aims to elucidate the molecular mechanisms underlying the plant's pharmacological potential and to identify promising natural lead compounds for therapeutic development.
:
9, v/v. Subsequently, 0.75 mL of 1% Na2CO3 solution was added to a mixture after allowing the reaction mixture for 3 minutes, it was incubated for two hours and then the sample absorbance was observed at 760 nm. For the estimation of total flavonoid content, 1 mL of the sample solution (1 mg mL−1) was combined with 1 mL of 2% aluminium chloride solution prepared in methanol, the mixture was incubated at room temperature for 10 minutes, and then the absorbance was recorded at 415 nm The results of total phenolic constituents were reported as equivalents of gallic acid (mg GAE per g extract), however the results of total flavonoid constituents were recorded as equivalents of quercetin (mg QE per g extract).1,2
:
MeOH (1
:
1, 8.5 g), was rechromatographed on sephadex LH-20 and elution was carried out with various mixtures of H2O–MeOH in decreasing order of polarity. The sub-fractions obtained with H2O–MeOH (6
:
4) (1.2 g), was re-chromatographed over polyamide resin eluting with CH2Cl2
:
MeOH to afford the semi-pure fractions, which were further purified by re-chromatographed over silica gel and eluted with CHCl3
:
MeOH (8.5
:
1.5) and CHCl3
:
MeOH (8.2
:
1.8) to afford known withnolideglycoside (WTH3) (Fig. 7).
ε). IR spectra was taken on JASCO. 1H and 13C-NMR spectra were noted on Bruker AV-500 spectrometer operating at 300, 500 MHz, respectively, and the data is given in δ (ppm). Bruker AMX 500 NMR spectrometer was used for 1 and 2D nuclear magnetic resonance (NMR) spectroscopy. Molecular form was confirmed through EI-MS at 70 eV on a Finnigan MAT-112 or MAT-312 presented as m/z (%). Glycerol matrix was used for FAB-MS on a JEOL HX-110 mass spectrometer. Recycling HPLC with columns ODS H-80 or L-80 (JAI, LC-908W, Japan Analytical Industry Co. Ltd) was used for final purification of compounds. The isolated compounds (WTH1–WTH3) were characterized by UV, IR, 1H and 13C NMR, 2D NMR (COSY, HSQC, HMBC, NOESY), and mass spectrometry (EI-MS, FAB-MS, HR-EIMS). Spectral data were compared with literature values to confirm structural assignments. Complete NMR and MS data are provided in the SI (Fig. S1–S8).
:
1
:
1 (v/v/v) and the absorbance was recorded at 593 nm 30 min. Milligrams of Trolox equivalents per gram of dry extract (TE per g extract) were the measurement unit.
| Inhibition (%) = [(A_control − A_sample)/A_control] × 100 |
The reaction mixture composed by 25 µL of the sample solution and 50 µL of the α-amylase solution (10 U mL−1) in phosphate buffer (pH 6.9 with 6 mM sodium chloride) was added to 50 µL of the starch solution (0.05%) and the reaction was stopped with the addition of 25 µL of HCl (1 M). Then 100 µL of the iodine-potassium iodide solution was added. After 10 min of incubation, the absorbance was recorded at 630 nm and the results expressed as millimoles of acarbose equivalents per gram of dry extract (ACAEs per g extract).
| Extracts | Total bioactive contents | Antioxidant activities | |||
|---|---|---|---|---|---|
| Total phenolic content (mg GAE per g) | Total flavonoid content (mg QE per g) | DPPH (mg TE per g extract) | FRAP (mg TE per g extract) | Phosphomolybdenum (mg TE per g extract) | |
| a Data from three repetitions, with mean ± standard deviation; means with different superscript letters in the same column are significantly (p < 0.05) different. GAE: gallic acid equivalent; QE: quercetin equivalent. | |||||
| Methanol | 71.53 ± 1.34 | 64.32 ± 0.52 | 113.45 ± 1.92 | 173.53 ± 0.91 | 32.83 ± 0.39 |
| n-Hexane | 12.74 ± 0.92 | 9.31 ± 1.12 | 13.54 ± 1.31 | 24.61 ± 0.92 | 73.69 ± 1.31 |
| Chloroform | 33.91 ± 1.02 | 45.71 ± 0.21 | 93.24 ± 0.87 | 81.04 ± 0.91 | 51.19 ± 1.31 |
| EA | 54.82 ± 1.37 | 39.87 ± 1.49 | 106.74 ± 1.52 | 95.49 ± 1.17 | 43.92 ± 0.83 |
| n-Butanol | 62.94 ± 0.43 | 49.87 ± 0.32 | 83.21 ± 0.96 | 89.42 ± 1.48 | 23.41 ± 0.93 |
The data, reflecting means from three repetitions with standard deviations, was statistically analyzed, and significant differences (p < 0.05) were noted among the extracts, indicating that the choice of solvent critically influences the extraction efficiency and the type of bioactive compounds obtained. These findings emphasize the necessity for selecting appropriate extraction solvents based on the desired chemical profile for applications in nutraceuticals and pharmaceutical formulations. Future studies should focus on identifying the specific antioxidant compounds extracted by n-hexane and exploring the biological implications of the extracts' antioxidant capacities in in vivo systems.
In the COSY 45°spectrum of WTH1, the C-2 methine proton coupling with C-3 methine, and C-6 methine proton with C-7 methine. C-22 methine proton showed vicinal coupling with the C-23 methylene proton. The 13C-NMR spectra showed the signal δ 128.2, 148.1, 119.6, 127.7 and 130.2 were assigned to vinylic carbon (C-2, C-3, C-4, C-6, and C-7, respectively). In ring A and B. while the peak at δ 149.0, 136.1 and 122.8 were attributed to the quaternary vinylic carbon (C-24, C-5 and C-25, respectively). The oxygen bearing methane carbon were at δ 83.5 (C-22) and at δ 58.5 for C-27 methylene carbon. The signal appearing at δ 20.2, 25.5, 18.5, and 12.5 were assigned to the C-28, C-21, C-19, C-18 methyle, respectively in Table 1. In the HMBC spectrum of WTH1, (C-19) methyl proton (δ 1.31) showed correlation with carbon resonating at δ 58.5 (C-10), δ 203.1 (C-1), and δ 148.1 (C-3), the C-3 proton (δ 6.89) indicated correlations with the down field carbon resonating at δ 203.1 (C-1) and 136.1 (C-5). The C-28 methyl proton (δ 1.98) exhibited coupling with C-24 methyl proton showed also interaction with the lactone carbonyl (C-26) resonating at δ 168.2. Whereas, C-28 methyl protons exhibited interaction with C-23 (δ 33.1) (Fig. 8). The stereo chemistry at C-22 was assigned on biogenetic ground and by chemical shift comparison of known withanolide data,21 the spectroscopy evidences led to structure of WTH1. The comprehensive structural and spectroscopic characterization of compound WTH1, including 1H and 13C NMR, DEPT experiments, 2D NMR (HSQC, HMBC, COSY, NOESY) for proton-carbon correlations and spatial interactions, as well as GC-MS analysis highlighting the molecular ion peak and fragmentation patterns, is provided in the SI (Fig. S1–S4).
The WTH2 new compound was isolated from chloroform extract as colourless powder. The electron impact (EI) mass spectrum fragment ion at m/z 436.3 corresponding to the formula C28H30O4, further peak was observed at m/z 418.3, 403.3 and 267, respectively. Unsaturated δ-lactone A and B α, β with an intense band at 222 nm was characterized through UV spectrum and supported by IR spectrum, which indicated absorption band at 1698, 1710 and 3500 cm−1 for a six membered cyclic kentone α, β-unsaturated δ-lactone and hydroxyl group. The 1H-NMR spectrum of WTH2 has close resemblance to WTH1, indicating the same substitution in ring A, and B side. Instead of a double bond at C-4 and ether bridge substituted at C 14/20 compound WTH2 in view of the shape of the signal of the 22-H, and while, its position also indicated the absence of a C-17-OH according to the molecular ion in the mass spectrum. The 1H-NMR spectral data of 2 displayed five methyl signals at δ 1.38 (3H, s, CH3-18) and δ 1.26 (3H, s, CH3-19), δ 1.40 (3H, s, CH3-21), δ 1.98 (3H, s, CH3-27), δ 1. 85 (3H, s, CH3-28). A downfield double doublet at δ 4.34 (1H, J22α,23α, = 12.7 Hz, J22α,23β = 3.4 Hz, H22). The multiplicity of H-22 indicated the absence of proton attached to vicinal C-20.22 Assignments of all functionality were achieved by HMBC and HMQC. The HMBC spectrum the 19-Me (δ 1.26), showed long range couplings (3J) with C-1 (δ 203.4), C-5 (δ 136.4), and C-9 (δ 37.1). Similarly, 3J the epoxy-bearing C-14 quaternary carbon (δ 84.4) showed 3J coupling with the 17 methine and 18-methyl protons. Likewise, 3H-21 showed 2J and 3J coupling with C-20 (δ 76.2) and C-17 (δ 50.3), respectively (Fig. 8). This spectroscopy data led to the structure for WTH2. The structural and spectroscopic data of compound WTH2, including 1H and 13C NMR, DEPT experiments, 2D NMR (HSQC, HMBC, COSY, NOESY) for detailed proton-carbon connectivity and spatial correlations, along with GC-MS analysis illustrating the molecular ion peak and fragmentation patterns, are provided in the SI (Fig. S5–S8).
| Extracts/fractions | Neurological problems | Inflammation | Diabetes | Skin problems | |
|---|---|---|---|---|---|
| AChE inhibition (mg GALAE per g extract) | BChE inhibition (mg GALAE per g extract) | Lipoxygenase (mg QE per g extract) | Glucosidase (mmol ACAE per g extract) | Tyrosinase (mg KAE per g extract) | |
| a GALAE: galatamine equivalent; QE: quercetin equivalent; ACAE: acarbose equivalent; KAE: kojic acid equivalent. All values expressed are means ± S.D. of three parallel measurements. Data marked with different letters within the same column indicate statistically significant differences for each species (p < 0.05). | |||||
| Methanol | 4.10 ± 0.91 | 3.71 ± 1.03 | 144.76 ± 2.13 | 1.94 ± 0.79 | 0.68 ± 0.68 |
| n-Hexane | 2.41 ± 1.31 | 1.70 ± 0.67 | 124.99 ± 1.14 | 1.87 ± 1.92 | 0.65 ± 1.13 |
| Chloroform | 3.51 ± 1.72 | 2.91 ± 1.12 | 97.63 ± 0.41 | 2.11 ± 0.67 | 0.83 ± 0.91 |
| EA | 3.91 ± 0.58 | 3.14 ± 1.42 | 137.84 ± 0.57 | 3.51 ± 0.52 | 0.18 ± 0.41 |
| n-Butanol | 2.87 ± 0.17 | 3.84 ± 0.41 | 129.04 ± 0.81 | 2.61 ± 1.12 | 0.94 ± 1.31 |
The findings highlight the complex interactions between the chemical profiles of W. coagulans extracts and their potential therapeutic applications. The varied enzyme inhibitory activities demonstrate the necessity for targeted research and development to optimize extraction methods that maximize health benefits, particularly for inflammation, neurological disorders, diabetes, and skin health. The correlations between bioactive content and biological activity also provide valuable insights, suggesting areas for further research in the extraction and application of natural products for health interventions.
![]() | ||
| Fig. 2 Chemical structures of the isolated phytoconstituents and the benchmark inhibitors of enzymes under study. | ||
In the case of AChE, the binding scores of the compounds ranged from −8.410 to −11.616 kcal mol−1, whereas the standard compound (physostigmine, PHY or eserine) scored −5.85 kcal mol−1. The WTH1, WTH2 and WTH3 have docking score −11.616, −8.410 and −11.438 respectively. The WTH1 has highest binding affinity with AChE. The BChE binding scores of the compounds ranged from −5.144 to −8.615 kcal mol−1, whereas the standard compound PHY scored −5.87 kcal mol−1. WTH3 demonstrated the highest binding affinity among the analyzed compounds, with a score of −9.30 kcal mol−1. The WTH1 and WTH2 have docking score −5.144 and −6.167 kcal mol−1 respectively. For alpha-glucosidase (GLU), the docking scores of the compounds ranged from −2.986 to −7.267 kcal mol−1. The WTH3 docking score −7.267 kcal mol−1 was comparable to that of the standard inhibitor ACB (−8.891 kcal mol−1). On other hand WTH1 and WTH2 have docking score −3.973 and −2.986 kcal mol−1 respectively. Apart from GLU enzyme, WTH1 also demonstrated highest binding affinity towards TYR enzyme (−5.042 kcal mol−1). Whereas, the standard inhibitor KJC also showed comparable binding potential (−6.81 kcal mol−1). The WTH2 and WTH3 have docking score −4.543 and −4.696 respectively for TYR. The docking scores of the compounds in LOX-bonded systems ranged from −5.867 to −6.139 kcal mol−1. With a highest docking scores of −6.139 kcal mol−1, WTH1 displayed significant binding potential towards LOX. The docking scores of the standard compound was −6.53 kcal mol−1, which is comparable to WTH3. WTH2 and WTH3 have docking score −5.922 and −5.867 kcal mol−1 which is also comparable to the standard baiclain docking score −6.53 kcal mol−1 as shown in Table 3. These findings clearly indicate that among isolated compounds WTH1, WTH2 and WTH3 are the main active ingredients responsible for the enzyme inhibitory potential of W. coagulans.
| Docking scores (kcal mol−1) | |||||
|---|---|---|---|---|---|
| Compounds | AChE | BChE | GLU | TYR | LOX |
| Standard | −5.85 | −5.87 | −8.89 | −6.81 | −6.53 |
| WTH1 | −11.616 | −5.144 | −3.973 | −5.042 | −6.139 |
| WTH2 | −8.410 | −6.167 | −2.986 | −4.543 | −5.922 |
| WTH3 | −11.438 | −9.30 | −7.267 | −4.696 | −5.867 |
To elucidate the variances in docking scores resultant from diverse interaction modalities, the optimal docking configurations for each molecule were documented and depicted graphically (Fig. 3). All ligands, along with benchmark inhibitors, occupy the same binding pocket. The graphical representation discloses that the AChE-ligand complexes share common interacting residues D72, Y70, Y121, W279, F290, F288, I287, and Y334 as demonstrated in (Fig. 3 superpose). In contrast, the AChE-PHY complex interacts with a distinct set of residues, including Q74, R289, F290, Y121, S286, and Y70, as illustrated in (Fig. 3 Image standard inhibitor). Furthermore, AChE and the three ligands integrate profoundly within the dual sub-sites (CAS and PAS) of AChE main active site. In the AChE-WTH1 complex, the binding involves key residues Y72, T83, W86, W286, F297, Y337, V294, F338, Y133, E202, and F295. WTH1 binds tightly through key hydrogen bonds with E202, Y72, and F338 (Image AI) and enhances hydrophobic interactions via residues F295 and F297, contributing to its specificity and stability (Image AI, Fig. S1 Image A1). The pyran-2-one ring of WTH1 penetrates deeply into the PAS region, where its hydroxyl group at the methyl-substituted carbon of the ring forms a hydrogen bond with the backbone oxygen of F338. Additionally, the hydroxyl group at carbon attach with ring D into the hydrophobic gorge between the CAS and PAS sub-sites, forming a hydrogen bond with the side chain hydroxyl group of Y72. The oxygen atom linked to the steroid backbone forms a hydrogen bond with the hydroxyl group of residues E202. PLIP analysis shows WTH1 binds tightly to AChE with key hydrophobic interactions and multiple contacts, indicating high specificity and effective binding with distances between 3.10 to 3.93 Å. In AChE-WTH2 bonded system have residues S81, W84, N85, H440, S122, Y121, F290, F288, F330, F331, H440, E199 and W279. It has one hydrogen bond with Y121 as shown in (Fig. 3 Image AII). The steroid backbone has oxygen atom make hydrogen bond with hydroxyl group of Y121. Residue F288, F290, F330, and F331, along with I287, strengthen the hydrophobic environment within the binding pocket. S81, S122, E199, and H440 play crucial roles in hydrogen bonding and ionic interactions, orienting the ligand for optimal inhibition (Fig. S1 Image B). WTH2 exhibits a robust binding profile as an AChE inhibitor, characterized by strategic hydrogen bonding and extensive hydrophobic contacts with a tight ligand–protein fit indicated by distances ranging from 3.32 to 3.88 Å. In AChE-WTH3 bonded system have residues W84, S286, F331, F330, H440, E199, S122, W279, S286, Y334, G335, Y121, I287, Y130 and S200. It has four hydrogen bonds, each H-bond bond with Y130, Y121, Y334 and G335 as shown in (Fig. 3 Image AIII). The hydroxyl group attach with steroid backbone make hydrogen bond with oxygen extended from hydroxyl group of Y130. The oxygen present in the ring D of steroid backbone also make hydrogen bond with oxygen extended from hydroxyl group of Y121. The hydroxyl groups of lactone group attach with the epoxide group make two hydrogen bond, one hydrogen bond with oxygen of aliphatic chain of Y334 and other hydrogen bond with oxygen of G335. In ACHE-WTH3 interaction, aromatic residues such as W84, Y121, and F330 engage in both π–π stacking and hydrogen bonding, enhancing ligand affinity (Fig. S1 Image C). Notably, H440 s imidazole side chain might stabilize the ligand through possibly π interactions. These additional interactions resulting from the structure of the evaluated compounds account for the second higher docking score of WTH3 compared to the WTH1. While PLIP interpret W279, L282, and F331 enhance ligand packing with hydrophobic interaction from 3.18 Å to 3.92 Å. AChE key residues G118, Y130, E199, S200, and D285 engage in hydrogen bonds ranging from 1.96 Å to 3.21 Å.
In BChE-ligand bonded systems, the residues surrounding the standard PHY and tested compounds are W82, S287, Y128, E197, G121, L125, H438, Y332 and F329 as shown in (Fig. 3 Image superpose). In BChE-PHY bonded system have residues W82, Y128, E197, A199, L286, V288, F398, W231, G115, T120, G115 and F329. In BChE-PHY bonded system, two hydrogen bonds with residue E197 (Fig. 3 Image standard inhibitor). In BChE-WTH1 have residues L286, V288, F398, S198, H438, G116, L286, Y332, P285, F329, D70, W231, T120, and S72. The oxygen of pyran-2-one ring structure make one hydrogen bond with hydroxyl group of S198 and one hydrogen bond with nitrogen of H438 (Fig. 3 Image BI). Hydrophobic interactions with residues F329, Y332, D70, L286, G117, and V288 form a snug pocket around the ligand, facilitating van der Waals interactions as depicted in (Fig. S1 Image D). Using PLIP, BChE-WTH1 interactions involve hydrophobic contacts with D70, G116, G117, L286, V288, F329, and Y332 ranging from 3.22 to 3.97 Å, and hydrogen bonds by G117 at 3.10 Å and 3.04 Å, S198 at 2.13 Å and 2.37 Å, and A199 at 3.18 Å, showing a tight fit and strong interactions in the binding site. H438 forms a salt bridge at 4.96 Å, highlighting precise ionic interactions with WTH1. In BChE-WTH2 bonded system have various different residues Q119, S198, G439, T120, P285, and I442 (Fig. 3 Image BII). It has hydrophobic interaction with residues W82, Y332, Y128, G116, E197, I442 and H438. This complex has no hydrogen bonding with any residue as shown in SI (Fig. S1 Image E). The presence of an aromatic ring, likely that of Y332, is poised to contribute to potential π–π interactions, adding another layer of specificity and binding affinity to the ligand. PLIP analysis reveals that WTH2 engages in key hydrophobic interactions with BCHE at 3.17 Å for W82 and 3.61 Å for Y332, and forms hydrogen bonds at 3.94 Å from W82 and 3.25 Å from G116, highlighting essential anchoring within the BCHE binding site. In BChE-WTH3 bonded system have various different residues W82, Y128, E197, V280, A277, F329, N289, T120, S198, T284 and H438. WTH3 has three hydrogen bonds, one hydrogen bond between hydrogen of one hydroxyl group of lactone ring and oxygen of hydroxyl group of Y128 and other hydrogen bond with other hydroxyl group of lactone ring and oxygen extended from residue E197 (Fig. 3 image BIII). The hydroxyl group of steroid backbone also make hydrogen bond with oxygen extended from residue A277. It also has hydrophobic interaction with residues W82, H438, F329 and E197 as shown in SI (Fig. S1 Image F). The indole ring of W82 are positioned to engage in stacking interactions, potentially stabilizing the ligand through π–π contacts. The network of surrounding residues, like T120, N289, and the aliphatic chains of A277 and V280, further sculpt the contours of this niche, possibly through hydrophobic contacts and van der Waals forces. PLIP analysis of the BCHE-WTH3 complex shows hydrophobic contacts with D70, T120, A277, S287, and F329 ranging from 3.37 Å to 3.79 Å, supporting ligand-binding pocket formation, and details hydrogen bonds from G116, T120, G121, Y128, E197, A277, and S287 between 1.96 Å to 3.87 Å, highlighting critical stabilizing interactions within the protein structure.
In GLU the binding cavity consists of various residues, including D68, D349, A278, D349, P309, F310, R312, H348, F157, R439 and F157 (Fig. 3 Image superpose). In GLU-Acarbose system have different residues from superpose docking are N241, F300, Q181, E276, H111, D214, F231, H239, H279, and F177. It has approximately 15 hydrogen bonds with the surrounding residues, including D214, D68, H111, H239, H279, E304, E276, P309, R212, H348, D349 and F158. There are two hydrogen bonds with each residue D68, D214 and E276 (Fig. 3 Image standard inhibitor). In GLU-WTH1 have residues H239, N241, W242, R312, S308, P309, E304, H279, F157, F311 and E304. It has three hydrogen bonds, one hydrogen bond between methyl–hydroxyl group of pyran-2-one ring and nitro group of H239, second hydrogen bond between oxygen of pyran-2-one ring and nitro group of N241. The WTH1 hydroxyl group at carbon attach with ring D of steroid backbone make hydrogen bond with extended oxygen of E304 (Fig. 3 Image CI). It has hydrophobic interaction with N241, H245, W242 and H279 as shown in SI (Fig. S1 Image G). The arrangement of aromatic rings, particularly the ring of H245 and H279 possibly stabilizing the ligand through π–π contacts. Hydrophobic contacts between N241, W242, and P307 with the ligand range from 3.46 to 3.94 Å, while hydrogen bonds with Q239 and G304 measure 2.16 Å and 1.68 Å, respectively, with a notable 167.09° donor angle for Q239. Additionally, a salt bridge with H253 at 4.30 Å enhances electrostatic complementarity with WTH1. In GLU-WTH2 have residues A278, S308, H239, H245, E276, F177, F157, R439, F300, P309, N241, H239, F158, T215, D349 and R312. It has one hydrogen bond with H239 (Fig. 2 Image CII). The steroid backbone ring D have oxygen atom make hydrogen bond with nitro group of H239. It has hydrophobic interaction with residues H245, R439 and R312 as shown in SI (Fig. S1 Image H). Additionally, other residues such as S308, P309, N241, H239, D349, E276, and T215 in proximity to the ligand, which may also play roles in binding through various interactions like van der Waals forces or additional hydrogen bonds. Hydrophobic contacts with F157 and A278 at around 3.5 Å, hydrogen bonds with N241 and E276 showing distances of 2.91 Å and 3.20 Å and angles above 106°, and a salt bridge with H239 at 3.47 Å collectively stabilize the ligand within the binding site. In GLU-WTH3 have residues P309, R312, D408, N412, F157, K155, H239, N241 and F231. It has six hydrogen bonds with various residues two hydrogen bonds with H239, one hydrogen bonds with N241, two hydrogen bonds with F157 and one hydrogen bond with D408 (Fig. 3 Image CIII). The lactone ring attaches with epoxide group make four hydrogen bonds, two hydrogen bond with oxygen of F157, one hydrogen bond with oxygen of D408 and oxygen present within the lactone ring make hydrogen bond with nitrogen of H239. The oxygen between lactone and epoxide group also make hydrogen bond with nitrogen of H239. It also has hydrophobic interactions with residues W242, F231, and R312 as shown in SI (Fig. S1 Image I). Hydrophobic interactions with P231A at 3.81 Å, W242A at 3.70 Å, and R312A at 3.91 Å suggest effective ligand packing, while hydrogen bonds with L155A at 3.03 Å, F157A at 1.99 Å, N241A at 2.65 Å, R312A at 2.43 Å, and D408A at 2.51 Å enhance the specificity and stability of ligand–protein binding.
In TYR, the binding cavity of kojic acid-tested compounds have common residues, including P134, Q133, R20, E16, H94, T84, C83, G62, P284, S282 and P290 (Fig. 3 image superpose). There are residues F264, H61, T261, M280 and A286 in the binding cavity standard TYR-kojic acid different from binding cavity of other compounds (Fig. 3 Image standard inhibitor). It has four H-bonds, two H-bonds with H263 and one H-bond with each residue, one H-bond with F264 and one with A286. In TYR-WTH1 bind cavity has residues H85, N81, P284, S282, H263, H259, F264, N260, V283, H244, V248 and R268. It has two hydrogen bonds, one with H259 and other hydrogen bond with H263 (Fig. 3 Image DI). Hydrogen bonds between the nitrogen in the imidazole ring of H263, H239 and a methyl–hydroxyl group pyran-2-one ring of WTH1. It also has hydrophobic interaction with H85, H244, P284 and F264 as shown in SI (Fig. S1 Image J). Hydrophobic interactions with H263, F264, and V283 range from 3.51 to 3.82 Å, stabilizing the complex through nonpolar contacts, while H263 forms a hydrogen bond at 2.56 Å and a 112.54° angle and a salt bridge at 5.32 Å, underscoring strong directional and electrostatic interactions that enhance binding specificity. In TYR-WTH2 has common residues F90, H61, H85, H259, S282, H263, V283, G281, N260, E256 and F264. It has five hydrogen bonds, three H-bonds with H85 and one H-bond with V283 and one H-bond with G281 (Fig. 3 Image DII). The oxygen of ring A makes two hydrogen bonds, one H-bond with nitrogen of V283 and other with hydroxyl group of G281. The oxygen within the ring of pyran-2-one make two hydrogen bonds with H85. The oxygen extended from epoxide group also make H-bond with H85. It also has hydrophobic interaction with N260, H263, H259, F264 and H61 as shown in SI (Fig. S1 Image K). Residue F90, F264 and V283 are likely the key players in the hydrophobic interactions. F264 has a bulky, nonpolar benzyl side chain, which can engage in van der Waals forces with hydrophobic regions on the ligand. V283 with its smaller aliphatic side chain, also contributes to the hydrophobic contact points. The TYR-WTH2 interaction includes hydrophobic contacts with H85A, E256A, N260A, H263A, and V283A between 2.74 and 3.47 Å, and hydrogen bonds by S282A at 2.21 Å and V283A at 2.45 Å, enhancing specificity and affinity. Salt bridges with H61A and H85A further stabilize the complex through electrostatic interactions. In TYR-WTH3 have residues F264, S282, P284, V283, A286, V248, H263, N260, F292, H61, T84, C83, T324, N81 and E322. It has five hydrogen bonds, two hydrogen bonds with H61, Two H-bonds with T84 and one hydrogen bond with C83 (Fig. 3 Image DIII). The imidazole side chain of H61 make two hydrogen bonds, one H-bond with oxygen within the lactone ring and other H-bond with hydroxyl group of lactone ring. T84 and C83 make H-bond with same hydroxyl group of lactone ring and oxygen T84 make another hydrogen bond with other hydroxyl group of lactone ring. It has hydrophobic interaction with resides A286, F264, F292, H263, P284, V286 and V283 as shown in SI (Fig. S1 Image L). The TYR-WTH3 complex features hydrophobic interactions with N260A, H263A, P264A, and V283A ranging from 3.23 Å to 3.78 Å for a tight ligand fit, and strong hydrogen bonds with A81A and C83A at 2.53 Å and 1.84 Å, respectively.
In the LOX binding cavity of Baicalain-compound have common residues including V671, S171, H372, H550, H367, N554, Q557, L607, F177, K409, I406 and A672 (Fig. 3 Image superpose). In LOX-Baicalain binding cavity have residues Q141, R138, R101, Y142, E134, K133 and T137. It has three hydrogen bonds one hydrogen bond with R138 and two hydrogen bonds with E134 (Fig. 3 Image standard inhibitor). In LOX-WTH1 have residues different from standard binding cavity are K409, A672, V671, H372, H550, H367, N554, I406, S171 and L607. It has three hydrogen bond, one hydrogen bond with H367, one hydrogen bond with F177 and one bond with S171 (Fig. 3 Image EI). The methyl-hydroxyl group of pyran-2-one ring make hydrogen bond with extended oxygen from S171. The hydroxyl at carbon attach with ring D of WTH1 make hydrogen bond with nitrogen of F177. It has hydrophobic interaction with the L607, F177, I406 and V671 which collectively create a nonpolar environment that helps secure the ligand in place through van der Waals forces. In the LOX-WTH1 complex, the hydrophobic interactions are delineated by the binding of F177 (2.94 Å), K409A (3.64 Å), and L607 (3.41 Å), which anchors the ligand within the hydrophobic core of the protein. In LOX-WTH2 binding cavity have residues L179, V178, L607, F177, Q413, H367, K409, I406 and S171. It has three hydrogen bonds, one hydrogen bond with Q413, two hydrogen bond with V178 and one bond with L179 (Fig. 3 Image EII). The nitrogen of V178, Q413 and L179 make hydrogen bond with oxygen extended from pyran-2-one ring. It has hydrophobic interaction with residues F177, I406, Q413 and V178. All residues present nonpolar side chains that interact with the hydrophobic parts of the ligand as shown in (Fig. S1 Image N). In the LOX-WTH2 complex, hydrophobic interactions with P177, L179, I406, and L607 range from 3.25 Å to 3.61 Å, securing the ligand in a stable, nonpolar region. Hydrogen bonds by V178, L179, and Q413, ranging from 2.58 Å to 2.70 Å with angles up to 163.83°, enhance specificity and affinity for the LOX enzyme. In LOX-WTH3 have residues H372, H367, Q363, N554, F555, H360, P569, T570, H600, R596, L607, F177, F610 and F359. It has six hydrogen bonds, two hydrogen bonds with R596, two hydrogen bonds with H600, one hydrogen bond with H367 and one hydrogen bond with Q363 (Fig. 3 Image EIII). The methyl–hydroxyl group of lactone ring make three hydrogen bonds, two hydrogen bonds with two nitrogen of R596 and other hydrogen bond with nitrogen of H600. Other hydroxyl attaches with lactone ring also make hydrogen bond with nitrogen of H600. The oxygen attach with ring A of steroid backbone makes hydrogen bond with oxygen attach with H367 and H372. The hydroxyl group attach with ring A of steroid backbone makes hydrogen bond with Q363. Meanwhile, the ligand is also ensconced within a hydrophobic pocket, with F177, F359, F610, L607, and P569 as shown in (Fig. S1 Image O) contributing to a nonpolar environment. In the LOX-WTH3 complex, P177, I406, A410, and L607 engage in hydrophobic interactions, anchoring the ligand with distances between 3.34 Å to 3.90 Å. Hydrogen bonds are formed with N554 and R596 at distances as close as 2.09 Å, and with H600 at 2.09 Å, establishing tight, specific contacts critical for stable ligand binding.
| Molecule | TPSA | Silicos-IT log Sw |
GI (% absorbed) | (SP) log Kp (cm s−1) |
BBB permeant | Pgp substrate | 1A2 inhibitor | 2C19 inhibitor | 2C9 inhibitor | 2D6 inhibitor | 3A4 inhibitor | Bioavailability | Synthetic accessibility | TC (log ml min−1 kg−1) | AMES toxicity | MTDH (log mg per kg per day) | LD50 (mol kg−1) | Hepatotoxicity | MT (log mM) |
|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
| WTH1 | 83.8 | −3.84 | 97.67 | −6.06 | No | Yes | No | No | Yes | No | Yes | 0.55 | 6.47 | 0.473 | No | −0.664 | 2.24 | No | 0.149 |
| WTH2 | 52.6 | −5.62 | 98.06 | −5.8 | Yes | No | No | No | Yes | No | No | 0.55 | 6.59 | 0.31 | No | −0.406 | 1.83 | No | 0.272 |
| WTH3 | 175.5 | −3.81 | 94.56 | −9.79 | No | Yes | No | No | No | No | No | 0.17 | 8.09 | 0.452 | No | −0.685 | 2.14 | No | 0.146 |
The reactivity of the compounds is strongly dependent on this analysis. This is especially true for the HOMO–LUMO energy gap, which is the lowest unoccupied and highest occupied molecular orbital energy difference. It is this gap that is necessary to understand the energy transfer of the molecule to itself. In Fig. 4, the molecular surface plots of HOMO and LUMO orbitals are shown for WTH1, WTH2 and WTH3. Contact values for the HOMO and LUMO indicate the compound's ability to accept or donate electrons. In Table 5, a summary is provided of the calculated quantum chemical descriptors for these compounds under aqueous conditions. DFT calculations have shed light on the AChE inhibitors' reactivity patterns and electronic structures. Notably, WTH2 exhibited an extraordinarily high dipole moment of 6.0750 debye, indicating a distinct electronic profile compared to WTH1 and WTH3, which demonstrated significantly lower dipole moments of 2.0757 and 2.3116 debye respectively. This suggests different electron-donating capabilities and interaction potentials with AChE. The HOMO–LUMO energy gap analysis revealed that WTH2 also exhibited the largest disparity, signifying enhanced reactivity and potential biological efficacy. These findings not only deepen our understanding of the molecular basis for the inhibitory activity against AChE but also highlight the critical need to optimize electronic and structural properties for the development of potent AChE inhibitors. This detailed analysis sets the foundation for subsequent structure–activity relationship studies, which will guide the design of novel inhibitors with improved therapeutic profiles.
| Parameters for DFT analysis | |||||||||||
|---|---|---|---|---|---|---|---|---|---|---|---|
| Ligand | Dipole moment (debye) | HOMO (a.u.) | LUMO (a.u.) | Energy gap (ΔEGap) | Ionization potential (eV) | Electron affinity (eV) | Electronegativity χ (eV) | Electrochemical potential µ (eV) | Hardness η (eV) | Softness S (eV) | Electrophilicity ω (eV) |
| WTH1 | 2.0757 | −7.944 | 1.979 | 9.923 | 7.944 | −1.979 | 2.983 | −2.983 | 9.923 | 0.101 | 0.448 |
| WTH2 | 6.0750 | −9.164 | 2.933 | 12.097 | 9.164 | −2.933 | 3.115 | −3.115 | 12.097 | 0.083 | 0.401 |
| WTHA | 2.3116 | −8.958 | −4.468 | 4.49 | 8.958 | 4.468 | 6.713 | −6.713 | 4.49 | 0.223 | 5.018 |
Employing Koopman's theorem, global reactivity parameters for ligands WTH1, WTH2, and WTH3 have been computed, demonstrating notable differences in their electronic properties as depicted in the data. WTH1 is characterized by an intermediate level of electronegativity (χ = 2.983 eV) and chemical potential (µ = −2.983 eV), with a substantial absolute hardness (η = 9.923 eV). The hardness obtained is very high and is consistent with a very strong resistance to deformation of the electron cloud, implying few degrees of freedom in adopting to changes in the electron distribution. This is further supported by its low global softness (σ = 0.101 eV−1) and moderate electrophilicity index (ω = 0.448 eV). On the other hand, WTH2 shows slightly higher electronegativity (χ = 3.115 eV), chemical potential (µ = −3.115 eV) and hardness (η = 12.097 eV) among the ligands under consideration. This high hardness is comparable to a striking degree of resistance to electronic rearrangement, and is a function of the lowest global softness (σ = 0.083 eV−1). As a result, WTH2 presents a rather less propensity to accept electrons, as its electrophilicity index (ω = 0.401 eV) suggests. Conversely, WTH3 stands out with significantly higher electronegativity (χ = 6.713 eV) and chemical potential (µ = −6.713 eV), signaling a potent ability to attract electrons compared to WTH1 and WTH2. It features the lowest hardness (η = 4.49 eV) among the ligands, which correlates with a higher global softness (σ = 0.223 eV−1). This suggests a greater flexibility in its electron cloud, facilitating easier electron mobility. WTH3 electrophilicity index (ω = 5.018 eV) is notably the highest, indicating a pronounced ability to stabilize additional electron density and thereby enhancing its reactivity under electrophilic attack. These distinct parameters illuminate the underlying electronic behavior of WTH1 and WTH2 as more resistant to electron cloud deformation, while WTH3 emerges as markedly adept at electron stabilization, presenting a stark contrast in reactivity profiles, which could influence their utility in various chemical environments.
The analysis of system dynamic stability included RMSD assessment from initial structures during 100 ns molecular dynamics (MD) simulations to validate the sampling method. An RMSD analysis confirmed that the WTH1-AChE and eserine-AChE complexes plus all other simulation systems established equilibrium during the first 5 nanoseconds of the MD simulation. The equilibrium period for the simulations produced stabilized RMSD values amounting to 1.7 Å for protein Cα atoms, 1.0 Å for binding pocket backbone atoms and 0.8 Å for ligand heavy atoms. The visualization of stability results in Fig. 5A and B confirms the system stability which enables following analyses that depend on conformations sampled between 5 and 100 ns. Monitoring the structural stability used snapshot coordinates from the MD simulation which were superposed against initial structures as shown in Fig. 5C and D. MD simulation results showed that eserine- as well as WTH1-AChE complexes maintained structural stability throughout the simulation while both ligands kept their original positions by sustaining critical hydrogen bonds with PAS and CAS region residues. Our MD simulation results show high reliability because the obtained findings help explain inhibitor-AChE interaction mechanisms while preparing for binding free energy studies.
The root-mean-square fluctuation (RMSF) analysis encompassing all ligand–protein interactions, as depicted in Fig. 5E, delineates the dynamic profiles and RMSF distribution across the protein structures of all evaluated systems. These dynamics manifest consistent patterns, with three areas within the acetylcholinesterase (AChE) structure, the catalytic active site (CAS), peripheral anionic site (PAS), and Mid Gorge (situated between the CAS and PAS), displaying pronounced fluctuations. In particular, the CAS (residues 119–126 and 447–458) showed similar fluctuations in nonbonded and bonded interaction with AChE. However, in the systems on passive binding, the peripheral anionic site (PAS), which covers residues 72–76 and 280–290, had slightly higher fluctuations. These changes imply increased mobility of the Mid Gorge (residues 294–297) in the presence of AChE due to ligand binding and thus increased flexibility between the CAS and PAS regions. Additionally, residues 480–490 in the CAS region exhibited increased fluctuations in the AChE-eserine and WTH1 complexes. Despite the variability, the Mid Gorge and PAS regions generally followed a similar fluctuation pattern, whereas AChE-eserine and WTH1 complexes noted fluctuations in these loops. In contrast, the CAS, PAS, and Mid Gorge regions displayed stability across both AChE systems, underscoring their essential role in maintaining structural integrity upon ligand binding. Notably, the CAS residues in receptor–ligand bonded systems exhibited a heightened fluctuation amplitude relative to those in the nonbonded AChE configuration, highlighting the significant dynamic influence of ligand binding on the mobility of this critical region.
The radius of gyration (Rg) of AChE was measured throughout the simulation in order to gauge the effect of ligand binding on the structural integrity of the protein, as shown in Fig. 5F. About 22.6 Å and 22.7 Å average Rg values for AChE bound to eserine and WTH1, respectively, were found to be in reasonably good agreement. They serve as an indication that ligand binding does not activate significant structural change in the protein and therefore the ligands must bind to AChE without inducing extensive alteration that includes unfolding or expansion. Also, the simulation allowed the Rg values to remain within a small range of 22.5 Å to 22.8 Å thus indicating the structural stability of the protein. These results indicate that the interaction between eserine or WTH1 and AChE does not perturb its native conformation and that vesicles are required for the inhibitor activity of both compounds due to their effectiveness maintaining AChE architecture appropriate for proper ligand recognition. Further understanding of the dynamics between AChE and the inhibitors was obtained by further analysis of the solvent accessible surface area (SASA) as shown in Fig. 5G. The WTH1 complex had an SASA of approximately 750 Å2 indicating stable interaction within the complex. Surprisingly, eserine was as stable as WTH1 with an average SASA of 730 Å2. The fact that these ligands exhibit similar performance across both complexes suggests that they are good AChE inhibitors.
000 randomly selected snapshots from 1. To assess the binding affinity of the selected compounds towards AChE, both MM/PBSA and MM/GBSA methods were used. Using the MM/PBSA method, the calculated binding free energies for the inhibitors WTH1 (−43.54 kcal mol−1) and Esorine (−42.70 kcal mol−1) were obtained. The results in Table 6 indicate that WTH1 possesses higher ΔGpred (GB) values in AChE bound systems, suggesting higher affinity for the AChE active site. This is consistent with the fact that MM/GB/PB/SA values give better ΔGpred values in comparison to those obtained from docking studies and confirms our methodology in assessing binding efficiency of putative AChE inhibitors. By decomposition of the total binding free energy to its individual components, the MM/GB/PBSA method allows for an increased understanding in the dynamics of the ligand–receptor interaction. Polar solvation energies of the complexes as shown in Fig. 6 are positive, partially cancelling the favorable electrostatic energies (ΔEele) in the gas phase for both complexes. Consequently, the formation of ligand–receptor complex is avoided as a result of unfavorable electrostatic contributions (ΔGele + ΔGele,sol). However, negative values of van der Waals interactions and the nonpolar solvation energy (ΔEvdW + ΔGnonpol,sol) contribute positively to the binding affinity of WTH1 toward AChE. Notably, the ΔEvdW values are larger than one does the ΔEele values in all systems, and in all systems the vdW and nonpolar interactions play an important role in improving the inhibitory effectiveness of AChE inhibitors.
| Complex | ACHE-WTH1 (kcal mol−1) | ACHE-ESE (kcal mol−1) |
|---|---|---|
| ΔEvdWa | −51.5212 | −37.4592 |
| ΔEelea | −5.1078 | −19.1437 |
| ΔGnonpol,sola | −5.726 | −4.4617 |
| Δggas | −56.629 | −56.6029 |
| ΔGsol | 13.089 | 13.9041 |
| ΔGele,sol (PB)a | 31.0876 | 44.6799 |
| ΔGele,sol (GB)a | 18.8149 | 18.3657 |
| ΔEvdW + ΔGnonpol,sola | −57.2472 | −41.9209 |
| ΔEele+ΔGele,sol (PB)a | 25.9798 | 25.5362 |
| ΔEele + ΔGele,sol (GB)a | 13.7071 | −0.778 |
| ΔGpred (PB)b | −29.731 | −15.0831 |
| ΔGpred (GB)b | −43.54 | −42.6988 |
| C no | Compound 1 | Compound 2 | Compound 3 | ||||||
|---|---|---|---|---|---|---|---|---|---|
| δC | Multi | δH J in Hz | δC | Multi | δH J in Hz | δC | Multi | δH J in Hz | |
| 1 | 214.4 | C | — | 213 | C | — | 214.2 | C | — |
| 2 | 46.8 | CH2 | 2.74 (1H, m) | 47.2 | CH2 | 2.85 (1H, dd, J = 13.0, 9.9), 2.70 (1H, dd, J = 13.0, 5.7) | 46.8 | CH2 | 2.72 (1H, d, J = 7.0) |
| 2.09 (1H, m) | 2.00 (1H, m) | ||||||||
| 3 | 76.9 | CH | 4.00 (1H, m) | 76.8 | CH | 3.88 (1H, m) | 76.8 | CH | 4.00 (1H, m) |
| 4 | 38.8 | CH2 | 2.68 (1H, dd, J = 6.2, 13.6) | 38.8 | CH2 | 2.67 (1H, dd, J = 13.6, 5.8) | 38.8 | CH2 | 2.68 (1H, dd, J = 6.3, 13.6) |
| 2.45 (1H, d, J = 13.6) | 2.58 (1H, m) | 2.48 (1H, d, J = 13.6) | |||||||
| 5 | 136.4 | C | — | 135.5 | C | — | 136.5 | C | — |
| 6 | 126.8 | CH | 5.69 (1H, br s) | 127.4 | CH | 5.71 (1H, d, J = 5.2) | 126.8 | CH | 5.67 (1H, br s) |
| 7 | 26.6 | CH2 | 2.09 (2H, m) | 27.5 | CH2 | 2.21 (1H, m), 1.81 (1H, m) | 26.9 | CH2 | 2.10 (1H, m), 1.92 (1H, m) |
| 8 | 35.4 | CH | 1.90 (1H, m) | 38 | CH | 1.77 (1H, m) | 37.1 | CH | 2.58 (1H, m) |
| 9 | 37.4 | CH | 2.04 (1H, m) | 39.5 | CH | 1.85 (1H, m) | 37.1 | CH | 2.58 (1H, m) |
| 10 | 54.3 | C | — | 54.1 | C | — | 54.3 | C | — |
| 11 | 22.3 | CH2 | 1.99 (1H, m) | 23.2 | CH2 | 1.68 (1H, m) | 23.2 | CH2 | 2.08 (1H, m) |
| 1.54 (1H, m) | 1.40 (1H, m) | 1.64 (1H, m) | |||||||
| 12 | 33.2 | CH2 | 2.01 (1H, m) | 34.6 | CH2 | 2.46 (1H, m) | 31.5 | CH2 | 2.34 (1H, ddd, J = 24.6, 12.6, 5.3), 1.27 (1H, m) |
| 1.65 (1H, m) | 1.66 (1H, m) | ||||||||
| 13 | 49.6 | C | — | 55.8 | C | — | 55.7 | C | — |
| 14 | 85.5 | C | — | 86 | C | — | 84 | C | — |
| 15 | 32.9 | CH2 | 1.94 (1H, m) | 32.1 | CH2 | 2.08 (1H, m), 1.49 (1H, m) | 33.2 | CH2 | 1.74 (2H, m) |
| 1.52 (1H, m) | |||||||||
| 16 | 21.7 | CH2 | 1.89 (2H, m) | 33.2 | CH2 | 2.47 (1H, m) | 37.5 | CH2 | 1.52 (1H, m) |
| 1.57 (1H, m) | 1.48 (1H, m) | ||||||||
| 17 | 50.7 | CH | 2.30 (1H, t, J = 9.4) | 89.2 | C | — | 88.8 | C | — |
| 18 | 18 | CH3 | 1.04 (3H, s) | 17.9 | CH3 | 1.28 (3H, s) | 21 | CH3 | 1.11 (3H, s) |
| 19 | 18.5 | CH3 | 1.27 (3H, s) | 19.6 | CH3 | 1.31 (3H, s) | 18.8 | CH3 | 1.27 (3H, s) |
| 20 | 76.5 | C | — | 79.2 | C | — | 79.9 | C | — |
| 21 | 20.2 | CH3 | 1.28 (3H, s) | 17.8 | CH3 | 1.43 (3H, s) | 19.5 | CH3 | 1.37 (3H, s) |
| 22 | 83.2 | CH | 4.23 (1H, dd, J = 13.2, 3.4) | 84.3 | CH | 4.58 (1H, dd, J = 13.2, 3.4) | 83 | CH | 4.82 (1H, d, J = 3.4) |
| 23 | 32.8 | CH2 | 2.51 (1H, dd, J = 18.0, 13.2) | 35.2 | CH2 | 2.64 (1H, dd, J = 18.8,13.2) | 35.7 | CH2 | 2.64 (1H, dd, J = 18.8, 13.2) |
| 2.39 (1H, dd, J = 18.0, 3.4) | 2.47 (1H, dd, J = 18.8, 3.4) | 2.50 (1H, dd, J = 18.8, 3.42) | |||||||
| 24 | 157.5 | C | — | 153.1 | C | — | 153.4 | C | — |
| 25 | 126.8 | C | — | 121.9 | C | — | 122 | C | — |
| 26 | 168 | C | — | 169 | C | — | 169.1 | C | — |
| 27 | 56.4 | CH2 | 4.38 (1H, d, J = 11.2) | 12.3 | CH3 | 1.84 (3H, s) | 12.3 | CH3 | 1.84 (3H, s) |
| 4.30 (1H, d, J = 11.2) | |||||||||
| 28 | 20.2 | CH3 | 2.09 (3H, s) | 20.5 | CH3 | 1.95 (3H, s) | 20.5 | CH3 | 1.95 (3H, s) |
| 1′ | 103.1 | CH | 4.36 (1H, d, J = 7.6) | 102.9 | CH | 4.37 (1H, d, J = 7.8) | 103 | CH | 4.35 (1H, d, J = 7.7) |
| 2′ | 75.1 | CH | 3.13 (1H, dd, J = 7.8, 8.6) | 75 | CH | 3.13 (1H, dd, J = 7.8, 9.0) | 75 | CH | 3.12 (1H, dd, J = 7.7, 7.9) |
| 3′ | 77.9 | CH | 3.33 (1H, m) | 77.9 | CH | 3.33 (1H, m) | 77.9 | CH | 3.33 (1H, m) |
| 4′ | 71.6 | CH | 3.29 (1H, m) | 71.6 | CH | 3.29 (1H, m) | 71.6 | CH | 3.24 (1H, m) |
| 5′ | 78.1 | CH | 3.24 (1H, d, J = 4.5) | 78 | CH | 3.25 (1H, d, J = 5.1) | 78 | CH | 3.25 (1H, d, J = 5.0) |
| 6′ | 62.8 | CH2 | 3.84 (1H, d, J = 11.2) | 62.7 | CH2 | 3.83 (1H, d, J = 11.8) | 62.7 | CH2 | 3.83 (1H, d, J = 11.8) |
| 3.63 (1H, dd, J = 11.2, 4.5) | 3.63 (1H, dd, J = 11.8, 5.1) | 3.63 (1H, dd, J = 11.8, 5.0) | |||||||
Multiple hydrophobic amino acids such as W286, V294, W86, Y72, Y337, Y133, F297, F295, and F338 determine the level of binding efficiency. The interactions between WTH1 inhibitors and AChE also depend significantly on electrostatic forces but show lower influence than van der Waals and nonpolar solvation effects. The electrostatic forces make eserine more effective in binding AChE than WTH1. The inhibitory potency of the AChE-WTH1 (−51.5212 kcal mol−1) as well as AChE-eserine (−37.4592 kcal mol−1) complexes depends predominantly on van der Waals interactions according to their substantial negative ΔEvdW results. The results demonstrate that van der Waals and hydrophobic forces play the most powerful role in determining the inhibitory properties of AChE inhibitors but electrostatic forces have a secondary influence. This study reveals critical information about binding interactions while showing that strengthening both van der Waals forces and hydrophobic bonding will lead to better AChE inhibitor design.
| This journal is © The Royal Society of Chemistry 2025 |