Open Access Article
Manjunath R.
a,
Ashwini Raob,
Mahesha
c,
Udaya Kumar A. H.d,
Sudarshan Acharyae,
Padmanabha Udupa E. G.b,
Abdul Ajees Abdul Salam
e,
Sushruta S. Hakkimanef,
Shashikala B. S.c,
Lokanath N. K.
g and
Santosh L. Gaonkar
*a
aDepartment of Chemistry, Manipal Institute of Technology, Manipal Academy of Higher Education, Manipal 576104, Karnataka, India. E-mail: sl.gaonkar@manipal.edu
bDepartment of Biochemistry, Kasturba Medical College, Manipal, Manipal Academy of Higher Education, Manipal 576104, Karnataka, India
cDepartment of Physics, Sri Jayachamarajendra College of Engineering, JSS Technical Institutions Campus, JSS Science and Technology University, Mysuru 570006, Karnataka, India
dDepartment of Physics, Seshadripuram Institute of Technology, Mysuru 571311, Karnataka, India
eManipal Institute of Applied Physics, Manipal Academy of Higher Education, Manipal 576104, Karnataka, India
fDepartment of Biotechnology, Manipal Institute of Technology Bengaluru, Manipal Academy of Higher Education, Manipal 576104, Karnataka, India
gDepartment of Studies in Physics, University of Mysore, Manasagangothri, Mysuru 570006, Karnataka, India
First published on 24th September 2025
Imidazole derivatives are prominent in medicinal chemistry because of their vast array of biological effects. They are particularly noteworthy in treating hypertension, as demonstrated by imidazole-based medications such as lisinopril and losartan, which are currently available on the market. In this context, six N-substituted-2-butyl-4-chloro-1H-imidazole derivatives were carefully designed and synthesized through an efficient two-step protocol, with good yields. The synthesized compounds (4a–f) were characterized via various analytical techniques, including FTIR, 1H NMR, 13C NMR, and mass spectrometry. An in vitro assessment of angiotensin-converting enzyme inhibition was conducted. The results showed that compound 4b exhibited an exceptional IC50 value in the micromolar range (1.31 ± 0.026 μM). Additionally, in silico studies were performed, including molecular docking to predict the spatial orientation of the compounds, molecular dynamics simulations to evaluate binding stability with the target protein, and drug likeness studies to ensure adherence to Lipinski's rule. Furthermore, DFT analysis was employed to explore the energy gap of the frontier molecular orbitals (FMOs) and the molecular electrostatic potential (MEP), facilitating the identification of potential nucleophilic and electrophilic attack sites. Comprehensive insights into the molecular structure and packing of compound 4c were obtained through crystallographic studies, Hirshfeld surface analysis, Cambridge Structural Database studies, and energy framework analysis.
Clinically, ACE inhibitors are often used in conjunction with medications such as eprosartan, losartan, and other angiotensin II antagonists to manage hypertension and cardiovascular disorders. In pursuit of developing a novel heterocyclic library with drug-like properties for evaluating their efficacy as ACE inhibitors, we incorporated the 2-butyl-4-chloro-1H-imidazole unit, derived from the structures of losartan and eprosartan, into a cohesive molecular framework, as illustrated in Fig. 1. The selection of the five-membered, nitrogen-containing substituted imidazole unit was guided by its structural resemblance to the proline unit found in lisinopril and captopril.10
In designing the current library, we took account of previously reported scaffolds, including 2-butyl-4-chloro-1H-imidazole-derived chalcones (IC50 = 2.24 μM) and pyrazole analogues (IC50 = 1.80 mM), peptidomimetics (IC50 = 0.100 μM), and our previously reported 1,3,4-oxadiazole derivatives (IC50 = 51.01 μM), all of which exhibited varying degrees of anti-ACE activity.11–13 These comparative insights lead to designing our current compounds, with the goal of enhancing potency and expanding the scope of ACE inhibition.
In our previous work, we synthesized 2-butyl-4-chloroimidazole-derived 1,3,4-oxadiazoles, which exhibited moderate ACE inhibitory activity. In the current study, we adopted a new strategy by reacting 2-butyl-4-chloro-formyl imidazole with phenacyl bromides, followed by the reduction of both the aldehyde and keto groups to yield derivatives with two hydroxyl groups. Unlike chalcones and pyrazoles, which primarily rely on conjugated π-systems for their activity, and peptidomimetics, which mimic natural peptide substrates of ACE but often lack desirable drug-like properties, our current scaffold presents a hybrid design that integrates both lipophilic substituents (butyl, chloro) and polar functionalities (two –OH groups). This balanced combination distinguishes our compounds from previously reported scaffolds and offers a unique framework for investigating ACE inhibition.
This study presents the successful synthesis and evaluation of the ACE inhibitory activity of N-substituted-2-butyl-4-chloro-1H-imidazole derivatives, accompanied by computational analysis. Additionally, we conducted crystallographic and Hirshfeld studies for compound 4C, along with studies based on the Cambridge Structural Database (CSD).
The FTIR data for the N-substituted-2-butyl-4-chloro-1H-imidazole derivatives (4a–f) exhibited distinctive bands at various wavenumbers, including 794 cm−1 (C–Cl), 1597 cm−1 (C
C), 1675 cm−1 (C
N), 2833 cm−1 (aliphatic C–H), 2954 cm−1 (aromatic C–H), and 3181 cm−1 (OH). The 1H NMR spectra revealed notable signals at δ 0.87 for the aliphatic methyl group and peaks at δ 1.2 and 1.5 for the aliphatic methylene protons. A signal was detected at δ 5.1 for the CH proton, whereas the OH protons appeared at δ 5.6 and 5.7. Additionally, aromatic protons were observed in the δ 7–8 ppm range. The 13C NMR spectrum displayed distinct chemical shifts attributed to the carbons within the imidazole ring, with significant signals at approximately 123, 147, and 148 ppm. The signal for the aliphatic methyl group was at 14.21 ppm, whereas the signals for the aliphatic methylene carbons were at 22.29, 26.16, and 29.42 ppm. Furthermore, a notable signal for CH carbon was found near 71 ppm. The mass of compounds 4a and 4f exhibited peaks at m/z values of 339.47 and 327.29, respectively.
| Sr. no | Compound name | IC50 (μM) |
|---|---|---|
| 1 | 4a | 2.47 ± 0.102 |
| 2 | 4b | 1.31 ± 0.026 |
| 3 | 4c | 1.56 ± 0.065 |
| 4 | 4d | 2.12 ± 0.015 |
| 5 | 4e | 4.89 ± 0.068 |
| 6 | 4f | 7.57 ± 0.085 |
| 7 | Lisinopril | 0.3 ± 0.135 |
Compounds 4a–f demonstrated superior in vitro ACE inhibitory activity compared to previously reported chalcones (IC50 = 2.24 μM) and pyrazole analogues (IC50 = 1.80 mM), with IC50 values ranging from 1.30 to 7.57 μM. Although their potency does not exceed that of peptidomimetics (IC50 = 0.100 μM), which are known to mimic the natural substrate of ACE, our compounds present a non-peptidic framework with enhanced drug-like properties. Additionally, the current series shows a significant improvement over our previously reported 1,3,4-oxadiazoles (IC50 = 51.01 μM), underscoring the benefits of incorporating two hydroxyl groups alongside lipophilic substituents (butyl and chloro) to achieve an optimal balance between hydrophobic and hydrogen-bonding interactions within the ACE active site. The detailed structure–activity relationships of compounds 4a–f were discussed in Section 2.7.
| Parameters | Compound |
|---|---|
| CCDC deposit no. | 2448889 |
| Empirical formula | C17H23Cl1N2O2 |
| Formula weight | 322.837 |
| Temperature (K) | 293 |
| Wavelength (Å) | 0.71073 |
| Crystal system, space group | Monoclinic, P21/c |
| Unit cell dimensions a (Å) | 11.2517(5) |
| b (Å) | 13.8607(6) |
| c (Å) | 11.2925(7) |
| β (°) | 94.920(5) |
| Volume (Å3) | 1754.65(15) |
| Z | 4 |
| Density(calculated) (Mg m−3) | 1.222 |
| Absorption coefficient (mm−1) | 0.226 |
| F000 | 689 |
| Crystal size (mm3) | 0.22 × 0.19 × 0.23 |
| 2θ range for data collection | 3.64° to 51° |
| Index ranges | −13 ≤ h ≤ 14 |
| −18 ≤ k ≤ 11 | |
| −14 ≤ l ≤ 14 | |
| Reflections collected | 15 638 |
| Unique reflections | 3267 [Rint = 0.0246] |
| Absorption correction | Multiscan |
| Refinement method | Full matrix least-squares on F2 |
| Data/restraints/parameters | 3267/55/233 |
| Goodness-of-fit on F2 | 1.062 |
| Final [I > 2σ(I)] | R1 = 0.0521, wR2 = 0.01319 |
| R Indices (all data) | R1 = 0.0658, wR2 = 0.1408 |
| Largest diff. Peak and hole | 0.22 and −0.41 eÅ−3 |
The structural conformation and crystal packing of the title compound (4c) are stabilized by supramolecular motifs formed through various noncovalent interactions. Intramolecular and intermolecular hydrogen bonds, along with C–H⋯π interactions, play crucial roles in the molecular packing of the structure. The carbonyl O1 atom participates in an intramolecular C–H⋯O interaction with the aromatic H1A atom, forming an S(5) supramolecular pseudo ring. Additionally, the hydroxyl (–OH) group involves an S(6) ring interaction with the aromatic H9A (–CH2) atom in the aliphatic chain.
In the crystal lattice, self-assembly is governed by strong, directional intermolecular O–H⋯N hydrogen bonds and dihydrogen (H⋯H) contacts. The carbonyl O1 atom forms an O2–H2⋯N2 hydrogen bond, generating an infinite one-dimensional zigzag chain along the crystallographic b-axis (Fig. 4, Table 3). This chain is further reinforced by H2⋯H15B dihydrogen interactions. Notably, the hydroxyl (–OH) group serves as both a hydrogen bond donor and acceptor, interconnecting the 1D chains via O–H⋯O interactions, which leads to the formation of a two-dimensional planar sheet, as illustrated in Fig. 4. Fig. 5 further illustrates the 2-D planar sheet architecture formed by the O–H⋯O hydrogen bonds.
![]() | ||
| Fig. 4 An infinite 1D chain is present in the crystal packing of the molecular fragments along the b-axis. | ||
| Type | D–H | H⋯A | D⋯A | D–H…A | Symmetry |
|---|---|---|---|---|---|
| C1–H1A⋯O1 | 0.930(4) | 2.418(3) | 2.743(3) | 100.4(2) | Intra |
| C1–H1B⋯O3 | 0.970(3) | 2.328(8) | 3.067(8) | 132.4(3) | Intra |
| C14–H14A⋯O1 | 0.970(8) | 2.376(5) | 3.190(5) | 141.2(5) | Intra |
| O1–H1⋯O2 | 0.818(13) | 1.881(15) | 2.673(3) | 163(3) | x, 1/2 − y, −1/2 + z |
| O1–H2⋯N2 | 0.822(14) | 1.967(12) | 2.777(2) | 168(2) | 1 − x, −1/2 + y, 3/2 − z |
Furthermore, weak C–H⋯π interactions offer additional stabilization for crystal packing. Specifically, the C4–H4B⋯π and C14–H14B⋯π interactions facilitate parallel stacking with adjacent molecules located at symmetric positions (−x, −y, 1 − z) and (x, 1/2 − y, 1/2 + z), respectively, reinforcing the total stability of the crystal structure (Fig. S25 provided in the SI Material).
![]() | ||
| Fig. 6 (a) and (b) Hirshfeld surface of the compound mapped over the dnorm surface, (c) shape index mapped surface, (d) 2D fingerprint plots of the compound. | ||
In the shape index surface, a blue-colored bulged area near the –CH atom of the phenyl ring and the –CH2 group, along with a red-colored dip on the ring system, corresponds to the weak C–H⋯π interactions seen in the crystal packing analysis (Fig. 6c). These interactions further contribute to the total stabilization of the molecular assembly in the crystal lattice.
Additionally, the horn-shaped symmetrical pattern observed at de + di ≈ 3.0 Å corresponds to H⋯Cl contacts, accounting for 14.8% of the Hirshfeld surface. The presence of C⋯H interactions, attributed to weak C–H⋯π interactions, is marked by two symmetrical wings at de + di ≈ 2.8 Å, contributing 14.1% of the total surface area. Notably, two symmetrical wings observed at de + di ≈ 2.9 Å in the (de, di) bin correspond to H⋯O/O⋯H contacts, contributing 7.3% and confirming the presence of O–H⋯O hydrogen bonds, as previously identified in the structural analysis.
Favored and disfavored contacts from a chemical element point of view can be emphasized via the enrichment ratios, as shown in Table 4. The di-hydrogen interaction (EHH = 0.94) is slightly underrepresented, indicating a minor tendency for self-interaction. An ECH value of 1.12 clearly indicates favorable interactions, likely due to aromatic C–H⋯π stacking interactions. N–N interactions are avoided, whereas N–H contacts are enriched because of hydrogen bonding (ENH = 1.18). Oxygen prevents self-interaction but favors hydrogen bonding via O–H⋯O interactions (EOH = 1.25). Overall, hydrogen bonding (O–H, N–H, and C–H) remains a key interaction mechanism in the crystal packing of a given organic structure.15
| H | C | N | O | Cl | |
|---|---|---|---|---|---|
| a (The enrichment ratios were not computed when the ‘random contacts’ were lower than 0.9%, as they are not meaningful.). | |||||
| H | 55.4 | Actual contacts | |||
| C | 14.1 | 0.9 | |||
| N | 6.8 | 0.4 | 0 | ||
| O | 7.3 | 0 | 0.3 | 0 | |
| Cl | 14.8 | 0 | 0 | 0 | 0 |
| Sx | 76.9 | 8.15 | 3.75 | 3.8 | 7.4 |
| H | 59.14 | Random contacts | |||
| C | 12.53 | 0.66 | |||
| N | 5.77 | 0.61 | 0.14 | ||
| O | 5.84 | 0.62 | 0.29 | 0.14 | |
| Cl | 11.38 | 1.21 | 0.56 | 0.56 | 0.55 |
| H | 0.94 | Enrichment ratio | |||
| C | 1.12 | — | |||
| N | 1.18 | — | — | ||
| O | 1.25 | — | — | — | |
| Cl | 1.30 | 0.00 | — | — | — |
To further evaluate structural consistency, the bond lengths and bond angles of molecule 4c were compared with the average values derived from 14 CSD structures sharing the common (2-butyl-4-chloro-1H-imidazol-5-yl)methanol framework. Overall, the bond lengths and angles in 4c were largely consistent with the CSD averages, with most values falling within the respective standard deviations. However, a notable deviation was observed in the C11–O2 bond length, which differed significantly from the mean (Fig. 9 and SI Table 1). This deviation can be attributed to structural variations among the CSD entries. Specifically, for the molecules QAKXAF, YALXOZ, PAJQUN, ILOPOO, OCAHAC, and VURTIL, the C11–O2 bond length ranges from 1.353 Å to 1.440 Å. These structures predominantly feature a methanol group as the terminal side chain, except for VURTIL, which contains a formaldehyde group. In contrast, the remaining CSD structures that possess a formaldehyde moiety exhibit shorter C11–O2 bond lengths, typically in the range of 1.193 Å to 1.219 Å. The elongation of the C–O bond in methanol-containing structures can be attributed to the electron-donating hydroxyl group, which introduces greater electron density and potential hydrogen bonding interaction factors known to lengthen the C–O bond.17,18 Similarly, significant variation was observed in the C10–C11–O2 bond angle, with a standard deviation of 6.52° across the 14 CSD structures (Fig. 10 and SI Table 2). For molecules such as VURTIL, ILOPOO, YALXOZ, OCAHAC, PAJQUN, QAKXAF, and ILOPOO01, the bond angles ranged from 111.79° to 115.77°. These molecules primarily contain a methanol terminal group, except for VURTIL. In contrast, structures with a formaldehyde side chain displayed larger C10–C11–O2 bond angles, varying from 124.10° to 127.74°. The discrepancy in bond length and angle for VURTIL, despite its formaldehyde moiety, may be influenced by π–π stacking interactions within the crystal lattice, which can locally alter the molecular conformation and geometry.19
The torsion angles were measured for both the methanol/formaldehyde side chains and the butyl chains across all 14 CSD structures (SI Table 3). The methanol side chains in molecules such as ILOPOO, YALXOZ, OCAHAC, PAJQUN, QAKXAF, and ILOPOO01 predominantly adopt a ± sc ± ac (synclinal/anticlinal) conformation, and molecule 4c follows the same pattern. In contrast, the formaldehyde side chains in molecules FEFNOU, FEFNUA, HAMRET, ILOPUU, ILOPUU01, KAMPUK, and WAKJEY exhibit a ± sp ± ap (synperiplanar/antiperiplanar) conformation. Interestingly, the formaldehyde side chain of VURTIL deviates from this trend and adopts a ± sc ± ac conformation, similar to that observed in methanol-bearing molecules. This deviation aligns with the bond length and bond angle anomalies observed for VURTIL, further suggesting that its geometry may be influenced by local structural factors, such as π–π stacking interactions within the crystal lattice. In the case of the butyl side chains, the molecules ILOPOO, ILOPOO01, FEFNUA, and KAMPUK adopt a ±ap ± sp ± ap ± ap (antiperiplanar/synperiplanar/antiperiplanar/antiperiplanar) conformation, a pattern that is also observed in molecule 4c. In contrast, the butyl chain of FEFNOU exhibited a ±ap ± sp ± ap–ac conformation, whereas VURTIL adopted a ±ap ± sp + sc ± ap arrangement. The HAMRET and QAKXAF molecules both display an -ac + sc ± ap ± ap conformation. ILOPUU and ILOPUU01 share a −ac + sc + sc + sc configuration, whereas OCAHAC adopts a +ac–ac ± ap–sc conformation. PAJQUN displays a +sc–ac + sc ± ap conformation, and both WAKJEY and YALXOZ exhibit a +sc-ac ± ap ± ap pattern. These diverse torsional conformations reflect the sensitivity of the butyl chains to local intramolecular and intermolecular environments. In particular, C–H⋯π interactions appear to be a key factor influencing the observed conformational variability of the butyl moieties.
The optimized energies of compounds 4a–f are depicted in Table 6, which indicates their relative stability. Among all the structures, 4f has the highest optimized energy (−1802.380 hartree), suggesting that it is the most stable configuration. Conversely, 4c has the lowest optimized energy (−1382.094 hartree), implying that it is the least stable among the given structures.
| Name | 4a | 4b | 4c | 4d | 4e | 4f |
|---|---|---|---|---|---|---|
| Optimized energy | −1457.322 | −1547.329 | −1382.094 | −1496.437 | −1442.034 | −1802.380 |
| Ehomo (hartree) | −0.22789 | −0.2392 | −0.2291 | −0.22968 | −0.23295 | −0.23345 |
| Elumo (hartree) | −0.02612 | −0.10967 | −0.02688 | −0.06277 | −0.03803 | −0.04009 |
| Ehomo (eV) | −6.2012 | −6.5090 | −6.2341 | −6.2499 | −6.3389 | −6.3525 |
| Elumo (eV) | −0.7108 | −2.9843 | −0.7314 | −1.7081 | −1.0348 | −1.0909 |
| Energy gap | 5.4904 | 3.5247 | 5.5022 | 4.5419 | 5.3040 | 5.2616 |
| Ionization energy (I) | 6.2012 | 6.5090 | 6.2341 | 6.2499 | 6.3389 | 6.3525 |
| Electron affinity (A) | 0.7108 | 2.9843 | 0.7314 | 1.7081 | 1.0348 | 1.0909 |
| Electronegativity(χ) | 3.4560 | 4.7466 | 3.4828 | 3.9790 | 3.6869 | 3.7217 |
| Chemical potential (μ) | −3.4560 | −4.7466 | −3.4828 | −3.9790 | −3.6869 | −3.7217 |
| Global hardness (η) | 2.7452 | 1.7623 | 2.7513 | 2.2709 | 2.6520 | 2.6308 |
| Global softness (s) | 0.3643 | 0.5674 | 0.3635 | 0.4403 | 0.3771 | 0.3801 |
| Electrophilicity index (ω) | 2.1754 | 6.3922 | 2.2043 | 3.4859 | 2.5628 | 2.6325 |
The highest occupied molecular orbital (HOMO) electron density is largely concentrated on the phenyl ring and hydroxyl (–OH) groups, highlighting their strong electron-donating nature. This localization suggests that these regions are highly reactive toward electrophilic attack. In contrast, the lowest unoccupied molecular orbital (LUMO) is distributed mainly over electronegative atoms such as nitrogen and oxygen, indicating their susceptibility to nucleophilic interactions. The charge separation between the HOMO and LUMO underscores a significant charge transfer capability, further supporting the observed stability and reactivity trends of the molecule.
The ΔE values of the FMOs of compounds 4a–f are depicted in Figs. 11(a) to 16(a). The wide HOMO–LUMO energy gap signifies high chemical stability and low reactivity. Additionally, the molecular electronegativity (χ) suggests a tendency to attract and donate electrons. The high ionization energy (I) reinforces the molecule's stability, whereas the electrophilicity index (ω) specifies a potential to accept electrons in chemical interactions.
The HOMO–LUMO energy gap (ΔE) manifests the electronic stability and reactivity. 4a (5.4904 eV), 4c (5.5022 eV), 4e (5.3040 eV), and 4f (5.2616 eV) have relatively large energy gaps, suggesting greater stability and lower chemical reactivity. On the other hand, 4b (3.5247 eV) has the smallest energy gap, indicating higher reactivity due to its greater electron acceptor–donor interaction potential.
The ionization energy (I) and electron affinity (A) values provide insights into the ability of these compounds to donate or accept electrons. 4b has the highest electron affinity (2.9843 eV), making it the most electrophilic and highly reactive among the given structures. The electronegativity (χ) trend follows a similar pattern, with 4b having the highest value (4.7466 eV), indicating its strong tendency to attract electrons. The chemical potential (μ) values indicate that 4b (−4.7466 eV) is the most reactive, whereas 4c (−3.4828 eV) is the least reactive.
The global hardness (η) and global softness (s) parameters indicate the resistance to electron transfer. 4a, 4c, 4e, and 4f exhibit higher hardness values (>2.6 eV), confirming their stability, whereas 4b (1.7623 eV) is the softest and most reactive. The electrophilicity index (ω) further supports this trend, with 4b having the highest value (6.3922 eV), suggesting that it is the most electrophilic and prone to reactions. The whole chemical reactive parameters of compounds (4a–f) are provided in Table 6.
From the MEP analysis (Fig. 11(b)–16(b)), the regions exhibiting negative electrostatic potential correspond to sites susceptible to nucleophilic attack, which are predominantly located on the hydroxyl groups and nitrogen atoms. These areas demonstrate a strong propensity for electron donation during chemical reactions. The hydroxyl group (−0.03724 a.u.) and the phenyl ring (−0.01467 a.u.) also display moderate nucleophilicity, indicating their potential role in charge transfer processes. On the other hand, the electrophilic sites, indicated by positive electrostatic potential values, are concentrated on the –CH and –CH2 groups (0.01885 and 0.02190 a.u.) as well as the hydroxyl (–OH) groups (0.03740 a.u). These regions serve as electron-deficient centers, making them more prone to nucleophilic attack. The MEP visualization confirms that the molecular surface has a well-defined electrostatic distribution, which influences intermolecular interactions, including hydrogen bonding and charge transfer reactions.
In all the compounds, the –OH group is red in color, indicating a high electron density region, making it a potential site for electrophilic attack. Conversely, the nitrogen atom in the five-membered ring appears blue, indicating a low electron density region, which suggests its tendency to attract electrons and act as an electrophile.
Additionally, the O–CH3 group is yellow in color, indicating moderate electron density, which may contribute to electron-donating or electron-withdrawing effects depending on the molecular environment. The C–H group appears light blue, suggesting slightly electron-deficient regions, making it less reactive than the other functional groups.
A unique feature observed in the MEP map is that halogens exhibit dual-color behavior, indicating an anisotropic charge distribution. This means that halogens can function as both nucleophiles and electrophiles, depending on their molecular interactions and external influences.
| Sl. no | Compound name | Docking score | Glide score | Interactions |
|---|---|---|---|---|
| 1 | 4a | −5.161 | −5.168 | H-bond: ALA 354, and LYS 511 |
| Polar: HIE 513, GLN 281, HIS 383, HIS 387 | ||||
| Halogen: H2O | ||||
| Hydrophobic: VAL 379, VAL 380, PHE 457, PHE 527 | ||||
| 2 | 4b | −5.508 | −5.515 | H-bond: TYR 520 |
| Salt bridge: ASP 415 | ||||
| Polar: HIE 513, GLN 281, HIS 353, HIS 383 | ||||
| Hydrophobic: VAL 380, VAL 379, PHE 457, VAL 518, TYR 523 | ||||
| 3 | 4c | −5.826 | −5.833 | H-bond: TYR 520 |
| Pi–Pi stacking: HIS 383, HIS 387, HIS 353 | ||||
| Hydrophobic: VAL 518, TYR 523, VAL 379, VAL 380, and PHE 527 | ||||
| 4 | 4d | −6.311 | −6.318 | H-bond: LYS 511, GLN 281, and TYR 520 |
| Hydrophobic: ALA 354, ALA 356, VAL 380, and VAL 379 | ||||
| 5 | 4e | −5.177 | −5.185 | H-bond: TYR 520, LYS 511, GLU 384 |
| Halogen: H2O | ||||
| Pi–Pi stacking: HIS 383 | ||||
| Hydrophobic: ALA 354, VAL 518, ALA356, TYR 523, VAL 379, VAL 380, PHE 457, PHE 527 | ||||
| 6 | 4f | −6.218 | −6.226 | H-bond: LYS 511 and GLN 281 |
| Polar: HIE 513, HIS 383 | ||||
| Hydrophobic: ALA 354, ALA 356, TYR 523, PHE 527 | ||||
| 7 | Lisinopril | −14.284 | −14.417 | H-bond: GLU 376, GLU 162, TYR 523, GLU 384 |
| Salt bridge: LYS 511 | ||||
| Polar: HIE 513, GLN 281, HIS 387 | ||||
| Hydrophobic: VAL 518, TYR 520, PHE 512, PHE 457, PHE 527, VAL 380 | ||||
| Metal coordination: Zn-701 |
The protein-ligand RMSD plots in Fig. 18 show that for compound 4b, a brief equilibration phase was observed, marked by slight fluctuations up to 50 ns. The RMSD stabilized between 50 and 150 ns, indicating that the ligand successfully achieved a binding conformation that was well suited to the protein's active site. During the final segment of the simulation (150–200 ns), only minor variations were noted, with the RMSD consistently remaining within the range of 1–3 Å. These findings underscore the potential of compound 4b as a leading candidate for further optimization.
In the case of compound 4c, the system displayed slight fluctuations between 0 and 30 ns, indicating an initial equilibration phase. The RMSD exhibited variability from 150 to 170 ns, indicating a period of moderate stability characterized by minimal oscillations. However, complete equilibration was not attained during the final phase of the simulation.
In the case of compound 4d, minor fluctuations were noted up to 25 ns, indicating an initial equilibration period. However, between 25 and 100 ns, significant differences emerged in the RMSD of both the ligand and the protein, suggesting that the ligand had shifted from its original binding position. Notably, the system appeared to stabilize at 100 ns, with the protein–ligand RMSD remaining constant until approximately 150 ns. After this point, fluctuations resumed, which may indicate either temporary binding interactions or ongoing conformational changes. These observations suggest that compound 4d may exhibit less stable binding characteristics.
For lisinopril, the early stabilization phase was characterized by slight changes in both the ligand and protein RMSD up to 50 ns. The system tended to reach equilibrium at 125 ns, with only minor deviations observed for the remainder of the simulation period. These findings indicate that lisinopril continues to bind to the target protein in a stable manner, reinforcing its role as a proven inhibitor. In contrast, compound 4b demonstrated rapid equilibration, achieving stability by 50 ns and maintaining that stability until 150 ns. The minor variations observed toward the end of the simulation indicate a consistent binding conformation for compound 4b. In comparison, lisinopril exhibited greater fluctuations, with RMSD deviations persisting until approximately 70 ns. Following this, the equilibration process commenced at approximately 125 ns and continued until the simulation's conclusion was reached. When comparing compound 4b with lisinopril, which requires a longer time to stabilize, it is clear that compound 4b exhibited faster equilibration and more robust stability. The minimal changes noted for compound 4b toward the end of the simulation further support the notion of a well-preserved binding conformation. The ligand–protein RMSD plots are depicted in Fig. 18.
To evaluate the flexibility and mobility of various amino acid residues in the presence of compounds 4b, 4c, and 4d and the standard lisinopril, we concentrated on the protein root mean square fluctuation (RMSF), as shown in Fig. 19. The findings indicate that a higher RMSF value is associated with increased flexibility, whereas a lower RMSF value signifies greater rigidity in the amino acid residues, and a green dash highlights the flexibility and mobility of amino acid residues in contact with compounds.
The protein RMSF plot reveals that compound 4b has higher RMSF values of 3.6, 3.8, and 4.0 Å, indicating enhanced flexibility among amino acid residues 120, 250, and 400, respectively. Conversely, an RMSF value below 0.5 Å implies that the amino acid residues exhibit greater rigidity. Similarly, amino acid residues affected by compound 4c presented increased protein RMSF values of 2.8 and 3.2 Å for residues 110 and 400, with the lowest protein RMSF observed at 0.4 Å. Compound 4d presented elevated protein RMSF values of 2.3 and 2.5 Å, alongside a minimal RMSF of 0.4 Å. The standard lisinopril exhibited greater flexibility in the amino acid residues, as evidenced by the higher protein RMSF values of 2.6, 2.4, and 4.5 Å, with the least flexibility occurring below 0.5 Å.
In summary, the results clearly indicate that amino acid residues are more flexible and mobile in compound 4b and lisinopril. The protein RMSF plots of compounds 4b, 4c, and 4d and lisinopril are depicted in Fig. 19.
As shown in Fig. 20, the contacts between proteins and ligands were analyzed. The data revealed that the standard lisinopril presented the greatest number of protein–ligand contacts, outperforming the other compounds. Like lisinopril, compound 4b demonstrated a significant number of contacts, whereas compound 4c exhibited moderate levels of contacts. In contrast, among the compounds evaluated, compound 4d presented the lowest number of protein–ligand contacts.
The protein–ligand contact histograms for compounds 4b, 4c, and 4d and lisinopril are presented in Fig. 21. This histogram effectively shows the various types of protein–ligand interactions, including hydrogen bonds (indicated in green), hydrophobic interactions (in gray), ionic interactions (in pink), water bridges (in blue), and halogen bond formations (in yellow). A value of 0.1 on the y-axis signifies that a specific interaction occurs for 10% of the simulation time; similarly, 0.2 corresponds to 20%, 0.5 to 50%, and 1.0 to 100%. These findings clearly indicate that lisinopril formed a greater number of interactions, followed by compound 4b, whereas compounds 4c and 4d displayed relatively similar interaction profiles.
| Compound | Molecular weight | Donor hydrogen bond | Acceptor hydrogen bond | Qp log Po/W |
nviolations | QP log PC16 |
QP log Poct |
|---|---|---|---|---|---|---|---|
a Qp log Po/W: octanol/water partition coefficient, QP log PC16: hexadecane/gas partition coefficient, QP log Poct: octanol/gas partition coefficient, n violations: violations from Lipinski's rule. |
|||||||
| Rule | <500 | ≤5 | ≤10 | ≤5 | ≤1 | 4–18 | 8–43 |
| 4a | 338.83 | 2 | 5.6 | 3.533 | 0 | 11.135 | 16.263 |
| 4b | 353.80 | 2 | 5.9 | 2.79 | 0 | 11.64 | 17.669 |
| 4c | 322.83 | 2 | 4.9 | 3.775 | 0 | 10.854 | 15.734 |
| 4d | 358.87 | 2 | 4.9 | 4.354 | 0 | 12.623 | 17.64 |
| 4e | 343.25 | 2 | 4.9 | 3.818 | 0 | 11.033 | 15.818 |
| 4f | 326.80 | 2 | 4.9 | 3.57 | 0 | 10.052 | 15.413 |
:
3) with visualization in a UV cabinet. 1H and 13C NMR spectra were recorded on a Bruker AM 400 MHz spectrometer, TMS was employed as an internal standard, DMSO-d6 was used as a solvent for NMR analysis, and a Waters ARC with a 2998 SQ detector 2 was used to collect the ESI + m/z fragments.
C), 1675 (C
N), 2833 (aliphatic C–H), 2954 (aromatic C–H), 3181 (OH). 1H NMR (400 MHz, DMSO-d6, ppm): δ 0.88 (t, 3H, J = 8.0 Hz), δ 1.26–1.35 (m, 2H), δ 1.47–1.60 (m, 2H), δ 2.36–2.47 (m, 2H), δ 3.75 (s, 3H), δ 3.97–4.05 (m, 2H), δ 4.34–4.44 (m, 2H), δ 4.83–4.87 (m, H), δ 5.18 (t, H, J = 4.0 Hz), δ 5.70 (d, H, J = 4.0 Hz). δ 6.92 (d, 2H, J = 8.0 Hz), δ 7.24 (d, 2H, J = 8.0 Hz). 13C NMR (100 MHz, DMSO-d6, ppm): δ 14.22, 22.33, 26.15, 29.43, 51.72, 51.88, 55.57, 71.85, 114.06, 125.47, 126.02, 127.57, 135.14, 148.11, 159.21. Molecular formula: [C17H23ClN2O3]. ESI mass (m/z): 339.47 (M + H)+.
C), 1662 (C
N), 2873 (aliphatic C–H), 2957 (aromatic C–H), 3162 (OH). 1H NMR (400 MHz, DMSO-d6, ppm): δ 0.85 (t, 3H, J = 8.0 Hz), δ 1.23–1.33 (m, 2H), δ 1.45–1.59 (m, 2H), δ 2.40–2.46 (m, 2H), δ 4.03–4.12 (m, 2H), δ 4.40–4.49 (m, 2H), δ 5.08–5.12 (m, H), δ 5.25 (t, H, J = 4.0 Hz), δ 6.08 (d, H, J = 4.0 Hz). δ 7.63 (d, 2H, J = 8.0 Hz), δ 8.23 (d, 2H, J = 8.0 Hz). 13C NMR (100 MHz, DMSO-d6, ppm): δ 14.21, 22.29, 26.17, 29.42, 51.04, 51.93, 71.60, 123.89, 125.66, 125.96, 127.79, 147.42, 148.28, 150.87. Molecular formula: [C16H20ClN3O4]. ESI mass (m/z): 354.29 (M + H)+.
C), 1681 (C
N), 2870 (aliphatic C–H), 2956 (aromatic C–H), 3304 (OH). 1H NMR (400 MHz, DMSO-d6, ppm): δ 0.85 (t, 3H, J = 8.0 Hz), δ 1.25–1.34 (m, 2H), δ 1.48–1.58 (m, 2H), δ 2.29 (s, 3H), δ 2.39–2.48 (m, 2H) δ 3.97–4.06 (m, 2H), δ 4.33–4.45 (m, 2H), δ 4.84–4.88 (m, H), δ 5.17 (t, H, J = 4.0 Hz), δ 5.71 (d, H, J = 4.0 Hz), δ 7.14–7.22 (m, 4H). 13C NMR (100 MHz, DMSO-d6, ppm): δ 14.22, 21.19, 22.32, 26.14, 29.42, 51.68, 51.88, 72.10, 125.46, 126.01, 126.31, 129.19, 137.08, 140.18, 148.14. Molecular formula: [C17H23ClN2O2]. ESI mass (m/z): 323.27 (M + H)+.
C), 1634 (C
N), 2930 (aliphatic C–H), 3050 (aromatic C–H), 3328 (OH). 1H NMR (400 MHz, DMSO-d6, ppm): δ 0.82 (t, 3H, J = 8.0 Hz), δ 1.20–1.29 (m, 2H), δ 1.41–1.57 (m, 2H), δ 2.41–2.47 (m, 2H), δ 4.10–4.20 (m, 2H), δ 4.40–4.51 (m, 2H), δ 5.08–5.12 (m, H), δ 5.24 (t, H, J = 4.0 Hz), δ 5.94 (d, H, J = 4.0 Hz). δ 7.49–7.55 (m, 3H), δ 7.87–7.93 (m, 4H). 13C NMR (100 MHz, DMSO-d6, ppm): δ 14.18, 22.29, 26.22, 29.42, 51.54, 51.95, 72.42, 124.80, 124.95, 125.56, 126.03, 126.36, 126.70, 128.04, 128.26, 128.27, 133.02, 133.24, 140.74, 148.22. Molecular formula: [C20H23ClN2O2]. ESI mass (m/z): 359.47 (M + H)+.
C), 1662 (C
N), 2867 (aliphatic C–H), 2956 (aromatic C–H), 3268 (OH). 1H NMR (400 MHz, DMSO-d6, ppm): δ 0.87 (t, 3H, J = 8.0 Hz), δ 1.24–1.34 (m, 2H), δ 1.45–1.59 (m, 2H), δ 2.35–2.48 (m, 2H), δ 4.0–4.08 (m, 2H), δ 4.36–4.46 (m, 2H), δ 4.92–4.96 (m, H), δ 5.21 (t, H, J = 4.0 Hz), δ 5.87 (d, H, J = 4.0 Hz). δ 7.35 (d, 2H, J = 8.0 Hz), δ 7.42 (d, 2H, J = 8.0 Hz). 13C NMR (100 MHz, DMSO-d6, ppm): δ 14.22, 22.32, 26.14, 29.41, 51.37, 51.89, 71.55, 125.57, 125.98, 128.32, 128.65, 132.52, 142.13, 148.19. Molecular formula: [C16H20Cl2N2O2]. ESI mass (m/z): 343.26 (M + H)+.
C), 1600 (C
N), 2864 (aliphatic C–H), 2963 (aromatic C–H), 3272 (OH). 1H NMR (400 MHz, DMSO-d6, ppm): δ 0.88 (t, 3H, J = 8.0 Hz), δ 1.26–1.35 (m, 2H), δ 1.49–1.60 (m, 2H), δ 2.41–2.49 (m, 2H), δ 3.99–4.09 (m, 2H), δ 4.35–4.46 (m, 2H), δ 4.91–4.96 (m, H), δ 5.21 (t, H, J = 4.0 Hz), δ 5.84 (d, H, J = 4.0 Hz). δ 7.17–7.22 (m, 2H), δ 7.35–7.38 (m, 2H). 13C NMR (100 MHz, DMSO-d6, ppm): δ 14.22, 22.31, 26.14, 29.43, 51.53, 51.88, 71.57, 115.33, 125.54, 125.99, 128.34, 139.34, 148.17, 160.85, 163.26. Molecular formula: [C16H20ClFN2O2]. ESI mass (m/z): 327.29 (M + H)+.
:
3) and vacuum-filtered using a Buchner funnel. Re-extraction was performed until the solid residue turned colorless. This residue was air-dried and sifted to obtain a uniform powder, which was stored in an airtight container at 0–4 °C.
000g for 60 minutes at 4 °C. The supernatant was collected and dialyzed against the same buffer without Triton X-100 (3 L, the buffer was changed 3 times) for 24 hours. This extract was stored at −20 °C until further use. The specific activity was 0.8234 units per mg of protein.| % inhibition = 100 − (Atest/Acontrol) × 100, |
Nonhydrogen atoms were identified on the basis of electron density and refined anisotropically to enhance structural accuracy. The hydrogen atoms were positioned geometrically on their respective parent atoms and refined using a riding model to ensure optimal placement.28 The geometrical parameters of the crystal structure were calculated using PLATON,29 and packing diagrams were generated using MERCURY 4.2.0.30 The crystal structure data and refinement parameters are summarized in Table 2.
In addition, 2D fingerprint plots derived from the dnorm surface help to evaluate individual molecular contacts quantitatively. The enrichment ratio (EXX) values derived from the Hirshfeld surface quantify the propensity of atomic contacts in crystal structures. The propensity of each contact was explored by calculating the enrichment ratio on the basis of the obtained actual contacts and derived random contacts.33,34
Pairwise intermolecular interaction energies were calculated for fragments within a 3.8 Å radius around the central independent fragment. The energy frameworks were constructed using these interaction energy values, allowing visualization of the 3D topology of the molecular structure.35–37
CCDC 2448889 contains the supplementary crystallographic data for this paper.39
The data underlying this study are available in the published article and its SI. See DOI: https://doi.org/10.1039/d5ra04675k.
| This journal is © The Royal Society of Chemistry 2025 |