Ranjana Aggarwal*ab,
Manisha Sharma
a,
Garima Sumran
c and
Parvin Kumar
a
aDepartment of Chemistry, Kurukshetra University, Kurukshetra-136119, Haryana, India
bCouncil of Scientific and Industrial Research-National Institute of Science Communication and Policy Research, New Delhi 110012, India. E-mail: ranjana67in@yahoo.com; ranjanaaggarwal67@gmail.com; Tel: +91-9896740740
cDepartment of Chemistry, D. A. V. College (Lahore), Ambala City, Haryana 134 003, India
First published on 14th May 2025
A facile and environmentally benign protocol for regioselective synthesis of diversely substituted imidazo[1,2-a]pyrimidines 5a–h has been described via multicomponent reaction of unsymmetrical β-diketones 1, N-bromosuccinimide 2 and 2-aminopyrimidine 4 in DCM. The reaction proceeds through in situ formation of α-bromo-β-diketones 3 and their ensuing condensation with 2-aminopyrimidine without the need of any organic or inorganic catalyst. The structure of the regioisomeric product was characterized by 1H, 13C NMR, heteronuclear 2D NMR and HRMS studies. The present protocol offers several advantages such as avoidance of metal-based and toxic catalysts, broad substrate scope with respect to substitutions on β-diketones, operational simplicity, easy work-up and high yields. Computational molecular docking studies were carried out to examine the interaction of imidazo[1,2-a]pyrimidines with bovine serum albumin (BSA). Moreover, different spectroscopic approaches viz. UV-visible, steady-state fluorescence and competitive displacement assays were carried out to investigate the binding mechanisms of imidazo[1,2-a]pyrimidines (5c, 5e and 5h) with BSA. The results thus obtained revealed that imidazo[1,2-a]pyrimidines showed moderate binding with BSA through a static quenching mechanism and compound 5e had more affinity to bind in site I of BSA.
Serum albumins, major constituent proteins of blood, are a type of biological receptor that are intricate in transport and distribution of endogenous and exogenous chemicals including thyroid hormones, metals, fatty acids, steroids, and other molecules throughout the bloodstream.19 Serum albumins also play a crucial role in maintaining the extracellular fluid by influencing the plasma oncotic pressure.20 The drug–protein interaction significantly influences pharmacokinetic properties such as absorption, delivery, metabolism, and excretion (ADME) and pharmacodynamics properties of drugs.21 In order to achieve the desired therapeutic effect, drugs must bind to serum albumin with optimum binding affinity. Therefore, investigation of the molecular interactions between proteins and small organic molecules and drugs is very crucial to design and develop novel and more efficient drugs with improved plasma solubility/stability.
Bovine serum albumin (BSA) is one of the most extensively researched proteins as primary molecular targets due to its capability to reversibly bind to a wide variety of drugs.22,23 BSA is more widely used as a crucial model protein to study the interaction of small molecules with serum albumins due to its cost effectiveness, wide availability, good stability and 76% structural homology with Human serum albumin (HSA).24 BSA's primary structure contains 583 amino acids with 17 disulfide bonds and a free cysteine residue.25 It has three homologous α-helical domains (I–III), each containing two subdomains A and B.26 BSA exhibits intrinsic fluorescence specifically due to Trp134 and Trp213 residues which are located on the surface of subdomain IB and in the hydrophobic cavity of subdomain IIA, respectively27 and the quenching of this fluorescence by small molecules is used in BSA-ligand interaction studies.
In view of their biological potential, to date different synthetic strategies have been developed for the construction of functionalized imidazo[1,2-a]pyrimidine derivatives.28 The traditional and most commonly used method for the synthesis of imidazo[1,2-a]pyrimidines is Hantzsch-type synthesis which involves the cyclo-condensation of α-functionalized carbonyl compounds with 2-aminopyrimidine in the presence of various catalysts such as neutral Al2O3,29 KI, K2CO3,30 PTSA,31 NaHCO3,32 etc. For α-bromination of 1,3-diketones, N-bromosuccinimide (NBS) was utilized over other reagents e.g. molecular bromine, sodium hypobromite, tetraalkylammonium tribromide, CuBr2 with Koser's reagent, HBr–H2O2, due to its easily accessible, easy to handle, inexpensive, and selective brominating property.33 In recent years, three component reactions of aldehydes, 2-aminopyrimidine and nitriles/isonitriles/alkyne or their derivatives have been reported to produce imidazo[1,2-a]pyrimidines.17,28 Other method involves the reaction between 2-aminopyrimidine, ketone and [hydroxy(tosyloxy)iodo]benzene (HTIB) in one pot.34 More recently, Chanda et al.35 described water and isopropanol mediated catalyst-free microwave assisted protocol for the synthesis of imidazo[1,2-a]pyrimidines which shows progress towards green synthetic methods. However, most of these methods often suffer from a drawback of harsh reaction conditions calling for high temperature, hazardous high boiling organic solvents, long reaction times, decomposition of substrates during overheating and lower yields of the products. The necessity of thermal activation for the imidazo[1,2-a]pyrimidines ring closure is one of the main and most disappointing limitations on the way to medicinal utilization of this heterocyclic motif. Therefore, the development of energy and environment-efficient greener synthetic methodologies for the construction of these versatile bicyclic heterocycles is always in demand.
Synthetic strategy for functionalized imidazo[1,2-a]pyrimidine through a classical Hanztsch reaction between the α-bromo-1,3-diketones and 2-aminopyrimidine is shown in Scheme 1. The reaction of unsymmetrical diketones 1 with 2-aminopyrimidine 4 in the presence of NBS may lead to the formation of two possible regioisomers: 3-aroyl-2-methyl-imidazo[1,2-a]pyrimidines 5 and 3-acetyl-2-aryl-imidazo[1,2-a]pyrimidines 6, by different permutations, as depicted in Scheme 1. Differently substituted acetophenones on reaction with ethyl acetate in the presence of sodium ethoxide were converted into corresponding unsymmetrical 1,3-diketones through Claisen condensation following the literature protocol.
Recently, multicomponent reactions (MCRs) have gained considerable interest as a sustainable approach offering several advantages such as efficiency, selectivity, high atom economy, reduced waste generation and short reaction time.36 Currently our research group is exploring regioselective reaction of substituted unsymmetrical α-bromo-1,3-diketones with various binucleophiles,37 therefore, it was planned to examine study of regioselectivity pattern of the reaction between α-bromo-1,3-diketones and 2-aminopyrimidine to construct the functionalized imidazo[1,2-a]pyrimidine scaffold using MCR approach. To the best of our knowledge, the reaction between 2-aminopyrimidine and unsymmetrical α-bromo-1,3-diketones has not been proposed yet. Intrigued by these findings and in continuation of our ongoing efforts towards the synthesis of fused nitrogen-bridged heterocyclic compounds using green synthetic protocols,37–39 herein we describe a highly efficient, catalyst-free regioselective approach for the synthesis of functionalized imidazo[1,2-a]pyrimidine derivatives involving in situ bromination of unsymmetrical 1,3-diketones by NBS and condensation with 2-aminopyrimidine under mild reaction condition and characterization of regioisomer on the basis of heteronuclear 2D NMR [(1H–13C) heteronuclear multiple bond correlation, HMBC, (1H–13C) heteronuclear multiple quantum coherence, HSQC, (1H–15N) HMBC] spectroscopic studies. The molecular docking studies have been executed to determine interaction mode of imidazo[1,2-a]pyrimidine derivatives with BSA. Binding studies of representative compounds with BSA have been investigated using UV-visible absorption spectroscopy and fluorescence emission spectroscopy.
Entry | Solventa | Temperaturea | Time | Yieldb (%) |
---|---|---|---|---|
a Reaction conditions: unsymmetrical β-diketone (1a, 1 mmol), NBS (2, 1 mmol), 2-aminopyrimidine (4, 1 mmol) and an appropriate solvent (10.0 mL).b Isolated yield.c NR, no reaction.d Domino reaction. | ||||
1 | Acetone | r.t. | 24 h | 25 |
2 | MeOH | Reflux | 6 h | 45 |
3 | EtOH | Reflux | 6 h | 52 |
4 | AcOH | Reflux | 4 h | 48 |
5 | DCM | r.t. | 6 h 45 min | 75 |
6 | DCM | Reflux | 2.5 h | 65 |
7 | MeCN | Reflux | 6 h | 55 |
8 | Toluene | Reflux | 5 h | NRc |
9 | Solvent-free | r.t. | 3 h | Trace |
10 | Solvent-free | 90 °C | 1.5 h | 60 |
11 | DCMd | r.t. | 4 h | 80 |
Further to optimize the reaction condition, we explored the model reaction in different polar protic (MeOH, EtOH, AcOH); polar aprotic (DCM, chloroform, acetone, MeCN) and nonpolar aprotic (toluene) solvents at different temperatures (room temperature up to reflux) to determine the most appropriate reaction conditions (entry 2–8, Table 1). Among the solvents screened, DCM improved the yield significantly to 75% at room temperature (entry 5, Table 1). To further improve the yield, reaction was also carried out in DCM by the gradual increase of temperature from r.t. to refluxing temperature. However, there was no further improvisation of the product yield. In addition, the reaction was performed at r.t. under solvent-free conditions which furnished the desired product in trace amount (entry 9, Table 1). However, on heating under solvent-free conditions, the desired product was obtained but only with 60% yield (entry 10, Table 1).
Having optimized conditions identified for sequential synthesis, we explored further one-pot multicomponent domino process by stirring an equimolar mixture of 2-aminopyrimidine 4, 1-phenylbutane-1,3-dione 1 and NBS 2 in DCM at r.t. The reaction proceeded smoothly to yield the single product in 4 h as indicated by TLC out of two possible regioisomers. It is noteworthy to mention that this domino approach increased the yield (entry 11 vs. 5) and the product was isolated in 80% yield in short reaction time (4 h vs. 6 h 45, entry 11 vs. 5) than the sequential synthesis protocol. Hence, this one-pot MCRs approach was chosen as the model method (Scheme 2).
![]() | ||
Scheme 2 Sequential and multicomponent regioselective synthesis of 3-aroylimidazo[1,2-a]pyrimidines. |
To get an insight into the extensibility and efficiency of the developed protocol, differently substituted unsymmetrical β-diketones containing electron donating and electron withdrawing substituents were reacted with 2-aminopyrimidine under optimized conditions. All the reaction combinations underwent successful condensation and yielded the corresponding products in high yields with high regioselectivity. Generally, β-diketones bearing electron-donating groups provided better yields of products than β-diketones bearing electron-withdrawing groups. Differently synthesized imidazo[1,2-a]pyrimidines are outlined in Table 2.
The successful conversion of reactants into products was confirmed through spectroscopic (1H-NMR, 13C-NMR, IR) data analysis. The IR spectrum of the attained product 5a displayed a sharp absorption band at 1648 cm−1 corresponding to carbonyl group stretch indicating the product formation. The 1H NMR spectrum of compound 5a displayed a singlet of three proton intensity at δ 2.21 ppm corresponding to the methyl group, three signals each as doublet of doublet of one proton intensity at 7.62, 9.02 and 9.65 ppm attributable to H-6, H-5, H-7 protons of pyrimidine nucleus, respectively, and three signals for two, one and two proton intensity at 7.62, 7.73, 7.80 ppm corresponding to phenyl protons. 13C NMR spectrum of compound 5a displayed signals at δ 15.44 and 186.20 ppm corresponding to methyl and carbonyl carbon, along with required number of signals for aromatic carbons, thus validating the successful condensation of 2-aminopyrimidine with α-bromo-β-diketones to give the desired product 5a. Additionally, the mass spectra of 5a exhibited a peak at m/z = 238.0890, representing the molecular ion, [M + H]+, which aligns with the expected composition for the title compound (C14H11N3O).
In order to unambiguously assign the correct regioisomeric structure to reaction product as 5 out of 5 and 6, multinuclear 2D NMR experiments [(1H–13C) HMBC, (1H–13C) HSQC and (1H–15N) HMBC] were carried out for compound 5a. The 2D NMR correlation results and 1H, 13C and 15N chemical shifts for compound 5a are shown in Fig. 2 and Table 3.
![]() | ||
Fig. 2 1H (in red), 13C (in blue) and 15N (in green) NMR chemical shift values and 2D NMR correlation illustration for compound 5a and 5c. |
Chemical shifts (δ in ppm) | gs-HMQC correlation | gs-HMBC correlation | Assignments |
---|---|---|---|
186.20 | — | 7.80 (H2′/H6′) | CO |
156.73 | 9.02 (5H) | 7.62 (H6) | C5 |
9.65 (H7) | |||
146.82 (C9) | |||
147.60 | — | 2.21 (2-CH3) | C2 |
146.82 | — | 9.02 (H5) | C9 |
9.65 (H7) | |||
138.81 | — | 7.62 (H3′/H5′) | C1′ |
138.07 | 9.65 (7H) | 7.62 (H6) | C7 |
9.02 (H5) | |||
146.82 (C9) | |||
133.52 | 7.73 (H4′) | 7.62 (H3′/H5′) | C4′ |
129.38 | 7.62 (H3′/H5′) | 138.81 (C1′) | C3′/C5′ |
129.27 | 7.80 (H2′/H6′) | 7.73 (H4′) | C2′/C6′ |
186.20 (CO) | |||
119.67 | — | 2.21(2-CH3) | C3 |
113.66 | 7.62 (6H) | 9.65 (7H) | C6 |
9.02 (5H) | |||
15.44 | 2.21 (2-CH3) | C2 (147.60) | 2-CH3 |
C3 (119.67) |
The (1H–13C) HMBC spectrum of 5a displayed cross peak of carbonyl carbon at δ 186.20 ppm with H2′/H6′ (δ 7.80) protons of aryl ring confirming the presence of CO with aryl/heteroaryl ring. Similarly (1H–13C) HMBC spectrum showed cross peaks of methyl protons (δ 2.21 ppm) with C-2 (δ 147.60 ppm) and C-3 (δ 119.67 ppm) which indicated the presence of methyl substituent at position-2 of imidazo[1,2-a]pyridine nucleus. Further, (1H–15N) HMBC of compound 5a also showed a cross peak of methyl protons (δ 2.21) with N-1 (δ −215.9), thus confirming the presence of methyl substituent at position-2 of imidazo[1,2-a]pyrimidinyl scaffold. Had the structure been 3-acetyl-2-arylimidazo[1,2-a]pyrimidine 6a, then the correlation between the methyl protons with N-1, C-3 and the cross peak of carbonyl carbon with the H2′/H6′ proton of the aryl part would have been absent. Thus, the structure can certainly be assigned as 1-phenyl-1-(2-methylimidazo[1,2-a]pyrimidin-3-yl)methanone 5a. Similar correlation results of (1H–13C) HMBC and (1H–15N) HMBC were observed for the compound 5c as shown in Fig. 2.
The plausible mechanism for regioselective synthesis of imidazo[1,2-a]pyrimidine is depicted in Scheme 3. In the first step, ring nitrogen of 2-aminopyrimidine attacks nucleophilically on α-bromo-1,3-diketones to give the intermediate that is cyclized by nucleophilic addition of imine nitrogen to either of the carbonyl carbons to form regioisomer 2-alkyl/arylimidazo[1,2-a]pyrimidine 5 or 6 (Route 1). Due to more electrophilic character and less steric hindrance on carbonyl carbon adjacent to methyl, imine nitrogen underwent nucleophilic addition on this carbonyl carbon (Path-a) followed by removal of water to give 2-alkylimidazo[1,2-a]pyrimidine 5 as the final product.
Outcomes from molecular docking simulation studies showed that imidazo[1,2-a]pyrimidine derivatives 5a–h exhibited binding interactions with the BSA protein mainly through hydrophobic interaction as well as through hydrogen bonding, electrostatic and van der Waals interaction with binding free energy in the range of −8.25 to −9.04 kcal mol−1 (Table 4). Among the synthesized imidazo[1,2-a]pyrimidine derivatives, compound 5e with 4-methoxy substitution at phenyl ring binds with the albumin protein (BSA, PDB ID: 4f5s) in AutoDock more efficiently than other derivatives with a maximum docking score of −9.04 kcal mol−1 (Table 4). Results of docking analysis using BIOVIA Discovery Studio Visualizer (DSV) have shown that 5e interacts with BSA in the active pocket of chain A. Compound 5e interacts with various amino acids (Tyr137, Lys136, Pro117, Tyr160, Met184, Ile181, Val188, Leu115, Phe133) through hydrogen bonding and various hydrophobic interactions as shown in the 2D and 3D plots (Fig. 3). Also, compound 5e formed a conventional hydrogen bond with Arg185 via its carbonyl group. The lowest binding free energy conformer of compounds 5a–h, PBZ and IBP with BSA is shown in Fig. 4a–j.
Compounds | Binding energy (kcal mol−1) | Interacted residues |
---|---|---|
a π-donor hydrogen bonding.b π–π T-shaped.c π–π stacked.d Alkyl.e π–alkyl.f van der Waals.g hydrogen bonding.h π–sulfur.i π–sigma.j Carbon–hydrogen bond.k π–cation.l Amide π-stacking. | ||
5a | −8.69 | Tyr160a,b,c,d, Tyr137b,c, Arg185e, Leu115e, Ile181d,e, Pro117d, Met184e, Val188e, Lys136f, Glu140f, Ile141f, Leu122f |
5b | −8.51 | Tyr160a,d, Tyr137b,g, Ile181d, Met184d,h, Arg185d,e, Val188e, Pro117e, Leu115e, Ile141i, Leu189f |
5c | −8.50 | Tyr137a,b, Met184c,h, Arg185k,e, Val188e, Ile181e, Leu115e, Pro117e, Lys114e, Leu122f, Tyr160f, Ile141f |
5d | −8.95 | Tyr160b,d, Met184d, Arg185d,e,g,i, Ile181d,e, Pro117e, Leu115e, Tyr137f, Ile141f, Val188f, Glu182f, Glu140f, Lys136f |
5e | −9.04 | Tyr137a,b, Lys136d, Pro117d, Tyr160b,d, Met184e, Ile181d,e, Arg185e,g, Val188e, Leu115e, Phe133j |
5f | −8.76 | Tyr160a,d, Tyr137c, Met184d, Ile181d,e, Arg185d,e,g,k, Pro117e, Glu140j, Lys136j, Leu115d,i, Leu122f, Ile141f, Glu182f |
5g | −8.28 | Tyr160a,c,d, Tyr137a,c,d, Val188e, Met184e, Arg185e,g, Ile181d,e, Ile141d, Pro117d,e, Leu115d,i, Leu122f, Glu140f |
5h | −8.25 | Tyr160b,g,h,l, Arg185e,g, Ile181d,e, Val188e, Tyr137b,h, Met184h, Leu115e, Pro117d,e |
PBZ | −8.81 | Tyr137a, Tyr160a, Pro117d, Ile181d, Val188e, Arg185e, Glu140f, Met184f, Leu178f, Glu182f, Leu115i, Ile141i |
IBP | −7.46 | Leu115d,e, Ile181d, Arg185d,g,k, Lys114d,j, Pro117e, Glu182f, Tyr160f, Tyr137f, Arg144f |
![]() | ||
Fig. 4 (a–j) 2D diagram showing of the interaction of ligand 5a–h, phenylbutazone (PBZ) and ibuprofen (IBP) with BSA protein. |
Docking analysis showed that almost all the derivatives exhibit strong interactions with the A chain of the protein with few exceptions where derivatives 5a and 5b were found to interact with the protein through chain B. Compound 5e was selected for further in vitro analysis with BSA protein using various spectroscopic techniques due to the best docking score. To compare the results, two compounds bearing substituents of different nature, 5c with 4-chlorophenyl moiety and 5h with 2-thienyl group, were chosen.
Among the synthesized compounds, 5e having electron donating 4-OMePh ring, 5h having 2-thienyl ring and 5c having electron withdrawing 4-ClPh ring were selected for UV-visible spectral studies with BSA. Absorption spectra of BSA were recorded in the absence and presence of 5e, 5h, and 5c at a varying concentration ranging from 0 to 40, 0 to 28 and 0 to 36 μM, respectively, in the wavelength range of 250–350 nm at room temperature. The UV-visible absorption spectra of BSA display characteristic peaks at about 278 nm, which could result from π–π* transition due to the presence of three aromatic amino acids (tryptophan, tyrosine and phenylalanine) on the protein chain surface.42 In the presence of increasing concentration of 5e, 5h and 5c, the intensity of the BSA spectra continues to rise due to the changes in the environment surrounding aromatic amino acids Trp, Phe and Tyr (Fig. 5a–c). This perturbation in the absorption spectra of BSA indicates the preferable interactions between BSA and compounds studied. The hyperchromic effect in the absorption spectra of BSA on subsequent addition of 5e, 5h and 5c indicated a possible static interaction between BSA and imidazo[1,2-a]pyrimidine ligands.
The stoichiometric ratio in which BSA protein interacts with a ligand was determined employing Job's method of continuous variation. Ten different solutions were prepared with different volumes of BSA and 5e fixed at a constant concentration of 15 μM. In Job's plot, the absorbance at λmax (278 nm) was measured and absorbance versus mole fraction of ligand 5e was plotted (Fig. 5d). The maxima in Job's plot was observed at 0.5 on the axis which indicates 1:
1 binding stoichiometry between BSA and ligand 5e.43
Further to understand the interaction between ligand 5e and BSA, a fixed concentration of compound 5e (0.1 mM) was titrated with increasing BSA concentration (0–40 μM) at room temperature (25 °C) and the effects on emission profiles was monitored following excitation at 367 nm (Fig. 6d). Free ligand (5e) exhibits maximum in absorption peak at 367 nm. Therefore, the emission spectra were recorded in the spectral range of 387–700 nm by fixing the excitation wavelength at 367 nm. Ligand 5e exhibited a strong fluorescence emission at 454 nm in the absence of BSA. As expected, emission intensity increases at 481 nm as the concentration of BSA is increased (Fig. 6d). The emission increases linearly as a function of BSA concentration until about 28 μM, where it reaches a plateau that may be attributed to a self-quenching phenomenon.44 Red shift in emission maxima was observed in the presence of BSA.
The static and dynamic quenching process can be identified by calculating the quenching constant Kq using the Stern–Volmer equation45 (eqn (1)).
![]() | (1) |
In order to find the effectiveness of binding of compounds with BSA, fluorescent data were utilized to compute quenching constant (Kq) and Stern–Volmer constant (KSV) by plotting emission spectral data at λmax (343 nm) at room temperature by employing the Stern–Volmer equation (eqn (1)). Based on the linear fit graph of fluorescence intensity (Fo) of BSA in the absence of ligand to the fluorescence intensity (F) of BSA in the presence of ligand at varying concentration was plotted against increasing concentration of ligand [Q] (Fig. 7), quenching constant (Kq) and Stern–Volmer constant (Ksv) were calculated and shown in Table 5. The values of quenching constant Kq are 2.0 × 1012, 4.9 × 1012, and 5.0 × 1012 M−1 s−1 for compounds (5c, 5e and 5h, respectively) which were found to be higher than maximum scatter collision quenching constant46 (2 × 1010 M−1 s−1), indicating the quenching of fluorophore (BSA) by our compounds is initiated by the static mechanism with formation complex at the ground state. According to Kq values, the capability of the compounds to quench the emission intensity of BSA follows the sequence 5h > 5e > 5c.
![]() | ||
Fig. 7 (a) Stern–Volmer plot of BSA quenching by compounds 5e, 5h and 5c. (b) Double logarithmic plot employed to determine binding parameters for compounds 5e, 5h and 5c. |
Compd | KSV × 104 (M−1) | Kq × 1012 (M−1 s−1) | Log![]() |
Kb (M−1) | n | ΔG° (kcal mol−1) |
---|---|---|---|---|---|---|
5c | 2.0 ± 0.03 | 2.0 ± 0.03 | 4.6 ± 0.1 | 4.7 × 104 | 1.0 | −6.27 |
5e | 4.9 ± 0.07 | 4.9 ± 0.07 | 4.8 ± 0.1 | 6.4 × 104 | 1.0 ± 0.02 | −6.54 |
5h | 5.0 ± 0.1 | 5.0 ± 0.1 | 6.6 ± 0.1 | 4.2 × 106 | 1.4 ± 0.03 | −9.00 |
![]() | (2) |
From the plot of log(Fo − F)/F against log[Q] (Fig. 8), the binding constant Kb and the number of binding sites n have been obtained from the intercept and slope, respectively. The results are mentioned in Table 5. The value of n unveils the presence of a single binding site in BSA for imidazopyrimidines. The desired value for binding constant is contemplated to be in the range 104–106 M−1. The value of Kb observed for 5c, 5e and 5h are in this range, suggesting that the ligand binding to BSA protein is reversible and allows the transport and release of the ligand to the target site. Results indicate that the compounds bind to BSA in the order of 5h > 5e > 5c. Furthermore, the negative value of standard Gibbs free energy change (ΔG°) from eqn (3) suggests the spontaneity of the binding process that resulted into BSA–imidazopyrimidine complex.
G° = −RT![]() ![]() | (3) |
In this experiment, the fluorescence spectra of BSA–PBZ and BSA–IBP complexes (1:
1) were recorded in the absence and presence of the subsequent increasing amounts of ligand 5e (0–52 μM) (Fig. 8a and b). A decrement in the emission intensity was observed by the increase in the concentration of 5e, and the data thus obtained was utilized to plot a graph of log(Fo − F)/F against log[Q] for complex 5e in the absence and presence of site markers as shown in Fig. 8c. The binding constant Kb for phenylbutazone and ibuprofen was found to be 0.0085 × 104 and 0.374 × 104, respectively (Table 6). It can be observed that the Kb and log Kb values of the compound 5e with BSA significantly decrease in the presence of phenylbutazone and ibuprofen, compared to the Kb value in the absence of any site markers, indicating that there are competitive interactions between compound 5e and the two site makers with BSA.48 Therefore, it can be inferred that the compound 5e may interact with BSA at both site I and site II. Nevertheless, the value of Kb is much lower in the presence of phenylbutazone, suggesting compound 5e mainly binds in Sudlow's Site I of BSA.
System | Log![]() |
Kb (M−1) |
---|---|---|
BSA + 5e | 4.790 | 6.2 × 104 |
BSA + PBZ + 5e | 1.932 | 0.0085 × 104 |
BSA + IBP + 5e | 2.573 | 0.374 × 104 |
1H NMR (400 MHz, DMSO-d6): δ 9.64–9.66 (d, 1H, J = 6.8 Hz, 7 H), 9.02 (d, 1H, J = 4.0 Hz, 5-H), 7.79–7.81 (d, 2H, J = 7.6 Hz, Ph-2′,6′-H), 7.73 (dd, 1H, Jo = 6.8, Jm = 4.4 Hz, Ph-4′-H), 7.60–7.64 (m, 3H, Ph-3′,5′, 6-H), 2.21 (s, 3H, 2-CH3).
13C NMR (100 MHz, DMSO-d6): δ 186.20, 156.99, 147.59, 146.82, 138.81, 138.08, 133.52, 129.38, 129.27, 119.67, 113.66, 15.44.
HRMS (ESI) m/z: 238.0890 [M + H]+.
Elemental analysis: calcd. for C14H11N3O: C, 70.87; H, 4.67; N, 17.71% found: C, 70.65; H, 4.74; N, 17.48%.
1H NMR (400 MHz, DMSO-d6): δ 9.60–9.62 (dd, 1H, J = 7.2, J = 2.0 Hz, 7-H), 8.99–9.01 (dd, 1H, J = 4.4, J = 2.0 Hz, 5-H), 7.87–7.91 (dt, 2H, Jo = 8.8, J(m)HF = 5.6 Hz, Ph-2′,6′-H), 7.59–7.62 (dd, 1H, J = 6.8 Hz, J = 4.4 Hz, 6-H), 7.43–7.48 (t, 2H, Jo = 8.8 Hz, Ph-3′,5′-H), 2.24 (s, 3H, 2-CH3).
13C NMR (100 MHz, DMSO-d6): δ 184.85, 156.91, 147.04, 138.05, 135.44, 135.41, 132.48, 132.38, 119.67, 116.68, 116.46, 113.56, 15.65.
HRMS (ESI) m/z: 256.0836 [M + H]+.
Elemental analysis: calcd. for C14H10FN3O: C, 65.88; H, 3.95; N, 16.46% found: C, 65.81; H, 3.81; N, 16.63%.
1H NMR (400 MHz, DMSO-d6): δ 9.61–9.63 (d, 1H, J = 6.8 Hz, 7-H), 9.01 (d, 1H, J = 3.2 Hz, 5-H), 7.81–7.83 (d, 2H, J = 8.4 Hz, Ph-2′,6′-H), 7.68–7.70 (d, 2H, J = 8.4 Hz, Ph-3′,5′-H), 7.61–7.64 (dd, 1H, J = 6.4, J = 4.4 Hz, 6-H), 2.24 (s, 3H, 2-CH3).
13C NMR (100 MHz, DMSO-d6): δ 184.96, 156.98, 147.98, 146.98, 138.31, 138.05, 137.50, 131.28, 129.53, 119.56, 113.63, 15.63.
HRMS (ESI) m/z: 272.0469 [M + H]+; 274.0463[M + 1 + 2]+; (3:
1).
Elemental analysis: calcd. for C14H10ClN3O: C, 61.89; H, 3.71; N, 15.47% found: C, 61.78; H, 3.63; N, 15.25%.
1H NMR (400 MHz, DMSO-d6): δ 9.61–9.63 (dd, 1H, J = 6.8, J = 1.2 Hz, 7-H), 8.97 (dd, 1H, J = 4.0, J = 1.2 Hz, 5-H), 7.82–7.84 (d, 2H, J = 8.4 Hz, Ph-2′,6′-H), 7.71–7.73 (d, 2H, J = 8.4 Hz, Ph-3′,5′-H), 7.56–7.58 (dd, 1H, J = 6.8, J = 4.4 Hz, 6-H), 2.21 (s, 3H, 2-CH3).
13C NMR (100 MHz, DMSO-d6): δ 185.19, 156.39, 149.45, 147.70, 138.11, 137.81, 132.42, 131.27, 127.15, 119.54, 113.15, 16.14.
HRMS (ESI) m/z: 316.9978 [M + H]+; 318.9959[M + 1 + 2]+; (1:
1).
Elemental analysis: calcd. for C14H10BrN3O: C, 53.19; H, 3.19; N, 13.29% found: C, 52.98; H, 3.24; N, 13.43%.
1H NMR (400 MHz, DMSO-d6): δ 9.52–9.54 (dd, 1H, J = 7.0, J = 2.4 Hz, 7-H), 9.01–9.03 (dd, 1H, J = 4.8, J = 2.0 Hz, 5-H), 7.82–7.84 (d, 2H, J = 8.8 Hz, Ph-2′,6′-H), 7.61–7.64 (dd, 1H, J = 6.0, J = 4.8 Hz, 6-H), 7.14–7.16 (d, 2H, J = 9.2 Hz, Ph-3′,5′-H), 3.90 (s, 3H, Ph-4′-OMe), 2.31 (s, 3H, 2-CH3).
13C NMR (100 MHz, DMSO-d6): δ 184.54, 164.08, 157.17, 146.21, 138.12, 132.26, 130.82, 119.82, 114.78, 113.76, 56.23, 15.07.
HRMS (ESI) m/z: 268.1003 [M + H]+.
Elemental analysis: calcd. for C15H13N3O2: C, 67.40; H, 4.90; N, 15.72% found: C, 67.28; H, 4.74; N, 15.59%.
1H NMR (400 MHz, DMSO-d6): δ 9.63–9.65 (dd, J = 6.8, J = 1.2 Hz, 1H, 7-H), 9.03–9.04 (dd, 1H, J = 4.4, J = 2.8 Hz, 5-H), 7.63–7.66 (dd, 1H, J = 6.8, J = 4.8 Hz, 6-H), 7.51–7.55 (t, 1H, Jo = 7.6 Hz, Ph-5′-H), 7.36–7.38 (d, 1H, Jo = 7.6 Hz, Ph-6′-H), 7.29–7.31 (m, 2H, Ph-2′,4′-H), 3.83 (s, 3H, Ph-3′-OMe), 2.23 (s, 3H, 2-CH3).
13C NMR (100 MHz, DMSO-d6): δ 185.89, 159.88, 157.17, 147.31, 146.60, 140.15, 138.12, 130.67, 121.53, 119.66, 119.54, 113.83, 113.74, 55.99, 15.25.
HRMS (ESI) m/z: 268.1003 [M + H]+.
Elemental analysis: calcd. for C15H13N3O2: C, 67.40; H, 4.90; N, 15.72% found: C, 67.52; H, 4.66; N, 15.97%.
1H NMR (400 MHz, DMSO-d6): δ 9.64–9.66 (dd, 1H, J = 6.8, J = 2.0 Hz, 7-H), 9.06–9.07 (dd, 1H, J = 4.4, J = 2.0 Hz, 5-H), 7.67–7.70 (dd, 1H, J = 6.8, J = 4.4 Hz, 6-H), 7.52–7.56 (t, 1H, J = 8.0 Hz, Ph-4′-H), 7.37–7.39 (d, 1H, J = 7.6 Hz, Ph-6′-H), 7.30–7.33 (m, 2H, Ph-3′,5′-H), 3.84 (s, 3H, Ph-2′-OMe), 2.25 (s, 3H, 2-CH3).
13C NMR (100 MHz, DMSO-d6): δ 185.89, 159.92, 157.59, 146.23, 140.03, 138.34, 130.76, 130.71, 121.68, 119.75, 119.70, 114.17, 113.81, 56.05, 15.05.
HRMS (ESI) m/z: 268.1003 [M + H]+.
Elemental analysis: calcd. for C15H13N3O2: C, 67.40; H, 4.90; N, 15.72% found: C, 67.71; H, 4.56; N, 15.46%.
1H NMR (400 MHz, DMSO-d6): δ 9.43–9.45 (dd, 1H, J = 6.8, J = 1.2 Hz, 7-H), 9.02–9.03 (dd, 1H, J = 4.0, J = 1.2 Hz, 5-H), 8.24–8.25 (d, 1H, J = 4.4 Hz, thienyl-5′-H), 7.93–7.94 (d, 1H, J = 3.2 Hz, thienyl-3′-H), 7.61–7.64 (dd, 1H, J = 6.8, J = 4.4 Hz, 6-H), 7.33–7.35 (t, 1H, J = 4.4 Hz, thienyl-4′-H), 2.51 (s, 3H, 2-CH3).
13C NMR (100 MHz, DMSO-d6): δ 177.02, 157.36, 145.97, 144.53, 142.68, 138.04, 137.19, 136.61, 129.26, 119.49, 113.74, 14.97.
HRMS (EI) m/z: 244.0467 [M + H]+.
Elemental analysis: calcd. for C12H9N3OS: C, 59.24; H, 3.73; N, 17.27% found: C, 59.12; H, 3.62; N, 17.49%.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ra01795e |
This journal is © The Royal Society of Chemistry 2025 |