5.2
Impact Factor
Generic selectors
Exact matches only
Search in title
Search in content
Post Type Selectors
Search in posts
Search in pages
Filter by Categories
Corrigendum
Current Issue
Editorial
Erratum
Full Length Article
Full lenth article
Letter to Editor
Original Article
Research article
Retraction notice
Review
Review Article
SPECIAL ISSUE: ENVIRONMENTAL CHEMISTRY
5.3
Impact Factor
Generic selectors
Exact matches only
Search in title
Search in content
Post Type Selectors
Search in posts
Search in pages
Filter by Categories
Corrigendum
Current Issue
Editorial
Erratum
Full Length Article
Full lenth article
Letter to Editor
Original Article
Research article
Retraction notice
Review
Review Article
SPECIAL ISSUE: ENVIRONMENTAL CHEMISTRY
View/Download PDF

Translate this page into:

Original Article
2025
:18;
5492025
doi:
10.25259/AJC_549_2025

Synthesis of new oxazole hybridized with pyrazole and/or thiazole compounds: Molecular modeling and cytotoxic activity

Department of Physical Sciences, Chemistry Division, College of Science, Jazan University, P.O. Box. 114, Jazan 45142, Saudi Arabia.
Department of Chemistry, College of Science, Taibah University, Madinah, Saudi Arabia.
Department of Chemistry, Faculty of Science, Umm Al Qura University, Makkah 24230, Saudi Arabia
Department of Physics, College of Science, Taibah University, Madinah, Saudi Arabia.
Department of Environment and Health Research, Umm Al-Qura University, Makkah, Saudi Arabia
Department of Chemistry, Faculty of Science, King Khalid University, 62529 Abha, Saudi Arabia

* Corresponding author: E-mail address: sabomlha@kku.edu.sa (S. Abu-Melha)

Licence
This is an open-access article distributed under the terms of the Creative Commons Attribution-Non Commercial-Share Alike 4.0 License, which allows others to remix, transform, and build upon the work non-commercially, as long as the author is credited and the new creations are licensed under the identical terms.

Abstract

Two series of oxazole-pyrazole hybrids 6a-c and oxazole-thiazole hybrids 7-8 were prepared and characterized by compatible results of spectroscopic analyses (infrared-IR, nuclear magnetic resonance-NMR, and mass spectrometry-MS). The density functional theory (DFT)/B3LYP optimization of the built hybrids 3-5 released a planar configuration, whereas the others 6-8 displayed varied spatial structures. The FMO’s shapes of the considered analogs indicated that the highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) of pyrazolyl hybrids 6a-c was centered on the oxazolylphenyl carboxamide and chlorophenylpyrazole portions, respectively. Conversely, the benzylidene conjugates 8a-b have displayed alternative configurations, mainly on the phenyl thiazolidinylidene cyanoacetamide segment. Cytotoxic assay was performed using three different human malignant cell lines and a normal lung fibroblast. Analog 8a exhibited the strongest anticancer effectiveness against HT-29 and MCF-7 cells, IC50 = 13.22±0.16 and 6.41±0.47 μM, respectively. However, furthermost analogs showed higher selectivity toward cancer cells compared to the normal WI-38 cells against the drug reference (dasatinib). Moreover, the synthesized hybrids revealed significant variations in inhibitory effectiveness against CA IX and CA XII. Derivative 6c emerged as the most active CA IX inhibitor (IC₅₀ = 0.011±0.015 μM), while analog 7 was the most effective against CA XII (IC₅₀ = 0.119±0.043 μM), showing the capability of these analogs as selective CA isoform levelling in anticancer therapy. Molecular docking was utilized to assign their binding affinity towards the human carbonic anhydrase II enzyme (PDB ID: 1V9E), where the conjugates 8b and 8a showed significant binding profiles, S = -7.0940 and -6.5463 kcal/mol, respectively. In addition, the SwissADME pharmacokinetic and drug-likeness showed that most analogs were within the acceptable drug-likeness range, with no breaches of Lipinski’s rule.

Keywords

Carbonic anhydrase
Cytotoxic
DFT
Docking
Oxazole-pyrazole
Pharmacokinetic

1. Introduction

Undoubtedly, heterocyclic compounds show a substantial function in the field of modern drug discovery owing to their structural resemblance to natural bioactive molecules and their wide spectrum of biological effectiveness [1-4]. Alongside these heterocyclic compounds, it was noticed that oxazole, pyrazole, and thiazole rings are important in the design of novel therapeutic drugs [5-8]. Their interbreeding into a collective molecular skeleton presents an amazing approach for developing their pharmacological effectiveness through synergistic effects [9,10]. However, oxazole moieties are classified as five-membered heteroaromatic analogs containing both nitrogen and oxygen atoms. They have been broadly deliberated as antimicrobial, anticancer, and anti-inflammatory effectiveness [11,12]. Their rigid skeletal and electronic properties allow them to cooperate obviously with a range of biological boards, mostly enzymes and receptors involved in tumor and inflammatory syndromes [13]. Likewise, pyrazole moieties are commonly used in the enhancement of the anti-inflammatory and anti-cancer treatments due to an eminent binding with the kinase domains and dissimilar protein pockets [14-18]. However, thiazole rings also improve the pharmacokinetic properties of hybrid compounds through the increases in both lipophilicity and membrane permeability [19-22]. So, by merging of more than bio-active heterocyclic moieties may enhance cytotoxic effectiveness against cancer cells by inhibiting the mechanism of enzyme actions [23]. Meanwhile, cytotoxic effectiveness alongside tumor cells is supposed to be improved by the structural combination between oxazole with pyrazole and/or thiazole moieties through a diversity of modes of action, such as enzyme inhibitions and metabolic pathway disturbances [24-26]. Along with in vitro tests, early-stage drug development currently mostly depends on computational methods such as molecular docking and pharmacokinetic modeling [27]. Moreover, carbonic anhydrases characterize a family of metalloenzymes that catalyze the reversible hydration of carbon dioxide, which is involved in several physiological and pathological processes. Owing to the extensive range of CA isoforms, selective inhibition has arisen as a talented strategy in the therapy of diseases such as cancer, epilepsy, and glaucoma. Recent studies have been provided thorough insights into the mechanism of action of CA inhibitors, principally those bearing moieties that bind to the active pocket catalytic zinc ion of the enzyme, thereby blocking its enzymatic activity showed potent inhibition of CA II and VII isoforms with good anti-epileptic activity, revealing the therapeutic promise of designed CAIs [28,29]. The current work intends to synthesis several new oxazole hybrids with pyrazole and/or thiazole analogs and assign their electronic structures through molecular modeling by density functional theory (DFT) simulation. To assess the synthetic analogs’ antiproliferative qualities, they were cytotoxically screened against a few human cancer cell lines. To investigate, their inhibitory activity against carbonic anhydrase was also evaluated. To clarify the interaction profile of active drugs with CA and other pertinent targets, molecular docking experiments were carried out. Lastly, SwissADME was used to forecast drug-likeness, bioavailability, and pharmacokinetic behavior, which helped choose potential candidates for additional research and modification.

2. Materials and Methods

2.1. Instruments

The melting temperature (°C) was measured using the open capillary method. Infrared (IR) spectral data (KBr discs, cm-1) were analyzed using a ThermoNicolet IS10-FTIR spectrophotometer. Nuclear magnetic resonance (NMR) spectral data were recorded in DMSO-d6 on a Jeol-500 MHz spectrometer. Mass spectra (MS) were recorded by a ThermoScientific GC-MS DSQII spectrometer at 70 eV. Microanalyses of the elements (C, H, and N) were recorded using a Perkin Elmer analyzer.

2.2. 2-Oxopropyl 4-(2-cyanoacetamido)benzoate (3)

A sodium carbonate solution (1.59 g, 15 mmol) in water (5 mL) was added dropwise to a solution of 4-(2-cyanoacetamido)benzoic acid (1) (3.00 g, 15 mmol) in 20 mL of DMSO in a 250 mL conical flask. The solution was stirred for 15 min, and then chloroacetone (1.38, 15 mmol) was added. The mixture was heated for 4 h on a water bath at 90°C. After cooling, the mixture was poured into ice-cold water, and the obtained solid was collected and crystallized from EtOH.

Yield = 85.8%, m.p. = 163-164°C. IR (ν/cm-1): 3348 (N-H), 2252 (C≡N), 1730, 1687 (C=O). 1H NMR (δ/ppm): 2.13 (s, 3H, -CH3 of acetyl), 3.88 (s, 2H, -CO-CH2-CN), 5.56 (s, 2H, -CO-CH2-O), 7.68 (d, J = 8.5 Hz, 2H), 7.85 (d, J = 8.5 Hz, 2H), 10.71 (s, 1H, N-H). 13C NMR (δ/ppm): 24.18, 30.89, 70.43, 116.76, 119.03 (2C), 124.37, 130.51 (2C), 142.60, 164.15, 167.21, 201.88. MS for C13H12N2O4 [M]+: m/z = 260 (37.25%). Analysis for C13H12N2O4 (260.08): Calcd.: C, 60.00; H, 4.65; N, 10.76%. Found: C, 60.13; H, 4.71; N, 10.67%.

2.3. 2-Cyano-N-(4-(4-methyloxazol-2-yl)phenyl)acetamide (4)

To a suspension of 2-oxopropyl 4-(2-cyanoacetamido)benzoate (3) (1.30 g, 5 mmol) and acetamide (1.47 g, 25 mol) in 50 mL xylene, 47% BF3/Et2O (0.35 mL) was added dropwise. The solution was refluxed for 8 h, and then the solution was quenched with cold water (50 mL), and the aqueous layer was extracted with diethyl ether. The combined organic layer was dried over anhydrous MgSO4 and evaporated under vacuum. The yellow solid was crystallized in 95% EtOH.

Yield = 71.4%, m.p. = 211-212°C. IR (ν/cm-1): 3284 (N-H), 2254 (C≡N), 1690 (C=O). 1H NMR (δ/ppm): 2.21 (s, 3H, oxazole-CH3), 3.96 (s, 2H, -CO-CH2-CN), 7.57 (s, 1H, oxazole-H5), 7.70 (d, J = 8.5 Hz, 2H), 7.81 (d, J = 8.5 Hz, 2H), 10.64 (s, 1H, N-H). 13C NMR (δ/ppm): 13.82, 31.16, 116.47, 120.33 (2C), 126.06, 127.74 (2C), 134.20, 138.63, 141.56, 162.38, 163.91. MS for C13H11N3O2 [M]+: m/z = 241 (30.08%). Analysis for C13H11N3O2 (241.09): Calcd.: C, 64.72; H, 4.60; N, 17.42%. Found: C, 64.55; H, 4.68; N, 17.54%.

2.4. N-(4-Chlorophenyl)-2-((4-(4-methyloxazol-2-yl)phenyl)amino)-2-oxoacetohydrazonoyl cyanide (5)

A diazonium solution derived from 4-chloroaniline (1.01 g, 8 mmol) was obtained by adding sodium nitrite solution (0.56 g in 10 mL H2O) drop by drop (for 10 min) to a suspension of 4-chloroaniline (1.01 g, 8 mmol) in conc. HCl (2.40 mL) at 0-5°C. The diazonium solution was added dropwise to a well-stirred suspension of 2-cyano-N-(4-(4-methyloxazol-2-yl)phenyl)acetamide (4) (1.92 g, 8 mmol) in 30 mL of pyridine at 0-5°C. The mixture was stirred at 0-5°C for 2 h and then diluted with ice water. The solid was filtered, dried, and then crystallized from an EtOH-DMF mixture (5:1).

Yield = 67.5%, m.p. = 242-243°C. IR (ν/cm-1): 3303, 3241 (N-H), 2221 (C≡N), 1667 (C=O). 1H NMR (δ/ppm): 2.23 (s, 3H, oxazole-CH3), 7.13 (d, J = 8.5 Hz, 2H), 7.36 (d, J = 8.5 Hz, 2H), 7.55 (s, 1H, oxazole-H5), 7.71 (d, J = 8.5 Hz, 2H), 7.83 (d, J = 8.5 Hz, 2H), 10.47 (s, 1H, N-H), 11.85 (s, 1H, C=N-N-H). 13C NMR (δ/ppm): 13.77, 110.54, 111.30, 117.35 (2C), 121.22 (2C), 126.01, 127.63 (2C), 128.28, 129.41 (2C), 133.87, 138.51, 140.46, 142.16, 162.14, 162.93. MS for C19H14ClN5O2 [M]+: m/z = 379 (13.77%). Analysis for C19H14ClN5O2 (379.08): Calcd.: C, 60.09; H, 3.72; N, 18.44%. Found: C, 60.25; H, 3.80; N, 18.55%.

2.5. 2-Oxo-2-(thiophen-2-yl)ethyl 4-(4-amino-1-(4-chlorophenyl)-1H-pyrazole-3-carboxamido)benzoate compounds 6a-c

A 150 mL RB-flask was charged with a solution of N-(4-chlorophenyl)-2-((4-(4-methyloxazol-2-yl)phenyl)amino)-2-oxoacetohydrazonoyl cyanide (5) (1.06 g, 2.8 mmol) in 40 mL tetrahydrofuran (THF) and 0.2 mL of triethylamine. The appropriate halogenated reagent, either chloroacetone, ethyl chloroacetate, or chloroacetonitrile (2.8 mmol), was added to the previous solution, and the mixture was refluxed for 6 h. The solid obtained after cooling was filtered and crystallized from EtOH to furnish the target oxazole-pyrazole hybrid compounds 6a-c.

2.5.1. 5-Acetyl-4-amino-1-(4-chlorophenyl)-N-(4-(4-methyloxazol-2-yl)phenyl)-1H-pyrazole-3-carboxamide (6a)

Yield = 60.1%, m.p. = 261-262°C. IR (ν/cm-1): 3323, 3281, 3205 (-NH2 and N-H), 1660, 1627 (C=O). 1H NMR (δ/ppm): 2.21 (s, 3H, oxazole-CH3), 2.30 (s, 3H, -CH3 of acetyl), 6.64 (s, 2H, -NH2), 7.34 (d, J = 8.5 Hz, 2H), 7.51 (d, J = 8.5 Hz, 2H), 7.58 (s, 1H, oxazole-H5), 7.70 (d, J = 8.5 Hz, 2H), 7.82 (d, J = 8.5 Hz, 2H), 10.11 (s, 1H, N-H). 13C NMR (δ/ppm): 13.72, 27.68, 118.10, 120.97 (2C), 122.39 (2C), 126.07, 127.64 (2C), 129.13 (2C), 130.56, 132.68, 133.36, 134.30, 137.74, 138.44, 140.26, 161.81, 162.52, 192.05. MS for C22H18ClN5O3 [M]+: m/z = 435 (28.14%). Analysis for C22H18ClN5O3 (435.11): Calcd.: C, 60.62; H, 4.16; N, 16.07%. Found: C, 60.48; H, 4.10; N, 16.18%.

2.5.2. Ethyl 4-amino-1-(4-chlorophenyl)-3-((4-(4-methyloxazol-2-yl)phenyl)carbamoyl)-1H-pyrazole-5-carboxylate (6b)

Yield = 62.4%, m.p. = 284-285°C. IR (ν/cm-1): 3336, 3267, 3195 (-NH2 and N-H), 1683, 1656, (C=O). 1H NMR (δ/ppm): 2.23 (s, 3H, oxazole-CH3), 1.29 (t, J = 7.0 Hz, 3H, COOCH2CH3), 4.27 (q, J = 7.0 Hz, 2H, COOCH2CH3), 6.52 (s, 2H, -NH2), 7.38 (d, J = 8.5 Hz, 2H), 7.51 (d, J = 8.5 Hz, 2H), 7.60 (s, 1H, oxazole-H5), 7.73 (d, J = 8.5 Hz, 2H), 7.88 (d, J = 8.5 Hz, 2H), 10.27 (s, 1H, N-H). 13C NMR (δ/ppm): 13.76, 14.55, 59.04, 120.69, 121.15 (2C), 122.26 (2C), 126.12, 127.67 (2C), 128.41, 129.18 (2C), 132.55, 133.21, 134.24, 137.85, 138.50, 139.98, 161.78, 162.46, 163.48. MS for C23H20ClN5O4 [M]+: m/z = 465 (21.63%). Analysis for C23H20ClN5O4 (465.12): Calcd.: C, 59.30; H, 4.33; N, 15.03%. Found: C, 59.12; H, 4.25; N, 15.16%.

2.5.3. 4-Amino-1-(4-chlorophenyl)-5-cyano-N-(4-(4-methyloxazol-2-yl)phenyl)-1H-pyrazole-3-carboxamide (6c)

Yield = 58.0%, m.p. = 270-271°C. IR (ν/cm-1): 3340, 3278, 3212 (-NH2 and N-H), 2217 (C≡N), 1670 (C=O). 1H NMR (δ/ppm): 2.21 (s, 3H, oxazole-CH3), 6.61 (s, 2H, -NH2), 7.35 (d, J = 8.5 Hz, 2H), 7.48 (d, J = 8.5 Hz, 2H), 7.57 (s, 1H, oxazole-H5), 7.70 (d, J = 8.5 Hz, 2H), 7.83 (d, J = 8.5 Hz, 2H), 10.13 (s, 1H, N-H). 13C NMR (δ/ppm): 13.82, 97.98, 117.83, 121.02 (2C), 122.34 (2C), 126.20, 127.71 (2C), 128.64, 129.23 (2C), 131.53, 132.59, 133.90, 137.83, 138.45, 139.87, 161.73, 162.38. MS for C21H15ClN6O2 [M]+: m/z = 418 (33.58%). Analysis for C21H15ClN6O2 (418.09): Calcd.: C, 60.22; H, 3.61; N, 20.07%. Found: C, 60.34; H, 3.56; N, 19.98%.

2.6. 2-Cyano-N-(4-(4-methyloxazol-2-yl)phenyl)-2-(4-oxo-3-phenylthiazolidin-2-ylidene)acetamide (7)

A mixture of 2-cyano-N-(4-(4-methyloxazol-2-yl)phenyl)acetamide (4) (0.96 g, 4 mmol) in DMF (20 mL) was stirred in a 250 mL conical flask with solid potassium hydroxide (0.22 g, 4 mmol) for 10 min, and phenyl isothiocyanate (0.48 g, 4 mmol) was then added. The reaction components were stirred for 6 h at 25-30°C. In situ, ethyl bromoacetate (0.67 mL, 4 mmol) was added, and the mixture was stirred for 6 h. The solid that formed upon dilution with 30 mL ice-water was collected and purified by crystallization from EtOH.

Yield = 64.8%, m.p. = 230-231°C. IR (ν/cm-1): 3388 (N-H), 2195 (C≡N), 1740, 1668 (C=O). 1H NMR (δ/ppm): 2.21 (s, 3H, oxazole-CH3), 4.01 (s, 2H, cyclic-CH2), 7.41 (d, J = 7.5 Hz, 2H), 7.48-7.53 (m, 3H), 7.58 (s, 1H, oxazole-H5), 7.71 (d, J = 8.5 Hz, 2H), 7.83 (d, J = 8.5 Hz, 2H), 9.54 (s, 1H, N-H). 13C NMR (δ/ppm): 13.76, 32.18, 77.97, 113.54, 120.85 (2C), 126.17, 127.73 (2C), 128.13, 129.40 (2C), 130.17 (2C), 133.82, 136.77, 138.39, 139.68, 162.05, 163.30, 170.91, 173.41. MS for C22H16N4O3S [M]+: m/z = 416 (25.93%). Analysis for C22H16N4O3S (416.09): Calcd.: C, 63.45; H, 3.87; N, 13.45%. Found: C, 63.34; H, 3.82; N, 13.51%.

2.7. 2-Cyano-2-(5-arylidene-4-oxo-3-phenylthiazolidin-2-ylidene)-N-(4-(4-methyloxazol-2-yl)phenyl)acetamide compounds 8a and 8b

To a suspension of 2-cyano-N-(4-(4-methyloxazol-2-yl)phenyl)-2-(4-oxo-3-phenylthiazolidin-2-ylidene)acetamide (7) (0.83 g, 2 mmol) in 30 mL EtOH and 0.1 mL of piperidine, either 4-aniladehyde or 4-chlorobenzaldehyde (2 mmol) was added. The mixture was refluxed for 4 h, and the solid obtained after cooling was filtered. The solid was subjected to purification by crystallization from an EtOH-DMF mixture (4:1) to furnish target oxazole-thiazole hybrids 8a and 8b.

2.7.1. 2-Cyano-2-(5-(4-methoxybenzylidene)-4-oxo-3-phenylthiazolidin-2-ylidene)-N-(4-(4-methyl oxazol-2-yl)phenyl)acetamide (8a):

Yield = 83.6%, m.p. = 255-56°C. IR (ν/cm-1): 3340 (N-H), 2202 (C≡N), 1708, 1665 (C=O). 1H NMR (δ/ppm): 2.23 (s, 3H, oxazole-CH3), 3.81 (s, 3H, OCH3), 7.11 (d, J = 9.0 Hz, 2H), 7.40 (d, J = 7.5 Hz, 2H), 7.47-7.51 (m, 3H), 7.57 (s, 1H, oxazole-H5), 7.62 (d, J = 9.0 Hz, 2H), 7.70 (d, J = 8.5 Hz, 2H), 7.78 (s, 1H, olefinic C=C-H), 7.83 (d, J = 8.5 Hz, 2H), 9.48 (s, 1H, N-H). 13C NMR (δ/ppm): 13.74, 55.91, 80.32, 113.76, 114.25 (2C), 120.96 (2C), 126.15, 126.87 (2C), 127.68 (2C), 128.21, 129.28 (2C), 130.04 (2C), 130.48 (2C), 133.80, 134.52, 135.50, 138.34, 139.59, 160.13, 162.15, 163.37, 163.84, 167.38. MS for C30H22N4O4S [M]+: m/z = 534 (21.68%). Analysis for C30H22N4O4S (534.14): Calcd.: C, 67.40; H, 4.15; N, 10.48%. Found: C, 67.56; H, 4.22; N, 10.58%.

2.7.2. 2-(5-(4-Chlorobenzylidene)-4-oxo-3-phenylthiazolidin-2-ylidene)-2-cyano-N-(4-(4-methyl oxazol-2-yl)phenyl)acetamide (8b):

Yield = 80.2%, m.p. = 271-272°C. IR (ν/cm-1): 3338 (N-H), 2201 (C≡N), 1713, 1671 (C=O). 1H NMR (δ/ppm): 2.21 (s, 3H, CH3), 7.41 (d, J = 7.5 Hz, 2H), 7.48-7.53 (m, 5H), 7.58 (s, 1H, oxazole-H5), 7.64 (d, J = 8.5 Hz, 2H), 7.72 (d, J = 8.5 Hz, 2H), 7.81 (d, J = 8.5 Hz, 2H), 7.88 (s, 1H, olefinic C=C-H), 9.56 (s, 1H, N-H). 13C NMR (δ/ppm): 13.74, 80.44, 113.81, 121.06 (2C), 126.18, 127.04, 127.70 (2C), 128.21, 128.92 (2C), 129.31 (2C), 129.78 (2C), 130.15 (2C), 132.83, 133.86 (2C), 134.40, 135.46, 138.20, 139.56, 162.32, 163.53, 164.35, 168.63. MS for C29H19ClN4O3S [M]+: m/z = 538 (16.43%). Analysis for C29H19ClN4O3S (538.09): Calcd.: C, 64.62; H, 3.55; N, 10.39%. Found: C, 64.44; H, 3.47; N, 10.52%.

2.8. DFT Computations

The Gaussian 09W built-in DFT/B3LYP/6-311++G(d,p) routine [30-33] was employed in the investigation of the designed derivatives’ optimum spatial configurations, frontier molecular orbitals (FMOs), and electronic properties, where the GaussView [34] was utilized for analyzing the outcomes.

2.9. Cytotoxic assay

The cytotoxic effects of the synthesized oxazole hybridized with pyrazole and/or thiazole analogs were assessed on three distinct human sarcoma cell lines: MCF-7 (breast cancer), Panc-1 (epithelioid carcinoma), and HT-29 (colorectal adenocarcinoma), besides a normal cell line (WI-38). The 3-(4,5-Dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromidefor (MTT) test was used to conduct the assessment, as described in the supporting information file. Dasatinib was used as the reference medication for the comparative analysis [35].

2.10. Carbonic anhydrase assay

The carbonic anhydrase inhibition activity against CA IX and CA XII of the synthesized oxazole derivatives hybridized with pyrazole and/or thiazole analogs was evaluated according to the protocol of the carbonic anhydrase technique [36] (see the supplementary file).

Supplementary file

2.11. Molecular docking

The docked analogs of all the hybrids 3-8 against carbonic anhydrase CA XII were analyzed separately and evaluated using M.O.E. 2019. The co-crystal structure of carbonic anhydrase CA XII was obtained from (PDB having PDBID 1V9E (www.rcsb.org) [37], and compared with the reference drug (acetazolamide). For protein preparation, hydrogen atoms were incorporated to correct protonation states, while water molecules and heteroatoms were removed to simplify the system. Furthermore, the structure underwent energy minimization applying the AMBER10: EHT force field to resolve steric clashes. Simultaneously, during ligand generation, energy was minimized using the MMFF94x force field, and partial charges were assigned. Additionally, the docking procedure was validated by re-docking the co-crystallized ligand into the binding site of the generated protein. The root-mean-square deviation (RMSD) between the docked and crystallographic conformations was 1.2 Å, demonstrating the reliability of the docking approach.

3. Results and Discussion

3.1. Synthesis of oxazole-pyrazole and oxazole-thiazole hybrids

The process of synthesizing oxazole-pyrazole and oxazole-thiazole hybrids was started by producing the versatile precursors, 2-cyano-N-(4-(4-methyloxazol-2-yl)phenyl)acetamide (4), and its corresponding N-(4-chlorophenyl)-hydrazone compound 5, according to the following sequence of reactions. The first step involved the esterification reaction of 4-(2-cyanoacetamido)benzoic acid (1) [38] with chloroacetone by heating in DMSO and sodium carbonate at 90°C to furnish the corresponding ester, 2-oxopropyl 4-(2-cyanoacetamido) benzoate (3) (Scheme 1). In the second step, the introduction of an oxazole ring into ester compound 3 was achieved by treatment with acetamide in boiling xylene and boron trifluoride. The product was identified as 2-cyano-N-(4-(4-methyloxazol-2-yl)phenyl)acetamide (4). The final step takes advantage of the methylene group’s reactivity in the cyanoacetamide part of compound 4, facilitating an electrophilic diazo-coupling reaction with a diazonium salt that originates from 4-chloroaniline. The reaction proceeds when 2-cyano-N-(4-(4-methyloxazol-2-yl)phenyl)acetamide (4) was treated with 4-chlorophenyl diazonium chloride at 0-5°C in pyridine to provide the conforming arylhydrazone compound (5), N-(4-chlorophenyl)-2-((4-(4-methyloxazol-2-yl)phenyl)amino)-2-oxoacetohydrazonoyl cyanide. The structures of compounds 4 and 5 were verified based on compatible spectral data. The IR spectrum showed the absorption frequency for the carbonyl group in cyanoacetamide compound 4 at 1690 cm-1, and at a lower wavenumber at 1667 cm-1 in its arylhydrazone compound 5 due to the extended conjugation with the hydrazone skeleton. The 1H NMR spectrum of compound 4 clearly indicated a singlet signal for the methylene protons at δ 3.96 ppm, while the 1H NMR spectrum of compound 5 exhibited no signal assignable to the methylene protons and displayed the hydrazone proton as a singlet at δ 11.85 ppm.

Synthesis of 2-cyano-N-(4-(oxazolyl)phenyl)acetamide derivatives 4 and 5.
Scheme 1.
Synthesis of 2-cyano-N-(4-(oxazolyl)phenyl)acetamide derivatives 4 and 5.

The first series of oxazole-pyrazole hybrids 6a-c was obtained by heating arylhydrazone compound 5 with either chloroacetone, ethyl chloroacetate, or chloroacetonitrile in THF and triethylamine (Scheme 2). The proposed mechanism begins with the nucleophilic displacement of the chlorine atom from the halogenated reagent to form the N-alkylated intermediate (A). This intermediate undergoes intramolecular cyclization via the addition of the methylene group to the nitrile function to form the pyrazole intermediate (B), which rapidly tautomerizes to produce the final aminopyrazole skeleton in hybrids 6a-c. The IR spectrum of 6a showed no absorption band assignable to the cyano function, while it displayed the absorptions of the –NH2 and N-H groups at stretching frequencies at 3323, 3281, and 3205 cm-1 for the NH groups, in addition to two absorptions for the carbonyl groups at 1660 (conjugate amidic carbonyl) and 1627 cm-1 (intramolecular hydrogen bonding carbonyl of acetyl group). The 1H NMR spectrum of compound 6a displayed two singlet signals at δ 2.21 and 2.30 ppm, which correspond to the protons of the methyl-linked oxazole and acetyl groups, respectively. The protons of the pyrazole amino group were detected as a singlet at δ 6.64 ppm. The aromatic protons appeared as four doublet signals at δ 7.34, 7.51, 7.70, and 7.82 ppm. Additionally, the protons of the oxazole-C5 and N-H groups were detected as singlet signals at δ 7.58 and 10.11 ppm, respectively.

Synthesis of oxazole-pyrazole hybrids 6a-c.
Scheme 2.
Synthesis of oxazole-pyrazole hybrids 6a-c.

The synthesis of the second targeting series of oxazole-thiazole hybrids 7 and 8 was initiated by the base-catalyzed addition of 2-cyano-N-(4-(4-methyloxazol-2-yl)phenyl)acetamide (4) to phenyl isothiocyanate in stirred DMF and potassium hydroxide to produce the corresponding thiocarbamoyl intermediate (C), which underwent in situ treatment with ethyl bromoacetate and cyclization into the thiazolidine-4-one ring in hybrid compound 7 by losing water and HBr molecules (Scheme 3). The reactivity of the methylene group of the thiazolidinone ring (hybrid 7) supports the condensation reaction with aromatic aldehydes (Knoevenagel reaction). Thus, oxazole-thiazolidinone hybrid 7 was heated with either 4-anisaldehyde or 4-chlorobenzaldehyde in EtOH and piperidine to furnish the corresponding 5-benzylidene-thiazolidin-4-one compounds 8a and 8b, respectively. Spectral studies were performed to elucidate the chemical structures of 7, 8a, and 8b. The introduction of a benzylidene group at position-5 of the thiazolidinone ring lowered the absorption frequency of the carbonyl group from 1740 cm-1 (IR spectrum of hybrid 7) to 1708 cm-1 (hybrid 8a) and 1713 cm-1 (hybrid 8b). The 1H NMR spectrum of hybrid 8b indicated the absence of any signal related to the methylene group and exhibited a singlet signal at δ 7.88 ppm assignable to the olefinic proton (benzylidene system).

Synthesis of oxazole-thiazole hybrids 7 and 8.
Scheme 3.
Synthesis of oxazole-thiazole hybrids 7 and 8.

3.2. Molecular modeling

The spital configuration for the cyanoacetamide conjugates 3 and 4, as well as the oxoacetohydrazonoyl cyanide 5, was planar (Figure 1). However, the resulting pyrazolyl carboxamide analogs 6a-b departed from planarity, whereas the cyanopyrazole derivative 6c conserved flatness. In 6a-b, a distortion from planarity was observed for the carboxamide group, which moved away from the plane of both phenyl and pyrazolyl rings, C3(Ph)-C4(Ph)-NH(Car)-CO(Car) = 2.9-3.6° and NH(Car)-CO(Car)-C3(Prz)-C4(Prz) = 24.1-24.9°, respectively. Further planarity deformation has been perceived for the chlorophenyl group, which strongly leans on the pyrazolyl’s plane, N2(Prz)-N1(Prz)-C1(PhPrz)-C2(PhPrz) = 48.9-52.1°. Moreover, the acetyl (6a) and carboxylate (6b) presented additional divergence from pyrazolyl’s level, e.g., N2(Prz)-N1(Prz)-C5(Prz)-CO(AcPrz) = 166.1° and N2(Prz)-N1(Prz)-C5(Prz)-CO(EstPrz) = -169.0°, respectively (Figure 1).

The DFT/B3LYP/6-311++G(d,p) optimized structures of derivatives 3-6.
Figure 1.
The DFT/B3LYP/6-311++G(d,p) optimized structures of derivatives 3-6.

Otherwise, the thiazolidinylidene derivatives 7-8 presented comparable non-planar configurations (Figure 2), where the cyanoacetamide group was tilted differentially on the phenyl and thiazolidinylidene rings, C2(Ph)-C1(Ph)-NH(Acm)-CO(Acm) = -37.4 - -41.3° and CO(Acm)-C(Acm)-C2(Tz)-S1(Tz) = 5.0-5.5°. Also, the cyano functional group has been moved away from the carbonyl of the cyanoacetamide joint, CO(Acm)-C(Acm)-CN(Acm)-NC(Acm) = -23.6 - -25.6°. Moreover, the N-substituent phenyl was approximately perpendicular to the thiazolidinylidene ring, C2(Tz)-N3(Tz)-C1(PhTz)-C2(PhTz) = 88.3-90.1 (Table S1).

Table S1
The DFT/B3LYP/6-311++G(d,p) optimized structures of derivatives 7-8.
Figure 2.
The DFT/B3LYP/6-311++G(d,p) optimized structures of derivatives 7-8.

Then, the studied hybrids’ DFT exhibited bond lengths and angles, in contrast with those belonging to analogous compounds, single crystal X-ray [39,40], divulged an appreciated congruence, since the observed inconsistencies were <0.13 Å and <14.0° (RMSD = 0.05-0.07 and 4.8-5.2), respectively. The discrepancy might have resulted from that calculations were performed on a single gaseous particle with no columbic interactions [41] (Tables S2-S3).

Table S2

Table S3

Furthermore, the highest occupied molecular orbital-lowest unoccupied molecular orbital (HOMO-LUMO), FMOs, shapes, and energies have outstanding weight owing to their substantial influence on the molecule’s performance to donate or acquire electrons [42-45]. The analogs 3 and 4 displayed comparable FMOs, where their HOMO and LUMO had been overlaid mainly on the phenylacetamide fragment with minor contribution of the oxopropyl and methyloxazolyl groups (π- and π*-orbitals, respectively). The chlorophenyl hydrazonyl 5 had FMOs spread over the entire skeleton, while the pyrazolyl derivatives 6a-c displayed analogous composition, wherein HOMO was centered on methyloxazolylphenyl carboxamide with minor involvement of pyrazole ring, and LUMO has been confined on the chlorophenylpyrazole portion. Along with, the phenyl thiazolidinylidene 7 presented equivalent HOMO configuration, on methyloxazolylphenyl carboxamide, however its LUMO was span over the whole molecule. Contrarywise, the benzylidene conjugates 8a-b have demonstrated unlike configurations, wherein the HOMO and LUMO were localized on the phenyl thiazolidinylidene cyanoacetamide segment (π- and π*-orbitals, respectively) (Figure 3 and Figure S1).

Figure S1
The DFT/B3LYP/6-311++G(d,p) FMO’s 3D-plots for hybrids 3, 4, 5, 6a, 7, and 8a.
Figure 3.
The DFT/B3LYP/6-311++G(d,p) FMO’s 3D-plots for hybrids 3, 4, 5, 6a, 7, and 8a.

Successively, the explored hybrids’ HOMO energies (EH) have been presented ranged from -5.81 (for 6b) to -7.15 eV (for 3), while these of LUMO (EL) were ranged from -1.75 (for 4) to -2.90 eV(for 8b). Also, the energy gap (ΔEH-L) was in 3.27-5.00 eV zone, obeying the array: 5 < 8a < 8b < 6a < 6c < 6b < 7 < 4 < 3, which designated that the thiazolidinylidene derivatives 8a-b have lower gap than the pyrazolyl conjugates 6a-c (Figure 4). Furthermore, the energies of FMOs were manipulated in the estimation of certain chemical reactivity factors [43]. As the electronegativity and hardness disclosed that compound 3 had the uppermost standards (χ = 4.65 and η = 2.50 eV), while derivatives 6b and 5 presented the lowest merits (3.93 and 1.63 eV), respectively. However, the electrophilicity (ω) ranged from 3.55 (for 4) to 6.24 eV (for 8b), endorsing the undoubted electrophilicity character of these conjugates, ω > 1.5 eV, with superior tendency to give than receiving electrons, as electron donating (ω+) power < accepting (ω-) [46,47] (Table 1).

Representation of explored hybrids’ FMO energies obtained from DFT/B3LYP/6-311++G(d,p) calculations.
Figure 4.
Representation of explored hybrids’ FMO energies obtained from DFT/B3LYP/6-311++G(d,p) calculations.
Table 1. FMO’s energies and reactivity descriptors (eV) of investigated compounds.
Compound EH EL ΔEH-L χ η δ ω ω+ ω-
3 -7.15 -2.15 5.00 4.65 2.50 0.40 4.33 2.31 6.96
4 -6.14 -1.75 4.38 3.95 2.19 0.46 3.55 1.85 5.80
5 -6.11 -2.84 3.27 4.47 1.63 0.61 6.13 4.09 8.57
6a -5.85 -2.32 3.54 4.09 1.77 0.57 4.72 2.90 6.99
6b -5.81 -2.04 3.77 3.93 1.88 0.53 4.09 2.36 6.29
6c -6.02 -2.34 3.68 4.18 1.84 0.54 4.76 2.90 7.08
7 -6.27 -1.90 4.37 4.08 2.19 0.46 3.82 2.05 6.13
8a -5.90 -2.59 3.31 4.24 1.66 0.60 5.43 3.52 7.76
8b -6.23 -2.90 3.33 4.56 1.67 0.60 6.24 4.17 8.73

Besides, the intramolecular charge-transfer and electronegativity might be empathized further from the Mulliken’s atomic charges [48]. The data showed that the cyanoacetamide nitrogen NC(CNM) has acquired a close negative charge in 3-5 and 7-8 derivatives, -0.133 - -0.179. As well, the corresponding nitrogen of cyanoacetamide NH(CNM) and carboxamide NH(Car) in 6a-c have caught comparable negative charge, but slightly higher in the latter, -0.505 - -0.532 and -0.548 - -0.567, respectively. Whereas, the carbonyl oxygen of these functionals OC(CNM) and OC(Car) were ranged from -0.337 to -0.389. Moreover, in compound 3, the oxopropyl oxygen atoms O2C1(Prp), O1C1(Prp) and OC2(Prp) displayed close negative charge values, -0.366, -0.380 and -0.308, respectively, whereas the conversion of this group into oxazolyl ring led to reduction of the oxazolyl oxygen O1(Oxz) to be -0.280 - -0.284. Also, the oxazolyl nitrogen N3(Oxz) were negatively charged with relative merits, -0.283 -0.292. Alternatively, in conjugate 5, the nitrogen atoms of hydrazonyl group, N(HZ) and NH(HZ), have diverse negative charge, -0.094 and -0.385, respectively. When they participated in formation of pyrazolyl ring in 6a-c, N2(Prz) and N1(Prz), their charges were -0.162 - -0.169 and -0.031 -0.085, respectively. However, the aminopyrazolyl nitrogen NH2(Prz) has higher negative charge, -0.802 -0.814. Further, the thiazolidinylidene sulfur S1(Tz) and nitrogen N3(Tz) atoms presented opposite charges, the former was positively charged (0.173-0.194), while the latter has negative charge (-0.298 - -0.313) (Table S4).

Table S4

Else, the molecular polarizability (αtotal), hyperpolarizabilities (βtotal), and dipole moment (μ) were evaluated [49-51] to demonstrate the distribution of electronic density and softness, which principally influence the intermolecular interactions [52]. The analogs exposed rampant dipole moment (μ) swung from 3.85 D (5) to 7.78 D (8a), implying 2.80-5.66 times the standard substance urea [53] (Table 2). Also, the conjugates 4 and 8b unveiled the least and greatest values of polarizability (αtotal = 1.61-3.44×10-23 esu), while the analog 7 showed the lowest first-order hyperpolarizability, however both of 8a and 6c displayed the highest value (βtotal = 2.07-5.03×10-30 esu). In contrast to urea [53], the considered hybrids offered larger hyperpolarizability, 5.53-13.45 times (Table 2).

Table 2. The dipole moment (μ), polarizability (αtotal), polarizability anisotropy (Δα), and first-order hyperpolarizability (βtotal) of explored hybrids.
Compound μ (Debye) μ/μurea αtotal (esu×10-23) Δα (esu×10-24) βtotal (esu×10-30) βtotalurea
3 7.34 5.34 1.95 1.24 2.35 6.28
4 7.56 5.51 1.61 7.44 4.21 11.27
5 3.85 2.80 2.40 2.58 3.68 9.85
6a 3.87 2.82 2.63 5.00 2.47 6.61
6b 5.87 4.27 2.76 4.56 2.15 5.75
6c 5.27 3.84 2.62 3.94 5.03 13.45
7 5.21 3.80 2.63 5.60 2.07 5.53
8a 7.78 5.66 3.28 7.41 5.03 13.45
8b 5.20 3.79 3.44 9.15 2.85 7.63

3.3. Cytotoxic assay

The cytotoxicity of the synthesized analogs demonstrated varying degrees of effectiveness across the three human cancer cell lines (Panc-1, HT-29, MCF-7) and a normal lung fibroblast line (WI-38) (Table 3). Hybrids 3, 4, and 5 showed weak effectiveness, IC50 valued between 18.69±0.04 - 40.61±0.19 μM (HT-29), 23.57±0.36 - 37.65±0.14 μM (MCF-7), and 38.05±0.28 - 46.32±0.21 μM (Panc-1), and less selectivity towards normal WI-38 cells. Meanwhile, the analogs of fused thiazole/pyrazole moieties 6a-c enhanced anticancer potency, mainly against MCF-7 (IC50 = 7.74±0.33 - 14.09±0.02 μM), Panc-1 (IC50 = 18.12±0.44 - 23.36±0.15 μM), and keeping good selectivity toward WI-38 (IC50 = 70.64±0.54 - 86.07±0.48 μM). Hybrid 7 exhibited sustained results with sensible effectiveness toward all cancer lines (IC50 = 11.61±0.11, 20.48±0.30 and 23.82±0.31 μM against MCF-7, HT-29 and Panc-1, respectively). Moreover, analogs 8a and 8b, which feature both oxazole and thiazole motifs, emerged as the most potent derivatives, e.g., analog 8a achieved IC50 values (μM) 23.52±0.04 (Panc-1), 13.22±0.16 (HT-29), and 6.41±0.47 (MCF-7), while showing high selectivity to normal cells (82.76±0.39), demonstrating an auspicious therapeutic index. In addition, analog 8b displayed comparable results. In comparison, the drug reference (dasatinib) showed more potency as an expected (IC50: 2.35-4.63 μM) over higher selectivity towards WI-38 cells (IC50 = 87.28 μM), close to those of analogs 6b, 8a, and 8b. These results underline the effective role of the heterocyclic amendments in enhancing cytotoxic effectiveness and selectivity.

Table 3. In vitro cytotoxic activities of synthesized oxazole hybridized with pyrazole and/or thiazole analogs.
Analogs IC50 (μM)a
Panc‐1 HT‐29 MCF-7 WI-38
3 46.32±0.21 40.61±0.19 37.65±0.14 52.42±0.29
4 41.77±0.10 29.03±0.43 32.81±0.23 56.31±0.02
5 38.05±0.28 18.69±0.04 23.57±0.36 61.29±0.18
6a 20.66±0.35 17.13±0.34 13.11±0.40 74.55±0.23
6b 18.12±0.44 15.62±0.17 7.74±0.33 86.07±0.48
6c 23.36±0.15 19.28±0.28 14.09±0.02 70.64±0.54
7 23.82±0.31 20.48±0.30 11.61±0.11 77.12±0.06
8a 23.52±0.04 13.22±0.16 6.41±0.47 82.76±0.39
8b 16.43±0.18 16.70±0.23 9.82±0.27 85.04±0.45
Dasatinib 3.29 ± 0.06 4.63±0.19 2.35±0.21 87.28±0.27

N.B: aIC50 values are the mean ± SD of three separate experiments, Dasatinib is the reference of cytotoxic examinations.

3.4. Structural activity relationship

Among the structures of the newly produced oxazole hybridized with pyrazole and/or thiazole analogs and their cytotoxic effectiveness, hybrid 3 has the cyano-acetamide and benzoate moiety; the ester-linked cyano-acetamide shows moderate activity, suggesting that the flexible 2-oxopropyl group may reduce its binding efficiency likened to more rigid analog. However, for hybrid 4, which possess the oxazole-ring, the direct binding between the cyano-acetamide group and the oxazole-phenyl core may improve its potency, likely due to stronger H-bonding or enhanced planarity. Meanwhile, for hybrid 5, with the hydrazonyl moiety, the appearance of both hydrazonoyl-cyanide and p-chlorophenyl groups improvements may enhanced its cytotoxic effectiveness, probably through the increase of electron-withdrawal or further hydrophobic contacts. For hybrids 6a-6c with the pyrazole ring, conjugate 6a contains the pyrazole ring and p-chlorophenyl group grows its rigidity and lipophilicity, leading to improved its cytotoxic effectiveness. Conjugate 6b has the ethyl carboxylate group likely improves its solubility or its metabolic stability, contributing to higher effectiveness, specifically against MCF-7 cells. Moreover, conjugate 6c with the nitrile group is expected to produce growth in solubility or metabolic stability, leading to its cytotoxic effectiveness. Hybrid 7 possess the thiazolidinone core, introducing a potential Michael acceptor, which may covalently modify cellular targets, increasing activity. Together with the hybrids 8a and 8b, which have the 4-oxo-3-phenylthiazolidin-2-ylidene moiety, conjugate 8a has the p-methoxy-benzylidene group, which enhances potency, possibly by improving the membrane permeability or the electronic stabilization. Furthermore, conjugate 8b with the p-chloro-benzylidene substitution maintains strong cytotoxic effectiveness, with the p-chloro group contributing to hydrophobic and electron-withdrawing effects.

3.5. Carbonic anhydrase assay

The carbonic anhydrase inhibition activity of the synthesized oxazole derivatives hybridized with pyrazole and/or thiazole analogs revealed significant variations in inhibitory potency against CA IX and CA XII, depending on the specific heterocyclic substitutions (Table S5 and Figure 5). Among the tested analogs, analog 6c exhibited the strongest inhibition towards CA IX (IC50 = 0.011±0.015 μM), comparable to the reference inhibitor acetazolamide (AZA, IC50 = 0.054±0.019 μM), followed by analog 6a (IC50 = 0.018±0.001 μM). However, analog 8b (IC50 = 0.079±0.019 μM) and analog 8a (IC50 = 0.095±0.043 μM) displayed good CA IX inhibition. Analogs 3, 4, and 5, had IC50 values ranging from 0.102±0.052 to 0.331±0.047 μM. Meanwhile, according to CA XII inhibition, analog 7 demonstrated the highest activity (IC50 = 0.119±0.043 μM), followed by analog 8a (IC50 = 0.102±0.009 μM), analog 8b (IC50 = 0.114±0.009 μM), and analog 6c (0.127 μM), comparable to or approaching AZA (IC50 = 0.080±0.033 μM). Excitingly, analog 6b exhibited moderate inhibition against both CA IX and CA XII isoforms, possibly due to the existence of a amino-pyrazole moiety, which affects its selectivity and potency. Moreover, analogs 3-5 still demonstrated low inhibition, indicating baseline activity. The results showed that hybridization with specific thiazole/pyrazole skeletons obviously enhances carbonic anhydrase inhibitory profiles, mainly in analogs 6a, 6c, and 8 hybrids.

Table S5
Carbonic anhydrase inhibition of the of synthesized analogs.
Figure 5.
Carbonic anhydrase inhibition of the of synthesized analogs.

3.6. Molecular docking

As the docking data show, unique binding modes are associated with structural properties of the newly synthesized thiophene hybridized with pyrazole, pyridine, and/or thiazole analogs (Table 4). Analog 3 exhibited a moderate binding energy S = -5.4630 kcal/mol (RMSD 1.2173), forming H-acceptor bonds with LYS168 (2.91 Å) and ALA241 (3.18 Å), through its acetyl and nitrile groups, respectively, stabilizing the ligand within the binding pocket of PDB ID: 1V9E (Figure S2). Meanwhile, analog 4 displayed a slightly enhancement in the docking score (S = -5.5212 kcal/mol, RMSD = 1.9559), indicating a more energetic fit, with five different bindings: i) H-donors between both nitrogen (N7) and acetamide carbon (C13) with GLU234; ii) H-acceptor from oxygen (O14) with Ser1, iii) π-H interaction of the oxazole-ring with ASN10, and iv) π-cation stacking of the phenyl-ring with LYS168. These various bindings endorse a more intricate and stable binding mode (Figure S2). While analog 5 presented a lower binding score (S = -5.1156 kcal/mol, RMSD = 1.0878), based on only π-H interaction between the oxazole-ring and Asn10 (3.63 Å). The minimal binding score may cause the relatively weaker bindings compared to the other analogs (Figure S2). In the interim, analog 6a demonstrated an improved score (S = -5.7659 kcal/mol, RMSD = 1.7577), forming a single H-acceptor binding amongst the acetyl group oxygen (O29) and GLN91 (3.00 Å). This contact, while singular, might contribute to its adequate binding energy (Figure 6). However, analog 6b showed a modest energy (S = -5.3282 kcal/mol, RMSD = 1.4752), motivated by lone π-H stacking amid the pyrazole ring and GLY7 (4.15 Å), with limited binding but potentially suitable binding stability (Figure S2). Also, analog 6c showed a considerable docking energy (S = -5.6831 kcal/mol, RMSD = 1.6365), featuring a H-bond from the amino-group to THR226 (3.09 Å) and a π-H stacking amongst the oxazole-ring and SER98 (3.92 Å). The sequence of the polar and π-type bindings likely contributes to its moderate binding (Figure S2). Similarly, derivative 7 disclosed a score S = -5.6063 kcal/mol (RMSD = 1.0789), forming a single H-acceptor bond between the nitrile nitrogen (N29) and LEU48 at 3.45 Å. This binding, though explicit, appears to bargain moderate stabilization (Figure S2). Moreover, analog 8a offered a significant increasing in binding affinity (S = -6.5463 kcal/mol, RMSD 1.7158), over three interactions: an H-donor with GLU212 via the thiazolidinone sulfur (3.6 Å), an H-acceptor interaction with ASN185 via the nitrile nitrogen (3.57 Å), and a π-H stacking with PRO153 (3.89 Å). The presence of both heteroatom-based and aromatic bindings enhances its binding strength (Figure 6). Furthermore, analog 8b yielded a strong docking (S = -7.0940 kcal/mol, RMSD =1.5035), forming a π-π stacking between the oxazole ring and the aromatic system of HIS93 (3.82 Å). This specific aromatic stacking recommends high binding affinity through effective π-delocalization (Figure 6). Finally, acetazolamide (reference) established an expected good binding energy (S = -8.6276 kcal/mol, RMSD = 1.3597. It formed five bindings: H-donor bond with GLY62 (3.16 Å), two H-acceptor bonds via the sulphonyl-group with LYS168 and ASN230 (3.53 and 2.98 Å), an H–π stacking between the amino-group and HIS2 (3.31 Å), and a π-H stacking of the thiadiazole-ring with PHE229 (4.26 Å). The diversity and number of interactions recommend that (Aza) is offering a stable and energetically good fit (Figure S2).

Figure S2
Table 4. Docking results of new oxazole hybridized with pyrazole and/or thiazole analogs.
No S (Kcal/mol) RMSD ligand bindings with the amino-acid residues Binding types Bond length (Å)
3 -5.4630 1.2173

O 16 of the acetyl-group with LYS168

N 18 of the nitrile group with ALA241

H-acceptor

H-acceptor

2.91

3.18

4 -5.5212 1.9559

N 7 of the amid-group with GLU234

C 13 of acetamide-group with GLU234

O 14 of the amid-group with SER1

The oxazole-ring with ASN10

The phenyl-ring with LYS168

H-donor

H-donor

H-acceptor

π-H

π-cation

2.98

3.48

3.60

4.21

3.86

5 -5.1156 1.0878 The oxazole-ring with ASN10 π-H 3.63
6a -5.7659 1.7577 O 29 of the acetyl-group with GLN91 H-acceptor 3.00
6b -5.3282 1.4752 The pyrazole-ring with GLY7 π-H 4.15
6c -5.6831 1.6365

N 15 of the amino-group with THR226

The oxazole-ring with SER98

H-donor

π-H

3,09

3.92

7 -5.6063 1.0789 N 29 of the nitrile-group LEU48 H-acceptor 3.45
8a -6.5463 1.7158

S19 of the thiazolidinone-ring with GLU212

N 36 of the nitrile-group with ASN185

Thiazolidinone-ring with PRO153

H-donor

H-acceptor

π-H

3.6

3.57

3.89

8b -7.0940 1.5035 The oxazole-ring with the 5-ring of HIS93 π-π 3.82
Aza -8.6276 1.3597

N 9 of the amino-group with GLY62

O 10 of the sulphonyl-group with LYS168

O 11 of the sulphonyl-group with ASN230

N 9 of the amino-group with 5-ring of HIS2

Thiadiazole-ring with PHE229

H-donor

H-acceptor

H-acceptor

H-π

π-H

3.16

3.53

2.98

3.31

4.26

Docking images of analogs 6a, 8a and 8b with PDB: 1V9E obtained from MOE
Figure 6.
Docking images of analogs 6a, 8a and 8b with PDB: 1V9E obtained from MOE

3.7. Pharmacokinetic properties

In addition to the pharmacokinetic properties of newly oxazole hybridized with pyrazole and/or thiazole analogs (Table S6 and Figure S3) using the Swiss ADME program. A detailed summary of the ADME parameters, including molecular weight, lipophilicity (iLOGP), solubility, topological polar surface area (TPSA), hydrogen bond donors/acceptors, gastrointestinal (GI) absorption, blood–brain barrier (BBB) permeability, P-glycoprotein (Pgp) substrate status, and Lipinski’s rule compliance, has been provided in the Supporting Information (Table S6 and Figure S3).

Table S6

Figure S3

Overall, the compounds exhibited favorable in silico pharmacokinetic properties and satisfactory drug-likeness profiles. All analogs showed high GI absorption, suggesting good oral bioavailability potential, and none were predicted to permeate the BBB or behave as Pgp substrates, traits that reduce CNS-related side effects and multidrug resistance risk. Most analogs complied fully with Lipinski’s rule of five, with acceptable values for TPSA and hydrogen bonding, supporting their potential as orally active agents. Notably, analogs 3 and 4, being highly soluble with low molecular weight, possessed optimal oral drug-likeness characteristics. Analogs of higher molecular weight, such as 6a–c, 7, and 8a–b, were within acceptable physicochemical parameters and possessed uniform bioavailability scores (0.55), moderate solubility, and balanced lipophilicity. These findings suggest that their pharmacokinetic potential was not compromised by structural class. Together, these in silico results maintain the efficacy of these synthesized hybrids as lead compounds for development. Comparative figures and extensive ADME data are presented in Supporting Information for reference (Table S6 and Figure S3).

4. Conclusions

The precursors 2-cyano-N-(4-oxazolyl-phenyl)acetamide compound 4 and its corresponding N-(4-chlorophenyl)-hydrazone compound 5 were prepared and utilized for the synthesis of oxazole hybridized with pyrazole or thiazole compounds. The aryl-hydrazone compound 5 was reacted readily with alpha-halogenated reagents to yield the corresponding oxazole-pyrazole hybrids 6a-c. The cyanoacetamide analog 4 was treated with phenyl isothiocyanate and ethyl bromoacetate to furnish the corresponding oxazole-thiazolidin-4-one hybrid 7, which undergoes condensation with 4-anisaldehyde or 4-chlorobenzaldehyde to produce the oxazole-thiazole hybrids 8a and 8b. The examined derivatives exhibited diverse FMO’s configurations. The energy gap data indicated that the conjugates 5 and 3 captured the bottommost and uppermost merits (ΔEH-L = 3.27 and 5.00 eV, respectively), besides the thiazolidinylidene derivatives 8a-b have lower gap than the pyrazolyl conjugates 6a-c. However, the in vitro cytotoxic screening revealed that analogs 8a and 8b, particularly 8a, exhibited an eminent anticancer profile, combining potent activity against HT-29 and MCF-7 cancer cell lines with high selectivity toward normal fibroblasts (WI-38). These results support additional preclinical development of oxazole-thiazole hybrids as potential anticancer agents. Meanwhile, analogs 6c and 6a showed the highest inhibitory activity against CA IX, while analogs 7, 8a, and 8 b were the most effective against CA XII, comparable potency to the standard inhibitor acetazolamide (AZA). Moreover, the docking study revealed that all synthesized analogs showed varying degrees of binding affinity towards the 1V9E protein, with analogs 8b and 8a demonstrated strong binding through π-π and multiple polar interactions (S = -7.0940 and -6.5463 kcal/mol), supporting the potential of oxazole-based hybrids as promising candidates for further development as carbonic anhydrase inhibitors. Besides, the SwissADME estimation of the analogs showed high GI absorption, without any BBB permeability, and non-substrates of Pgp. Analogs 3 and 4 displayed low molecular weight and high solubility, recommending optimal oral bio-availability. These insights support that the synthesized analogs possess auspicious pharmacokinetic and drug-likeness properties suitable for extra preclinical progress.

Acknowledgment

The authors extend their appreciation to the Deanship of Research and Graduate Studies at King Khalid University for funding this work through Small Research Project under grant number RGP1/7/46

CRediT authorship contribution statement

Hana M. Abumelha, Gadeer R. S. Ashour: Data curation, formal analysis, methodology, and software; Matokah M. Abualnaja, Abeer Mogadem: Investigation and writing – review & editing; Renad Almughathawi, Arwa Alharbi: formal analysis, investigation, writing-original draft. Majid A. Bamaga, Nashwa M. El-Metwaly: Supervision and administration of research group.

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Data availability

All relevant data are within the manuscript and available from the corresponding author upon request.

Declaration of Generative AI and AI-assisted technologies in the writing process

The authors confirm that there was no use of artificial intelligence (AI)-assisted technology for assisting in the writing or editing of the manuscript and no images were manipulated using AI.

Supplementary data

Supplementary material to this article can be found online at https://dx.doi.org/10.25259/AJC_549_2025

References

  1. . Secondary metabolites: Secondary metabolic products consisting of C and H; C, H, and O; N, S, and P Elements; and O/N Heterocycles. In: Therapeutic Use of Medicinal Plants and their Extracts: Volume 2. Progress in Drug Research. Cham: Springer International Publishing; p. :165-309.
    [Google Scholar]
  2. , , , . Recent advances in the synthesis of five‐ and six‐membered heterocycles as bioactive skeleton: A concise overview. ChemistrySelect, 2022. 2022;7:e202103139. https://doi.org/10.1002/slct.202103139
    [Google Scholar]
  3. , , , . Catalytic synthesis and application of heterocyclic and heteroatom compounds: Recent advances. In Sustainable Green Catalytic Processes, Wiley Online Library 2024:79-96. https://doi.org/10.1002/9781394212767.ch4
    [Google Scholar]
  4. , , . Diversity of chemical skeletons. In: New Horizons in Natural Compound Research. Elsevier; p. :75-132. https://doi.org/10.1016/B978-0-443-15232-0.00023-0
    [Google Scholar]
  5. , , , , , . Recent advances in the synthesis and utility of thiazoline and its derivatives. RSC Advances. 2024;14:902-953. https://doi.org/10.1039/d3ra06444a
    [Google Scholar]
  6. . Targeting drug-resistant bacteria: The promise of oxazole-based antibacterial agents and their structure-activity relationships (SAR) Palestinian Medical and Pharmaceutical Journal. 2024;10:113-130. http://doi.org/10.59049/2790-0231.10.2.2308
    [Google Scholar]
  7. , , , , . Pharmaceutical and medicinal significance of sulfur (SVI)-Containing motifs for drug discovery: A critical review. European Journal of Medicinal Chemistry. 2019;162:679-734. https://doi.org/10.1016/j.ejmech.2018.11.017
    [Google Scholar]
  8. , , . Thiazole: A promising heterocycle for the development of potent CNS active agents. European Journal of Medicinal Chemistry. 2015;92:1-34. https://doi.org/10.1016/j.ejmech.2014.12.031
    [Google Scholar]
  9. , , , , . Synthesis, biological evaluation and molecular docking of new triphenylamine-linked pyridine, thiazole and pyrazole analogues as anticancer agents. BMC Chemistry. 2022;16:88. https://doi.org/10.1186/s13065-022-00879-x
    [Google Scholar]
  10. , , , , . Synthesis, antimicrobial evaluation, and molecular docking of new azole, azine, thiazole, and chromene derivatives. Polycyclic Aromatic Compounds. 2023;43:3429-3449. https://doi.org/10.1080/10406638.2022.2069135
    [Google Scholar]
  11. , , , , , , , , , . An overview on biological activities of oxazole, isoxazoles and 1,2,4-oxadiazoles derivatives. In: Nanostructured Biomaterials. Materials Horizons: From Nature to Nanomaterials. Singapore: Springer Singapore; p. :379-400. https://doi.org/10.1007/978-981-16-8399-2_10
    [Google Scholar]
  12. , , , , . Review on Therapeutic Diversity of Oxazole Scaffold: An Update. ChemistrySelect, 2024. 2024;9:e202403179.. https://doi.org/10.1002/slct.202403179
    [Google Scholar]
  13. , . A comprehensive review on biological activities of oxazole derivatives. BMC Chemistry. 2019;13:16. https://doi.org/10.1186/s13065-019-0531-9
    [Google Scholar]
  14. , . A review on potential heterocycles for the treatment of glioblastoma targeting receptor tyrosine kinases. Oncology Research. 2024;32:849-875. https://doi.org/10.32604/or.2024.047042
    [Google Scholar]
  15. , , , , , , , , , , . Recent advancement in drug design and discovery of pyrazole biomolecules as cancer and inflammation therapeutics. Molecules (Basel, Switzerland). 2022;27:8708. https://doi.org/10.3390/molecules27248708
    [Google Scholar]
  16. , , , , , , . Structural insights into pyrazoles as agents against anti‐inflammatory and related disorders. ChemistrySelect. 2022;7:e202104429. https://doi.org/10.1002/slct.202104429
    [Google Scholar]
  17. , , , , . Design and synthesis of new pyrazole hybrids linked to oxime and nitrate moieties as COX-2, EGFRL858R/T790M inhibitors and nitric oxide donors with dual anti-inflammatory/anti-proliferative activities. Bioorganic Chemistry. 2025;161:108563. https://doi.org/10.1016/j.bioorg.2025.108563
    [Google Scholar]
  18. , , , , , , , , . Novel pyrazole carboxylate derivatives as lonazolac bioisosteres with selective COX-2 inhibition: Design, synthesis and anti-inflammatory activity. Molecular Diversity. 2025;2025 https://doi.org/10.1007/s11030-025-11220-8
    [Google Scholar]
  19. , , , , , , , . Curcumin based pyrazole-thiazole hybrids as antiproliferative agents: Synthesis, pharmacokinetic, photophysical properties, and docking studies. Journal of Molecular Structure. 2023;1275:134633. https://doi.org/10.1016/j.molstruc.2022.134633
    [Google Scholar]
  20. , , , , . Thiazole derivatives in medicinal chemistry: Recent advancements in synthetic strategies, structure activity relationship and pharmacological outcomes. Journal of Molecular Structure. 2022;1266:133479. https://doi.org/10.1016/j.molstruc.2022.133479
    [Google Scholar]
  21. , , , , , , . Structure-activity relationship studies of thiazole agents with potential anti methicillin-resistance Staphylococcus aureus (MRSA) activity. Process Biochemistry. 2023;132:13-29. https://doi.org/10.1016/j.procbio.2023.06.013
    [Google Scholar]
  22. , , , , , , . New pyrazolyl-thiazolidinone/thiazole derivatives as celecoxib/dasatinib analogues with selective COX-2, HER-2 and EGFR inhibitory effects: Design, synthesis, anti-inflammatory/anti-proliferative activities, apoptosis, molecular modelling and ADME studies. Journal of Enzyme Inhibition and Medicinal Chemistry. 2023;38:2281262. https://doi.org/10.1080/14756366.2023.2281262
    [Google Scholar]
  23. , , , , , , , . Synthesis of new thiazole‐pyrazole analogues: Molecular modelling, antiproliferative/antiviral activities, and ADME studies. Chemical Biology Drug Design. 2025;105:e70090. https://doi.org/10.1111/cbdd.70090
    [Google Scholar]
  24. , . Recent advances of heterocycle based anticancer hybrids. Physical Sciences Reviews. 2023;8:3043-3065. https://doi.org/10.1515/psr-2021-0217
    [Google Scholar]
  25. , , , , . Therapeutic potential of oxazole scaffold: A patent review (2006–2017) Expert Opinion on Therapeutic Patents. 2018;28:783-812. https://doi.org/10.1002/slct.202403179
    [Google Scholar]
  26. , , . Biologically active pyrazole derivatives. New Journal of Chemistry. 2017;41:16-41. https://doi.org/10.1039/C6NJ03181A
    [Google Scholar]
  27. , , , , , , . Accelerating therapeutic protein design with computational approaches toward the clinical stage. Computational and Structural Biotechnology Journal. 2023;21:2909-2926. https://doi.org/10.1016/j.csbj.2023.04.027
    [Google Scholar]
  28. , , , , , . Novel benzenesulfonamide derivatives as potential selective carbonic anhydrase IX, XII inhibitors with anti-proliferative activity: Design, synthesis and in silico studies. Bioorganic Chemistry. 2024;153:107881. https://doi.org/10.1016/j.bioorg.2024.107881
    [Google Scholar]
  29. , , , , , , , , , , . Novel benzenesulfonamide derivatives linked to diaryl pyrazole tail as potential carbonic anhydrase II/VII inhibitors with anti-epileptic activity. European Journal of medicinal Chemistry. 2025;291:117619. https://doi.org/10.1016/j.ejmech.2025.117619
    [Google Scholar]
  30. , , , , , , , , , . Gaussian 09W, Revision A. 1. Wallingford, CT, USA: Gaussian. Inc.; . https://doi.org/10.1007/s42250-023-00674-8
  31. . Density‐functional thermochemistry. III. The role of exact exchange. Journal of Chemical Physics. 1993;98:5648-5652. https://doi.org/10.1063/1.464913
    [Google Scholar]
  32. , , . Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Physical Review. B, Condensed Matter. 1988;37:785-789. https://doi.org/10.1103/physrevb.37.785
    [Google Scholar]
  33. , . Pair-distribution function and its coupling-constant average for the spin-polarized electron gas. Physical Review. B, Condensed Matter. 1992;46:12947-12954. https://doi.org/10.1103/physrevb.46.12947
    [Google Scholar]
  34. , , . GaussView. Version. 2009;5
    [Google Scholar]
  35. , , , , , , , . Synthesis of bis‐(2‐thiazolyl)amine analogues and evaluation of their antibacterial, antioxidant and cytotoxic activities. ChemistrySelect. 2019;4:11726-11734. https://doi.org/10.1002/slct.201902272
    [Google Scholar]
  36. , , , , , , , , , , , , , , , . New sulfamethoxazole derivatives as selective carbonic anhydrase IX and XII inhibitors: Design, synthesis, cytotoxic activity and molecular modeling. Pharmaceuticals (Basel, Switzerland). 2022;15:1134. https://doi.org/10.3390/ph15091134
    [Google Scholar]
  37. , , , , , , . Synthesis, carbonic anhydrase inhibition, anticancer activity, and molecular docking studies of 1,3,4-oxadiazole derivatives. Molecular Diversity. 2023;27:193-208. https://doi.org/10.1007/s11030-022-10416-6
    [Google Scholar]
  38. , , , . α-Cyanocinnamide derivatives: A new family of non-peptide, non-sulfhydryl inhibitors of ras farnesylation. Bioorganic & Medicinal Chemistry. 1999;7:1727-1736. https://doi.org/10.1016/s0968-0896(99)00118-2
    [Google Scholar]
  39. , , , . Current development of 1,2,3-triazole derived potential antimalarial scaffolds: Structure- activity relationship (SAR) and bioactive compounds. European Journal of Medicinal Chemistry. 2023;259:115699. https://doi.org/10.1016/j.ejmech.2023.115699
    [Google Scholar]
  40. , , , , , , , . Synthesis, molecular modeling, and anticancer activity of new thiophene and thiophene-pyrazole analogues incorporating benzene-sulfonamide moiety as carbonic anhydrase isozymes (CA-IX and CA-XII) Journal of Molecular Structure. 2024;1295:136609. https://doi.org/10.1016/j.molstruc.2023.136609
    [Google Scholar]
  41. , , , . Natural bond orbital analysis, electronic structure, non-linear properties and vibrational spectral analysis of L-histidinium bromide monohydrate: A density functional theory. Spectrochimica Acta. Part A, Molecular and Biomolecular Spectroscopy. 2011;81:85-98. https://doi.org/10.1016/j.saa.2011.05.052
    [Google Scholar]
  42. , , , . Condensation of frontier molecular orbital fukui functions. The Journal of Physical Chemistry A. 2004;108:342-349. https://doi.org/10.1021/jp036416r
    [Google Scholar]
  43. , , . NBO, conformational, NLO, HOMO–LUMO, NMR and electronic spectral study on 1-phenyl-1-propanol by quantum computational methods. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy. 2015;137:306-320. https://doi.org/10.1016/j.saa.2014.08.039
    [Google Scholar]
  44. , , . Experimental and DFT insights into molecular structure and optical properties of new chalcones as promising photosensitizers towards solar cell applications. Applied Surface Science. 2018;452:337-351. https://doi.org/10.1016/j.apsusc.2018.05.007
    [Google Scholar]
  45. , , , , . Synthesis and characterization of new complexes of nickel (II), palladium (II) and platinum(II) with derived sulfonamide ligand: Structure, DFT study, antibacterial and cytotoxicity activities. Journal of Molecular Structure. 2018;1161:345-355. https://doi.org/10.1016/j.molstruc.2018.02.057
    [Google Scholar]
  46. , , , . Quantum study on the optoelectronic properties and chemical reactivity of phenoxazine-based organic photosensitizer for solar cell purposes. Theoretical Chemistry Accounts, 2022. 2022;141:1-14. https://doi.org/10.1007/s00214-022-02882-w
    [Google Scholar]
  47. , , . Applications of the conceptual density functional theory indices to organic chemistry reactivity. Molecules (Basel, Switzerland). 2016;21:748. https://doi.org/10.3390/molecules21060748
    [Google Scholar]
  48. , , , , , , , . Vibrational spectroscopic (FT-IR, FT-Raman, (1)H NMR and UV) investigations and computational study of 5-nitro-2-(4-nitrobenzyl) benzoxazole. Spectrochimica Acta. Part A, Molecular and Biomolecular Spectroscopy. 2013;102:99-113. https://doi.org/10.1016/j.saa.2012.09.032
    [Google Scholar]
  49. , , , , . Nanoring structure and optical properties of Ga8As8. Chemical Physics Letters. 2003;381:397-403. https://doi.org/10.1016/j.cplett.2003.09.115
    [Google Scholar]
  50. , , , . Molecular structure, spectroscopic studies and first-order molecular hyperpolarizabilities of p-amino acetanilide. Spectrochimica Acta. Part A, Molecular and Biomolecular Spectroscopy. 2008;71:355-367. https://doi.org/10.1016/j.saa.2008.01.010
    [Google Scholar]
  51. , , . Structure, stability, dipole polarizability and differential polarizability in small gallium arsenide clusters from all-electron ab initio and density-functional-theory calculations. Physical Review A: Atomic, Molecular and Optical Physics. 2008;77:013201. https://doi.org/10.1103/PhysRevA.77.013201
    [Google Scholar]
  52. , , , , , , , . Identification of potent inhibitors of NEK7 protein using a comprehensive computational approach. Scientific Reports. 2022;12:6404. https://doi.org/10.1038/s41598-022-10253-5
    [Google Scholar]
  53. , , , , . Structural, vibrational and theoretical studies of l-histidine bromide. Journal of Molecular Structure. 2008;888:180-186. https://doi.org/10.1016/j.molstruc.2007.11.056
    [Google Scholar]
Show Sections