Next Article in Journal
Retinal BMI1 Expression Preserves Photoreceptors in Sodium-Iodate-Induced Oxidative Stress Models
Next Article in Special Issue
Transcriptomic Landscape of Paclitaxel-Induced Multidrug Resistance in 3D Cultures of Colon Cancer Cell Line DLD1
Previous Article in Journal
Diclofenac Immune-Mediated Hepatitis: Identification of Innate and Adaptive Immune Responses at Clinically Relevant Doses
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Exploring Carboxamide Derivatives as Promising Anticancer Agents: Design, In Vitro Evaluation, and Mechanistic Insights

by
Manal M. Al-Najdawi
1,2,
Maysaa M. Saleh
2,
Dima A. Sabbah
3,
Rima Hajjo
3,4,5,
Hiba Zalloum
6,
Suha M. Abudoleh
7,
Duaa A. Abuarqoub
8,9,
Yusuf M. Al-Hiari
10,
Mohammad Yasin Mohammad
11,
Husam ALSalamat
6,12,
Hebah Mansour
13,
Nawzat D. Aljbour
1 and
Aktham H. Mestareehi
2,14,*
1
Department of Pharmaceutical Technology and Cosmetics, Faculty of Pharmacy Middle East University, Amman 11831, Jordan
2
Department of Applied Pharmaceutical Sciences and Clinical Pharmacy, Faculty of Pharmacy, Isra University, Amman 11622, Jordan
3
Department of Pharmacy, Faculty of Pharmacy, Al-Zaytoonah University of Jordan, Amman 11733, Jordan
4
Laboratory for Molecular Modeling, Division of Chemical Biology and Medicinal Chemistry, Eshelman School of Pharmacy, The University of North Carolina, Chapel Hill, NC 27599, USA
5
Advisory Board, Jordan CDC, Amman 11118, Jordan
6
Hamdi Mango Center for Scientific Research, The University of Jordan, Amman 11942, Jordan
7
Department of Basic Pharmaceutical Sciences, Faculty of Pharmacy, Middle East University, Amman 11831, Jordan
8
Department of Pharmaceutical and Biomedical Sciences, Faculty of Pharmacy and Medical Sciences, University of Petra, Amman 11196, Jordan
9
Cell Therapy Center, The University of Jordan, Amman 11942, Jordan
10
Department of Pharmaceutical Sciences, Faculty of Pharmacy, The University of Jordan, Amman 11942, Jordan
11
Pharmacological and Diagnostic Research Center, Faculty of Pharmacy, Al-Ahliyya Amman University, Amman 19328, Jordan
12
Department of Pharmacology, School of Medicine, Al-Balqa Applied University, Al-Salt 19117, Jordan
13
Department of Basic Pharmaceutical Sciences, Faculty of Pharmacy, Isra University, Amman 11622, Jordan
14
Department of Pharmaceutical Sciences, Eugene Applebaum College of Pharmacy and Health Sciences, Wayne State University, Detroit, MI 48202, USA
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2025, 26(12), 5903; https://doi.org/10.3390/ijms26125903
Submission received: 1 May 2025 / Revised: 11 June 2025 / Accepted: 12 June 2025 / Published: 19 June 2025
(This article belongs to the Special Issue Biological Hallmarks and Therapeutic Strategies in Cancer)

Abstract

Carboxamide derivatives are a promising class of compounds in anticancer drug discovery, owing to their ability to interact with multiple oncogenic targets and their favorable pharmacological profiles. In this study, we report the design, synthesis, and biological evaluation of a series of N-substituted 1H-indole-2-carboxamides as potential anticancer agents. The synthesized compounds were assessed for antiproliferative activity using the MTT assay against MCF-7 (breast cancer), K-562 (leukemia), and HCT-116 (colon cancer) cell lines, with normal human dermal fibroblasts included as a non-cancerous control. Several compounds demonstrated notable cytotoxicity and selectivity. Compounds 12, 14, and 4 exhibited potent activity against K-562 cells, with IC50 values of 0.33 µM, 0.61 µM, and 0.61 µM, respectively. Compound 10 showed the most significant activity against HCT-116 cells (IC50 = 1.01 µM) with a high selectivity index (SI = 99.4). Moderate cytotoxicity was observed against MCF-7 cells. To elucidate the mechanism of action, molecular docking and induced-fit docking studies were conducted against key cancer-related targets, including topoisomerase–DNA (PDB ID: 5ZRF), PI3Kα (4L23), and EGFR (3W32), revealing favorable binding interactions. Additionally, principal component analysis of molecular descriptors indicated that the compounds possess promising drug-like and lead-like properties, particularly compound 10. Overall, this study highlights N-substituted indole-2-carboxamides as promising scaffolds for further optimization. The integration of synthetic chemistry, biological assays, and computational modeling provides a robust foundation for the continued development of these compounds as potential anticancer agents.

1. Introduction

The global burden of cancer is expected to rise dramatically, with cases increasing from 14.1 million in 2012 to 21.6 million in 2023 and projected to reach 28.4 million by 2040 (a 47% surge from 2020). This rapid escalation solidifies cancer as the second leading cause of death worldwide following cardiovascular diseases. In 2020, the most frequently diagnosed cancers included breast, lung, and colorectal cancers (CRC), accounting for approximately 12.5%, 12.2%, and 10.7% of all confirmed cases, respectively [1,2]. Additionally, CRC is the second leading cause of cancer-related deaths globally [2,3]. Leukemia remains the most common cancer in children and continues to contribute significantly to cancer-related morbidity and mortality worldwide [4].
Cancer treatment encompasses a range of therapeutic approaches, including surgery, radiotherapy, hormonal therapies, monoclonal antibody-based treatments, and conventional cytotoxic agents [5,6]. Among these, cytotoxic drugs remain a cornerstone of oncology care, effectively inhibiting cell division and suppressing tumor growth. However, despite significant advancements in anticancer therapy, several challenges persist, including severe adverse effects, limited bioavailability, off-target cytotoxicity, and the development of drug resistance. Overcoming these obstacles necessitates the design and discovery of potent, selective chemotherapeutic agents with enhanced efficacy and reduced toxicity. Leveraging diverse molecular scaffolds and targeting multiple therapeutic pathways are crucial strategies for advancing cancer treatment and achieving these goals [7].
Tyrosine Kinase (TK) Vascular Endothelial Growth Factor Receptor-2 (VEGFR-2) is a cell surface protein that functions as a receptor for growth factors from the vascular endothelial growth factor (VEGF) family [8,9]. These growth factors and their receptors play a pivotal role in regulating blood vessel formation. Among them, VEGFR-2 is the primary mediator of angiogenesis, the process of new blood vessel development [10]. The inhibition of VEGFR-2 has emerged as a promising therapeutic strategy in cancer treatment, driving extensive research efforts in this area [11]. To disrupt the VEGF/VEGFR signaling pathway, various inhibitors have been developed, including small-molecule compounds and monoclonal antibodies [12]. These inhibitors are frequently used in combination with other cancer therapies to enhance treatment efficacy, reduce toxicity, and overcome drug resistance [13]. Notable VEGFR-2 inhibitors include small-molecule drugs such as Sunitinib, Sorafenib, and Pazopanib, as well as monoclonal antibodies like Ramucirumab and Bevacizumab [14,15]. These agents have demonstrated significant clinical promise and have been approved for treating multiple cancer types, including renal cell carcinoma, hepatocellular carcinoma, and colorectal cancer. While VEGFR-2 inhibitors represent a significant advancement in oncology, ongoing research aims to further improve their safety and therapeutic efficacy for better patient outcomes [16].
Topoisomerase inhibitors remain a cornerstone and highly promising in cancer therapy due to their ability to disrupt DNA replication and repair mechanisms, ultimately triggering cancer cell death. Among these, Camptothecin, a well-established topoisomerase I inhibitor, has demonstrated remarkable efficacy in both preclinical and clinical studies [17]. Correspondingly, Etoposide, which targets topoisomerase II, is widely used in the treatment of various malignancies, including small-cell lung cancer, testicular cancer, and ovarian cancer [18,19]. Other clinically approved inhibitors, such as Irinotecan and Topotecan, have also shown significant effectiveness against multiple cancer types [20]. Despite their therapeutic potential, these inhibitors are often associated with adverse effects, including myelosuppression and an increased risk of infections, which can pose challenges to their clinical application.
To address these challenges, ongoing research is focused on developing next-generation topoisomerase inhibitors with enhanced efficacy and reduced toxicity. Efforts are directed toward designing novel structural frameworks and optimizing pharmacokinetic properties to improve therapeutic outcomes. These advancements aim to refine the pharmacological profile of topoisomerase inhibitors while mitigating their associated adverse effects, ultimately improving their clinical applicability [21].
Privileged structures are molecular frameworks known for their versatility, exhibiting diverse biological activities and favorable pharmacokinetic properties. These structures play a crucial role in drug discovery, serving as effective starting points for optimizing structure–activity relationships in new compounds [22]. Their significance has been extensively documented, with examples such as the imidazole scaffold, which has demonstrated antimicrobial, antiviral, and anticancer properties [23]. By leveraging privileged structures, researchers can streamline the drug discovery process, reducing time and resource investment while utilizing well-established molecular frameworks. Furthermore, these structures inspire the design of novel drugs with enhanced activity and specificity. For instance, the benzodiazepine scaffold has contributed to the development of various therapeutics, including anxiolytics, hypnotics, and anticonvulsants, guiding the creation of innovative compounds with improved pharmacological profiles [24].
Furan and indole are well-established privileged structures extensively investigated for their potential in anticancer drug development. Indole, a highly versatile scaffold, plays a crucial role in the design of cancer therapeutics due to its ability to interact with target receptors through hydrogen bonding, hydrophobic interactions, and π–π stacking [25]. Likewise, furan has emerged as a promising framework for anticancer drug discovery, attributed to its diverse pharmacological activities and capacity to form covalent bonds with nucleophilic molecules. The incorporation of an amide bond into indole and furan derivatives has been shown to enhance their binding affinity to target receptors, thereby strengthening their anticancer efficacy [26]. Notably, clinically approved drugs such as Imatinib and Sunitinib, which feature indole or furan scaffolds, have demonstrated significant anticancer activity, further underscoring the therapeutic relevance of these privileged structures [25].
Indole-2-carboxamide derivatives have garnered significant interest in anticancer drug discovery due to their structural versatility and ability to interact with diverse biological targets involved in tumor progression. The indole moiety is a privileged scaffold known for its role in modulating critical signaling pathways such as apoptosis, cell cycle regulation, and topoisomerase inhibition. Moreover, the carboxamide linkage enhances molecular flexibility and offers opportunities for hydrogen bonding, which can improve target affinity and pharmacokinetic properties. In this study, we focused on the synthesis of N-substituted indole-2-carboxamides as potential anticancer agents based on their promising drug-like characteristics and previously reported cytotoxic activities. To further enhance their selectivity and potency, particularly against the MCF-7 cancer cell line, various substituents were introduced to explore structure–activity relationships. Future modifications may include the incorporation of electron-donating or sterically bulky groups, which have been reported to influence cell permeability and target specificity, thereby offering avenues for improved therapeutic performance.
Human cancer cell lines, including HCT-116 (colon cancer), MCF-7 (breast cancer), and K-562 (leukemia), serve as invaluable models in preclinical cancer research, offering critical insights into the biological activity and therapeutic potential of novel anticancer compounds. The MCF-7 cell line, derived from breast cancer and characterized by hormone receptor expression, is widely utilized to study breast cancer biology and evaluate the efficacy of potential therapeutic agents [27].
Research using MCF-7 cells in vitro provides a platform for assessing the mechanisms of action of novel compounds in breast cancer treatment [28]. Likewise, HCT-116, a colorectal cancer cell line, is frequently utilized as a model for studying colorectal cancer and assessing the effectiveness of new therapeutic agents. Data obtained from studies using HCT-116 cells can complement in vitro findings and provide deeper insights into compound activity against colorectal cancer. Additionally, the K-562 cell line, originating from chronic myeloid leukemia (CML), is extensively used to explore CML pathophysiology and assess the potential of emerging treatments. Studies utilizing K-562 cells aid in identifying promising therapeutic candidates and optimizing structure–activity relationships to improve drug efficacy [29]. Collectively, these cancer cell lines play a vital role in evaluating experimental compounds, providing essential insights that contribute to the advancement of bioactive compound revelation and therapeutic improvement.
In computational chemistry research, various proteins are extensively studied to assess the impact of novel compounds on cancers such as breast and colorectal cancers and leukemia. For breast cancer, estrogen receptors (ERs), particularly ERα and ERβ, serve as key targets due to their role in promoting the growth of hormone-dependent breast cancer. In colorectal cancer, proteins such as the epidermal growth factor receptor (EGFR), vascular endothelial growth factor receptor (VEGFR), and components of the mitogen-activated protein kinase (MAPK) signaling pathway are of significant interest, as they play crucial roles in tumor progression and metastasis [30].
Likewise, for leukemia, tyrosine kinase receptors, notably BCR-ABL, are critical therapeutic targets given their involvement in the proliferation and survival of chronic myeloid leukemia (CML) cells [31]. These proteins are essential targets in evaluating the potential of experimental compounds as anticancer agents. Computational techniques such as molecular docking are widely employed to predict the binding affinity and interaction of novel compounds with these target proteins, offering valuable insights into their therapeutic potential [32,33]. These findings contribute to the identification of promising drug candidates and the optimization of their structure–activity relationships, facilitating the advancement of targeted cancer therapies.
The integration of in vitro assays and molecular modeling techniques, such as docking studies, provides critical insights into the anticancer potential of indole carboxamide derivatives. The MTT assay, a widely used method for evaluating antiproliferative activity, enables the assessment of these compounds against various cancer cell lines. Testing synthesized derivatives on HCT-116, MCF-7, and K-562 cell lines yield essential data on their efficacy and potential mechanisms of action. These findings play a pivotal role in guiding further investigations aimed at optimizing the structure–activity relationship (SAR) of indole carboxamide derivatives, thereby advancing their development as promising candidates for anticancer therapeutics.
This study aims to investigate the anticancer potential of indole carboxamide derivatives against MCF-7, colon HCT-116, and K-562 cancer cell lines, with a particular emphasis on understanding their mechanisms of action. Through a comprehensive evaluation of their cytotoxicity, selectivity, and molecular interactions, the study seeks to establish a detailed structure–activity relationship (SAR) that will aid in the rational design of more potent and selective anticancer agents. Additionally, advanced computational modeling, including molecular docking and pharmacokinetic profiling, will be employed to predict binding affinities and optimize lead compounds for further preclinical development. The ultimate goal is to contribute to the development of novel therapeutic agents with enhanced efficacy and reduced toxicity, addressing the challenges of drug resistance and adverse effects associated with current treatments.

2. Results and Discussion

The synthetic approach described in this study demonstrates the efficient preparation of a series of N-substituted 1H-indole-2-carboxamides, featuring diverse aromatic moieties appended to the indole core. This synthetic strategy involved the initial synthesis of the key intermediate, indole-2-carbonyl chloride (2), which acts as a versatile precursor for the subsequent derivatization steps, facilitating the incorporation of various substituents. Indole-2-carbonyl chloride (2) was synthesized by refluxing indole-2-carboxylic acid with thionyl chloride (SOCl2) in the presence of a catalytic amount of N,N-dimethylformamide (DMF). This reaction efficiently converted the carboxylic acid group into the corresponding acid chloride, yielding a notably high conversion rate of 84%. The structural identity of the product was confirmed through characteristic spectroscopic data, ensuring the successful formation of the desired intermediate.
The acid chloride (2) served as a key building block for the synthesis of a diverse serious of N-substituted 1H-indole-2-carboxamides. The derivatization reactions proceeded efficiently, yielding the target compounds with reasonable to excellent yields (32% to 71%). These transformations were achieved through the reaction of acid chloride (2) with various aromatic amines under optimized conditions, as seen in Scheme 1, leading to the successful formation of the desired carboxamide derivatives.
Compound 16, N-(9,10-Dihydro-9,10-dioxoanthracen-1-yl)-2-furamide, was synthesized through the reaction of 1-aminoanthraquinone with furan-2-carbonyl chloride. This transformation proceeded smoothly and efficiently and yielded carboxamide 16 as a dark red solid, indicating its successful synthesis as shown in Scheme 2. The structural elucidation was supported by comprehensive spectroscopic analysis.
In conclusion of these syntheses, this study establishes an efficient and versatile synthetic strategy for the preparation of N-substituted 1H-indole-2-carboxamides with diverse aromatic functionalities, including the structurally distinctive compound 16. The successful synthesis and characterization of these derivatives highlight their potential as promising candidates for further investigation in drug discovery, particularly in the development of novel anticancer agents and other biologically relevant molecules.
The proliferation inhibitory activities of the carboxamide derivatives 4-16 were tested in vitro using the MTT assay [34,35] against the human cancer MCF-7, K-562, and HCT-116 cell lines in addition to normal human cells (dermal fibroblasts) to evaluate their putative selective cytotoxic activities. The concentrations at 50% cell proliferation inhibition (IC50) were determined from the dose–response curves, as illustrated in Figure 1, after 72 h exposure of the cells to carboxamide derivatives 4-16 and are presented in Table 1.
We have elaborated on the rationale behind selecting electron-withdrawing groups (EWGs) such as Cl, F, and NO2 for compounds 3, 5, and 7. These substituents were chosen based on their known ability to enhance binding interactions with biological targets through increased polarity and potential hydrogen bonding. Specifically, EWGs can influence the electronic distribution within the molecule, potentially improving interactions with active sites of oncogenic proteins. However, we acknowledge the importance of exploring the effects of electron-donating groups (EDGs) like CH3 or OCH3 on anticancer activity. EDGs can alter the electron density of the molecule differently, which may affect binding affinity and selectivity toward cancer cells. We have highlighted this as a limitation of the current study and propose it as a direction for future research. Furthermore, we have analyzed the differences in anticancer activity among the synthesized compounds, correlating the nature of the substituents with their observed cytotoxic effects. This analysis provides insights into how electronic properties of substituents influence biological activity, thereby guiding the design of more potent anticancer agents in subsequent studies.
The majority of the carboxamide compounds exhibited significant antiproliferative activity against the three carcinoma cell lines. Notably, the K-562 cell line demonstrated the highest sensitivity to carboxamides 12 (featuring a 1-anthraquinone moiety), 14 (with a 2-anthraquinone moiety), and 4 (bearing a p-chlorobenzene group), with sub-micromolar IC50 values of 0.33 µM, 0.61 µM, and 0.61 µM, respectively. The HCT-116 and MCF-7 cell lines displayed moderate sensitivity, with IC50 values ranging from 2.64 to 3.98 µM and 7.16 to 10.6 µM, respectively. Interestingly, the K-562 cell line exhibited considerable resistance to the antiproliferative effects of carboxamides 8 (featuring a p-nitrobenzene moiety), 10 (bearing a pyridine ring), and 16 (containing a furyl group), with IC50 values exceeding 100 µM, indicating weaker anticancer potential, and may require structural modifications for improved efficacy. Additionally, the MCF-7 cell line demonstrated exceptional resistance to carboxamide 8.
The pyridinyl carboxamide 10 exhibited the most potent antiproliferative activity against the HCT-116 cell line, with an IC50 = 1.01 µM and a high selectivity index (SI = 99.4), suggesting strong specificity toward colon cancer cells. Meanwhile, the anthraquinone-based carboxamides 12 (IC50 = 0.33 µM) and 14 (IC50 = 7.16 µM) demonstrated the highest proliferation inhibitory effects against the MCF-7and K-562 cell lines, respectively. Notably, compounds 10 and 16 displayed selective and potent antitumor activity against HCT-116, showing 69.6- and 10.6-fold selectivity over MCF-7, respectively. In contrast, the p-fluorobenzene carboxamide 6 exhibited the lowest antitumor activity against both the MCF-7 and K-562 cell lines, with IC50 values of 80.8 µM and 84.6 µM, respectively. Similarly, the weakest antiproliferative effect against the HCT-116 cell line was observed for p-nitrobenzene carboxamide 8, with an IC50 value of 32.0 µM.
The results revealed that the seven synthesized carboxamides (4-16) exhibited notable selectivity towards cancer cells compared to normal fibroblasts (IC50 > 100 µM), with selectivity indices (SIs) ranging from 3.1 to 99.4 (fibroblasts/colon cancer), 1.0 to 14.0 (fibroblasts/breast cancer), and 1.0 to 303 (fibroblasts/leukemia). In particular, carboxamides 12, 14, and 4 demonstrated the highest antitumor activities against K-562 cells, showing the best safety margins with selectivity indices (SIs) of 303 and 164, respectively. Additionally, the MTT assay results revealed that carboxamides 4-16 exhibit significant and selective inhibition of cell proliferation, as well as potent cytotoxic effects against a variety of tumor cell lines. These findings highlight their potential as promising therapeutic candidates for the treatment of several types of carcinomas, including leukemia and colon and breast cancers, suggesting they could be developed into effective anticancer agents.
To explain the anticancer activity of the verified compounds 4-16, we utilized the human topoisomerase–DNA (PDB ID: 5ZRF) complex (resolution = 2.30 Å) [36], phosphoinositide 3-kinase (PI3Kα) (PDB ID: 4L23) (RES = 2.50 Å) [37], and epidermal growth factor receptor (EGFR) (PDB ID: 3W32) (RES = 1.80 Å) complex [38], all obtained from the Protein Data Bank (PDB). These structures were analyzed to understand the molecular interactions between ligands 4–16 and the binding sites of 5ZRF, 4L23, and 3W32. Additionally, to explore the complexes of 5ZRF, 4L23, and 3W32/ligands 4-16 in the binding sites, we carried out induced-fit docking (IFD) [39,40,41,42] studies against 5ZRF, 4L23, and 3W32. The IFD results revealed that ligands 4-16 occupy the binding domains enclosed by topoisomerase residues and DNA nucleotide backbones (Figure 2A, Figure 3A and Figure 4A), as well as the PI3Kα and EGFR binding sites. Notably, the scaffold of ligand 16 closely resembles the template of co-crystalized ligands EVP (in 5ZRF), X6K (in 4L23), and W32 (in 3W32) (Figure 2B, Figure 3B and Figure 4B).
The IFD explores the conformational changes in the coordinates of 5ZRF, 4L23, and 3W32. Ligands are docked into the binding domains of these structures using Glide docking [39,40,41,42], followed by energetic minimization of the top ligand geometry along with the binding sites of 5ZRF, 4L23, and 3W32 using the Prime algorithm [39]. A redocking protocol is then applied to the minimized conformations of 5ZRF, 4L23, and 3W32. The flexibility of these proteins is assessed throughout the docking process. The synthesized carboxamide derivatives bind to the 5ZRF backbone, particularly to residues R503, Q778, DG7, DC8, DA12, and DG13, forming hydrogen bonds, as shown in Table 2. Furthermore, recorded computational [43,44,45,46] and experimental [36] data highlight the contribution of Q778, DG13, and DC8 in the interaction with the topoisomerase II-DNA complex. The docking data revealed that the synthesized derivatives bind with PI3Kα residues, specifically W780, E798, Y836, E849, V851, S854, and Q859, as illustrated in Table 3. Reported computational [47,48,49,50,51,52,53,54,55,56,57,58,59,60,61] and experimental [37,62] studies highlight the significance of W780, E798, Y836, E849, V851, S854, and Q859 in PI3Kα complex formation. Moreover, the derivatives bind with EGFR residues, distinctively K745, M793, T854, and D855 (Table 4). Disclosed computational [51,63,64,65] and experimental [38] studies emphasize the significance role of K745, M793, T854, and D855 in the formation of the EGFR complex.
It is important to note that a more negative binding score indicates a better binding affinity. The docking results against 5ZRF show that the three fused rings in ligands 12, 14, and 16 enhance the binding affinity. This finding aligns with previous studies [45,46] and recommends that incorporating at least four fused rings (like in Doxorubicin) or three fused rings combined with bulky or branched alkyl groups (such as in Mitoxantrone) is key to eliciting activity. Interestingly, tailoring the scaffold with p-NO2 (ligand 8) demonstrates different binding modes, suggesting that the rigidity of NO2 might help orient the ligand more effectively in the binding cleft. Overall, the results suggest that a system with three or four fused rings, coupled with a rigid moiety, is necessary for inducing activity. Specifically, DNA intercalation requires planar, rigid, and fused rings to disrupt the H-bond between DNA strands or to intercalate between them. Docking results for PI3Kα reveal that ligands 12 and 16 exhibit comparable affinities to the X6K co-crystallized ligand. However, the docking studies against EGFR indicate a better binding affinity, anticipating that derivatives may be selective for EGFR. In fact, ligands 8, 12, 14, and 16 show comparable binding affinities to W32.
To evaluate the performance of the IFD program, we compared the docked conformations of EVP in 5ZRF, X6K in 4L23, and W32 in 3W32 with their native poses in the crystal structures, as seen in Figure 5. The RMSD (Root Mean Square Deviation) for heavy atoms (all atoms except hydrogen) of EVP, X6K, and W32 between the IFD-generated poses and their native orientations were 0.1403, 0.1131, and 0.4937 Å, respectively. These results indicate that the IFD program is capable of accurately predicting the native pose within the crystal structure and effectively defining the ligand binding conformation.
All 2D AlvaDesc molecular descriptors [66] were calculated for the synthesized molecules 4-16 (Supplementary Table S1). These descriptor values were subsequently analyzed using principal component analysis (PCA). The PCA results showed that the molecules clustered into four main groups within the 2D space defined by the first two principal components, which explained the variance in the descriptors, as illustrated in Figure 6A. This clustering demonstrated that the synthesized compounds covered a broad range of physicochemical properties while maintaining acceptable drug-like and lead-like scores. Further evaluation using eight drug-like and two lead-like molecular descriptors from Kode Cheminformatics [66] indicated that compound 10 emerged as the most promising molecule, exhibiting the best drug-like and lead-like properties, as illustrated in the radar plot (Figure 6B). All drug-like scores (DLSs) and lead-like scores (LLSs) ranged from 0 to 1, where higher scores indicate greater similarity to the properties of actual drugs and leads.

3. Materials and Methods

3.1. Synthesis of Target Compounds and Materials

All chemicals, reagents and solvents were purchased from commercial sources ((SigmaAldrich (St. Louis, MO, USA), Riedel-de Haen (Seelze, Lower Saxony, Germany), Fluka (Buchs, St. Gallen, Switzerland), Ubichem (Budapest, Hungary), Aldrich (Milwaukee, Wisconsin, USA), Tedia (Fairfield, OH, USA), and Merck Corporation (Darmstadt, Hesse, Germany)) and were used in the experimentation without further purification. 1H-indole-2-carboxylic acid, 1H-furoyl-2-carboxylic acid, 2-aminopyridine, p-chloroaniline, p-fluoraniline, 1-aminoanthraquinone, phenyl magnesium chloride, thionyl chloride, and silica gel were bought from (Aldrich, USA). p-Nitroaniline, hydrochloric acid and potassium carbonate were bought from (Riedel-De Haen, Germany). Benzene, chloroform, ethanol, acetone, dimethyl sulfoxide, ethyl acetate, cyclohexane, pyridine, methanol, and dichloromethane were bought from (Tedia, USA). Dimethylformamide was bought from (Merck, Germany). Magnesium sulfate anhydrous was bought from (Scharlau, Spain). Filter paper was bought from (Whatman, England).

3.2. Spectroscopic and Characterization Techniques

Nuclear Magnetic Resonance (NMR) spectra were recorded using a Varian Oxford-300 (300 MHz) spectrometer (Palo Alto, CA, USA) at the Faculty of Pharmacy, University of Jordan. Chemical shifts are reported in parts per million (ppm) relative to tetramethylsilane (TMS) as an internal standard. Deuterated dimethyl sulfoxide (DMSO-d6) and deuterated chloroform (CDCl3) were used as solvents for sample preparation, unless otherwise specified. 1H NMR data are presented in the following format: chemical shift (ppm), multiplicity, coupling constant (J value) in hertz (Hz), number of protons, and the corresponding proton (s).
Infrared (IR) spectra were recorded using a Shimadzu 8400F FT-IR spectrophotometer (Kyoto, Japan) at the Faculty of Pharmacy, Al-Zaytoonah Private University. Samples were dissolved in chloroform (CHCl₃) and analyzed as thin solid films using a potassium bromide (KBr) disc (Merck, Darmstadt, Germany).
Melting points (m.p.) were determined using a Stuart Scientific Electrothermal Melting Point Apparatus (CA, USA) at the Faculty of Pharmacy, University of Jordan. Measurements were conducted in open capillaries and are reported as uncorrected values.
High-resolution mass spectra (HRMS) were acquired in positive or negative ion mode using the ElectroSpray Ionization (ESI) technique with collision-induced dissociation on a Bruker APEX-4 (7 Tesla), (Billerica, MA, USA) instrument at the Department of Chemistry, University of Jordan. Samples were dissolved in acetonitrile, diluted in spray solution (methanol/water 1:1 v/v + 0.1% formic acid), and infused via a syringe pump at a flow rate of 572 µL/min. External calibration was performed using an arginine cluster within the mass range m/z 175–871.
Elemental analysis (EA) of carbon (C), hydrogen (H), and nitrogen (N) was conducted using a Euro Elemental Analyzer (Milan, Italy) at the Faculty of Pharmacy, University of Jordan. Moreover, thin-layer chromatography (TLC) was performed on aluminum plates pre-coated with fluorescent silica gel GF254 (ALBET, Bremen, Germany). Silica gel (70–230 Mesh) from GCC (London, United Kingdom) was used for preparative TLC. The plates were visualized under a UV lamp using a Spectroline cabinet, Model CX-20 (Westbury, NY, USA).

3.3. Synthesis of Target Compounds and Data Analysis

  • Compound (2), Indole-2-carbonyl chloride
In a round-bottomed flask, 25 mL of anhydrous chloroform (CHCl3) was added, followed by the addition of indole-2-carboxylic acid (1 g, 6.2 mmol), thionyl chloride (3.6 mL, 50 mmol), and a small quantity of N,N-dimethylformamide (DMF). The reaction mixture was subjected to reflux conditions, maintaining a temperature within the range of 70 to 80 °C for a duration of 18 h to ensure the reaction was complete. Following the reaction, the excess thionyl chloride (SOCl2) was removed by suction, and the solvent was evaporated, resulting in the formation of acid chloride (2, 0.95 g, 84%) as a greenish-yellow crystalline precipitate. The product was characterized with an Rf value of 0.86 (cyclohexane/ethyl acetate, 7:3) and melting point of 115–120 °C. The 1H-NMR (300 MHz, DMSO-d6) showed the following peaks: δ = 11.90 (s, 1H, NH-indole), 7.65 (d, J = 7.8 Hz, 1H, H-7), 7.47 (d, J = 8.1 Hz, 1H, H-4), 7.26 (dd, J = 7.2, 7.8 Hz, 1H, H-6), 7.16 (s, 1H, H-3), 7.07 (dd, J = 7.5, 7.2 Hz, 1H, H-5) ppm (Hopkins et al., 2006) [67].
  • Compound (4), N-(4-Chlorophenyl)-1H-indole-2-carboxamide
A controlled addition of p-chloroaniline (0.41 g, 3.22 mmol) in 15 mL of dichloromethane (DCM) was carefully carried out into a solution containing indole-2-carbonyl chloride (2, 0.40 g, 2.23 mmol) and triethylamine (TEA, 1.52 g, 15.00 mmol). The reaction mixture was gently stirred at room temperature for 3 h. Following this, the solution was diluted with 15 mL of 1 N hydrochloric acid (HCl), leading to the formation of distinct liquid phases. The organic layer was successively washed with 15 mL of 2 N sodium hydroxide (NaOH) solution and 15 mL of brine (saturated sodium chloride solution). The solvents were then removed under reduced pressure, and the resulting residue was recrystallized from a mixture of methanol (CH3OH) and chloroform (CHCl3), yielding an off-white crystalline product (4, 0.36 g, 60%). The compound was extensively characterized, with the following analytical data providing comprehensive details on its structure and properties:
Rf value: 0.60 (cyclohexane/ethyl acetate, 6:4).
Melting point: 259 °C.
1H -NMR (300 MHz, DMSO-d6) showed the following peaks, δ ppm: 11.79 (br s, 1H, NH-indole), 10.35 (s, 1H, NHCO), 7.86 (d, J = 9.0 Hz, 2H, Ar-H), 7.69 (d, J = 8.1 Hz, 1H, Ar-H), 7.49 (s, 1H, Ar-H), 7.47–7.39 (m, 3H, Ar-H), 7.23 (dd, J = 8.1, 7.2 Hz, 1H, Ar-H), 7.08 (dd, J = 7.2, 7.8 Hz, 1H, Ar-H).
13C -NMR (DMSO-d6, δ ppm): 160.26 (CONH), 138.43, 137.36, 131.66, 129.09 (C-3′, 5′), 127.57, 127.44, 124.37 (CH-Ar), 122.26 (CH-Ar), 122.05 (C-2′, 6′), 120.44 (CH-Ar), 112.88 (CH-Ar), 104.60 (CH-Ar).
IR (thin film, cm−1): 3406 (NH-indole), 3318 (NHCO), 1651 (CO), 1593, 1539, 1493, 1400, 1315, 1238, 1192, 1099, 1015, 937, 826, 748.
Elemental analysis (C15H11N2OCl):
Theoretical values (calculated): C, 66.55%; H, 4.10%; N, 10.35%.
Experimental values (found): C, 66.40%; H, 3.92%; N, 10.12%.
  • Compound (6), N-(4-Florophenyl)-1H-indole-2-carboxamide
A blend of p-fluoroaniline (5, 0.35 g, 2.12 mmol) and triethylamine (TEA, 1.52 g, 15.00 mmol) was meticulously added dropwise to a solution containing indole-2-carbonyl chloride (2, 0.38 g, 2.12 mmol) in dichloromethane (DCM, 15 mL). The resulting mixture was stirred at room temperature (RT) for 3 h, after which it was diluted with 15 mL of 1 N hydrochloric acid (HCl), leading to the formation of distinct liquid layers. The organic layer was then sequentially washed with 15 mL of 2 N sodium hydroxide (NaOH) solution and 15 mL of brine. Any residual moisture was removed using anhydrous magnesium sulfate. The solvent was evaporated under reduced pressure, and the resulting crude product was recrystallized from a mixture of methanol (CH₃OH) and chloroform (CHCl₃), affording the title compound (6, 0.25 g, 46%) as a faint yellow solid. The compound was characterized by the following analytical data:
Rf value: 0.54 (cyclohexane/ethyl acetate, 7:3).
Melting point: 245 °C (decomposed).
1H -NMR (300 MHz, DMSO-d6, δ ppm): 11.78 (br s, 1H, NH-indole), 10.80 (s, 1H, NHCO), 7.89–7.79 (m, 2H, Ar-H), 7.67 (d, J = 7.8 Hz, 1H, Ar-H), 7.48 (d, J = 8.0 Hz, 1H, Ar-H), 7.44–7.39 (m, 1H, Ar-H), 7.28–7.15 (m, 3H, Ar-H), 7.11–7.01 (m, 1H, Ar-H).
13C-NMR (DMSO-d6, δ ppm, rotamers observed): 160.28, 157.09 (d, 1J C-F = 283 Hz, C-4′, CH-Ar), 160.14 (CONH), 157.09, 137.30, 135.82, 131.82, 127.47, 124.25 (CH-Ar), 122.44, 122.19 (d, 2J C-F = 19 Hz, C-3′, 5′, CH-Ar), 122.34 (CH-Ar), 120.38 (CH-Ar), 115.89, 115.59 (d, 3J C-F = 22.13 Hz, C-2′, 6’, CH-Ar), 112.86 (CH-Ar), 104.40 (CH-Ar).
IR (thin film, cm−1): 3422 (NH-indole), 3337 (NHCO), 3271, 1639 (CO), 1543, 1508, 1408, 1339, 1315, 1223, 1188, 1096, 833, 818, 748.
HRMS (ESI, negative mode): m/z (M+-H+) 253.07826 (C15H10N2OF) requires 253.07772.
Elemental analysis (C15H11N2OF):
Theoretical values: C, 70.86%; H, 4.36%; N, 11.02%.
Experimental values: C, 70.97%; H, 4.15%; N, 10.90%.
  • Compound (8), N-(4-Nitrophenyl)-1H-indole-2-carboxamide
A solution of indole-2-carbonyl chloride (2, 0.45 g, 2.51 mmol) in dichloromethane (DCM, 15 mL) was gradually added to a mixture containing p-nitroaniline (7, 0.45 g, 3.26 mmol) and triethylamine (TEA, 1.52 g, 15.00 mmol). The mixture was stirred at room temperature for 3 h and then diluted with 15 mL of 1 N hydrochloric acid (HCl), resulting in the formation of two distinct liquid phases. The organic layer was sequentially washed with 15 mL of 2 N sodium hydroxide (NaOH) solution and 15 mL of brine, followed by drying with anhydrous magnesium sulfate. The solvent was removed under reduced pressure, and the crude product was recrystallized from a mixture of methanol (CH₃OH) and chloroform (CHCl₃), yielding the title compound (8, 0.50 g, 71%) as a green-brown solid. The characterization data of the compound are listed below:
Rf value: 0.47 (CHCl3/CH3OH/CH3CO2H, 94:5:1).
Melting point: > 290 °C (decomposed).
1H-NMR (300 MHz, DMSO-d6, δ ppm): 11.84 (br s, 1H, NH-indole), 10.74 (br s, 1H, NHCO), 8.28 (d, J = 9.2 Hz, 2H, H-3′, 5′, Ar-H), 8.11 (d, J = 9.2 Hz, 2H, H-2′, 6′, Ar-H), 7.71 (d, J = 8.0 Hz, 1H, Ar-H), 7.58–7.45 (m, 2H, Ar-H), 7.76 (dd, J = 7.4, 7.5 Hz, 1H, Ar-H), 7.09 (dd, J = 7.6, 7.3 Hz, 1H, Ar-H).
13C-NMR (DMSO-d6, δ ppm): 160.67 (CONH), 145.84, 142.79, 137.68, 131.15, 127.38, 125.31 (CH-Ar), 124.80 (CH-Ar), 122.45 (CH-Ar), 120.92 (CH-Ar), 120.02 (CH-Ar), 112.97 (CH-Ar), 105.69 (CH-Ar).
Elemental analysis (C15H11N3O3):
Theoretical values: C, 64.05%; H, 3.94%; N, 14.94%.
Experimental values: C, 63.97%; H, 3.89%; N, 14.78%.
  • Compound (10), N-(Pyridin-2-yl)-1H-indole-2-carboxamide
Indole-2-carbonyl chloride (2, 0.42 g, 2.34 mmol) was combined with 2-aminopyridine (9, 0.42 g, 4.51 mmol) in a flask containing 25 mL of N,N-dimethylformamide (DMF) and 2.0 mL of triethylamine (TEA, 7.5 mmol). The reaction mixture was heated at 150 °C for 2 days. Afterward, the solvent was removed under reduced pressure, and the residue was dissolved in chloroform (CHCl3) and extracted with acidic water. The organic layer was dried over anhydrous magnesium sulfate (MgSO4), and the solvent was evaporated. The product was further purified by column chromatography, using chloroform as the eluent, yielding the target compound (10, 180 mg, 32%) as a faint yellow solid. The compound was thoroughly characterized, and the results of the analyses are summarized below:
Rf: 0.53 (CHCl3/CH3OH/CH3CO2H, 94:5:1).
Melting point: 260 °C (decomposed).
1H-NMR (300 MHz, DMSO-d6): rotamers, δ = 13.75 (br s, 1H, NH-indole), 11.8 + 10.91 (2 br s, 1H, NHCO), 8.61–7.78 (m, 3H, Ar-H), 7.75–7.35 (m, 3H, Ar-H), 7.30–6.80 (m, 3H, Ar-H) ppm.
13C-NMR (DMSO-d6): δ = 160.11 (CONH), 144.10, 137.60, 137.24, 133.32, 127.56, 124.76, 113.56, 112.45, 111.90, 108.37, 104.99.
IR (thin film): ν = 3449 (NH-indole), 3422 (NHCO), 3144, 2924, 1701 (CO), 1558, 1450, 1366, 1346, 1200, 1080, 748 cm−1.
HRMS (ESI, negative mode): m/z (M+-H+) 236.08294 (C14H11N3O) requires 236.08239.
Elemental analysis (C14H11N3O):
Theoretical values (calculated): C, 70.87%; H, 4.67%; and N, 17.71%.
Experimental values (found): C, 70.62%; H, 4.88%; and N, 17.55%.
  • Compound (12), N-(9,10-Dihydro-9,10-dioxoanthracen-1-yl)-1H-indole-2-carboxamide
A mixture of 1-aminoanthraquinone (11, 0.40 g, 1.8 mmol) and indole-2-carbonyl chloride (2, 0.40 g, 2.2 mmol) was refluxed at 120 °C for 18 h using an air condenser. After completion, 15 mL of 1,4-dioxane was added, and the reaction mixture was stirred at room temperature for an additional 24 h. This procedure was adapted from Al-Najdawi et al. [68]. The product was isolated by filtration, washed with ethanol, dried, and purified through recrystallization from chloroform, yielding the title compound (12, 0.30 g, 45%) as a brown-green solid. The compound was characterized, and the following analytical data were obtained:
Rf: 0.59 (cyclohexane/ethyl acetate, 6:4).
Melting point: > 350 °C (decomposed).
1H-NMR (300 MHz, DMSO-d6) rotamers: δ = 13.23 (br s, 1H, NH-indole), 12.08 (s, 1H, NHCO), 9.15 + 9.13 (2s, 1H, Ar-H), 8.50–7.76 (m, 5H, Ar-H), 7.62–6.89 (m, 5H, Ar-H) ppm.
13C-NMR (DMSO-d6): δ = 182.60 (CO-ketone), 176.31 (CO-ketone), 163.21 (CONH), 134.37 (CH-Ar), 133.90 (CH-Ar), 132.90, 131.68, 129.37, 128.06 (CH-Ar), 127.69, 127.13, 126.48 (CH-Ar), 125.25, 124.47, 123.85 (CH-Ar), 122.37 (CH-Ar), 121.83, 120.90 (CH-Ar), 118.16 (CH-Ar), 116.20 (CH-Ar), 115.24 (CH-Ar), 112.94 (CH-Ar), 104.18 (CH-Ar) ppm.
IR (thin film): ν = 3314 (NH-indole), 3213 (NHCO), 3125, 1667 (br CO), 1636 (CONH), 1578, 1531, 1412, 1312, 1269, 1238, 1173, 1015, 806, 741, 705 cm⁻¹.
HRMS (ESI, negative mode): m/z (M+-H+) 365.09317 (C23H13N2O3) requires 365.09262.
Elemental analysis (C23H₁4N2O3):
Theoretical values (calculated): C, 75.40%; H, 3.85%; and N, 7.65%.
Experimental values (found): C, 75.19%; H, 4.01%; and N, 7.96%.
  • Compound (14), N-(9,10-Dihydro-9,10-dioxoanthracen-2-yl)-1H-indole-2-carboxamide
A mixture of 2-aminoanthraquinone (13, 0.45 g, 2.0 mmol) and indole-2-carbonyl chloride (2, 0.50 g, 2.8 mmol) was refluxed at 120 °C for 18 h using an air condenser. After completion, 15 mL of 1,4-dioxane was added, and the reaction mixture was stirred at room temperature for an additional 24 h. The resulting product was isolated by filtration, thoroughly washed with ethanol, dried, and further purified via recrystallization from chloroform, yielding the title compound (14, 0.46 g, 63%) as a pale green solid. The compound was thoroughly characterized, and the following analytical data were recorded:
Rf: 0.41 (cyclohexane/ethyl acetate, 6:4).
Melting point: > 350 °C (decomposed).
1H-NMR (300 MHz, DMSO-d6): δ = 11.88 (br s, 1H, NH-indole), 10.79 (s, 1H, NHCO), 8.68 (d, J = 3.0 Hz, 1H, Ar-H), 8.39 (d, J = 9.0 Hz, 1H, Ar-H), 8.31–8.11 (m, 3H, Ar-H), 8.0–7.83 (m, 2H, Ar-H), 7.75 (d, J = 6.0 Hz, 1H, Ar-H), 7.56 (s, 1H, Ar-H), 7.46 (d, J = 6.0 Hz, 1H, Ar-H), 7.37–7.03 (m, 2H, Ar-H) ppm.
13C-NMR (DMSO-d6): δ = 183.0 (CO-ketone), 181.86 (CO-ketone), 160.67 (CONH), 145.19, 137.65, 135.03 (CH-Ar), 134.68 (CH-Ar), 134.58, 133.67, 131.20, 128.89 (CH-Ar), 128.50, 127.43, 127.21 (CH-Ar), 127.13 (CH-Ar), 124.94 (CH-Ar), 124.77 (CH-Ar), 122.48 (CH-Ar), 120.58 (CH-Ar), 117.21 (CH-Ar), 112.96 (CH-Ar), 105.59 (CH-Ar) ppm.
IR (thin film): ν = 3402 (NH-indole), 3321 (NHCO), 3117, 3067, 1667 (br CO), 1651 (CONH), 1589, 1535, 1416, 1331, 1292, 1238, 954, 840, 755, 710 cm−1.
HRMS (ESI, negative mode): m/z (M+-H+) 365.09317 (C23H13N2O3) requires 365.09262.
Elemental analysis (C23H14N2O3):
Theoretical values (calculated): C, 75.40%; H, 3.85%; and N, 7.65%.
Experimental values (found): C, 75.72%; H, 3.74%; and N, 7.87%.
  • Compound (16), N-(9,10-Dihydro-9,10-dioxoanthracen-1-yl)-2-furamide
A mixture of 1-aminoanthraquinone (11, 0.40 g, 1.8 mmol) and furan-2-carbonyl chloride (15, 0.45 g, 3.5 mmol) was refluxed at 120 °C for 18 h under an air-cooled condenser. Afterward, 15 mL of 1,4-dioxane was added, and the reaction mixture was stirred at room temperature for an additional 24 h. The resulting solid product was purified through filtration, washed with water, and dried. Further purification via recrystallization from chloroform yielded the title compound (16, 0.18 g, 32%) as a dark red solid. The compound underwent comprehensive characterization, with the following analytical data obtained:
Rf: 0.67 (CHCl3/CH3OH, 9:1).
Melting point: 280 °C (decomposed).
1H-NMR (300 MHz, DMSO-d6) rotamers: δ = 13.18 (br s, 1H, NHCO), 9.15 (br s, 1H, Ar-H, H3-furan), 8.35–8.10 (m, 2H, Ar-H), 7.94–7.72 (m, 4H, Ar-H), 7.46–7.12 (m, 2H, Ar-H), 6.95–6.61 (m, 1H, Ar-H) ppm.
13C-NMR (DMSO-d6) rotamers: δ = 192.82 (CO-ketone), 186.44 (CO-ketone), 161.22 (CONH), 148.09, 147.51, 147.22, 136.33, 135.25, 127.62, 126.96, 125.71, 122.59, 118.83, 118.18, 116.91, 113.52, 112.77, 112.56 ppm.
Elemental analysis (C19H11NO4):
Theoretical values (calculated): C, 71.92%; H, 3.49%; and N, 4.41%.
Experimental values (found): C, 71.55%; H, 3.63%; and N, 4.66%.

3.4. Cell Proliferation Study

3.4.1. Cells and Cell Culture Conditions

The human colorectal carcinoma (HCT-116), human breast adenocarcinoma (MCF-7), human chronic myelogenous leukemia (K-562), and human dermal fibroblast normal cell lines were purchased from were purchased from the American Type Culture Collection (ATCC) (Manassas, VA, USA) and stored in liquid nitrogen. Cells were passaged twice weekly upon reaching 70–80% confluency, and only low passage numbers (≤ 10) were used in all experiments [6]. Mycoplasma contamination was monitored using the Lonza MycoAlert™ mycoplasma detection kit. Cells were sub-cultured and maintained with appropriate media: RPMI-1640 medium (Buffalo, New York, USA) was used to grow MCF-7, K-562 and HCT-116 cancer cells, while fibroblast cells were cultured in Iscove’s Modified Dulbecco’s medium (IMDM) (EuroClone S.p.A., Pero, MI, Italy). Both media were supplemented with 10% (v/v) heat-inactivated fetal bovine serum (FBS) (Gibco, CA, USA), 2 mM of L-Glutamine, 100 units/mL of Penicillin, and 100 μg/mL of Streptomycin (EuroClone, Italy). Cells were maintained at 37 °C in a humidified incubator with 5% CO2 (NuAire, Plymouth, MN, USA) [69,70,71].

3.4.2. MTT Assay

The MTT colorimetric assay utilized in this study was previously described [69,70,71,72,73,74] and adapted from Mosmann [34]. The MTT regent [3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyl tetrazolium bromide] was employed to evaluate the antiproliferative activities of carboxamides 4-16 against MCF-7, K-562, HCT-116, and dermal normal fibroblast cell lines. Cells (6–8 × 103 per well, 180 µL) were seeded into 96-well plates (Corning, New York, USA) and allowed to adhere overnight at 37 °C in a humidified 5% CO₂ incubator. Fresh top stock solutions of the test compounds (10 mM in DMSO) were then freshly prepared, followed by serial dilutions in the appropriate culture media before being added to the seeded cells. Control wells received media alone (20 µL per well), ensuring that the final DMSO concentration in the wells did not exceed 1%. Initial screening for potential antiproliferative activity was carried out using two concentration points (10 µM and 50 µM) in triplicate over a 72-hour incubation period. Compounds demonstrating more than 50% inhibition of cell proliferation were further evaluated across a broader concentration range for all tested cell lines. Following 72 h of exposure, MTT solution (2 mg/mL in PBS, 50 µL per well) was added, and plates were incubated for an additional 4 h. The supernatant in each well was then removed, and the resulting formazan crystals were solubilized by adding DMSO (150 µL per well). Absorbance, measured as the optical density (OD), was recorded at 570 nm using a microplate reader (Thermo Fisher Scientific, Waltham, MA, USA). The detected signal intensity corresponds to metabolic activity, which directly correlates with the viable cell number.

3.4.3. Statistical Analysis

All experiments were performed in triplicate, and the results are presented as the mean ± standard deviation (SD). The IC50 values for antiproliferative activity were calculated using nonlinear regression analysis in GraphPad Prism software (GraphPad Prism version 9.0.0 for Windows, GraphPad Software, San Diego, California USA). Additionally, a two-way ANOVA test was conducted to analyze significant variations in IC50 values, with a significant level set at α = 0.05.

3.5. Molecular Modeling and Cheminformatics Methods

3.5.1. Protein Preparation

The X-ray crystal structures of human topoisomerase–DNA (PDB ID: 5ZRF) [36], phosphoinositide 3-kinase (PI3Kα) (PDB ID: 4L23) (resolution = 2.50 Å) [37], and epidermal growth factor receptor (EGFR) (PDB ID: 3W32) (resolution = 1.80 Å) complex [38] were downloaded from the Protein Data Bank (PDB). Protein preparation was carried out using MAESTRO software (version 2022.2) [39], which involved filling in missing residues, capping the N- and C-termini, minimize hydrogen atoms, and optimizing the hydrogen bond network. Furthermore, side-chain perturbations were applied to reduce steric clashes and enhance structural accuracy.

3.5.2. Ligand Structure Preparation

The validated ligand molecules were constructed using the Build panel in MAESTRO, based on the template of the co-crystallized ligand (EVP) in the 5ZRF structure. Further energy minimization and conformational refinement of the ligands were performed using the LigPrep module in MAESTRO [39] to ensure optimal geometry and stability for subsequent docking studies.

3.5.3. Induced-Fit Docking (IFD)

The co-crystallized ligands EVP/5ZRF, X6K/4L23, and W32/3W32 were designated as centroids within their respective binding sites. To ensure optimal ligand–receptor interactions, the van der Waals scaling factors for both the receptor and ligand were adjusted to 0.5, allowing for appropriate relaxation and flexibility during docking. All other docking parameters were maintained at their default settings. The ligand pose with the highest XP Glide binding score was selected for further analysis. Docking scores were reported in kcal/mol, with more negative values indicating stronger binding affinities.

3.5.4. Molecular Descriptors

All molecular structures were sketched in ChemDraw Professional 16.0 [75] and saved in SDF file format. The structures were then standardized according to the methods described by Hajjo et al. [76]. Subsequently, 2D molecular descriptors were calculated using alvaDesc software (version 1.0.22) from Kode Cheminformatics to facilitate further computational analysis [66].

3.5.5. Principal Component Analysis (PCA)

Principal component analysis (PCA) was conducted on all synthesized active molecular structures utilizing the calculated 2D alvaDesc molecular descriptors. This analysis was performed using alvaDesc software (version 1.0.22) from Kode Cheminformatics, enabling a systematic exploration of molecular diversity, the identification of structural patterns, and the assessment of correlations among key molecular properties [66].

4. Conclusions

This study successfully designed, synthesized, and evaluated a series of N-substituted 1H-indole-2-carboxamides, demonstrating their potential as novel anticancer agents. The synthesized compounds exhibited significant antiproliferative activity across various cancer cell lines, with a notable selectivity for cancerous cells over normal fibroblasts. Among the compounds, 4, 12, and 14 were particularly effective against the K-562 cell line, showing strong cytotoxicity, while compound 10 exhibited the highest potency against the HCT-116 cell line. The MCF-7 cells showed moderate sensitivity, highlighting the compounds’ broad applicability across different types of cancer. Molecular modeling and computational studies further enhanced our understanding of the mechanisms of action, revealing the critical binding interactions between these derivatives and key oncogenic targets, including topoisomerase–DNA, PI3Kα, and EGFR. These computational predictions were consistent with experimental findings, providing a strong basis for the observed anticancer activity. The data suggest that these derivatives effectively disrupt the function of these targets, which are integral to cancer cell proliferation and survival, thereby presenting a multi-target approach to cancer treatment. Furthermore, the principal component analysis of the molecular descriptors revealed that the synthesized compounds occupy a diverse space of physicochemical properties, indicating their potential for drug-like and lead-like characteristics. Among them, compound 10 emerged as the most promising candidate, displaying both excellent drug-like and lead-like properties, positioning it as a favorable candidate for further development.
This work not only supports the potential of these indole carboxamide derivatives as anticancer agents but also underscores the importance of combining experimental and computational strategies in drug discovery. The strong correlation between in vitro results and computational predictions demonstrates that these compounds may offer a novel therapeutic approach with fewer off-target effects compared to traditional chemotherapies. The findings from this study provide a foundation for future research aimed at optimizing the efficacy, specificity, and pharmacokinetic properties of these derivatives, with the ultimate goal of advancing their clinical development as effective and less toxic anticancer agents. This work contributes significantly to the ongoing global efforts to discover and develop new, effective, and less toxic anticancer therapies. By focusing on selective inhibition of key molecular pathways involved in cancer progression, these compounds may pave the way for more personalized and targeted cancer treatments in the future.

Limitation and Future Investigations

Although the synthesized N-substituted indole-2-carboxamides demonstrated promising anticancer activity in vitro, we acknowledge the necessity of in vivo validation to further substantiate their therapeutic potential. Preliminary in vivo studies, including xenograft models to assess pharmacokinetics and toxicity profiles, are currently underway and will be reported in a subsequent publication. These future investigations will provide deeper insights into the compounds’ efficacy, safety, and translational relevance. We believe this continued work will strengthen the preclinical development of these derivatives as viable anticancer candidates.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms26125903/s1.

Author Contributions

Performed research, M.M.A.-N., M.M.S., D.A.S., R.H., H.Z., H.A. and H.M.; Conceptualization, M.M.A.-N., M.M.S., D.A.S., R.H. and A.H.M.; Data curation, M.M.A.-N., M.M.S., D.A.S., R.H., H.Z. and H.M.; Formal analysis, M.M.A.-N., M.M.S., D.A.S., R.H., H.Z. and D.A.A.; Funding acquisition, M.M.A.-N., M.M.S., D.A.S., H.Z., Y.M.A.-H. and M.Y.M.; Investigation, M.M.A.-N., M.M.S., D.A.S. and R.H.; Methodology, M.M.A.-N., M.M.S., D.A.S. and R.H.; Project administration, M.M.A.-N., M.M.S., D.A.S. and R.H.; Resources, M.M.A.-N., M.M.S., D.A.S., R.H., H.Z. and Y.M.A.-H.; Software, M.M.S., D.A.S., R.H., H.Z., S.M.A., D.A.A. and A.H.M.; Supervision, M.M.A.-N., M.M.S., Y.M.A.-H. and A.H.M.; Validation, M.M.A.-N., M.M.S., S.M.A., N.D.A. and A.H.M.; Visualization, M.M.A.-N., M.M.S., D.A.S., R.H. and A.H.M.; Writing—original draft, M.M.A.-N., M.M.S., D.A.S., R.H., D.A.A. and A.H.M.; Writing—review and editing, M.M.A.-N., M.M.S., D.A.S., R.H., M.Y.M., N.D.A. and A.H.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was sponsored and financially supported by research funding from Isra University, Amman, Jordan, to Maysaa M. Saleh. The project was conducted under Isra University Deans’ Council Decision No. 51/3-2018/2019 (date of Grant: 18 November 2018). The authors acknowledge the generous support from the Deanship of Scientific Research at Al-Zaytoonah University of Jordan for the computational resources and databases (Grant number: 2023-2022/17/50).

Institutional Review Board Statement

This study’s dataset is entirely derived from results of samples we collected. Ethical approval and informed consent were granted by the relevant institutional review board.

Informed Consent Statement

This work does not report any research conducted on human or animal subjects by the authors.

Data Availability Statement

The datasets used and analyzed during the current study are available upon reasonable request from the corresponding author. Please note that due to the ongoing experiment’s requirements in the later stages, the data will be released at this time.

Acknowledgments

The authors would like to thank The University of Jordan, Amman, Jordan; Isra University, Amman, Jordan; Al-Zaytoonah University of Jordan, Amman, Jordan; and Hamdi Mango Center for Scientific Research at The University of Jordan, Amman, Jordan, for their facilities and financial support.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

ANOVAAnalysis of variance
HCT-116Human colorectal carcinoma cell line
IC5050% Cell proliferation inhibition
K-562Human chronic myeloid leukemia cell line
MTT3-(4,5-Dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide
MCF-7Human breast adenocarcinoma
ODOptical density
PCAPrincipal component analysis
SISelectivity index
RESResolution
PI3KPhosphoinositide 3-kinase
EGFREpidermal growth factor receptor

References

  1. Bray, F.; Ferlay, J.; Soerjomataram, I.; Siegel, R.L.; Torre, L.A.; Jemal, A. Global cancer statistics 2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2018, 68, 394–424. [Google Scholar] [CrossRef] [PubMed]
  2. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA Cancer J. Clin. 2020, 71, 209–249. [Google Scholar] [CrossRef] [PubMed]
  3. Tarawneh, N.; Hamadneh, L.; Alshaer, W.; Al Bawab, A.Q.; Bustanji, Y.; Abdalla, S. Downregulation of aquaporins and PI3K/AKT and upregulation of PTEN expression induced by the flavone scutellarein in human colon cancer cell lines. Heliyon 2024, 10, e39402. [Google Scholar] [CrossRef] [PubMed]
  4. Cancer Facts & Figures. Available online: https://www.cancer.org/content/dam/cancer-org/research/cancer-facts-and-statistics/annual-cancer-facts-and-figures/2021/cancer-facts-and-figures-2021.pdf (accessed on 14 September 2023).
  5. Alsamarat, R.; Sunoqrot, S. A Glucose Oxidase-Curcumin Composite Nanoreactor for Multimodal Synergistic Cancer Therapy. ACS Appl. Bio Mater. 2024, 7, 4611–4621. [Google Scholar] [CrossRef]
  6. Mestareehi, A.; Zhang, X.; Seyoum, B.; Msallaty, Z.; Mallisho, A.; Burghardt, K.J.; Kowluru, A.; Yi, Z. Metformin Increases Protein Phosphatase 2A Activity in Primary Human Skeletal Muscle Cells Derived from Lean Healthy Participants. J. Diabetes Res. 2021, 2021, 9979234. [Google Scholar] [CrossRef]
  7. Khan, S.U.; Fatima, K.; Aisha, S.; Malik, F. Unveiling the mechanisms and challenges of cancer drug resistance. J. Cell Commun. Signal. 2024, 22, 109. [Google Scholar] [CrossRef]
  8. Holmes, K.; Roberts, O.L.; Thomas, A.M.; Cross, M.J. Vascular endothelial growth factor receptor-2: Structure, function, intracellular signalling and therapeutic inhibition. Cell. Signal. 2007, 19, 2003–2012. [Google Scholar] [CrossRef]
  9. Mestareehi, A.; Abu-Farsakh, N. Impact of Protein Phosphatase Expressions on the Prognosis of Hepatocellular Carcinoma Patients. ACS Omega 2024, 9, 10299–10331. [Google Scholar] [CrossRef]
  10. Shibuya, M. VEGF-VEGFR Signals in Health and Disease. Biomol. Ther. 2014, 22, 1–9. [Google Scholar] [CrossRef]
  11. Liu, Y.; Li, Y.; Wang, Y.; Lin, C.; Zhang, D.; Chen, J.; Ouyang, L.; Wu, F.; Zhang, J.; Chen, L. Recent progress on vascular endothelial growth factor receptor inhibitors with dual targeting capabilities for tumor therapy. J. Hematol. Oncol. 2022, 15, 89. [Google Scholar] [CrossRef]
  12. Carmeliet, P.; Jain, R.K. Molecular mechanisms and clinical applications of angiogenesis. Nature 2011, 473, 298–307. [Google Scholar] [CrossRef] [PubMed]
  13. Ferrara, N.; Adamis, A.P. Ten years of anti-vascular endothelial growth factor therapy. Nat. Rev. Drug Discov. 2016, 15, 385–403. [Google Scholar] [CrossRef] [PubMed]
  14. Barata, P.C.; Rini, B.I. Treatment of renal cell carcinoma: Current status and future directions. CA Cancer J. Clin. 2017, 67, 507–524. [Google Scholar] [CrossRef] [PubMed]
  15. Mestareehi, A. Global Gene Expression Profiling and Bioinformatics Analysis Reveal Downregulated Biomarkers as Potential Indicators for Hepatocellular Carcinoma. ACS Omega 2024, 9, 26075–26096. [Google Scholar] [CrossRef]
  16. Patel, S.A.; Nilsson, M.B.; Le, X.; Cascone, T.; Jain, R.K.; Heymach, J.V. Molecular Mechanisms and Future Implications of VEGF/VEGFR in Cancer Therapy. Clin. Cancer Res. 2023, 29, 30–39. [Google Scholar] [CrossRef]
  17. Pommier, Y.; Leo, E.; Zhang, H.; Marchand, C. DNA topoisomerases and their poisoning by anticancer and antibacterial drugs. Chem. Biol. 2010, 17, 421–433. [Google Scholar] [CrossRef]
  18. Hande, K.R. Etoposide: Four decades of development of a topoisomerase II inhibitor. Eur. J. Cancer 1998, 34, 1514–1521. [Google Scholar] [CrossRef]
  19. Mestareehi, A.H. Optimized and Validated Stability-Indicating RP-HPLC Method for Comprehensive Profiling of Process-Related Impurities and Stress-Induced Degradation Products in Rivaroxaban (XARELTO)(®). Int. J. Mol. Sci. 2025, 26, 4744. [Google Scholar] [CrossRef]
  20. Salerno, S.; Da Settimo, F.; Taliani, S.; Simorini, F.; La Motta, C.; Fornaciari, G.; Marini, A.M. Recent advances in the development of dual topoisomerase I and II inhibitors as anticancer drugs. Curr. Med. Chem. 2010, 17, 4270–4290. [Google Scholar] [CrossRef]
  21. Chen, A.Y.; Liu, L.F. DNA topoisomerases: Essential enzymes and lethal targets. Annu. Rev. Pharmacol. Toxicol. 1994, 34, 191–218. [Google Scholar] [CrossRef]
  22. Mestareehi, A. Development and validation of a stability-indicating reversed phase high-performance liquid chromatography method for quantifying rivaroxaban (XARELTO). J. Chromatogr. B Analyt. Technol. Biomed. Life Sci. 2025, 1257, 124573. [Google Scholar] [CrossRef] [PubMed]
  23. Gaba, M.; Mohan, C. Development of drugs based on imidazole and benzimidazole bioactive heterocycles: Recent advances and future directions. Med. Chem. Res. 2016, 25, 173–210. [Google Scholar] [CrossRef]
  24. Calcaterra, N.E.; Barrow, J.C. Classics in chemical neuroscience: Diazepam (valium). ACS Chem. Neurosci. 2014, 5, 253–260. [Google Scholar] [CrossRef] [PubMed]
  25. Dhuguru, J.; Skouta, R. Role of Indole Scaffolds as Pharmacophores in the Development of Anti-Lung Cancer Agents. Molecules 2020, 25, 1615. [Google Scholar] [CrossRef]
  26. Asif, M.; Srivastava, R.; Fatima, A.; Shakeel, M.; Hassan, F.; Nasibullah, M. A Perspective of the Amide Group Containing FDA Approved Anticancer Drugs from 2021–2022 (A Review). Russ. J. Bioorganic Chem. 2023, 49, 1165–1176. [Google Scholar] [CrossRef]
  27. Comşa, Ş.; Cîmpean, A.M.; Raica, M. The Story of MCF-7 Breast Cancer Cell Line: 40 years of Experience in Research. Anticancer. Res. 2015, 35, 3147–3154. Available online: https://ar.iiarjournals.org/content/35/6/3147 (accessed on 15 June 2024).
  28. Hassan, A.S.; Moustafa, G.O.; Awad, H.M.; Nossier, E.S.; Mady, M.F. Design, Synthesis, Anticancer Evaluation, Enzymatic Assays, and a Molecular Modeling Study of Novel Pyrazole-Indole Hybrids. ACS Omega 2021, 6, 12361–12374. [Google Scholar] [CrossRef]
  29. Kumar, A.; Anand, T.; Bhattacharyya, J.; Sharma, A.; Jaganathan, B.G. K562 chronic myeloid leukemia cells modify osteogenic differentiation and gene expression of bone marrow stromal cells. J. Cell Commun. Signal. 2018, 12, 441–450. [Google Scholar] [CrossRef]
  30. Rubin, E.; Shan, K.S.; Dalal, S.; Vu, D.U.D.; Milillo-Naraine, A.M.; Guaqueta, D.; Ergle, A. Molecular Targeting of the Human Epidermal Growth Factor Receptor-2 (HER2) Genes across Various Cancers. Int. J. Mol. Sci. 2024, 25, 1064. [Google Scholar] [CrossRef]
  31. Miziak, P.; Baran, M.; Błaszczak, E.; Przybyszewska-Podstawka, A.; Kałafut, J.; Smok-Kalwat, J.; Dmoszyńska-Graniczka, M.; Kiełbus, M.; Stepulak, A. Estrogen Receptor Signaling in Breast Cancer. Cancers 2023, 15, 4689. [Google Scholar] [CrossRef]
  32. Li, L.; Liu, S.; Wang, B.; Liu, F.; Xu, S.; Li, P.; Chen, Y. An Updated Review on Developing Small Molecule Kinase Inhibitors Using Computer-Aided Drug Design Approaches. Int. J. Mol. Sci. 2023, 24, 13953. [Google Scholar] [CrossRef] [PubMed]
  33. Mestareehi, A.; Li, H.; Zhang, X.; Venkata, S.P.M.; Jaiswal, R.; Yu, F.S.; Yi, Z.; Wang, J.M. Quantitative Proteomics Reveals Transforming Growth Factor β Receptor Targeted by Resveratrol and Hesperetin Coformulation in Endothelial Cells. ACS Omega 2023, 8, 16206–16217. [Google Scholar] [CrossRef] [PubMed]
  34. Mosmann, T. Rapid colorimetric assay for cellular growth and survival: Application and cytotoxicity assays. J. Immunol. Methods 1983, 65, 55–63. [Google Scholar] [CrossRef]
  35. Denizot, F.; Lang, R. Rapid colorimetric assay for cell growth and survival. Modifications to the tetrazolium dye procedure giving improved sensitivity and reliability. J. Immunol. Methods 1986, 89, 271–277. [Google Scholar] [CrossRef] [PubMed]
  36. Chen, S.-F.; Huang, N.-L.; Lin, J.-H.; Wu, C.-C.; Wang, Y.-R.; Yu, Y.-J.; Gilson, M.K.; Chan, N.-L. Structural insights into the gating of DNA passage by the topoisomerase II DNA-gate. Nat. Commun. 2018, 9, 3085. [Google Scholar] [CrossRef]
  37. Zhao, Y.; Zhang, X.; Chen, Y.; Lu, S.; Peng, Y.; Wang, X.; Guo, C.; Zhou, A.; Zhang, J.; Luo, Y.; et al. Crystal structures of PI3Kalpha complexed with PI103 and Its derivatives: New directions for inhibitors design. ACS Med. Chem. Lett. 2013, 5, 138–142. [Google Scholar] [CrossRef]
  38. Kawakita, Y.; Seto, M.; Ohashi, T.; Tamura, T.; Yusa, T.; Miki, H.; Iwata, H.; Kamiguchi, H.; Tanaka, T.; Sogabe, S.; et al. Design and synthesis of novel pyrimido [4,5-b]azepine derivatives as HER2/EGFR dual inhibitors. Bioorganic Med. Chem. 2013, 21, 2250–2261. [Google Scholar] [CrossRef]
  39. Schrödinger. Protein Preparation Wizard, Maestro, Macromodel, QPLD-Dock, and Pymol; Schrödinger, LLC: Portland, OR, USA, 2022. [Google Scholar]
  40. Friesner, R.A.; Banks, J.L.; Murphy, R.B.; Halgren, T.A.; Klicic, J.J.; Mainz, D.T.; Repasky, M.P.; Knoll, E.H.; Shelley, M.; Perry, J.K.; et al. Glide: A new approach for rapid, accurate docking and scoring. 1. Method and assessment of docking accuracy. J. Med. Chem. 2004, 47, 1739–1749. [Google Scholar] [CrossRef]
  41. Friesner, R.A.; Murphy, R.B.; Repasky, M.P.; Frye, L.L.; Greenwood, J.R.; Halgren, T.A.; Sanschagrin, P.C.; Mainz, D.T. Extra precision glide: Docking and scoring incorporating a model of hydrophobic enclosure for protein-ligand complexes. J. Med. Chem. 2006, 49, 6177–6196. [Google Scholar] [CrossRef]
  42. Sherman, W.; Day, T.; Jacobson, M.P.; Friesner, R.A.; Farid, R. Novel procedure for modeling ligand/receptor induced fit effects. J. Med. Chem. 2006, 49, 534–553. [Google Scholar] [CrossRef]
  43. Jadhav, A.K.; Karuppayil, S.M. Molecular docking studies on thirteen fluoroquinolines with human topoisomerase II a and b. Silico Pharmacol. 2016, 5, 4. [Google Scholar] [CrossRef] [PubMed]
  44. Sharma, K.K.; Singh, B.; Mujwar, S.; Bisen, P.S. Molecular Docking Based Analysis to Elucidate the DNA Topoisomerase IIβ as the Potential Target for the Ganoderic Acid; A Natural Therapeutic Agent in Cancer Therapy. Curr. Comput.-Aided Drug Des. 2020, 16, 176–189. [Google Scholar] [CrossRef] [PubMed]
  45. Yu, C.; Hu, J.; Luyten, W.; Sun, D.; Jiang, T. Identification of novel topoisomerase II alpha inhibitors by virtual screening, molecular docking, and bioassay. Chem. Biol. Drug Des. 2022, 99, 92–102. [Google Scholar] [CrossRef] [PubMed]
  46. Arthur, D.E. Molecular docking studies of some topoisomerase II inhibitors: Implications in designing of novel anticancer drugs. Radiol. Infect. Dis. 2019, 6, 68–79. [Google Scholar] [CrossRef]
  47. Sabbah, D.A.; Samarat, H.H.; Al-Shalabi, E.; Bardaweel, S.K.; Hajjo, R.; Sweidan, K.; Khalaf, R.A.; Al-Zuheiri, A.M.; Abushaikha, G. Design, Synthesis, and Biological Examination of N-Phenyl-6-fluoro-4-hydroxy-2-quinolone-3-carboxamides as Anticancer Agents. ChemistrySelect 2022, 7, e202200662. [Google Scholar] [CrossRef]
  48. Sweidan, K.; Elfadel, H.; Sabbah, D.A.; Bardaweel, S.K.; Hajjo, R.; Anjum, S.; Sinoj, J.; Nair, V.A.; Abu-Gharbieh, E.; El-Huneidi, W. Novel Derivatives of 4,6-Dihydroxy-2-Quinolone-3-Carboxamides as Potential PI3Kα Inhibitors. ChemistrySelect 2022, 7, e202202263. [Google Scholar] [CrossRef]
  49. Sweidan, K.; Sabbah, D.A.; Engelmann, J.; Halim, H.A.; Sheikha, G.A. Computational docking studies of novel heterocyclic carboxamides as potential PI3Kα inhibitors. Lett. Drug Des. Discov. 2015, 12, 856–863. [Google Scholar] [CrossRef]
  50. Sweidan, K.; Zalloum, H.; Sabbah, D.A.; Idris, G.; Abudosh, K.; Mubarak, M.S. Synthesis, characterization, and anticancer evaluation of some new N1-(anthraquinon-2-yl) amidrazone derivatives. Can. J. Chem. 2018, 96, 1123–1128. [Google Scholar] [CrossRef]
  51. Sweidan, K.; Sabbah, D.A.; Bardaweel, S.; Dush, K.A.; Sheikha, G.A.; Mubarak, M.S. Computer-aided design, synthesis, and biological evaluation of new indole-2-carboxamide derivatives as PI3Kα/EGFR inhibitors. Bioorganic Med. Chem. Lett. 2016, 26, 2685–2690. [Google Scholar] [CrossRef]
  52. Sabbah, D.A.; Ibrahim, A.H.; Talib, W.H.; Alqaisi, K.M.; Sweidan, K.; Bardaweel, S.K.; Sheikha, G.A.; Zhong, H.A.; Al-Shalabi, E.; Khalaf, R.A. Ligand-based drug design: Synthesis and biological evaluation of substituted benzoin derivatives as potential antitumor agents. Med. Chem. 2019, 15, 417–429. [Google Scholar] [CrossRef]
  53. Sabbah, D.A.; Hasan, S.E.; Abu Khalaf, R.; Bardaweel, S.K.; Hajjo, R.; Alqaisi, K.M.; Sweidan, K.A.; Al-Zuheiri, A.M. Molecular modeling, synthesis and biological evaluation of N-phenyl-4-hydroxy-6-methyl-2-quinolone-3-carboxamides as anticancer agents. Molecules 2020, 25, 5348. [Google Scholar] [CrossRef] [PubMed]
  54. Sabbah, D.A.; Haroon, R.A.; Bardaweel, S.K.; Hajjo, R.; Sweidan, K. N-phenyl-6-chloro-4-hydroxy-2-quinolone-3-carboxamides: Molecular Docking, Synthesis, and Biological Investigation as Anticancer Agents. Molecules 2021, 26, 73. [Google Scholar] [CrossRef] [PubMed]
  55. Sabbah, D.A.; Al-Tarawneh, F.; Talib, W.H.; Sweidan, K.; Bardaweel, S.K.; Al-Shalabi, E.; Zhong, H.A.; Sheikha, G.A.; Khalaf, R.A.; Mubarak, M.S. Benzoin schiff bases: Design, synthesis, and biological evaluation as potential antitumor agents. Med. Chem. 2018, 14, 695–708. [Google Scholar] [CrossRef] [PubMed]
  56. Sabbah, D.A.; Vennerstrom, J.L.; Zhong, H.A. Binding selectivity studies of phosphoinositide 3-kinases using free energy calculations. J. Chem. Inf. Model. 2012, 52, 3213–3224. [Google Scholar] [CrossRef]
  57. Sabbah, D.A.; Vennerstrom, J.L.; Zhong, H. Docking studies on isoform-specific inhibition of phosphoinositide-3-kinases. J. Chem. Inf. Model. 2010, 50, 1887–1898. [Google Scholar] [CrossRef]
  58. Sabbah, D.A.; Simms, N.A.; Wang, W.; Dong, Y.; Ezell, E.L.; Brattain, M.G.; Vennerstrom, J.L.; Zhong, H.A. N-Phenyl-4-hydroxy-2-quinolone-3-carboxamides as selective inhibitors of mutant H1047R phosphoinositide-3-kinase (PI3Kα). Bioorganic Med. Chem. 2012, 20, 7175–7183. [Google Scholar] [CrossRef]
  59. Sabbah, D.A.; Simms, N.A.; Brattain, M.G.; Vennerstrom, J.L.; Zhong, H. Biological evaluation and docking studies of recently identified inhibitors of phosphoinositide-3-kinases. Bioorganic Med. Chem. Lett. 2012, 22, 876–880. [Google Scholar] [CrossRef]
  60. Sabbah, D.A.; Saada, M.; Khalaf, R.A.; Bardaweel, S.; Sweidan, K.; Al-Qirim, T.; Al-Zughier, A.; Halim, H.A.; Sheikha, G.A. Molecular modeling based approach, synthesis, and cytotoxic activity of novel benzoin derivatives targeting phosphoinostide 3-kinase (PI3Kα). Bioorganic Med. Chem. Lett. 2015, 25, 3120–3124. [Google Scholar] [CrossRef]
  61. Sabbah, D.A.; Hishmah, B.; Sweidan, K.; Bardaweel, S.; AlDamen, M.; Zhong, H.A.; Khalaf, R.A.; Ibrahim, A.H.; Al-Qirim, T.; Sheikha, G.A. Structure-based design: Synthesis, X-ray crystallography, and biological evaluation of N-substituted-4-hydroxy-2-quinolone-3-carboxamides as potential cytotoxic agents. Anticancer Agents Med. Chem. 2018, 18, 263–276. [Google Scholar] [CrossRef]
  62. Huang, C.-H.; Mandelker, D.; Schmidt-Kittler, O.; Samuels, Y.; Velculescu, V.E.; Kinzler, K.W.; Vogelstein, B.; Gabelli, S.B.; Amzel, L.M. The structure of a human p110 alpha/p85 alpha complex elucidates the effects of oncogenic PI3K alpha mutations. Science 2007, 318, 1744–1748. [Google Scholar] [CrossRef]
  63. Liao, Q.-H.; Gao, Q.-Z.; Wei, J.; Chou, K.-C. Docking and molecular dynamics study on the inhibitory activity of novel inhibitors on epidermal growth factor receptor (EGFR). Med. Chem. 2011, 7, 24–31. [Google Scholar] [CrossRef] [PubMed]
  64. Celik, İ.; Ayhan-Kılcıgil, G.; Guven, B.; Kara, Z.; Gurkan-Alp, A.S.; Karayel, A.; Onay-Besikci, A. Design, synthesis and docking studies of benzimidazole derivatives as potential EGFR inhibitors. Eur. J. Med. Chem. 2019, 173, 240–249. [Google Scholar] [CrossRef] [PubMed]
  65. Bhatia, P.; Sharma, V.; Alam, O.; Manaithiya, A.; Alam, P.; Alam, M.T.; Imran, M. Novel quinazoline-based EGFR kinase inhibitors: A review focussing on SAR and molecular docking studies (2015–2019). Eur. J. Med. Chem. 2020, 204, 112640. [Google Scholar] [CrossRef] [PubMed]
  66. alvaDesc 2.0. Available online: http://dev.kode-datacenter.net:10050/pf/alvadesc-2-0/ (accessed on 10 October 2023).
  67. Hopkins, C.R.; O’neil, S.V.; Laufersweiler, M.C.; Wang, Y.; Pokross, M.; Mekel, M.; Evdokimov, A.; Walter, R.; Kontoyianni, M.; Petrey, M.E.; et al. Design and synthesis of novel N-sulfonyl-2-indole carboxamides as potent PPAR-gamma binding agents with potential application to the treatment of osteoporosis. Bioorganic Med. Chem. Lett. 2006, 16, 5659–5663. [Google Scholar] [CrossRef]
  68. Al-Najdawi, M.; Al-Hiari, Y.; Al-Qirim, T.; Shattat, G.; Al-Zweri, M.; Sheikha, G.A. Synthesis and pharmacological evaluation of novel unsubstituted indole-anthraquinone carboxamide derivatives as potent antihyperlipidemic agents. Z. Naturforsch. C 2014, 69, 21–28. [Google Scholar] [CrossRef]
  69. Saleh, M.M.; Abuarqoub, D.A.; Hammad, A.M.; Hossan, M.S.; Ahmed, N.; Aslam, N.; Naser, A.Y.; Moody, C.J.; Laughton, C.A.; Bradshaw, T.D. In Vitro Anticancer Properties of Novel Bis-Triazoles. Curr. Issues Mol. Biol. 2023, 45, 175–196. [Google Scholar] [CrossRef]
  70. Saleh, M.M.; Abuirmeileh, A.N.; Al-Rousan, R.M.; Abudoleh, S.M.; Hassouneh, L.K.; Zihlif, M.A.; Taha, M.O.; Abutayeh, R.F.; Mansour, H.; Abu-Irmaileh, B. Biological Evaluation and Reverse Pharmacophore Mapping of Innovative Bis-Triazoles as Promising Anticancer Agents. Open J. Med. Chem. 2022, 16, e187410452207200. [Google Scholar] [CrossRef]
  71. Saleh, M.M.; Laughton, C.A.; Bradshaw, T.D.; Moody, C.J. Development of a series of bis-triazoles as G-quadruplex ligands. RSC Adv. 2017, 7, 47297–47308. [Google Scholar] [CrossRef]
  72. Al-Sammarra’e, A.; Al-Najdawi, M.M.; Saleh, M.M.; Al-Hiari, Y.M.; Al-Bashiti, R. Synthesis and Biological Evaluation of Furyl-Carboxamide Derivatives as Potential Anticancer Agents. J. Turk. Chem. Soc. A Chem. 2022, 9, 909–918. [Google Scholar] [CrossRef]
  73. Al-Ali, L.; Al-Ani, R.J.; Saleh, M.M.; Hammad, A.M.; Abuarqoub, D.A.; Abu-Irmaileh, B.; Naser, A.Y.; Najdawi, M.M.; Abbas, M.M.; Alkrad, J.A. Biological evaluation of combinations of tyrosine kinase inhibitors with Inecalcitol as novel treatments for human chronic myeloid leukemia. Saudi Pharm. J. 2024, 32, 101931. [Google Scholar] [CrossRef]
  74. Zalloum, H.; Zalloum, W.; Hameduh, T.; ALSalamat, H.; Zihlif, M. Anti-Proliferative Effect of Potential LSD1/CoREST Inhibitors Based on Molecular Dynamics Model for Treatment of SH-SY5Y Neuroblastoma Cancer Cell Line. Asian Pac. J. Cancer Prev. 2022, 23, 3533–3540. [Google Scholar] [CrossRef] [PubMed]
  75. Cousins, K.R. Computer review of ChemDraw Ultra 12.0. J. Am. Chem. Soc. 2011, 133, 8388. [Google Scholar] [CrossRef] [PubMed]
  76. Hajjo, R.; Grulke, C.M.; Golbraikh, A.; Setola, V.; Huang, X.P.; Roth, B.L.; Tropsha, A. Development, validation, and use of quantitative structure-activity relationship models of 5-hydroxytryptamine (2B) receptor ligands to identify novel receptor binders and putative valvulopathic compounds among common drugs. J. Med. Chem. 2010, 53, 7573–7586. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. The synthetic pathway for the preparation of N-substituted 1H-indole-2-carboxamides, beginning with indole-2-carboxylic acid and leading to the formation of aromatic substitutions 4, 6, 8, 10, 12, and 14. i: Thionyl chloride (SOCl2), 70–80 °C, chloroform (CHCl3); ii: triethylamine (TEA), dichloromethane (DCM), RT.
Scheme 1. The synthetic pathway for the preparation of N-substituted 1H-indole-2-carboxamides, beginning with indole-2-carboxylic acid and leading to the formation of aromatic substitutions 4, 6, 8, 10, 12, and 14. i: Thionyl chloride (SOCl2), 70–80 °C, chloroform (CHCl3); ii: triethylamine (TEA), dichloromethane (DCM), RT.
Ijms 26 05903 sch001
Scheme 2. The synthetic pathway for the preparation of N-substituted 1H-furan-2-carboxamides, highlighting the formation of compound 16 through the reaction of 1-aminoanthraquinone with furan-2-carbonyl chloride. ii: triethylamine (TEA), dichloromethane (DCM), RT.
Scheme 2. The synthetic pathway for the preparation of N-substituted 1H-furan-2-carboxamides, highlighting the formation of compound 16 through the reaction of 1-aminoanthraquinone with furan-2-carbonyl chloride. ii: triethylamine (TEA), dichloromethane (DCM), RT.
Ijms 26 05903 sch002
Figure 1. Determination of IC50 values of carboxamide compounds 4, 6, 8, 10, 12, 14, and 16 against different types of human cancer cells: (A) colon cancer (HCT-119); (B) breast cancer (MCF-7); (C) chronic myeloid leukemia (K-562); and (D) normal fibroblast cell lines. The IC50 value of each compound was calculated using the nonlinear regression of the log concentration as inhibition percentage values using GraphPad Prism 9.0.
Figure 1. Determination of IC50 values of carboxamide compounds 4, 6, 8, 10, 12, 14, and 16 against different types of human cancer cells: (A) colon cancer (HCT-119); (B) breast cancer (MCF-7); (C) chronic myeloid leukemia (K-562); and (D) normal fibroblast cell lines. The IC50 value of each compound was calculated using the nonlinear regression of the log concentration as inhibition percentage values using GraphPad Prism 9.0.
Ijms 26 05903 g001
Figure 2. (A) The X-ray structure of the topoisomerase II–DNA complex (5ZRF) binding pocket accommodating the co-crystallized ligand EVP (red color) and IF-docked poses of carboxamides 416 depicted in yellow, pink, green, purple, and cyan. (B) Superposing of EVP (red color) and ligand 16 (purple color) at 5ZRF binding site. DNA strands are blue colored, some of the key binding residues are depicted, and H atoms are hidden for clarification purposes. Picture visualized using PYMOL [39].
Figure 2. (A) The X-ray structure of the topoisomerase II–DNA complex (5ZRF) binding pocket accommodating the co-crystallized ligand EVP (red color) and IF-docked poses of carboxamides 416 depicted in yellow, pink, green, purple, and cyan. (B) Superposing of EVP (red color) and ligand 16 (purple color) at 5ZRF binding site. DNA strands are blue colored, some of the key binding residues are depicted, and H atoms are hidden for clarification purposes. Picture visualized using PYMOL [39].
Ijms 26 05903 g002
Figure 3. (A) The X-ray structure of the PI3Kα (4L23) binding pocket accommodating the co-crystallized ligand X6K (red color) and IF-docked poses of carboxamides 416 depicted in yellow, pink, green, purple, and cyan. (B) Superposing of X6K (red color) and ligand 16 (green color) at 4L23 binding site. Some of the key binding residues are depicted; C atoms are grey colored, O (red), and N (blue). H atoms are hidden for clarification purposes, and the picture is visualized using PYMOL [39].
Figure 3. (A) The X-ray structure of the PI3Kα (4L23) binding pocket accommodating the co-crystallized ligand X6K (red color) and IF-docked poses of carboxamides 416 depicted in yellow, pink, green, purple, and cyan. (B) Superposing of X6K (red color) and ligand 16 (green color) at 4L23 binding site. Some of the key binding residues are depicted; C atoms are grey colored, O (red), and N (blue). H atoms are hidden for clarification purposes, and the picture is visualized using PYMOL [39].
Ijms 26 05903 g003
Figure 4. (A) The X-ray structure of the EGFR (3W32) binding pocket accommodating the co-crystallized ligand W32 (red color) and IF-docked poses of carboxamides 416 depicted in yellow, pink, green, purple, and cyan. (B) Superposing of W32 (red color) and ligand 16 (green color) at 3W32 binding site. Some of the key binding residues are depicted; C atoms are grey colored, O (red), and N (blue). H atoms are hidden for clarification purposes, and the picture is visualized using PYMOL [39].
Figure 4. (A) The X-ray structure of the EGFR (3W32) binding pocket accommodating the co-crystallized ligand W32 (red color) and IF-docked poses of carboxamides 416 depicted in yellow, pink, green, purple, and cyan. (B) Superposing of W32 (red color) and ligand 16 (green color) at 3W32 binding site. Some of the key binding residues are depicted; C atoms are grey colored, O (red), and N (blue). H atoms are hidden for clarification purposes, and the picture is visualized using PYMOL [39].
Ijms 26 05903 g004
Figure 5. Superposition of the (A) EVP IF-docked pose (purple) and its native geometry (green) in 5ZRF; (B) X6K IF-docked pose (sand) and its native geometry (red) in 4L23; and (C) W32 IF-docked pose (blue) and its native geometry (beige) in 3W32. Picture is visualized using PYMOL.
Figure 5. Superposition of the (A) EVP IF-docked pose (purple) and its native geometry (green) in 5ZRF; (B) X6K IF-docked pose (sand) and its native geometry (red) in 4L23; and (C) W32 IF-docked pose (blue) and its native geometry (beige) in 3W32. Picture is visualized using PYMOL.
Ijms 26 05903 g005
Figure 6. Molecular descriptor analysis. (A) Graphical 2D representation of the principal component analysis (PCA) of all 2D alvaDesc molecular descriptors for the synthesized active carboxamides 4,6,8,10,12,14, and 16. PC1 explained 77.29% of the variation in descriptor values on the x-axis, and PC2 explained 9.09% of the variation in descriptor values on the y-axis. The dots corresponding to compounds on the figure are colored based on their lead-like scores (LLS_01). (B) Radar plot of drug-like scores (DLSs) and lead-like scores (LLSs) for all synthesized active compounds.
Figure 6. Molecular descriptor analysis. (A) Graphical 2D representation of the principal component analysis (PCA) of all 2D alvaDesc molecular descriptors for the synthesized active carboxamides 4,6,8,10,12,14, and 16. PC1 explained 77.29% of the variation in descriptor values on the x-axis, and PC2 explained 9.09% of the variation in descriptor values on the y-axis. The dots corresponding to compounds on the figure are colored based on their lead-like scores (LLS_01). (B) Radar plot of drug-like scores (DLSs) and lead-like scores (LLSs) for all synthesized active compounds.
Ijms 26 05903 g006
Table 1. The proliferation inhibitory activities of the carboxamide derivatives 4-16 against human colorectal carcinoma (HCT-116), breast adenocarcinoma (MCF-7), chronic myelogenous leukemia (K-562), and dermal normal fibroblast cell lines. Experiments were carried out for a 72 h treatment duration. IC50 values are expressed in micromolar as the mean ± standard deviation, based on at least three independent trials (n = 6 per trial). SI: Selectivity index (IC50 normal fibroblast/IC50 cancer cell line). The yellow highlights represent the most potent antiproliferative activities of carboxamides 4–16 against the three types of cancer cell lines, while the highest SI values are green highlighted.
Table 1. The proliferation inhibitory activities of the carboxamide derivatives 4-16 against human colorectal carcinoma (HCT-116), breast adenocarcinoma (MCF-7), chronic myelogenous leukemia (K-562), and dermal normal fibroblast cell lines. Experiments were carried out for a 72 h treatment duration. IC50 values are expressed in micromolar as the mean ± standard deviation, based on at least three independent trials (n = 6 per trial). SI: Selectivity index (IC50 normal fibroblast/IC50 cancer cell line). The yellow highlights represent the most potent antiproliferative activities of carboxamides 4–16 against the three types of cancer cell lines, while the highest SI values are green highlighted.
Compound No.IC50 (µM) ± S.D.Selectivity Indices
(SIs)
HCT-116
Colon Cancer
MCF-7
Breast Cancer
K-562
Leukemia
Dermal Fibroblast Normal
43.94 ± 0.0310.6 ± 0.530.61 ± 0.03>10025.4, 9.4, 164
625.4 ± 0.3880.8 ± 4.0484.6 ± 4.23>1003.9, 1.2, 1.2
832.0 ± 0.83>100>100>1003.1, 1.0, 1.0
101.01 ± 0.0470.3 ± 3.52>100>10099.4, 1.4, 1.0
123.98 ± 0.069.07 ± 0.450.33 ± 0.02>10025.1, 11.0, 303
142.64 ± 0.077.16 ± 0.360.61 ± 0.03>10037.9, 14.0, 164
163.11 ± 0.04933.1 ± 1.66>100>10032.2, 3.0, 1.0
Table 2. The induced-fit docking (IFD) scores (Kcal/mol) and H-bond interactions against 5ZRF. NA stands for not available.
Table 2. The induced-fit docking (IFD) scores (Kcal/mol) and H-bond interactions against 5ZRF. NA stands for not available.
Compound No. (Moiety)Docking ScoreBinding ResidueBinding Nucleotide
4−7.35NADG13
6−7.83Q778NA
8−8.88NADC8, DG13
10−6.79Q778NA
12−10.16R503NA
14−9.35NADC8, DA12
16−8.95H2O-Q778DG13
EVP−13.76D479DG7, DG13
Table 3. The induced-fit docking (IFD) scores (Kcal/mol) and H-bond interactions against 4L23. NA stands for not available.
Table 3. The induced-fit docking (IFD) scores (Kcal/mol) and H-bond interactions against 4L23. NA stands for not available.
Compound No. (Moiety)Docking ScoreBinding Residue
4−6.85V851
6−6.71Y836
8−7.11E849, V851
10−7.29V851
12−10.10E849, V851, Q859
14−8.85W780, E798, V851, S854
16−9.11Y836, V851
X6K−10.16S773, Y836, V851
Table 4. The induced-fit docking (IFD) scores (Kcal/mol) and H-bond interactions against 3W32. NA stands for not available.
Table 4. The induced-fit docking (IFD) scores (Kcal/mol) and H-bond interactions against 3W32. NA stands for not available.
Compound No. (Moiety)Docking ScoreBinding Residue
4−8.50T854
6−8.60T854
8−9.04M793, T854
10−8.05K745, T854
12−9.99K745, D855
14−11.74M793, T854
16−10.12M793, T854
W32−10.71R841
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Al-Najdawi, M.M.; Saleh, M.M.; Sabbah, D.A.; Hajjo, R.; Zalloum, H.; Abudoleh, S.M.; Abuarqoub, D.A.; Al-Hiari, Y.M.; Mohammad, M.Y.; ALSalamat, H.; et al. Exploring Carboxamide Derivatives as Promising Anticancer Agents: Design, In Vitro Evaluation, and Mechanistic Insights. Int. J. Mol. Sci. 2025, 26, 5903. https://doi.org/10.3390/ijms26125903

AMA Style

Al-Najdawi MM, Saleh MM, Sabbah DA, Hajjo R, Zalloum H, Abudoleh SM, Abuarqoub DA, Al-Hiari YM, Mohammad MY, ALSalamat H, et al. Exploring Carboxamide Derivatives as Promising Anticancer Agents: Design, In Vitro Evaluation, and Mechanistic Insights. International Journal of Molecular Sciences. 2025; 26(12):5903. https://doi.org/10.3390/ijms26125903

Chicago/Turabian Style

Al-Najdawi, Manal M., Maysaa M. Saleh, Dima A. Sabbah, Rima Hajjo, Hiba Zalloum, Suha M. Abudoleh, Duaa A. Abuarqoub, Yusuf M. Al-Hiari, Mohammad Yasin Mohammad, Husam ALSalamat, and et al. 2025. "Exploring Carboxamide Derivatives as Promising Anticancer Agents: Design, In Vitro Evaluation, and Mechanistic Insights" International Journal of Molecular Sciences 26, no. 12: 5903. https://doi.org/10.3390/ijms26125903

APA Style

Al-Najdawi, M. M., Saleh, M. M., Sabbah, D. A., Hajjo, R., Zalloum, H., Abudoleh, S. M., Abuarqoub, D. A., Al-Hiari, Y. M., Mohammad, M. Y., ALSalamat, H., Mansour, H., Aljbour, N. D., & Mestareehi, A. H. (2025). Exploring Carboxamide Derivatives as Promising Anticancer Agents: Design, In Vitro Evaluation, and Mechanistic Insights. International Journal of Molecular Sciences, 26(12), 5903. https://doi.org/10.3390/ijms26125903

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop