Next Article in Journal
Factor XIa Inhibitors as a Novel Anticoagulation Target: Recent Clinical Research Advances
Next Article in Special Issue
In Vitro and In Silico Evaluation of Antiproliferative Activity of New Isoxazolidine Derivatives Targeting EGFR: Design, Synthesis, Cell Cycle Analysis, and Apoptotic Inducers
Previous Article in Journal
Binary Polymeric Surfactant Mixtures for the Development of Novel Loteprednol Etabonate Nanomicellar Eyedrops
Previous Article in Special Issue
7-Chloroquinolinehydrazones as First-in-Class Anticancer Experimental Drugs in the NCI-60 Screen among Different Investigated Series of Aryl, Quinoline, Pyridine, Benzothiazole and Imidazolehydrazones
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Exploring Pyrrolo-Fused Heterocycles as Promising Anticancer Agents: An Integrated Synthetic, Biological, and Computational Approach

by
Roxana-Maria Amărandi
1,
Maria-Cristina Al-Matarneh
2,3,*,
Lăcrămioara Popovici
3,
Catalina Ionica Ciobanu
4,
Andrei Neamțu
1,
Ionel I. Mangalagiu
3 and
Ramona Danac
3,*
1
TRANSCEND Research Center, Regional Institute of Oncology Iasi, 2-4 General Henri Mathias Berthelot Street, 700483 Iasi, Romania
2
“Petru Poni” Institute of Macromolecular Chemistry of Romanian Academy, 41A Grigore Ghica Voda Alley, 700487 Iasi, Romania
3
Faculty of Chemistry, Alexandru Ioan Cuza University of Iasi, 11 Carol I, 700506 Iasi, Romania
4
Institute of Interdisciplinary Research-CERNESIM Centre, Alexandru Ioan Cuza University of Iasi, 11 Carol I, 700506 Iasi, Romania
*
Authors to whom correspondence should be addressed.
Pharmaceuticals 2023, 16(6), 865; https://doi.org/10.3390/ph16060865
Submission received: 14 April 2023 / Revised: 17 May 2023 / Accepted: 9 June 2023 / Published: 11 June 2023
(This article belongs to the Special Issue Novel Heterocyclic Compounds for Drug Discovery)

Abstract

:
Five new series of pyrrolo-fused heterocycles were designed through a scaffold hybridization strategy as analogs of the well-known microtubule inhibitor phenstatin. Compounds were synthesized using the 1,3-dipolar cycloaddition of cycloimmonium N-ylides to ethyl propiolate as a key step. Selected compounds were then evaluated for anticancer activity and ability to inhibit tubulin polymerization in vitro. Notably, pyrrolo[1,2-a]quinoline 10a was active on most tested cell lines, performing better than control phenstatin in several cases, most notably on renal cancer cell line A498 (GI50 27 nM), while inhibiting tubulin polymerization in vitro. In addition, this compound was predicted to have a promising ADMET profile. The molecular details of the interaction between compound 10a and tubulin were investigated through in silico docking experiments, followed by molecular dynamics simulations and configurational entropy calculations. Of note, we found that some of the initially predicted interactions from docking experiments were not stable during molecular dynamics simulations, but that configurational entropy loss was similar in all three cases. Our results suggest that for compound 10a, docking experiments alone are not sufficient for the adequate description of interaction details in terms of target binding, which makes subsequent scaffold optimization more difficult and ultimately hinders drug design. Taken together, these results could help shape novel potent antiproliferative compounds with pyrrolo-fused heterocyclic cores, especially from an in silico methodological perspective.

1. Introduction

Pyrrole and its hetero-fused derivatives have sparked great interest in the field of medicinal chemistry as valuable scaffolds for generating new drugs with biological activity, including anticancer, antimicrobial, or antiviral compounds [1]. Semisynthetic pyrroloquinolines are currently used as first-line agents in cancer therapy, including camptothecin analogs irinotecan and topotecan, with many others currently being investigated as potent antiproliferative drugs [2]. Pyrrolo[2,1-a]isoquinoline is a common motif in various bioactive alkaloids such as lamellarin D, which is known for its topoisomerase I inhibitory properties [3], or the antiproliferative alkaloid crispine A [4]. Pyrrolopyrimidine-based derivatives such as pemetrexed, immucillin H, or variolin B are also used as anticancer drugs. In addition, several other pyrrolo[1,2-a]pyrazine alkaloids, such as dibromofakelin and longamide B, exhibit multiple biological activities, including cytotoxicity [5]. Moreover, our group has reported pyrrole-fused heterocyclic derivatives with an indolizine, pyrrolopyridazine, pyrrolo(iso)quinoline, and pyrrolophenathroline core, with part of them showing considerable anticancer properties [6,7,8,9,10,11,12,13,14].
Of note, many pyrrole-fused heterocyclic ligands exhibit antitumor activity by binding to the colchicine site of tubulin, which is particularly known for its ability to accommodate a wide variety of structurally diverse ligands [15]. Compounds that target the colchicine site mainly exert their biological effects by inhibiting the formation of the cell mitotic spindle, causing arrest in the G2/M phase of the cell cycle, eventually triggering apoptotic cell death [16]. Since the identification of the colchicine binding site in 2004 through the 3.5 Å resolution crystal structure of the α,β-tubulin heterodimer and N-deacetyl-N-(2-mercaptoacetyl) colchicine (DAMA-colchicine) [17], several other colchicine site-binding agents have been co-crystallized with tubulin [18], paving the way to structure-based drug design campaigns aimed at identifying novel antitumor agents with various scaffolds [19,20,21]. At the same time, the abundance of published protein structures with diverse inhibitors allows for the direct comparison of co-crystallized and predicted conformations for a multitude of biologically active agents. However, results do not always agree, often leading to very different directions in structural optimization strategies [22,23]. In the case of tubulin, the recently published crystal structures of several heterocyclic-fused pyrimidines revealed 180° flipped binding poses when compared to previous molecular docking predictions [24]. As such, it is increasingly evident that simple molecular docking approaches for ligand binding pose prediction by using X-ray structures of tubulin co-crystallized with structurally different molecular scaffolds can be challenging and have limited reliability. Although useful for medicinal chemistry, structural insights obtained from molecular docking studies alone regarding the interaction between pyrrole/pyrrole-fused heterocyclic derivatives and tubulin [6,25,26] can ultimately confuse novel design and development efforts in this class of tubulin inhibitors due to the possibility of questionable interpretation [27].
Molecular dynamics (MD) can be a useful tool for estimating the stability of ligand-receptor conformations predicted by molecular docking and can be used as a utensil for docking pose validation [28,29]. As such, when the conformation of a ligand after an MD simulation deviates by more than a specific RMSD value from the initial docking prediction, the docking pose can be regarded as unstable even though it has a high docking score [27,30]. Since MD has been described as useful for assessing the stability of several tubulin colchicine site binders, including combretastatin analogs [31] and benzimidazoles [32], as well as for investigating DAMA-colchicine binding to different tubulin isotypes [33], using such a strategy to validate molecular docking poses in the case of pyrrole-fused derivatives could be useful in the future development of more potent tubulin inhibitors containing this scaffold.
This study presents the development, synthesis, and assessment of the anticancer potential of newly designed pyrrolo-fused heterocycles, which include quinoline, isoquinoline, benzo[f]quinoline, pyrazine, and pyrimidine. The rationale behind compound design was based on the observation that by substituting the 3-hydroxy-4-methoxyphenyl ring of phenstatin (a tubulin polymerization inhibitor) with pyrrolo-fused heterocycles such as indolizine, pyrrolopyridazine, or pyrroloquinoline (Figure 1), compounds with excellent anticancer properties can be obtained [6,11,14,34,35].
The tubulin polymerization inhibition mechanism can be partly responsible for the anticancer activity of these compounds [6,11,34,35]. The best compound in the series, pyrrolo[1,2-a]quinoline 10a, was further chosen for a series of in silico investigations, including global and local docking, molecular dynamics simulations, and configurational entropy calculations. This thorough in silico evaluation was conducted in order to describe the interaction of this compound with the colchicine binding site in detail, as well as investigate the stability of identified ligand-protein interactions through molecular docking experiments.
In the above context, the herein study was designed in order to identify novel anticancer tubulin-targeting agents that bind to the colchicine site, as well as describe the molecular details of hit binding through extensive in silico investigations. The broader goal of our investigation is to deepen the molecular understanding of biological activity and help shape future rational drug design campaigns aimed at the colchicine site of tubulin.

2. Results

2.1. Chemistry

The strategy to build the five series of target phenstatin analogs (Figure 1) consisted of two main synthetic steps, starting from the desired heterocycle to be fused to the pyrrole ring.
Four derivatives were synthesized in each series of compounds, differing by the substituent at the phenyl ring. First, monoquaternary salts 3, 6, 9, 12, and 15 were prepared by the direct reaction of pyrazine 1, 4-(4-chlorophenyl)-pyrimidine 5, quinoline 8, benzo[f]quinoline 11, or isoquinoline 14, respectively, with 2-bromo-acetophenones 2 in acetone, at room temperature (r.t.) (Scheme 1, Scheme 2, Scheme 3, Scheme 4 and Scheme 5). The reaction of pyrimidine in similar conditions did not produce the desired salts, and that is why we used the substituted pyrimidine derivative 5 (previously synthesized by our group [36]) as the starting material for the target compounds 7ad.
We employed the 1,3-dipolar cycloaddition method to synthesize pyrrolopyrazine/pyrrolopyrimidine/indolizine rings. In this method, we generated heterocyclic ylides in situ from the monoquaternary salts 3, 6, 9, 12, and 15 and reacted them with ethyl propiolate in a basic medium. The reaction schemes for each series are illustrated in Scheme 1, Scheme 2, Scheme 3, Scheme 4 and Scheme 5. As expected, cycloadditions occurred with the selective formation of the regioisomers 4, 7, 10, and 13, which is in agreement with the electronic effects within ethyl propiolate and ylide species.
All 4-bromo-substituted intermediate salts (3d [37], 6d [36], 9d [6], 12d [38], and 15d [6]) and 4-bromobenzoyl-cycloadducts (7d [36], 10d [39], 13d [40], and 16d [39]) have been previously reported in the literature. However, the compounds were synthesized and subsequently tested for their anticancer activity in order to provide a more detailed understanding of the structure-activity relationships (SAR).
The fused target compounds were obtained in moderate yields ranging from 30% to 80%. The structures of all intermediate and final compounds were fully confirmed by spectral and physicochemical data (Supplementary Figures S1–S19). The structures of the cycloadducts presented in Scheme 1, Scheme 2, Scheme 3, Scheme 4 and Scheme 5 were confirmed by NMR and IR spectra. These spectra are consistent with those of similar compounds previously reported in the literature [6,9], where clear structural evidence was emphasized, including X-ray diffraction.

2.2. Biological Activity

2.2.1. Anticancer Activity

All forty synthesized compounds (monoquaternary salts and cycloadducts) were electronically submitted to the National Cancer Institute (NCI) platform, and twenty-five compounds (4bd, 7bc, 9a, 9d, 10ad, 12ad, 13ad, 15a, 15d, and 16ad) were selected for single dose (10−5 M) screening against a panel of 60 human tumor cell lines, representing leukemia, melanoma, and cancers of the lung, colon, central nervous system, ovary, kidney, prostate, and breast [41]. Selected representative results for best hits and phenstatin are summarized in Table 1, all the other tested compounds being inactive against the NCI cancer cells at 10−5 M (results provided by NCI for all tested compounds can be found in Supplementary Figures S20–S44).
Compound 10a, which contains a pyrrolo[1,2-a]quinoline scaffold, showed excellent growth inhibitory properties against almost all tested cell lines, with an average GP (growth percent—growth of treated cultures relative to untreated cultures) for all tested cell lines of 25.7% (74.3% GP inhibition). This compound exhibited the best inhibitory activity on the growth of leukemia cell lines, with an average GP inhibition of 87.33%, prostate cancer lines (78.0% GP inhibition), and breast cancer lines (76.4% GP inhibition). Compound 10a also had a cytotoxic effect on several cell lines, most notably on melanoma MDA-MB-435 cells, where 54.59% of cultured cells were killed at a 10−5 M concentration. Even if phenstatin was superior in terms of average GP inhibition, compound 10a showed similar or even better inhibitory properties against several cell lines. Interestingly, the other three compounds from this series, 10b10d, did not exert any relevant inhibitory activity, suggesting that the 3,4,5-trimethoxyphenyl group plays a very important role in the observed anticancer activity in this series of compounds.
From the other series of synthesized cycloadducts, only pyrrolo[2,1-a]isoquinoline 16b exhibited specific cytotoxic activity against the SNB-75 CNS cancer cell line, while compound 13b, a benzo[f]pyrrolo[1,2-a]quinoline, was moderately active in inhibiting the growth of SR leukemia cells (71% GP inhibition).
Among the intermediate salts, benzo[f]quinolin-4-ium bromides 12 showed moderate inhibitory properties against the growth of colon HCT-116 cells and breast MDA-MB-468 cells, with growth inhibition values of approximately 50%.
Compounds that fulfilled predetermined threshold inhibition criteria were then chosen by the NCI for the second testing step to determine GI50, TGI, and LC50 parameters. Showing the most significant growth inhibition, compound 10a was the only one selected for evaluation against the sixty tumor cell lines in five-dose assays [41,42,43]. The most representative results from the NCI-60 five-dose screen are shown in Table 2 (full results can be found in Figure S45 from the Supplementary Materials).
Notably, compound 10a displayed almost all GI50 values in the nanomolar concentration range. This pyrrolo[1,2-a]quinoline derivative outperformed control phenstatin in several cases, including renal cancer cell line A498, colon cancer cell line COLO 205, and breast cancer cell line T-47D. Furthermore, compound 10a demonstrated good performance against various cancer cell lines (including melanoma MDA-MB-435, renal cancer A498, ovarian cancer OVCAR-3, breast cancer MDA-MB-468, and non-small cell lung cancer NCI-H522 cell lines), exhibiting complete growth inhibition at submicromolar concentrations (Table 2). However, LC50 values (the molar concentration of tested compound causing 50% death of tumor cells) for compound 10a and phenstatin, respectively, were found to be >100 μM against all tested lines.
NCI’s in silico platform, PRISM, enables investigators to compare their active molecules with other molecules that have been previously tested through NCI from both a biological and chemical perspective [44]; this evaluation is performed with two components: COMPARE and PILOT. Interestingly, when analyzing the most active compound 10a with PRISM, we found a best-fitting profile with N-(4-bromophenyl)-4-(2-cyclopropyloxazol-5-yl)benzenesulfonamide [45], which is a compound from a series of reported anticancer derivatives and is also able to inhibit tubulin polymerization, as we expect for our compounds.

2.2.2. In Vitro Tubulin Polymerization Inhibition

To confirm that the observed anticancer activity of compound 10a is conferred by a microtubule-targeting mechanism, we evaluated its effect on the assembly of tubulin, using paclitaxel (a tubulin stabilizer) and phenstatin (a tubulin polymerization inhibitor) as controls. As expected and shown in Figure 2, paclitaxel was found to stimulate tubulin polymerization, while phenstatin and compound 10a inhibited tubulin polymerization, although 10a had a slightly weaker effect. The obtained data indicate that compound 10a effectively inhibits tubulin polymerization in vitro, suggesting that the observed cytotoxicity of this compound is related to a microtubule-targeting mechanism.

2.3. Molecular Modeling

2.3.1. Blind Docking

Since compound 10a is similar to phenstatin in inhibiting tubulin polymerization in vitro, we hypothesize that it binds to the colchicine site of the α,β-tubulin heterodimer, the known site for phenstatin [46]. In addition, recent tubulin crystal structures co-crystalized with quinazolinone and tetrahydroisoquinoline anticancer agents reveal the preference of these heterocyclic compounds to the colchicine binding site [47,48] and support our previous hypothesis that quinoline derivatives could exert their anticancer activity by binding to the colchicine site of tubulin [6]. However, tubulin binding sites other than colchicine have been described, including the vinca, taxol, and peloruside/laulimalide sites [49], and other novel sites are being actively discovered and/or investigated for the rational design of tubulin modulators [50,51] or compounds targeting drug resistant cancer phenotypes [19,52]. Thus, we performed blind docking experiments for compound 10a, in order to investigate its relative preference towards the colchicine binding site, while validating our docking protocol using colchicine and phenstatin (Figure S46 from Supplementary Material).
The generated poses for compound 10a had estimated binding energies between −8.2 and −2.0 kcal/mol, with an overall score distribution and RMSD from best-scoring conformation for all docked poses resembling both colchicine and phenstatin (Supplementary Figure S47). Clustering with a 2.0 Å RMSD tolerance gave 127 unique cluster representatives for 10a, 13 of which scored lower than −7 kcal/mol. Out of these, the lowest-scoring conformation and an additional 10 cluster representatives were positioned in the colchicine binding site. All colchicine cluster representatives with estimated binding energies lower than −7 kcal/mol were positioned in the colchicine binding site of tubulin, as expected. RMSD between co-crystallized colchicine and lowest-scoring cluster representative from the blind docking was 1.082 Å, which is lower than the 2.0 Å cutoff generally used as a criterion for correct bound structure prediction [53], suggesting that the used docking protocol is suitable for the studied system. None of the phenstatin conformations scored less than −7 kcal/mol, but the four lowest-scoring cluster representatives were also positioned in the colchicine binding site. These results suggest that compound 10a prefers the colchicine binding site of tubulin and is likely able to inhibit tubulin polymerization by binding to the colchicine binding site of tubulin while having low affinity for other tubulin sites.

2.3.2. Local Docking

As the blind docking experiments revealed that the most energetically favorable poses for compound 10a are positioned in the colchicine binding site of tubulin, we performed local docking on this site in order to investigate the molecular nature of these preferential conformations. A clear improvement in the overall docking score distribution can be observed when compared to blind docking due to the decrease in conformational search space, as expected (Supplementary Figure S47). Interestingly, three low scoring clusters of comparable energies (−8.95 kcal/mol, −8.77 kcal/mol, and −8.69 kcal/mol—conformations I, II, and III, respectively) were revealed through local docking and were accommodated in the colchicine binding site in three geometrically different modes (BM I, BM II, and BM III). Of note, BM III greatly overlapped with the lowest-scoring cluster representative from blind docking for compound 10a (RMSD 1.034 Å). These three lowest-scoring cluster representatives were further chosen for analysis, as there have been cases of false positives in top-scoring poses from docking experiments, and subsequent analysis of multiple low-scoring docking poses is strongly recommended to correctly identify the most likely ligand conformation [54,55]. The RMSD between these modes was 7.51 Å and 6.408 Å (using BM I as a reference, as it is the lowest-scoring cluster), and the distances between their centers of mass were 3.197 Å (BM I/BM II), 3.294 Å (BM I/BM III), and 5.145 Å (BM II/BM III), indicating that the three conformations greatly differ from one another. The three best-scoring docking solutions from local docking also varied in terms of key amino acids involved in ligand stabilization at the binding site (Figure 3). For docking protocol validation, colchicine was re-docked in the same manner as 10a. RMSD between co-crystallized colchicine and the lowest-scoring cluster representative after colchicine re-docking was 1.110 Å.
In the case of BM I, the pyrrolo[1,2-a]quinoline ring is buried in the hydrophobic pocket lined by βCys241, βLeu242, βLeu252, βLeu255, βMet259, βVal315, and βAla316, while the trimethoxy-substituted ring is positioned towards the α subunit, in a manner similar to what we have previously seen for moderately active isoquinoline cycloadducts [6], yet very dissimilar to colchicine (Supplementary Figure S48). This conformation is further stabilized through hydrogen bonding with the sidechain of βLys254, an amino acid that has been previously identified through molecular docking experiments as an interaction partner for other tubulin polymerization inhibitors with anticancer activity [56,57,58].
However, to our knowledge, none of the co-crystallized tubulin inhibitors that bind to the colchicine site have been shown to interact with this residue, although it has been highlighted that this amino acid is likely involved in microtubule assembly through interaction with the γ-phosphate group of the N-site GTP [59]. Furthermore, alanine scan mutations have shown that D251A-R253A-K254A is lethal in yeast [60]. In addition, a recent molecular dynamics study focused on the dynamics of the βT7 loop concluded that an interaction between the backbone N atom of βLys254 and the side chain O atom of βAsp251 is essential for the structural rearrangement of the βT7 loop [61], which has been shown to play an important role in colchicine binding [62] and prevents the ‘curved-to-straight’ structural transition of tubulin from its free form, a process that is necessary for microtubule formation [63]. Therefore, βLys254 could be a likely interaction partner for compound 10a.
BM II was stabilized in the colchicine binding site exclusively through hydrophobic interactions, having an orientation similar to what we have previously predicted for active isoquinoline derivatives [6]. However, no hydrogen bond interaction with βCys241, the presumed anchor point for the trimethoxy-substituted ring, was observed. Nevertheless, the trimethoxy-substituted ring roughly occupies the same hydrophobic pocket as the trimethoxy-substituted moiety of colchicine (Figure S48). Since some colchicine site binders, including 2-aroylindoles, have been shown to be stabilized in the colchicine site exclusively through hydrophobic interactions and water-mediated polar interactions that do not include βCys241 [64,65], such a conformation could also be probable for compound 10a. In fact, this conformation occupies hydrophobic center II according to a previously described structure-based colchicine binding-site inhibitor model [21] and greatly overlaps with 2-aroylindole tubulin polymerization inhibitor D64131 [65], with an additional hydrophobic extension towards the α subunit (Figure S49).
In BM III, the trimethoxy-substituted ring was oriented towards the α subunit and engaged in hydrogen bonding with many amino acid sidechains known or thought to be involved in colchicine binding site inhibitor interaction, including αAsn101 [51], αThr179 [65], αSer178 [31], βLys352 [66], and βGln247 [67]. However, to our knowledge, no other colchicine site inhibitors possessing a trimethoxy-substituted ring have been co-crystallized in a similar conformation [18]. This conformation is also the most dissimilar from colchicine (Figure S48).
Importantly, all three lowest cluster representatives of 10a had similar theoretical binding energies. While scoring functions take into account a wide range of contributions, including electrostatic interactions, van der Waals contacts, and desolvation effects [68], they are limited in reflecting the conformational changes induced by ligand binding, as well as the stability of identified amino acid interactions in time.

2.3.3. Molecular Dynamics Simulations

To investigate the detailed dynamics and interaction stability of ligand-tubulin complexes, the tubulin heterodimer and the three best-scoring local docking poses were subjected to 10 ns MD simulations.
Overall, all three systems reached equilibrium, according to root mean square deviation (RMSD) analysis (Figure 4a), even though RMSD was higher for all investigated binding modes when compared to reference colchicine. In addition, BM I shifted from the initial frame more than BM II or BM III (Figure 4b), having a mean RMSD of 1.154 Å, while BM II and BM III had mean RMSD of 0.721 Å and 0.66 Å, respectively. All binding modes had a higher RMSD than colchicine (mean deviation 0.39 Å), which would be expected since the starting tubulin heterodimer structure is co-crystallized with colchicine. The ‘Lig fit Prot’ (Figure 4c) represents the RMSD of ligand-heavy atoms after first aligning the protein-ligand complex on the protein backbone. If the observed values differ greatly from the protein backbone RMSD, it is likely that the ligand diffuses away from the binding site [69]. In the case of BM I, this parameter shows drift during the simulation, particularly at the beginning of the trajectory, but is stabilized until the end of the simulation. This suggests that the initially found local docking conformation does not maintain close contact with surrounding amino acids and that another conformation is stabilized during the trajectory.
Indeed, for BM I, the 10 ns MD simulation revealed that the interaction with βLys254, which was the main anchor point for this docking conformation, was only transient, being maintained in less than 1% of the simulation time (Figure 5a). However, BM I maintained contacts in more than 25% of simulation time with αTyr224 (44.96%), βAsn249 (58%), βLys352 (53.35%), βAla250 (59.44%), βLeu255 (61.49%), βAla316 (37.91%), and the co-crystalized GTP molecule (39.51%). The ligand RMSD with regards to the initial docking conformation had the highest value of all simulated binding modes (3.543 Å). Superimposition between the docked solution and the final frame of the simulation revealed major differences in terms of amino acid contacts within the colchicine binding site (Figure S50). Of note, the carboxylate moiety flips after 3.57 ns and remains flipped throughout the simulation. The flexibility of this group is also evident in the ligand root mean square fluctuation—RMSF (Figure S51). Overall, the BM I-containing simulated system converged to the least similar conformation from the starting docking solution.
BM II maintained all the initially identified hydrophobic contacts, while slightly drifting from the initial docked conformation (RMSD 1.746 Å). In addition, a cation-pi interaction with βLys352 was observed in 27.32% of the simulation time. This binding mode was also the only one to engage in favorable interactions with βCys241, maintaining contact with this amino acid in 20.53% of simulation time (Figure 5b), mainly through the central methoxy moiety from its trimethoxy-substituted ring. The flexibility of this substituent in BM II is most evident from the abrupt increase or decrease in the main RMSD profile with respect to the first frame of the simulation (Figure 4b), as well as from the ligand root mean square fluctuation (Figure S51). Of note, RMSF of the same substituent in colchicine shows a similar behavior throughout the simulation (Figure S51).
In the case of BM III, the H-bonds with αAsn101, αSer178, and βGln247 identified from the docking solution were mostly not maintained during the MD simulation (Figure 5c), except for the initial frames (Figure S52). The H-bond with αThr179 was maintained in 15.03% of simulated time, either directly or mediated through a water molecule, while interaction with βLys352 was maintained throughout the entire simulation (Figure 5c). RMSD between the initial docked solution and the final frame of the MD simulation was 3.525 Å.
It is important to note that from all final frame MD simulation conformations, BM II was the only one that largely overlapped with co-crystallized colchicine binding site inhibitors with a trimethoxy-substituted ring [25,65,70], with most hydrophobic contacts being maintained throughout the simulation, even though the number of polar contacts was the smallest from all simulations (Figures S53 and S54). In addition, all simulated BMs for compound 10a maintained various kinds of contacts with βLys352 in more than 25% of the simulated time, suggesting that this amino acid could be relevant for the binding of this compound in the colchicine site, regardless of the initially identified docking solution. The importance of this residue in 10a binding could be further investigated through alanine scanning, as has been previously performed for the anticancer natural product pironetin [71].
Since receptor-ligand binding does not rely solely upon specific interactions between amino acids and particular chemical moieties but also on reduction of molecular flexibility upon binding [72], we further investigated the configurational entropy changes of all three binding modes upon interacting with the colchicine binding site of tubulin. This gave us an estimation of the entropic cost upon target binding. Since longer MD simulations are preferred in order to fully assess the stability of the investigated systems [30], and sufficient sampling should be reached to estimate configurational entropies [73], we extended our simulations to 25 ns, for which a good convergence of the configurational entropy profile was achieved. We found that the internal configurational entropy is reduced upon binding from 156.02 J/mol K to 99.52 J/mol K, 99.55 J/mol K, and 98.28 J/mol K, for BM I, BM II, and BM III, respectively (ΔSconf = −56.50 J/mol K; −56.47 J/mol K; −57.74 J/mol K, respectively). Of note, BM II exhibited a markedly lower internal configurational entropy than BM I and BM III in the first quarter of the simulation but steadily increased to comparable values as the other two binding modes throughout the rest of the simulated time (Figure S55). Our calculated values are similar in magnitude to other receptor-ligand systems [73].
Overall, the molecular modeling outcomes enabled us to predict, from a structural perspective, the interactions between our novel compound and tubulin with atomic resolution. Our results suggest that all three binding modes to tubulin described here are characterized by both positive enthalpic and entropic contributions, highlighting that all three binding modes are equally probable in a biological context. The molecular mechanics energies combined with the Poisson–Boltzmann or generalized Born and surface area continuum solvation (MM/PBSA and MM/GBSA) methods [74] could also have been of particular use in assessing ligand binding affinities for each of the three poses, but they have not been explored in the current study.

2.3.4. In Silico ADME and Toxicity Predictions

The result of the predicted parameters, including molecular properties, pharmacokinetics, drug-likeness, and medicinal chemistry are presented in Table 3.
Compound 10a follows both Lipinski’s rule of five (without any violations) and Veber’s rule, having less than 10 rotatable bonds and a topological polar surface area smaller than 140 Å2. In addition, compound 10a displays a moderately soluble behavior, with a Log S (ESOL) value of −6.00 and no PAINS or Brenk alerts in its structure. Derivative 10a also scores well in terms of bioavailability and ease of synthetic accessibility (Table 3). This compound is predicted to have a high gastrointestinal absorption but is unable to permeate through the blood-brain barrier (BBB). Compound 10a does not appear to be a P-glycoprotein (P-gp) substrate.
The chart generated from the Swiss ADME QSAR web tool, regarding the accessibility of compound 10a to be orally bioavailable, is presented in Figure 6.
This radar involves six parameters, lipophilicity (LIPO), size, polarity (POLAR), insolubility (INSOLU), unsaturation (INSATU), and flexibility (FLEX), of the tested compound and is represented by a red line integrated into a pink area. Molecules that fall within the pink region of the radar are considered drug-like. Compound 10a exhibited compliance to only four of the six rules, with violations to INSATU (ratio of hybridized sp3 atoms to the total number of C atoms) and LIPO (XLOGP3 between −0.7 and +5.0). Taken together, these predicted data show a promising ADME and drug-likeness profile for compound 10a.
The predicted toxicity spectrum is represented by a list of activities with probabilities “to be active” (Pa) and “to be inactive” (Pi). The obtained results presented in Table 4 show predicted cytotoxicity (Pa > Pi and Pa > 0.3) against several cancer cell lines, including four of the tested NCI cell lines: T-47D, HT-29, DU-145, and MCF7 (Table 3). The fact that no normal human cell lines appeared on the list could be an indication for a good selectivity of compound 10a against cancer cell lines.

3. Materials and Methods

3.1. Chemistry

All commercially available reagents and solvents employed were used without further purification. Melting points were recorded on an A. Krüss Optronic Melting Point Meter KSP1 and are uncorrected. Analytical thin-layer chromatography was performed with commercial silica gel plates 60 F254 (Merck Darmstadt, Germany) and visualized under UV light (λmax = 254 or 365 nm). The NMR spectra were recorded on a Bruker Avance III 500 MHz spectrometer or a Bruker Avance 400 DRX (400 MHz) (Bruker, Vienna, Austria). Chemical shifts are reported in delta (δ) units, part per million (ppm), and coupling constants (J) in Hz. The following abbreviations are used to designate chemical shift multiplicities: s = singlet, d = doublet, t = triplet, q = quartet, m = multiplet, and bs = broad singlet. Infrared (IR) spectra were recorded as films on potassium bromide (KBr) pellets on an FTIR Prestige 8400 s spectrophotometer (Shimadzu, Kyoto, Japan) or a Jasco 660 FTIR spectrophotometer. Elemental analyses indicated by the symbols of the elements were within ± 0.4% of the theoretical values. HR-MS experiments were recorded on a HESI Orbitrap Exploris 120 Mass Spectrometer (Thermo Fisher, Walthan, MA, USA) in positive mode.

3.1.1. General Procedure for Monoquaternary Salts 3, 6, 9, 12, and 15

The corresponding heterocycle (pyrazine 1, 4-(4-chlorophenyl)pyrimidine 5, quinoline 8, benzo[f]quinoline 11 or isoquinoline 14) (1 mmol, 1 equiv.) was dissolved in 5–7 mL acetone. Then, the corresponding reactive 2-bromoacetophenone 2a-d (1.1 mmol, 1.1 equiv.) was added and the resulting mixture was stirred overnight at room temperature (r.t.). The formed precipitate was filtered and washed with diethyl ether to obtain the desired product, which was used in the next reaction without any further purification.

3.1.2. General Procedure for Compounds 4, 7, 10, 13, and 16

The cycloimmonium salt (3, 6, 9, 12, and 15) (1 mmol, 1 equiv.) and ethyl propiolate (1.1 mmol, 1.1 equiv.) were added to dichloromethane (DCM). Then, a solution of triethylamine (TEA) (3 mmol, 3 equiv.) in DCM (3 mL) was added dropwise over 1 h (magnetic stirring), and the resulting mixture was stirred overnight at r.t. Methanol (10 mL) was added, and the resulting solid was collected by filtration and washed with 5 mL methanol. The product was then purified by crystallization from dichloromethane/methanol (1/1, v/v) and/or column chromatography using dichloromethane/methanol (99.5/0.5, v/v). Compounds 4, 7, 10, 13, and 16 were the only pure compounds obtained from the reaction mixture fractions. While spectral evidence for some decomposition products of the ylides was observed in the other fractions, no evidence was found for the presence of the other possible regioisomer cycloadducts.

3.1.3. Spectral Data

1-(2-oxo-2-(3,4,5-Trimethoxyphenyl)ethyl)pyrazin-1-ium Bromide 3a

Brown solid; 50% yield; mp 160–162 °C; IR ν(cm−1): 3009, 2986, 1682, 1630, 1584, 1445, 1416, 1186, 1124; 1H NMR (400 MHz, DMSO-d6) δppm: 3.80 (s, 3H, OMe), 3.90 (s, 6H, 2 × OMe), 6.67 (s, 2H, H7), 7.38 (s, 2H, H10, H14), 9.19 (d, J = 4.0 Hz, 2H, H3, H5), 9.72 (d, J = 4.0 Hz, 2H, H2, H6). 13C NMR (100 MHz, DMSO-d6) δppm: 56.4 (2 × OMe), 60.4 (OMe), 67.0 (C7), 106.2 (C10, C14), 128.4 (C9), 138.3 (C3, C5), 143.3 (C12), 150.9 (C2, C6), 153.0 (C11, C13), 188.2 (C8). Anal. Calcd. for C15H17BrN2O2: C, 48.80; H, 4.64; N, 7.59. Found: C, 48.90; H, 4.58; N, 7.68.

1-(2-(3,5-Dimethoxyphenyl)-2-oxoethyl)pyrazin-1-ium Bromide 3b

Beige solid; 48% yield; mp 200–202 °C; IR ν(cm−1): 3086, 3009, 2978, 1694, 1601, 1454, 1420, 1346, 1304, 1194, 1180, 1161, 1018; 1H NMR (500 MHz, DMSO-d6) δppm: 3.86 (s, 6H, 2 × OMe), 6.64 (s, 2H, H7), 6.94 (t, J = 2.0 Hz, 1H, H12), 7.19 (d, J = 2.5 Hz, 2H, H10, H14), 9.20 (d, J = 4.0 Hz, 2H, H3, H5), 9.71 (d, J = 4.0 Hz, 2H, H2, H6). 13C NMR (125 MHz, DMSO-d6) δppm: 55.8 (2 × OMe), 67.1 (C7), 106.2 (C10, C14), 106.2 (C12), 135.1 (C9), 138.4 (C3, C5), 150.9 (C2, C6), 160.9 (C11, C13), 189.1 (C8). Anal. Calcd. for C14H15BrN2O3: C, 49.57; H, 4.46; N, 8.26. Found: C, 49.55; H, 4.58; N, 8.28.

1-(2-(3,4-Dimethoxyphenyl)-2-oxoethyl)pyrazin-1-ium Bromide 3c

Beige solid; 52% yield; mp 196–198 °C; IR ν(cm−1): 3092, 3009, 2918, 1678, 1591, 1518, 1449, 1350, 1279, 1213, 1177, 1134; 1H NMR (500 MHz, DMSO-d6) δppm: 3.85 (s, 3H, OMe), 3.91 (s, 3H, OMe), 6.58 (s, 2H, H7), 7.23 (d, J = 8.5 Hz, 1H, H13), 7.52 (d, J = 1.5 Hz, 1H, H10), 7.78 (dd, J = 8.5; 1.5 Hz, 1H, H14), 9.18 (d, J = 3.0 Hz, 2H, H3, H5). 9.69 (d, J = 3.0 Hz, 2H, H2, H6). 13C NMR (125 MHz, DMSO-d6) δppm: 55.8 (OMe), 56.1 (OMe), 66.7 (C7), 110.4 (C10), 111.3 (C13), 123.6 (C14), 125.9 (C9), 138.4 (C3, C5), 148.9 (C11), 150.8 (C2, C6), 154.5 (C12), 187.5 (C8). Anal. Calcd. for C14H15BrN2O3: C, 49.57; H, 4.46; N, 8.26. Found: C, 49.63; H, 4.39; N, 8.30.

1-(2-(4-Bromophenyl)-2-oxoethyl)pyrazin-1-ium Bromide 3d

White solid; 57% yield; mp 240–243 °C; IR ν(cm−1): 3023, 2915, 1690, 1586, 1447, 1397, 1240, 986; 1H NMR (500 MHz, DMSO-d6) δppm: 6.57 (s, 2H, H7), 7.91 (d, J = 8.4 Hz, 2H, H11, H13), 8.02 (d, J = 8.4 Hz, 2H, H10, H14), 9.15 (d, J = 4.5 Hz, 2H, H3, H5), 9.70 (d, J = 4.4 Hz, 2H, H2, H6). 13C NMR (100 MHz, DMSO-d6) δppm: 66.8 (C7), 129.0 (C12), 130.2 (C10, C14), 132.2 (C11, C13, C9), 138.3 (C3, C5), 150.7 (C2, C6), 188.6 (C8). Anal. Calcd. for C12H10Br2N2O: C, 40.26; H, 2.82; N, 7.82. Found: C, 40.22; H, 2.78; N, 7.88.

Ethyl 6-(3,4,5-Trimethoxybenzoyl)pyrrolo[1,2-a]pyrazine-8-carboxylate 4a

Beige solid; 40% yield; mp 162–164 °C; IR ν(cm−1): 2999, 2947, 1713, 1632, 1581, 1518, 1469, 1323, 1221, 1204; 1H NMR (400 Mz, CDCl3) δppm: 1.44 (t, J = 7.2 Hz, 3H, CH3), 3.91 (s, 6H, 2 × OMe), 3.97 (s, 3H, OMe), 4.45 (q, J = 7.2 Hz, 2H, CH2), 7.10 (s, 2H, H12, H16),7.91 (s, 1H, H7), 8.15 (d, J = 4.8 Hz, 1H, H4), 9.59 (dd, J = 4.8; 1.6 Hz, 1H, H3), 9.77 (d, J = 1.6 Hz, 1H, H1). 13C NMR (100 MHz, CDCl3) δppm: 14.6 (CH3), 56.6 (2 × OMe), 61.0 (CH2), 61.2 (OMe), 106.9 (C12, C16), 120.6 (C3), 123.3 (C8), 127.3 (C7), 132.1 (C6), 132.8 (C4), 134.0 (C11), 136.2 (C9), 142.1 (C14), 146.1 (C1), 153.3 (C13, C15), 163.2 (COO), 185.5 (C10). Anal. Calcd. for C20H20N2O6: C, 62.49; H, 5.24; N, 7.29. Found: C, 62.44; H, 5.20; N, 7.30.

Ethyl 6-(3,5-Dimethoxybenzoyl)pyrrolo[1,2-a]pyrazine-8-carboxylate 4b

Beige solid; 35% yield; mp 160–162 °C; IR ν(cm−1): 2984, 2940, 1691, 1630, 1586, 1516, 1468, 1340, 1315, 1219; 1H NMR (400 Mz, CDCl3) δppm: 1.43 (t, J = 7.0 Hz, 3H, CH3), 3.86 (s, 6H, 2 × OMe), 4.40 (q, J = 7.0 Hz, 2H, CH2), 6.69 (t, J = 2.5 Hz, 1H, H14), 6.94 (d, J = 2.5 Hz, 2H, H12, H16), 7.91 (s, 1H, H7), 8.15 (d, J = 5.0 Hz, 1H, H4), 9.63 (dd, J = 4.5; 1.5 Hz, 1H, H3), 9.77 (bs, 1H, H1). 13C NMR (100 MHz, CDCl3) δppm: 14.6 (CH3), 56.8 (2 × OMe), 61.0 (CH2), 104.5 (C14), 107.1 (C12, C16), 109.3 (C6), 120.7 (C3), 123.3 (C8), 127.7 (C7), 132.2 (C9), 132.9 (C4), 140.1 (C11), 146.1 (C1), 160.9 (C13, C15), 163.2 (COO), 186.1 (C10). Anal. Calcd. for C19H18N2O5: C, 64.40; H, 5.12; N, 7.91. Found: C, 64.44; H, 5.15; N, 7.89.

Ethyl 6-(3,4-Dimethoxybenzoyl)pyrrolo[1,2-a]pyrazine-8-carboxylate 4c

Beige solid; 30% yield; mp 170–172°C; IR ν(cm−1): 2926, 2853, 1705, 1674, 1663, 1595, 1466, 1269, 1209, 1136, 1022; 1H NMR (500 Mz, CDCl3) δppm: 1.43 (t, J = 7.0 Hz, 3H, CH3), 3.96 (s, 3H, OMe), 3.98 (s, 3H, OMe), 4.42 (q, J = 7.0 Hz, 2H, CH2), 6.97 (d, J = 8.5 Hz, 1H, H15), 7.44 (s, 1H, H12), 7.50 (d, J = 8.0 Hz, 1H, H16), 7.88 (s, 1H, H7), 8.10 (d, J = 4.0 Hz, 1H, H4), 9.54 (d, J = 4.0 Hz, 1H, H3), 9.74 (s, 1H, H1). 13C NMR (125 MHz, CDCl3) δppm: 14.5 (CH3), 56.1 (OMe), 56.2 (OMe), 60.8 (CH2), 110.2 (C15), 111.6 (C12), 120.4 (C3), 123.4 (C8), 123.9 (C16), 126.9 (C7), 131.4 (C11), 131.8 (C6), 132.4 (C4), 134.5 (C9), 145.9 (C1), 149.3 (C13), 153.0 (C14), 163.2 (COO), 184.9 (C10). Anal. Calcd. for C19H18N2O5: C, 64.40; H, 5.12; N, 7.91. Found: C, 64.45; H, 5.09; N, 7.92.

Ethyl 6-(4-Bromobenzoyl)pyrrolo[1,2-a]pyrazine-8-carboxylate 4d

Beige solid; 35% yield; mp 167–169 °C; IR ν(cm−1): 2924, 2866, 1699, 1630, 1587, 1522, 1466, 1350, 1269, 1225; 1H NMR (400 Mz, CDCl3) δppm: 1.44 (t, J = 7.2 Hz, 3H, CH3), 4.44 (q, J = 7.2 Hz, 2H, CH2), 7.69 (d, J = 8.4 Hz, 2H, H12, H16), 7.73 (d, J = 8.4 Hz, 2H, H13, H15), 7.83 (s, 1H, H7), 8.17 (d, J = 4.4 Hz, 1H, H4), 9.64 (dd, J = 4.8; 1.6 Hz, 1H, H3), 9.78 (d, 1H, J = 1.2 Hz, H1). 13C NMR (100 MHz, CDCl3) δppm: 14.5 (CH3), 60.9 (CH2), 120.5 (C3), 122.9 (C8), 127.3 (C14), 127.4 (C7), 130.6 (C13, C15), 132.0 (C12, C16), 132.1 (C6), 132.9 (C4), 134.5 (C9), 137.5 (C11, C9), 145.9 (C1), 163.0 (COO), 185.1 (C10). Anal. Calcd. for C17H13BrN2O3: C, 54.71; H, 3.51; N, 7.51. Found: C, 54.75; H, 4.48; N, 7.49.

4-(4-Chlorophenyl)-1-(2-oxo-2-(3,4,5-trimethoxyphenyl)ethyl)pyrimidin-1-ium Bromide 6a

Yellow solid; 53% yield; mp 205–207 °C; IR ν(cm−1): 3009, 2997, 2958, 1674, 1630, 1593, 1452, 1416, 1339, 1314, 1198, 1165, 1128, 1093; 1H NMR (400 MHz, DMSO-d6) δppm: 3.82 (s, 3H, OMe), 3.92 (s, 6H, 2 × OMe), 6.53 (s, 2H, H7), 7.42 (s, 2H, H10, H14), 7.81 (d, J = 8.4 Hz, 2H, H17, H19), 8.51 (d, J = 8.4 Hz, 2H, H16, H20), 9.04 (d, J = 6.8 Hz, 1H, H5), 9.43 (dd, J = 6.8; 0.8 Hz, 1H, H6), 9.81 (s, 1H, H1). 13C NMR (100 MHz, DMSO-d6) δppm: 56.3 (2 × OMe), 60.3 (OMe), 62.4 (CH2), 106.0 (C10, C14), 118.2 (C5), 128.5 (C9), 129.9 (C17, C19), 130.9 (C16, C20), 131.7 (C15), 139.9 (C18), 143.1 (C12), 153.0 (C11, C13), 153.2 (C6), 154.5 (C2), 167.6 (C4), 189.1 (C8). Anal. Calcd. for C21H20BrClN2O4: C, 52.57; H, 4.20; N, 5.84. Found: C, 51.55; H, 4.18; N, 5.88.

4-(4-Chlorophenyl)-1-(2-(3,5-dimethoxyphenyl)-2-oxoethyl)pyrimidin-1-ium Bromide 6b

Yellow solid; 53% yield; mp 200–203 °C; IR ν(cm−1): 3026, 2988, 2930, 1692, 1628, 1593, 1549, 1454, 1341, 1316, 1298, 1206, 1155, 1009; 1H NMR (500 MHz, DMSO-d6) δppm: 3.87 (s, 6H, 2 × OMe), 6.40 (s, 2H, H7), 6.95 (t, J = 2.0 Hz, 1H, H12), 7.21 (d, J = 2.5 Hz, 2H, H10, H14), 7.81 (d, J = 8.5 Hz, 2H, H17, H19), 8.50 (d, J = 8.5 Hz, 2H, H16, H20), 8.99 (d, J = 7.0 Hz, 1H, H5), 9.36 (dd, J = 7.0; 1.5 Hz, 1H, H6), 9.74 (s, 1H, H1). 13C NMR (125 MHz, DMSO-d6) δppm: 55.8 (2 × OMe), 62.6 (CH2), 106.2 (C10, C14), 106.3 (C12), 118.3 (C5), 130.0 (C17, C19), 131.1 (C16, C20), 131.8 (C15), 135.3 (C9), 140.0 (C18), 153.3 (C6), 154.5 (C2), 149.0 (C11, C13), 167.7 (C4), 190.0 (C8). Anal. Calcd. for C20H18BrClN2O3: C, 53.41; H, 4.03; N, 6.23. Found: C, 53.40; H, 4.00; N, 6.19.

4-(4-Chlorophenyl)-1-(2-(3,4-dimethoxyphenyl)-2-oxoethyl)pyrimidin-1-ium Bromide 6c

Yellow solid; 53% yield; mp 208–210 °C; IR ν(cm−1): 3028, 2916, 1670, 1634, 1593, 1348, 1275, 1177, 1136, 1019, 839; 1H NMR (500 MHz, DMSO-d6) δppm: 3.86 (s, 3H, OMe), 3.92 (s, 3H, OMe), 6.38 (s, 2H, H7), 7.24 (d, J = 8.5 Hz, 1H, H13), 7.54 (d, J = 2.0 Hz, 1H, H10), 7.80–7.82 (m, 3H, H17, H19, H14), 8.50 (d, J = 8.5 Hz, 2H, H16, H20), 8.98 (d, J = 6.5 Hz, 1H, H5), 9.36 (dd, J = 7.0; 1.0 Hz, 1H, H6), 9.75 (s, 1H, H1). 13C NMR (125 MHz, DMSO-d6) δppm: 55.8 (OMe), 56.1 (OMe), 62.2 (CH2), 110.3 (C10), 111.3 (C13), 118.2 (C5), 123.6 (C14), 126.1 (C9), 130.0 (C17, C19), 131.0 (C16, C20), 131.8 (C15), 140.0 (C18), 149.0 (C11), 153.4 (C6), 154.4 (C12), 154.5 (C2), 167.7 (C4), 188.5 (C8). Anal. Calcd. for C20H18BrClN2O3: C, 53.41; H, 4.03; N, 6.23. Found: C, 53.45; H, 4.02; N, 6.28.

1-(2-(4-Bromophenyl)-2-oxoethyl)-4-(4-chlorophenyl)pyrimidin-1-ium Bromide 6d

Yellow solid; 55% yield; mp 240–243 °C; IR ν(cm−1): 2990, 2918, 1688, 1626, 1584, 1449, 1208, 981; 1H NMR (500 MHz, DMSO-d6) δppm: 6.41 (s, 2H, H7), 7.82 (d, J = 8.5 Hz, 2H, H17, H19), 7.92 (d, J = 8.5 Hz, 2H, H11, H13), 8.04 (d, J = 8.5 Hz, 2H, H10, H14), 8.50 (d, J = 8.5 Hz, 2H, H16, H20), 8.99 (d, J = 7.0 Hz, 1H, H5), 9.35 (d, J = 7.0 Hz, 1H, H6), 9.73 (s, 1H, H1). 13C NMR (125 MHz, DMSO-d6) δppm: 62.4 (CH2), 118.3 (C5), 129.0 (C12), 130.0 (C17, C19), 130.3 (C10, C14), 131.0 (C16, C20), 131.8 (C15), 132.4 (C11, C13), 135.5 (C9), 140.0 (C18), 153.4 (C6), 154.5 (C2), 167.7 (C4), 189.6 (C8). Anal. Calcd. for C18H13Br2ClN2O: C, 46.14; H, 2.80; N, 5.98. Found: C, 46.15; H, 2.80; N, 6.00.

Ethyl 3-(4-Chlorophenyl)-7-(3,4,5-trimethoxybenzoyl)pyrrolo[1,2-c]pyrimidine-5-carboxylate 7a

Brown solid; 36% yield; mp 180–181 °C; IR ν(cm−1): 3088, 2936, 1705, 1622, 1584, 1478, 1412, 1340, 1209, 1128; 1H NMR (400 MHz, CDCl3) δppm: 1.43 (t, J = 7.2 Hz, 3H, CH3), 3.94 (s, 6H, 2 × OMe), 3.97 (s, 3H, OMe), 4.42 (q, J = 7.2 Hz, 2H, CH2), 7.12 (s, 2H, H18, H22), 7.50 (d, J = 8.0 Hz, 2H, H12, H14),7.89 (s, 1H, H6), 8.15 (d, J = 8.5 Hz, 2H, H11, H15), 8.61 (s, 1H, H4), 10.53 (s, 1H, H1). 13C NMR (100 MHz, CDCl3) δppm: 14.7 (CH3), 56.6 (2 × OMe), 60.7 (CH2), 61.2 (OMe), 106.8 (C18, C22), 108.7 (C4), 122.4 (C5), 128.3 (C11, C15), 129.4 (C12, C14), 129.5 (C17), 129.7 (C6), 134.1 (C7), 135.2 (C10), 136.4 (C13), 140.7 (C8), 141.0 (C1), 142.0 (C20), 148.6 (C3), 153.3 (C19, C21), 163.7 (COO), 184.5 (C16). Anal. Calcd. for C26H23ClN2O6: C, 63.10; H, 4.68; N, 5.66. Found: C, 63.14; H, 4.70; N, 5.69.

Ethyl 3-(4-Chlorophenyl)-7-(3,5-trimethoxybenzoyl)pyrrolo[1,2-c]pyrimidine-5-carboxylate 7b

Yellow solid; 40% yield; mp 222–225 °C; IR ν(cm−1): 3088, 2982, 2936, 1690, 1620, 1593, 1524, 1474, 1350, 1207, 1161, 1049; 1H NMR (500 Mz, CDCl3) δppm: 1.43 (t, J = 7.0 Hz, 3H, CH3), 3.87 (s, 6H, 2 × OMe), 4.42 (q, J = 7.0 Hz, 2H, CH2), 6.70 (t, J = 2.0 Hz, 1H, H20), 6.97 (d, J = 2.0 Hz, 2H, H18, H22), 7.49 (d, J = 8.5 Hz, 2H, H12, H14), 7.89 (s, 1H, H6), 8.12 (d, J = 8.5 Hz, 2H, H11, H15), 8.62 (d, J = 1.5 Hz, 1H, H4), 10.58 (d, J = 1.0 Hz, 1H, H1). 13C NMR (125 MHz, CDCl3ppm: 14.7 (CH3), 55.8 (2 × OMe), 60.7 (CH2),104.3 (C20), 107.1 (C18, C22), 107.4 (C5), 108.7 (C4), 122.5 (C7), 128.3 (C11, C15), 129.4 (C12, C14), 130.1 (C6), 135.2 (C10, C17), 136.4 (C13), 140.9 (C8), 141.0 (C1), 148.8 (C3), 160.9 (C19, C21), 163.7 (COO), 185.2 (C16). Anal. Calcd. for C26H23ClN2O6: C, 63.10; H, 4.68; N, 5.66. Found: C, 63.12; H, 4.65; N, 5.64.

Ethyl 3-(4-Chlorophenyl)-7-(3,4-trimethoxybenzoyl)pyrrolo[1,2-c]pyrimidine-5-carboxylate 7c

Beige solid; 35% yield; mp 150–153 °C; IR ν(cm−1): 2976, 2930, 1703, 1620, 1516, 1476, 1267, 1211, 1175, 1089; 1H NMR (500 Mz, CDCl3) δppm: 1.43 (t, J = 7.0 Hz, 3H, CH3), 3.98 (s, 3H, OMe), 4.00 (s, 3H, OMe), 4.42 (q, J = 7.0 Hz, 2H, CH2), 6.99 (d, J = 8.0 Hz, 1H, H21), 7.44–7.54 (m, 6H, H18, H22, H12, H14, H22, H18), 7.87 (s, 1H, H6), 8.12 (d, J = 8.0 Hz, 2H, H11, H15), 8.60 (s, 1H, H4), 10.52 (s, 1H, H1). 13C NMR (125 MHz, CDCl3) δppm: 14.7 (CH3), 56.2 (OMe), 56.3 (OMe), 60.7 (CH2), 107.1 (C5), 108.7 (C4), 110.3 (C21), 111.7 (C18), 122.7 (C7), 123.8 (C22), 128.3 (C11, C15), 129.4 (C12, C14), 129.5 (C6), 131.6 (C17), 135.3 (C10), 136.3 (C13), 140.5 (C8), 141.0 (C1), 148.4 (C3), 149.4 (C19), 153.0 (C20), 163.8 (COO), 184.3 (C16). Anal. Calcd. for C26H23ClN2O6: C, 63.10; H, 4.68; N, 5.66. Found: C, 63.13; H, 4.69; N, 5.60.

Ethyl 7-(4-Bromobenzoyl)-3-(4-chlorophenyl)pyrrolo[1,2-c]pyrimidine-5-carboxylate 7d

Yellow solid; 40% yield; mp 162–164 °C; IR ν(cm−1): 3094, 2986, 1722, 1699, 1613, 1471, 1265, 1234, 1202; 1H NMR (500 Mz, CDCl3) δppm: 1.43 (t, J = 7.0 Hz, 3H, CH3), 4.42 (q, J = 7.0 Hz, 2H, CH2), 7.49 (d, J = 8.0 Hz, 2H, H12, H14), 7.70 (d, J = 8.0 Hz, 2H, H18, H22), 7.74 (d, J = 8.5 Hz, 2H, H19, H21), 7.81 (s, 1H, H6), 8.13 (d, J = 8.5 Hz, 2H, H11, H15), 8.63 (s, 1H, H4), 10.58 (s, 1H, H1). 13C NMR (125 MHz, CDCl3) δppm: 14.7 (CH3), 60.8 (CH2), 107.5 (C5), 108.7 (C4), 122.2 (C7), 127.2 (C20), 128.3 (C11, C15), 129.4 (C12, C14), 129.9 (C6), 130.7 (C19, C21), 132.1 (C18, C22), 135.1 (C10), 137.8 (C17), 136.5 (C13), 140.2 (C8), 141.0 (C1), 149.0 (C3), 163.6 (COO), 184.2 (C16). Anal. Calcd. for C23H16BrClN2O3: C, 57.11; H, 3.33; N, 5.79. Found: C, 57.13; H, 3.39; N, 5.76.

1-(2-oxo-2-(3,4,5-Trimethoxyphenyl)ethyl)quinolin-1-ium Bromide 9a

Yellow solid; 80% yield; mp 278–279 °C; IR ν(cm−1): 3064, 2964, 1693, 1589, 1416, 1320, 1118, 995; 1H NMR (500 MHz, DMSO-d6) δppm: 3.96 (s, 3H, OMe), 4.00 (s, 6H, 2 × OMe), 7.58 (s, 2H, H14, H18), 7.67 (s, 2H, H11), 7.96 (t, J = 7.0 Hz, 1H, H6), 8.06–8.15 (overlapped signals, 3H, H7, H3, H5), 8.27 (d, J = 8.5 Hz, 1H, H8), 9.00 (d, J = 7.5 Hz, 1H, H4), 10.34 (d, J = 6.0 Hz, 1H, H2). 13C NMR (125 MHz, DMSO-d6) δppm: 57.1 (2 × OMe), 61.2 (OMe), 64.3 (C11), 106.7 (C14, C18), 119.0 (C5), 122.1 (C3), 128.6 (C13), 129.9 (C9), 130.4 (C6), 130.6 (C8), 136.3 (C7), 139.5 (C10), 144.4 (C1), 147.5 (C4), 151.3 (C2), 153.7 (C15, C17), 189.1 (C12). Anal. Calcd. for C20H20BrNO4: C, 57.43; H, 4.82; N, 3.35. Found: C, 57.42; H, 4.80; N, 3.36.

1-(2-oxo-2-(3,5-Trimethoxyphenyl)ethyl)quinolin-1-ium Bromide 9b

Beige solid; 63% yield; mp 265–268 °C; IR ν(cm−1): 3007, 2871, 1680, 1649, 1597, 1483, 1352, 1290, 1203, 1166, 1022; 1H NMR (500 MHz, DMSO-d6) δppm: 3.87 (s, 6H, 2 × OMe), 6.97 (s, 1H, H16), 7.01 (s, 2H, H11), 7.28 (s, 2H, H14, H18), 8.07 (t, J = 7.0 Hz, 1H, H6), 8.24 (t, J = 7.5 Hz, 1H, H7), 8.33 (t, J = 6.5 Hz, 1H, H3), 8.40 (d, J = 8.5 Hz, 1H, H5), 8.55 (d, J = 7.5 Hz, 1H, H8), 9.45 (d, J = 7.5 Hz, 1H, H4), 9.50 (d, J = 5.0 Hz, 1H, H2). 13C NMR (125 MHz, DMSO-d6) δppm: 55.8 (2 × OMe), 63.3 (C11), 106.5 (C14, C18, C16) 119.1 (C5), 122.2 (C3), 129.4 (C9), 130.0 (C6), 130.7 (C8), 135.4 (C13), 136.0 (C7), 138.6 (C10), 148.7 (C4), 151.0 (C2), 160.8 (C15, C17), 190.6 (C12). Anal. Calcd. for C19H8BrNO3: C, 58.78; H, 4.67; N, 3.61. Found: C, 58.77; H, 4.65; N, 3.63.

1-(2-oxo-2-(3,4-Trimethoxyphenyl)ethyl)quinolin-1-ium bromide 9c

Beige solid; 51% yield; mp 270–271 °C; IR ν(cm−1): 3042, 3010, 2934, 1676, 1644, 1592, 1515, 1280, 1160, 1021; 1H NMR (500 MHz, DMSO-d6) δppm: 3.86 (s, 3H, OMe), 3.93 (s, 3H, OMe), 7.00 (s, 2H, H11), 7.27 (d, J = 8.5 Hz, 1H, H17), 7.58 (d, J = 1.5 Hz, 1H, H14), 7.89 (dd, J = 8.5; 1.5 Hz, 1H, H18), 8.07 (t, J = 7.5 Hz, 1H, H6), 8.23 (t, J = 7.5 Hz, 1H, H7), 8.32 (dd, J = 8.5; 6.0 Hz, 1H, H3), 8.38 (d, J = 8.0 Hz, 1H, H5), 8.55 (d, J = 8.0 Hz, 1H, H8), 9.45 (d, J = 8.5 Hz, 1H, H4), 9.53 (d, J = 5.5 Hz, 1H, H2). 13C NMR (125 MHz, DMSO-d6) δppm: 55.8 (OMe), 56.1 (OMe), 62.9 (C11), 110.6 (C14), 111.3 (C17), 119.1 (C5), 122.2 (C3), 123.8 (C18), 126.3 (C13), 129.4 (C9), 130.0 (C6), 130.7 (C8), 136.0 (C7), 138.6 (C10), 148.5 (C4), 148.9 (C15), 151.0 (C2), 154.5 (C16), 189.0 (C12). Anal. Calcd. for C19H8BrNO3: C, 58.78; H, 4.67; N, 3.61. Found: C, 58.76; H, 4.67; N, 3.62.

1-(2-(4-Bromophenyl)-2-oxoethyl)quinolin-1-ium Bromide 9d

Physical and spectral data are similar to the ones reported in ref. [6].

Ethyl 1-(3,4,5-Trimethoxybenzoyl)pyrrolo[1,2-a]quinoline-3-carboxylate 10a

Beige solid; 39% yield; mp 231–234 °C; IR ν(cm−1): 2947, 1711, 1628, 1584, 1459, 1361, 1192, 1144, 759; 1H NMR (500 Mz, CDCl3) δppm: 1.41 (t, J = 7.0 Hz, 3H, CH2CH3), 3.93 (s, 6H, 2 × OMe), 3.99 (s, 3H, OMe), 4.39 (q, J = 7.0 Hz, 2H, CH2CH3), 7.36 (s, 2H, H15,H19), 7.49 (t, J = 7.5 Hz, 1H, H7), 7.57 (dt, J = 7.5; 1.5 Hz, 1H, H8), 7.66 (s, 1H, H2), 7.69 (d, J = 9.0 Hz, 1H, H5), 7.82 (d, J = 7.5 Hz, 1H, H6), 8.00 (d, J = 8.5 Hz, 1H, H9), 8.34 (d, J = 9.0 Hz, 1H, H4). 13C NMR (125 MHz, CDCl3) δppm: 14.7 (CH2CH3), 56.5 (2 × OMe), 60.4 (CH2CH3), 61.2 (OMe), 107.8 (C15, C19), 107.9 (C3), 117.9 (C4), 120.2 (C9), 125.3 (C11), 125.6 (C7), 128.0 (C1), 128.8 (C2), 128.9 (C5), 129.0 (C8), 129.1 (C6), 133.3 (C12), 133.6 (C14), 140.3 (C10), 142.6 (C17), 153.2 (C16, C18), 164.3 (COO), 184.5 (C13). Anal. Calcd. for C25H23NO6: C, 69.27; H, 5.35; N, 3.23. Found: C, 69.28; H, 5.33; N, 3.25. HRMS (HESI+): m/z calcd for C25H23NO6 (M+): 433.1525; found 433.1519.

Ethyl 1-(3,5-Trimethoxybenzoyl)pyrrolo[1,2-a]quinoline-3-carboxylate 10b

Yellow solid; 43% yield; mp 240–241 °C; IR ν(cm−1): 2976, 1710, 1629, 1595, 1457, 1360, 1192, 760; 1H NMR (500 Mz, CDCl3) δppm: 1.40 (t, J = 7.0 Hz, 3H, CH2CH3), 3.88 (s, 6H, 2 × OMe), 4.38 (q, J = 7.0 Hz, 2H, CH2CH3), 6.75 (t, J = 2.0 Hz, 1H, H17), 7.22 (d, J = 2.5 Hz, 2H, H15, H19), 7.49 (t, J = 7.5 Hz, 1H, H7), 7.57 (t, J = 7.5 Hz, 1H, H8), 7.68 (s, 1H, H2), 7.69 (d, J = 9.0 Hz, 1H, H5), 7.82 (d, J = 7.0 Hz, 1H, H6), 8.04 (d, J = 9.0 Hz, 1H, H9), 8.33 (d, J = 9.0 Hz, 1H, H4). 13C NMR (125 MHz, CDCl3) δppm: 14.7 (CH2CH3), 55.8 (2 × OMe), 60.4 (CH2CH3), 105.6 (C17), 107.9 (C3), 108.9 (C15, C19), 117.9 (C4), 120.3 (C9), 125.3 (C11), 125.6 (C7), 128.2 (C1), 128.9 (C5), 129.0 (C8), 129.1 (C6), 129.7 (C2), 133.3 (C12), 140.5 (C10, C14), 160.9 (C16, C18), 164.3 (COO), 184.8 (C13). Anal. Calcd. for C24H21NO5: C, 71.45; H, 5.25; N, 3.47. Found: C, 71.43; H, 5.24; N, 3.49.

Ethyl 1-(3,4-Trimethoxybenzoyl)pyrrolo[1,2-a]quinoline-3-carboxylate 10c

Beige solid; 42% yield; mp 259–261 °C; IR ν(cm−1): 2919; 1709, 1638, 1514, 1461, 1384, 1268, 1140, 1081, 1020; 1H NMR (500 Mz, CDCl3) δppm: 1.40 (t, J = 7.0 Hz, 3H, CH2CH3), 3.99 (s, 3H, OMe), 4.01 (s, 3H, OMe), 4.38 (q, J = 7.0 Hz, 2H, CH2CH3), 6.99 (d, J = 8.5 Hz, 1H, H18), 7.47 (dt, J = 8.0; 1.0 Hz, 1H, H7), 7.55 (dt, J = 7.0; 1.5 Hz, 1H, H8), 7.62 (s, 1H, H2), 7.65–7.69 (overlapped signals, 2H, H5, H15), 7.77 (dd, J = 8.0; 2.0 Hz, 1H, H19), 7.80 (dd, J = 8.0; 1.5 Hz, 1H, H6), 7.98 (d, J = 8.5 Hz, 1H, H9), 8.32 (d, J = 9.5 Hz, 1H, H4). 13C NMR (125 MHz, CDCl3) δppm: 14.7 (CH2CH3), 56.2 (OMe), 56.3 (OMe), 60.3 (CH2CH3), 107.6 (C3), 110.1 (C18), 112.0 (C15), 117.9 (C4), 120.1 (C9), 125.2 (C11), 125.5 (C7), 125.6 (C19), 128.1 (C1), 128.4 (C2), 128.6 (C5), 128.8 (C8), 129.1 (C6), 131.2 (C14), 133.3 (C12), 139.9 (C10), 149.3 (C16), 153.6 (C17), 164.4 (C20), 184.6 (C13). Anal. Calcd. for C24H21NO5: C, 71.45; H, 5.25; N, 3.47. Found: C, 71.46; H, 5.26; N, 3.48.

Ethyl 1-(4-Bromobenzoyl)pyrrolo[1,2-a]quinoline-3-carboxylate 10d

Yellow solid; 60% yield; mp 179–180 °C; 1H NMR (CDCl3, 500 MHz) δppm: 1.40 (t, J = 7.0 Hz, 3H, CH2CH3), 4.38 (q, J = 7.0 Hz, 2H, CH2CH3), 7.50 (dt, J = 8.0; 1.0 Hz, 1H, H7), 7.59 (dt, J = 7.0; 1.5 Hz, 1H, H8), 7.62 (s, 1H, H2), 7.69–7.72 (overlapped signals, 3H, H5, H16,H18), 7.82 (dd, J = 8.0; 1.5 Hz, 1H, H6), 7.97 (d, J = 8.5 Hz, 2H, H15, H19), 8.03 (d, J = 8.5 Hz, 1H, H9), 8.34 (d, J = 9.0 Hz, 1H, H4). 13C NMR (CDCl3, 125 MHz) δppm: 14.7 (CH2CH3), 60.4 (CH2CH3), 108.1 (C3), 117.8 (C4), 120.3 (C9), 125.3 (C11), 125.7 (C7), 127.8 (C1), 128.1 (C17), 129.0 (C8), 129.1 (C6), 129.8 (C2), 129.4 (C5), 131.8 (C15, C19), 132.0 (C16, C18), 133.3 (C14), 137.4 (C12), 140.7 (C10), 164.1 (C20), 183.8 (C13). Anal. Calcd. for C22H16BrNO3: C, 62.57; H, 3.82; N, 3.32. Found: C, 62.56; H, 3.80; N, 3.33.

4-(2-oxo-2-(3,4,5-Trimethoxyphenyl)ethyl)benzo[f]quinolin-4-ium Bromide 12a

Beige solid; 54% yield; mp 212–214 °C; IR ν(cm−1): 3092, 2980, 2933, 1690, 1584, 1416, 1323, 1234, 1159, 1122, 1064, 993, 762; 1H NMR (500 MHz, DMSO-d6) δppm: 3.82 (s, 3H, OMe), 3.93 (s, 6H, 2 × OMe), 7.18 (s, 2H, H11), 7.50 (s, 2H, H14, H18), 7.99 (t, J = 7.0 Hz, 1H, H7), 8.05 (t, J = 7.0 Hz, 1H, H6), 8.27 (d, J = 9.5 Hz, 1H, H10), 8.31 (d, J = 8.0 Hz, 1H, H8), 8.48 (dd, J = 8.5; 6.0 Hz, 1H, H3), 8.66 (d, J = 10.0 Hz, 1H, H9), 9.18 (d, J = 8.5 Hz, 1H, H5), 9.50 (d, J = 5.5 Hz, 1H, H2), 10.28 (d, J = 8.5 Hz, 1H, H4). 13C NMR (125 MHz, DMSO-d6) δppm: 56.4 (2 × OMe), 60.4 (OMe), 64.0 (C11), 106.4 (C14, C18), 116.3 (C10), 122.9 (C3), 124.3 (C5), 127.9 (C4a), 128.1 (C4b), 128.7 (C13), 129.5 (C8), 130.2 (C6), 130.2 (C7), 130.9 (C8a), 138.2 (C9), 140.1 (C10a), 142.6 (C4), 143.3 (C16), 148.2(C2), 153.0 (C15, C17), 189.7 (C12). Anal. Calcd. for C24H22BrNO4: C, 61.55; H, 4.73; N, 2.99. Found: C, 61.54; H, 4.70; N, 3.00.

4-(2-oxo-2-(3,5-Trimethoxyphenyl)ethyl)benzo[f]quinolin-4-ium Bromide 12b

Yellow solid; 51% yield; mp 163–165 °C; IR ν(cm−1): 3062, 2978, 2903, 1692, 1593, 1423, 1344, 1304, 1200, 1155, 1061, 810, 750; 1H NMR (400 MHz, DMSO-d6) δppm: 3.89 (s, 6H, 2 × OMe), 6.98 (t, J = 2.0 Hz, 1H, H16), 7.10 (s, 2H, H11), 7.31 (d, J = 2.0 Hz, 2H, H14, H18), 8.01 (t, J = 7.2 Hz, 1H, H7), 8.06 (t, J = 7.0 Hz, 1H, H6), 8.29 (d, J = 9.6 Hz, 1H, H10), 8.32 (d, J = 8.4 Hz, 1H, H8), 8.48 (dd, J = 8.4; 6.0 Hz, 1H, H3), 8.66 (d, J = 9.6 Hz, 1H, H9), 9.18 (d, J = 8.4 Hz, 1H, H5), 9.49 (d, J = 6.0 Hz, 1H, H2), 10.28 (d, J = 8.8 Hz, 1H, H4). 13C NMR (100 MHz, DMSO-d6) δppm: 55.8 (2 × OMe), 63.9 (C11), 106.5 (C14, C18, C16), 116.2 (C10), 122.8 (C3), 124.2 (C5), 127.9 (C4a), 128.0 (C4b), 129.4 (C8), 130.1 (C6), 130.2 (C7), 130.8 (C8a), 135.4 (C13), 138.1 (C9), 140.0 (C10a), 142.5 (C4), 148.2 (C2), 160.8 (C15, C17), 190.5 (C12). Anal. Calcd. for C23H20BrNO3: C, 63.02; H, 4.60; N, 3.20. Found: C, 63.01; H, 4.59; N, 3.22.

4-(2-oxo-2-(3,4-Trimethoxyphenyl)ethyl)benzo[f]quinolin-4-ium Bromide 12c

Beige solid; 49% yield; mp 170–172 °C; IR ν(cm−1): 3069, 2974, 1684, 1591, 1508, 1420, 1344, 1260, 1150, 1013, 816, 764; 1H NMR (500 MHz, DMSO-d6) δppm: 3.87 (s, 3H, OMe), 3.94 (s, 3H, OMe), 7.09 (s, 2H, H11), 7.28 (d, J = 8.5 Hz, 1H, H17), 7.60 (s, 1H, H14), 7.91 (d, J = 8.0 Hz, 1H, H18), 7.99 (t, J = 7.0 Hz, 1H, H7), 8.05 (t, J = 7.0 Hz, 1H, H6), 8.26 (d, J = 9.5 Hz, 1H, H10), 8.30 (d, J = 7.5 Hz, 1H, H8), 8.47 (at, J = 7.0 Hz, 1H, H3), 8.65 (d, J = 9.5 Hz, 1H, H9), 9.18 (d, J = 8.0 Hz, 1H, H5), 9.50 (d, J = 5.0 Hz, 1H, H2), 10.27 (d, J = 8.5 Hz, 1H, H4). 13C NMR (125 MHz, DMSO-d6) δppm: 55.8 (OMe), 56.1 (OMe), 63.6 (C11), 110.6 (C14), 111.3 (C17), 116.3 (C10), 122.9 (C3), 123.9 (C18), 124.3 (C5), 126.3 (C13), 127.9 (C4a), 128.0 (C4b), 129.5 (C8), 130.1 (C6), 130.2 (C7), 130.9 (C8a), 138.1 (C9), 140.1 (C10a), 142.5 (C4), 148.3 (C2), 148.8 (C15), 154.5 (C16), 189.0 (C12). Anal. Calcd. for C23H20BrNO3: C, 63.02; H, 4.60; N, 3.20. Found: C, 63.03; H, 4.58; N, 3.21.

4-(2-(4-Bromophenyl)-2-oxoethyl)benzo[f]quinolin-4-ium Bromide 12d

Yellow solid; 50% yield; mp 246–248 °C; IR ν(cm−1): 3018, 2822, 1695, 1582, 1406, 1354, 1225, 991, 804, 754; 1H NMR (500 MHz, DMSO-d6) δppm: 7.05 (s, 2H, H11), 7.95 (d, J = 8.5 Hz, 2H, H15, H17), 8.00 (t, J = 7.0 Hz, 1H, H7), 8.06 (t, J = 7.0 Hz, 1H, H6), 8.09 (d, J = 8.5 Hz, 2H, H14, H18), 8.31 (d, J = 8.0 Hz, 1H, H8), 8.36 (d, J = 10.0 Hz, 1H, H10), 8.47 (dd, J = 8.5; 6.0 Hz, 1H, H3), 8.64 (d, J = 9.5 Hz, 1H, H9), 9.18 (d, J = 8.5 Hz, 1H, H5), 9.45 (d, J = 5.0 Hz, 1H, H2), 10.28 (d, J = 8.5 Hz, 1H, H4). 13C NMR (125 MHz, DMSO-d6) δppm: 63.8 (C11), 116.4 (C10), 122.9 (C3), 124.3 (C5), 127.9 (C4a), 128.0 (C4b), 129.0 (C16), 129.5 (C8), 130.1 (C6), 130.2 (C7), 130.6 (C14, C18), 130.9 (C8a), 132.2 (C15, C17), 132.7 (C13), 138.1 (C9), 140.2 (C10a), 142.6 (C4), 148.3 (C2), 190.2 (C12). Anal. Calcd. for C21H15BrNO: C, 66.86; H, 4.01; N, 3.71. Found: C, 66.85; H, 4.00; N, 3.73.

Ethyl 3-(3,4,5-Trimethoxybenzoyl)benzo[f]pyrrolo[1,2-a]quinoline-1-carboxylate 13a

Yellow solid; 80% yield; mp 239–240 °C; IR ν(cm−1): 1703, 1632, 1579, 1499, 1416, 1344, 1227, 1128, 1078, 744; 1H NMR (500 Mz, CDCl3) δppm: 1.43 (t, J = 7.0 Hz, 3H, CH2CH3), 3.95 (s, 6H, 2 × OMe), 4.00 (s, 3H, OMe), 4.42 (q, J = 7.0 Hz, 2H, CH2CH3), 7.41 (s, 2H, H15, H19), 7.65 (t, J = 8.0 Hz, 1H, H8), 7.75 (dd, J = 8.5; 7.0 Hz, 1H, H9), 7.78 (s, 1H, H2), 7.96–7.98 (overlapped signals, 3H, H7, H11, H12), 8.55 (d, J = 9.5 Hz, 1H, H6), 8.62 (d, J = 9.5 Hz, 1H, H5), 8.64 (d, J = 8.5 Hz, 1H, H10). 13C NMR (125 MHz, CDCl3) δppm: 14.7 (CH2CH3), 56.6 (2 × OMe), 60.4 (CH2CH3), 61.2 (OMe), 107.2 (C1), 107.7 (C15, C19), 117.9 (C6), 119.7 (C12), 121.0 (C10b), 123.0 (C5), 123.8 (C10), 126.8 (C8), 127.6 (C3), 127.9 (C9), 129.0 (C2), 129.7 (C7), 129.8 (C11), 130.0 (C10a), 130.9 (C6a), 132.1 (C14), 133.6 (C4a), 140.3 (C12a), 142.6 (C17), 153.2 (C16, C18), 164.3 (COO), 183.9 (C13). Anal. Calcd. for C29H25NO6: C, 72.04; H, 5.21; N, 2.90. Found: C, 72.02; H, 5.20; N, 2.92.

Ethyl 3-(3,5-Dimethoxybenzoyl)benzo[f]pyrrolo[1,2-a]quinoline-1-carboxylate 13b

Yellow solid; 70% yield; mp 246–247 °C; IR ν(cm−1): 1699, 1632, 1593, 1498, 1427, 1354, 1300, 1230, 1153, 1082, 744; 1H NMR (400 MHz, CDCl3) δppm: 1.42 (t, J = 7.2 Hz, 3H, CH2CH3), 3.88 (s, 6H, 2 × OMe), 4.40 (q, J = 7.2 Hz, 2H, CH2CH3), 6.76 (bs, 1H, H17), 7.26 (d, J = 2.0 Hz, 2H, H15, H19), 7.62 (t, J = 7.6 Hz, 1H, H8), 7.71 (t, J = 7.6 Hz, 1H, H9), 7.79 (s, 1H, H2), 7.93–7.98 (overlapped signals, 3H, H7, H11, H12), 8.51 (d, J = 9.6 Hz, 1H, H6), 8.56–8.60 (overlapped signals, 2H, H5, H10). 13C NMR (100 MHz, CDCl3) δppm: 14.6 (CH2CH3), 55.7 (2 × OMe), 60.2 (CH2CH3), 105.4 (C17), 107.1 (C1), 107.9 (C15, C19), 117.7 (C6), 119.7 (C12), 120.9 (C10b), 122.8 (C5), 123.7 (C10), 126.6 (C8), 127.6 (C3), 127.7 (C9), 128.8 (C7), 129.5 (C11), 129.8 (C10a), 130.2 (C2), 130.8 (C6a), 132.0 (C4a), 140.3 (C12a), 140.4 (C14), 160.8 (C16, C18), 164.1 (COO), 184.0 (C13). Anal. Calcd. for C28H25NO5: C, 74.16; H, 5.11; N, 3.09. Found: C, 74.15; H, 5.10; N, 3.11.

Ethyl 3-(3,4-Dimethoxybenzoyl)benzo[f]pyrrolo[1,2-a]quinoline-1-carboxylate 13c

Yellow solid; 70% yield; mp 244–245 °C; IR ν(cm−1): 1699, 1630, 1512, 1427, 1344, 1232, 1137, 1078, 1022, 806, 740; 1H NMR (500 Mz, CDCl3) δppm: 1.43 (t, J = 7.0 Hz, 3H, CH2CH3), 4.00 (s, 3H, OMe), 4.03 (s, 3H, OMe), 4.41 (q, J = 7.0 Hz, 2H, CH2CH3), 7.02 (d, J = 8.0 Hz, 1H, H19), 7.63 (t, J = 7.5 Hz, 1H, H8), 7.69 (d, J = 2.0 Hz, 2H, H15), 7.73 (t, J = 7.0 Hz, 1H, H9), 7.75 (s, 1H, H2), 6.76 (dd, J = 8.0; 2.0 Hz, 1H, H18), 7.95–7.97 (overlapped signals, 3H, H7, H11, H12), 8.54 (d, J = 9.5 Hz, 1H, H6), 8.59 (d, J = 9.5 Hz, 1H, H5), 8.63 (d, J = 8.5 Hz, 1H, H10). 13C NMR (125 MHz, CDCl3) δppm: 14.7 (CH2CH3), 56.3 (OMe), 56.4 (OMe), 60.3 (CH2CH3), 107.0 (C1), 110.2 (C19), 112.0 (C15), 117.9 (C6), 119.7 (C12), 120.9 (C10b), 123.0 (C10), 123.5 (C5), 125.4 (C18), 126.7 (C8), 127.7 (C3), 127.8 (C9), 129.0 (C7), 129.2 (C2), 129.7 (C11), 130.0 (C10a), 130.9 (C6a), 131.2 (C14), 132.0 (C4a), 140.0 (C12a), 149.4 (C16), 153.5 (C17), 164.4 (COO), 184.0 (C13). Anal. Calcd. for C28H25NO5: C, 74.16; H, 5.11; N, 3.09. Found: C, 74.14; H, 5.10; N, 3.10.

Ethyl 3-(4-Bromobenzoyl)benzo[f]pyrrolo[1,2-a]quinoline-1-carboxylate 13d

Yellow solid; 65% yield; mp 258–259 °C; IR ν(cm−1): 1708, 1636, 1502, 1425, 1339, 1231, 1078, 742; 1H NMR (500 Mz, CDCl3) δppm: 1.42 (t, J = 7.0 Hz, 3H, CH2CH3), 4.41 (q, J = 7.0 Hz, 2H, CH2CH3), 7.64 (t, J = 7.5 Hz, 1H, H8), 7.72–7.75 (overlapped signals, 4H, H15, H19, H9, H2), 7.96–7.97 (overlapped signals, 3H, H7, H11, H12), 8.00 (d, J = 8.5 Hz, 2H, H15, H19), 8.54 (d, J = 9.6 Hz, 1H, H6), 8.61–8.63 (overlapped signals, 2H, H5, H10). 13C NMR (125 MHz, CDCl3) δppm: 14.7 (CH2CH3), 60.4 (CH2CH3), 107.4 (C1), 117.8 (C6), 119.7 (C12), 121.1 (C10b), 123.0 (C5), 124.2 (C10), 126.9 (C8), 127.4 (C17), 127.9 (C9), 128.0 (C3), 129.0 (C7), 129.8 (C11), 129.9 (C10a), 130.5 (C2), 130.9 (C6a), 131.7 (C16, C18), 132.0 (C15, C19), 132.1 (C4a), 137.4 (C14), 140.6 (C12a), 164.1 (COO), 183.2 (C13). Anal. Calcd. for C26H18BrNO3: C, 66.11; H, 3.84; N, 2.97. Found: C, 66.10; H, 3.83; N, 2.96.

2-(2-oxo-2-(3,4,5-Trimethoxyphenyl)ethyl)isoquinolin-2-ium Bromide 15a

White solid; 49% yield; mp 270–271 °C; IR ν(cm−1): 3019, 1675, 1633, 1581, 1415, 1319, 1161, 1122; 1H NMR (500 MHz, DMSO-d6) δppm: 3.81 (s, 3H, OMe), 3.91 (s, 6H, 2 × OMe), 6.67 (s, 2H, H11), 7.42 (s, 2H, H14, H18), 8.13 (t, J = 7.0 Hz, 1H, H6), 8.34 (t, J = 7.5 Hz, 1H, H5), 8.43 (d, J = 8.5 Hz, 1H, H4), 8.56 (d, J = 8.5 Hz, 1H, H7), 8.70 (d, J = 6.5 Hz, 1H, H3), 8.74 (d, J = 7.0 Hz, 1H, H2), 10.03 (s, 1H, H8). 13C NMR (125 MHz, DMSO-d6) δppm: 56.4 (2 × OMe), 60.4 (OMe), 66.1 (C11), 106.1 (C14, C18), 125.6 (C3), 126.9 (C10), 127.5 (C4), 128.8 (C13), 130.7 (C7), 131.5 (C6), 136.3 (C2), 137.3 (C5), 137.5 (C9), 143.2 (C16), 151.7 (C8), 153.1 (C15, C17), 189.9 (C12). Anal. Calcd. for C20H20BrNO4: C, 57.43; H, 4.82; N, 3.35. Found: C, 57.42; H, 4.80; N, 3.36.

2-(2-oxo-2-(3,5-Trimethoxyphenyl)ethyl)isoquinolin-2-ium Bromide 15b

Beige solid; 61% yield; mp 265–268 °C; IR ν(cm−1): 3009, 2841, 1685, 1639, 1597, 1463, 1356, 1294, 1208, 1156, 1026; 1H NMR (500 MHz, DMSO-d6) δppm: 3.87 (s, 6H, 2 × OMe), 6.66 (s, 2H, H11), 6.95 (t, J = 2.0 Hz, 1H, H16),7.23 (d, J = 2.0 Hz, 2H, H14, H18), 8.12 (t, J = 8.0 Hz, 1H, H6), 8.34 (t, J = 7.5 Hz, 1H, H5), 8.42 (d, J = 8.0 Hz, 1H, H4), 8.55 (d, J = 8.0 Hz, 1H, H7), 8.70 (d, J = 7.0 Hz, 1H, H3), 8.75 (d, J = 6.5 Hz, 1H, H2), 10.05 (s, 1H, H8). 13C NMR (125 MHz, DMSO-d6) δppm: 55.8 (2 × OMe), 66.3 (C11), 106.2 (C14, C18), 106.3 (C16), 125.5 (C3), 126.9 (C10), 127.4 (C4), 130.6 (C7), 131.5 (C6), 135.5 (C13), 136.3 (C2), 137.3 (C5), 137.5 (C9), 151.7 (C8), 160.9 (C15, C17), 190.8 (C12). Anal. Calcd. for C19H8BrNO3: C, 58.78; H, 4.67; N, 3.61. Found: C, 58.76; H, 4.67; N, 3.62.

2-(2-oxo-2-(3,4-Trimethoxyphenyl)ethyl)isoquinolin-2-ium Bromide 15c

Beige solid; 43% yield; mp 259–261 °C; IR ν(cm−1): 3039, 3008, 2944, 1675, 1641, 1592, 1510, 1279, 1153, 1011; 1H NMR (500 MHz, DMSO-d6) δppm: 3.86 (s, 3H, OMe), 3.92 (s, 3H, OMe), 6.62 (s, 2H, H11), 7.25 (d, J = 8.5 Hz, 1H, H17), 7.55 (d, J = 1.5 Hz, 1H, H14), 7.82 (dd, J = 8.5; 1.5 Hz, 1H, H18), 8.12 (t, J = 8.0 Hz, 1H, H6), 8.33 (t, J = 7.5 Hz, 1H, H5), 8.42 (d, J = 8.5 Hz, 1H, H4), 8.55 (d, J = 8.5 Hz, 1H, H7), 8.68 (d, J = 7.0 Hz, 1H, H3), 8.75 (d, J = 7.0 Hz, 1H, H2), 10.03 (s, 1H, H8). 13C NMR (125 MHz, DMSO-d6) δppm: 55.8 (OMe), 56.0 (OMe), 65.8 (C11), 110.3 (C14), 111.3 (C17), 123.5 (C18), 125.5 (C3), 126.3 (C13), 126.9 (C10), 127.4 (C4), 130.6 (C7), 131.4 (C6), 136.4 (C2), 137.2 (C5), 137.5 (C9), 148.9 (C15), 151.7 (C8), 154.4 (C16), 189.2 (C12). Anal. Calcd. for C19H8BrNO3: C, 58.78; H, 4.67; N, 3.61. Found: C, 58.77; H, 4.65; N, 3.63.

2-(2-(4-Bromophenyl)-2-oxoethyl)isoquinolin-2-ium 15d

Physical and spectral data are similar to the ones reported in [6].

Ethyl 3-(3,4,5-Trimethoxybenzoyl)pyrrolo[2,1-a]isoquinoline-1-carboxylate 16a

White solid; 56% yield; mp 223–225 °C; IR ν(cm−1): 2945, 1707, 1631, 1579, 1456, 1370, 1187, 1134, 761; 1H NMR (500 Mz, CDCl3) δppm: 1.41 (t, J = 7.0 Hz, 3H, CH2CH3), 3.93 (s, 6H, 2 × OMe), 3.97 (s, 3H, OMe), 4.41 (q, J = 7.0 Hz, 2H, CH2CH3), 7.15 (s, 2H, H15, H19), 7.28 (d, J = 7.5 Hz, 1H, H8), 7.66–7.68 (overlapped signals, 2H, H7, H6), 7.77 (m, 1H, H5), 7.90 (s, 1H, H2), 9.57 (d, J = 7.5 Hz, 1H, H9), 9.84 (dd, J = 7.0; 2.0 Hz, 1H, H4). 13C NMR (125 MHz, CDCl3) δppm: 14.6 (CH2CH3), 56.5 (2 × OMe), 60.8 (CH2CH3), 61.2 (OMe), 107.1 (C15, C19), 110.4 (C3), 115.7 (C8), 123.3 (C1), 124.8 (C11), 125.1 (C9), 126.9 (C5), 128.0 (C6), 128.3 (C4), 129.5 (C7), 129.9 (C2), 130.7 (C12), 135.0 (C14), 137.1 (C10), 141.6 (C17), 153.1 (C16, C18), 164.8 (C20), 185.1 (C13). Anal. Calcd. for C25H23NO6: C, 69.27; H, 5.35; N, 3.23. Found: C, 69.28; H, 5.33; N, 3.25.

Ethyl 3-(3,5-Trimethoxybenzoyl)pyrrolo[2,1-a]isoquinoline-1-carboxylate 16b

Yellow solid; 63% yield; mp 227–229 °C; IR ν(cm−1): 2974, 1717, 1631, 1596, 1458, 1370, 1189, 754; 1H NMR (500 Mz, CDCl3) δppm: 1.41 (t, J = 7.0 Hz, 3H, CH2CH3), 3.87 (s, 6H, 2 × OMe), 4.41 (q, J = 7.0 Hz, 2H, CH2CH3), 6.70 (t, J = 2.5 Hz, 1H, H17), 6.99 (d, 2H, J = 2.5 Hz, H15, H19), 7.29 (d, J = 7.5 Hz, 1H, H8), 7.66–7.68 (overlapped signals, 2H, H7, H6), 7.78 (m, 1H, H5), 7.88 (s, 1H, H2), 9.64 (d, J = 7.5 Hz, 1H, H9), 9.83 (dd, J = 7.0; 2.5 Hz, 1H, H4). 13C NMR (125 MHz, CDCl3) δppm: 14.6 (CH2CH3), 55.8 (2 × OMe), 60.8 (CH2CH3), 104.3 (C17), 107.3 (C15, C19), 110.5 (C3), 115.7 (C8), 123.4 (C1), 124.8 (C11), 125.3 (C9), 126.9 (C5), 128.0 (C4), 128.3 (C6), 129.5 (C7), 130.2 (C2), 130.7 (C12), 137.2 (C10), 141.9 (C14), 160.8 (C16, C18), 164.8 (C20), 185.7 (C13). Anal. Calcd. for C24H21NO5: C, 71.45; H, 5.25; N, 3.47. Found: C, 71.46; H, 5.26; N, 3.48.

Ethyl 3-(3,4-Trimethoxybenzoyl)pyrrolo[2,1-a]isoquinoline-1-carboxylate 16c

Yellow solid; 41% yield; mp 219–220 °C; IR ν(cm−1): 2979, 1698, 1629, 1598, 1503, 1368, 1190, 750; 1H NMR (500 Mz, CDCl3) δppm: 1.41 (t, J = 7.0 Hz, 3H, CH2CH3), 3.98 (s, 3H, OMe), 4.00 (s, 3H, OMe), 4.41 (q, J = 7.0 Hz, 2H, CH2CH3), 6.99 (d, J = 8.0 Hz, 1H, H18), 7.25 (d, J = 7.5 Hz, 1H, H8), 7.49 (d, J = 2.0 Hz, 1H, H15), 7.54 (dd, J = 8.5; 2.0 Hz, 1H, H19), 7.63–7.68 (overlapped signals, 2H, H7, H6), 7.76 (dd, J = 8.5; 2.0 Hz, 1H, H5), 7.85 (s, 1H, H2), 9.53 (d, J = 7.5 Hz, 1H, H9), 9.85 (dd, J = 7.5; 1.5 Hz, 1H, H4). 13C NMR (125 MHz, CDCl3) δppm: 14.6 (CH2CH3), 56.2 (OMe), 56.3 (OMe), 60.7 (CH2CH3), 110.1 (C3), 110.2 (C18), 112.0 (C15), 115.5 (C8), 123.7 (C1), 124.0 (C19), 124.9 (C11), 125.2 (C9), 126.9 (C5), 127.9 (C6), 128.2 (C4), 129.3 (C7), 129.4 (C2), 130.6 (C12), 132.5 (C14), 136.9 (C10), 149.1 (C16), 152.7 (C17), 164.9 (C20), 184.9 (C13). Anal. Calcd. for C24H21NO5: C, 71.45; H, 5.25; N, 3.47. Found: C, 71.43; H, 5.24; N, 3.49.

Ethyl 3-(4-Bromobenzoyl)pyrrolo[2,1-a]isoquinoline-1-carboxylate 16d

Yellow solid; 61% yield; 1H NMR (500 Mz, CDCl3) δppm: 1.41 (t, J = 7.0 Hz, 3H, CH2CH3), 4.41 (q, J = 7.0 Hz, 2H, CH2CH3), 7.28 (d, J = 7.5 Hz, 1H, H8), 7.66–7.69 (overlapped signals, 4H, H16, H18, H7, H6), 7.73–7.76 (overlapped signals, 3H, H5, H15, H19), 7.77 (s, 1H, H2), 9.61 (d, J = 8.5 Hz, 1H, H9), 9.83 (dd, J = 7.0; 2.5 Hz, 1H, H4). 13C NMR (125 MHz, CDCl3) δppm: 14.6 (CH2CH3), 60.9 (CH2CH3), 110.7 (C3), 115.9 (C8), 123.1 (C1), 124.7 (C11), 125.1 (C9), 126.8 (C17), 126.9 (C5), 128.0 (C4), 128.3 (C6), 129.6 (C7), 130.0 (C2), 130.7 (C12), 130.9 (C15, C19), 131.8 (C16, C18), 137.3 (C10), 138.8 (C14), 164.6 (C20), 184.8 (C13). Anal. Calcd. for C22H16BrNO3: C, 62.57; H, 3.82; N, 3.32. Found: C, 62.55; H, 3.81; N, 3.34.

3.2. Biological Activity

3.2.1. Anticancer Activity

Compounds were tested against a panel of 60 human cancer cell lines at the National Cancer Institute, Rockville, MD. Cytotoxicity experiments were performed using a 48 h exposure protocol consisting of a sulforhodamine B assay [41,42,43], described in detail in our previous work [6].

3.2.2. Tubulin Polymerization Assay

A tubulin polymerization assay kit (Cytoskeleton Inc. Denver, CO, USA, Cat. # BK006P) was used to study microtubule assembly according to the manufacturer’s instructions [75,76]. Tubulin polymerization was monitored using a FLUOstar Omega multi-mode microplate reader (BMG LABTECH). The final buffer for tubulin polymerization contained 80 mM PIPES (piperazine-N,N’-bis(2-ethanesulfonic acid)sesquisodium salt) pH = 6.9, 2 mM MgCl2, 0.5 mM EGTA (ethylene glycol-bis(β-aminoethyl ether)-N,N,N’,N’-tetraacetic acid), 1 mM GTP, and 10.2% glycerol. Test compounds were added in one single concentration (10 μM) and all compounds except purified tubulin were heated to 37 °C. The reaction was initiated by the addition of tubulin to a final concentration of 3.0 mg/mL. Paclitaxel and phenstatin were used as positive controls under the same conditions. Absorbance was measured at 340 nm for 1 h at 1 min intervals at 37 °C.

3.3. Molecular Modeling

3.3.1. Molecular Docking

Flexible-ligand docking experiments were performed in two steps. First, blind docking was performed using Autodock 4.2 [77] using a 90 × 118 × 114 gridbox with 0.825 Å spacing (total search volume of 998,811 Å3, as to include the entire heterodimer, PDB ID: 4O2B [63]) and was centered on the entire structure (x = 18.997, y = 68.705, z = 46.491). Co-crystallized GTP, GDP, Ca2+, and Mg2+ ions were kept during receptor preparation, and the target protein was kept rigid during all docking experiments. The 3D structures of colchicine, phenstatin, and compound 10a were constructed in Avogadro v1.2.0 [78] and were energetically optimized in the MMFF94 force field until a local energy minimum was reached. A total of 8 jobs of 2000 runs were performed for each of the 3 ligands using the Lamarckian Genetic Algorithm (LGA) for a total of 16,000 runs per ligand. Conformations having theoretical binding energies lower than −5.5 kcal/mol were clustered and ranked using an RMSD tolerance of 2.0 Å in order to select the lowest-binding representatives for further binding pose refinement. Cluster representatives with binding energies lower than −7.0 kcal/mol were visually inspected for binding orientation.
A smaller gridbox centered on the lowest binding cluster of compound 10a from blind docking (58 × 58 × 58 points, 0.375 Å spacing, x = 14.316, y = 67.184, z = 44.464) was chosen for local docking. In this case, 1,000 poses were generated per ligand in a single run, and results were clustered and ranked based on theoretical binding energy. RMSD between re-docked and co-crystallized colchicine ligand was used as quality control for the docking protocol. Visual inspection and molecular graphics were made in the PyMOL Molecular Graphics System, Version 2.5.0. (Schrödinger, LLC, New York, NY, USA). Furthermore, 2D ligand interaction diagrams were generated in Maestro, release 2018-4 (Schrödinger, LLC, New York, NY, USA).

3.3.2. Molecular Dynamics Simulations

Each of the three lowest-scoring cluster representatives of compound 10a obtained from local docking experiments was subjected to MD simulations in the colchicine binding site of the α,β-tubulin heterodimer (chains A and B of PDB ID: 4O2B). The BioLuminate® (Schrödinger, Inc., New York, NY, USA) graphical interface [79] was used for initial system construction, while the subsequent MD simulations were performed with the Desmond (D.E. Shaw Research) software [80]. The missing loop of tubulin chain B (residues 276–281) was modeled using MODELLER [81], as implemented in UCSF Chimera release 1.15 [82]. Proteins, ligands, and ions were described using the OPLS-AA force field [83,84]. The 10a/tubulin complexes were solvated using a TIP3P water model [85] in orthorhombic simulation boxes while allowing for a minimum distance of 10 Å between the complex and the walls of the simulation box. Ions were added to electrostatically neutralize the systems and to attain a salt concentration of 0.15 M NaCl. After initial geometry optimization, the standard Desmond solute relaxation protocol was applied. This protocol involves sequential short MD simulations using restraints of decreasing magnitude on atomic positions, followed by short restraint-free equilibration simulations. In the production runs, each complex was simulated for 10 ns, and system configurations were saved every 5 ps for analysis. Throughout the simulations, the NPT ensemble (i.e., constant number of particles, pressure, and temperature) was employed, using the Nosé–Hoover chain thermostat (300 K) [86] and Martyna–Tobias–Klein barostat (1 atm) [87]. The MD trajectories were analyzed with respect to protein backbone and ligand RMSD, as well as protein-ligand contacts. Protein-ligand interaction fraction plots were generated in R using the ggplot2 package [88]. Configurational entropy variations were evaluated from 25 ns long MD simulations of the free compound in solution and bound to tubulin, respectively. Schlitter’s formula was used for the configurational entropy calculation [89]. Prior to entropy calculations using the GROMACS 2020.3 modelling suite [90], all the structures in the MD trajectories were geometrically fitted to the first frame in order to remove the rotational and translational degrees of freedom. The fitting procedure included all heavy atoms of compound 10a. To evaluate the convergence of the entropy profile, the calculations included steeply increasing time intervals along the entire MD trajectories.

3.3.3. In Silico ADME and Toxicity Predictions

The ADME in silico evaluation for the most active compound 10a was performed using the Swiss ADME web tool (http://swissadme.ch/index.php (accessed on 29 March 2023)) in terms of molecular properties, pharmacokinetics, drug-likeness, and medicinal chemistry.
The in silico toxicological evaluation for the most active compound 10a was performed using the web-service Cell-Line Cytotoxicity Predictor (CLC-Pred), which screens for in silico cytotoxicity on a panel of 278 tumor cells and 27 normal human cell lines from different tissues [91].

4. Conclusions

Five series of potential microtubule-targeting anticancer pyrrolo-fused heterocycles were designed as analogs of phenstatin and synthesized through 1,3-dipolar cycloaddition. Their antiproliferative activity was tested against NCI’s panel of 60 cancer cell lines. Pyrrolo[1,2-a]quinoline 10a proved to be the best in terms of growth inhibition properties, with almost all GI50 values in the submicromolar range. The best anticancer behavior was displayed on the A498 renal cell line (GI50 = 27 nM), melanoma MDA-MB-435 cell line (GI50 = 35 nM), and non-small cell lung cancer NCI-H522 line (GI50 = 50 nM). In vitro, tubulin polymerization inhibitory properties were found for compound 10a, while PRISM analysis indicated a strong profile correlation to other microtubule inhibitors (even if with a totally different structure). QSAR simulation of ADMET properties showed a promising drug-likeness profile for compound 10a with very good selectivity for tumor cells, good physicochemical properties, good GI absorption, moderate solubility, and adequate bioavailability. The binding mode of 10a was investigated through molecular docking and molecular dynamics simulations, as well as configurational entropy calculations. This compound was confirmed to bind to the colchicine site of tubulin through global and local docking experiments, but molecular dynamics simulations revealed that some of the predicted interactions in the docking experiments were not stable. As such, the lowest-scoring conformation from local docking, BM I, drifted away from its original orientation during MD simulations, while BM III, which was also identified through global docking experiments, was the most stable throughout the MD simulation. The second-lowest scoring conformation, BM II, which was the most similar in terms of relative orientation and hydrophobic amino acid interactions to other co-crystallized trimethoxy-substituted ring-containing tubulin binders including colchicine, was also stable throughout the MD simulation. However, longer MD simulations are necessary in order to demonstrate the full stability of the investigated systems. Configurational entropy calculations indicated similar values for configurational entropy loss upon binding for all three docked solutions, suggesting that all three binding modes are equally favorable in terms of entropic contribution. Taken together, our results suggest that for compound 10a, docking experiments alone are not sufficient for the adequate description of molecular interaction details in terms of target binding, which makes subsequent scaffold optimization more difficult to achieve. This highlights that a thorough in silico investigation should be performed for novel pyrrole-fused agents in order to have more meaningful insights into the molecular details of anticancer activity. Hopefully, our efforts will guide the discovery of novel potent antiproliferative compounds with pyrrolo-fused heterocyclic cores, especially from an in silico perspective.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ph16060865/s1; Figure S1: HRMS spectrum of compound 10a; Figures S2–S19: NMR-spectra of active compounds; Figures S20–S45: NCI data for tested compounds; Figure S46: conformation distribution after blind docking of compounds (a) colchicine, (b) phenstatin, and (c) 10a; Figure S47: binding score distribution and RMSD from the lowest-scoring solution from docking experiments. Figure S48: local docking to α,β-tubulin and superimposition with colchicine for (a,d) BM I, (b,e) BM II, and (c,f) BM III; Figure S49: superimposition of BM II of 10a from local docking and D64131 (PDB ID 6K9V); Figure S50: superimposition of BM I of 10a from local docking and the last frame of the MD simulation; Figure S51: ligand root mean square fluctuation (RMSF) throughout the simulations for (A) 10a and (B) colchicine; Figure S52: timeline representation of the number of H-bond interactions throughout the MD simulations; Figure S53: timeline representation of the number of polar interactions throughout the MD; Figure S54. timeline representation of the number of all interactions throughout the MD simulations; Figure S55. configurational entropy for free 10a, BM I, BM II, and BM III when bound to tubulin, and the difference between bound and free state.

Author Contributions

Conceptualization, R.D. and M.-C.A.-M.; writing—original draft preparation, R.D., M.-C.A.-M. and R.-M.A.; visualization, R.-M.A.; funding acquisition, I.I.M. and R.D. Biological data analysis was performed by M.-C.A.-M., R.D., I.I.M. and R.-M.A. Synthesis and structure elucidation were performed by L.P., M.-C.A.-M., C.I.C. and R.D. Molecular modeling experiments were performed by R.-M.A. and A.N. Writing-review and editing was performed by R.-M.A., M.-C.A.-M. and R.D. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by a grant from the Romanian Ministry of Education and Research, CNCS-UEFISCDI, project number PN-III-P4-ID-PCE-2020-0371, within PNCDI III.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article and Supplementary Material.

Acknowledgments

The authors are thankful to the Romanian Ministry of Research, Innovation and Digitization, within Program 1—Development of the national RD system, Subprogram 1.2—Institutional Performance—RDI excellence funding projects, contract no.11PFE/30.12.2021, for financial support. The authors also thank the National Cancer Institute, US, for the anticancer evaluation of the compounds on their 60-cell panel (the testing was performed by the Developmental Therapeutics Program, Division of Cancer Treatment and Diagnosis) and the CERNESIM Research Centre from “Al. I. Cuza” University of Iași for NMR infrastructure.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Petri, G.L.; Spano, V.; Spatola, R.; Holl, R.; Raimondi, M.V.; Barraja, P.; Montalbano, A. Bioactive pyrrole-based compounds with target selectivity. Eur. J. Med. Chem. 2020, 208, 112783. [Google Scholar] [CrossRef]
  2. Mao, Y.; Soni, K.; Sangani, C.; Yao, Y. An Overview of Privileged Scaffold: Quinolines and Isoquinolines in Medicinal Chemistry as Anticancer Agents. Curr. Top. Med. Chem. 2020, 20, 2599–2633. [Google Scholar] [CrossRef] [PubMed]
  3. Lv, H.N.; Zeng, K.W.; Zhao, M.B.; Jiang, Y.; Tu, P.F. Pyrrolo[2,1-a]isoquinoline and pyrrole alkaloids from Sinomenium acutum. J. Asian Nat. Prod. Res. 2017, 20, 195–200. [Google Scholar] [CrossRef]
  4. Allin, S.M.; Gaskell, S.N.; Towler, J.M.R.; Bulman Page, P.C.; Saha, B.; McKenzie, M.J.; Martin, W.P. A New Asymmetric Synthesis of the Anti-Tumor Alkaloid (R)-(+)-Crispine A. J. Org. Chem. 2007, 72, 8972–8975. [Google Scholar] [CrossRef] [PubMed]
  5. Yaremchuk, I.O.; Muzychka, L.V.; Smolii, O.B.; Kucher, O.V.; Shishkina, S.V. Synthesis of novel 1,2-dihydropyrrolo[1,2-a]pyrazin-1(2H)-one derivatives. Tetrahedron Lett. 2017, 59, 442–444. [Google Scholar] [CrossRef]
  6. Al-Matarneh, M.C.; Amarandi, R.-M.; Mangalagiu, I.I.; Danac, R. Synthesis and Biological Screening of New Cyano-Substituted Pyrrole Fused (Iso)Quinoline Derivatives. Molecules 2021, 26, 2066. [Google Scholar] [CrossRef]
  7. Amariucai-Mantu, D.; Antoci, V.; Sardaru, M.C.; Matarneh, C.M.A.; Mangalagiu, I.; Danac, R. Fused pyrrolo-pyridines and pyrrolo-(iso)quinoline as anticancer agents. Phys. Sci. Rev. 2022. [Google Scholar] [CrossRef]
  8. Sardaru, M.-C.; Carp, O.; Ursu, E.-L.; Craciun, A.-M.; Cojocaru, C.; Silion, M.; Kovalska, V.; Mangalagiu, I.; Danac, R.; Rotaru, A. Cyclodextrin Encapsulated pH Sensitive Dyes as Fluorescent Cellular Probes: Self-Aggregation and In Vitro Assessments. Molecules 2020, 25, 4397. [Google Scholar] [CrossRef] [PubMed]
  9. Al Matarneh, C.M.; Mangalagiu, I.I.; Shova, S.; Danac, R. Synthesis, structure, antimycobacterial and anticancer evaluation of new pyrrolo-phenanthroline derivatives. J. Enzym. Inhib. Med. Chem. 2016, 31, 470–480. [Google Scholar] [CrossRef] [Green Version]
  10. Al-Matarneh, C.M.; Amarandi, R.-M.; Craciun, A.M.; Mangalagiu, I.I.; Zbancioc, G.; Danac, R. Design, Synthesis, Molecular Modelling and Anticancer Activities of New Fused Phenanthrolines. Molecules 2020, 25, 527. [Google Scholar] [CrossRef] [Green Version]
  11. Sardaru, M.C.; Craciun, A.M.; Al Matarneh, C.M.; Sandu, I.A.; Amarandi, R.M.; Popovici, L.; Ciobanu, C.I.; Peptanariu, D.; Pinteala, M.; Mangalagiu, I.I.; et al. Cytotoxic substituted indolizines as new colchicine site tubulin polymerisation inhibitors. J. Enzym. Inhib. Med. Chem. 2020, 35, 1581–1595. [Google Scholar] [CrossRef] [PubMed]
  12. Leontie, L.; Danac, R.; Apetroaei, N.; Rusu, G.I. Study of electronic transport properties of some new N-(p-R-phenacyl)-1,7-phenanthrolinium bromides in thin films. Mater. Chem. Phys. 2011, 127, 471–478. [Google Scholar] [CrossRef]
  13. Marangoci, N.L.; Popovici, L.; Ursu, E.L.; Danac, R.; Clima, L.; Cojocaru, C.; Coroaba, A.; Neamtu, A.; Mangalagiu, I.I.; Pinteala, M.; et al. Pyridyl-indolizine derivatives as DNA binders and pH-sensible fluorescent dyes. Tetrahedron 2016, 72, 8215–8222. [Google Scholar] [CrossRef]
  14. Popovici, L.; Amarandi, R.M.; Mangalagiu, I.I.; Mangalagiu, V.; Danac, R. Synthesis, molecular modelling and anticancer evaluation of new pyrrolo[1,2-b]pyridazine and pyrrolo[2,1-a]phthalazine derivatives. J. Enzym. Inhib. Med. Chem. 2019, 34, 230–243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Lu, Y.; Chen, J.; Xiao, M.; Li, W.; Miller, D.D. An overview of tubulin inhibitors that interact with the colchicine binding site. Pharm. Res. 2012, 29, 2943–2971. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Kulshrestha, A.; Katara, G.K.; Ibrahim, S.A.; Patil, R.; Patil, S.A.; Beaman, K.D. Microtubule inhibitor, SP-6-27 inhibits angiogenesis and induces apoptosis in ovarian cancer cells. Oncotarget 2017, 8, 67017–67028. [Google Scholar] [CrossRef] [Green Version]
  17. Ravelli, R.B.; Gigant, B.; Curmi, P.A.; Jourdain, I.; Lachkar, S.; Sobel, A.; Knossow, M. Insight into tubulin regulation from a complex with colchicine and a stathmin-like domain. Nature 2004, 428, 198–202. [Google Scholar] [CrossRef]
  18. Wang, J.; Miller, D.D.; Li, W. Molecular interactions at the colchicine binding site in tubulin: An X-ray crystallography perspective. Drug Discov. Today 2022, 27, 759–776. [Google Scholar] [CrossRef]
  19. Pallante, L.; Rocca, A.; Klejborowska, G.; Huczynski, A.; Grasso, G.; Tuszynski, J.A.; Deriu, M.A. In silico Investigations of the Mode of Action of Novel Colchicine Derivatives Targeting β-Tubulin Isotypes: A Search for a Selective and Specific β-III Tubulin Ligand. Front. Chem. 2020, 8, 108. [Google Scholar] [CrossRef]
  20. Bueno, O.; Estevez Gallego, J.; Martins, S.; Prota, A.E.; Gago, F.; Gomez-SanJuan, A.; Camarasa, M.J.; Barasoain, I.; Steinmetz, M.O.; Diaz, J.F.; et al. High-affinity ligands of the colchicine domain in tubulin based on a structure-guided design. Sci. Rep. 2018, 8, 4242. [Google Scholar] [CrossRef] [Green Version]
  21. Wang, Y.; Zhang, H.; Gigant, B.; Yu, Y.; Wu, Y.; Chen, X.; Lai, Q.; Yang, Z.; Chen, Q.; Yang, J. Structures of a diverse set of colchicine binding site inhibitors in complex with tubulin provide a rationale for drug discovery. FEBS J. 2016, 283, 102–111. [Google Scholar] [CrossRef] [Green Version]
  22. Wang, Z.; Sun, H.; Yao, X.; Li, D.; Xu, L.; Li, Y.; Tian, S.; Hou, T. Comprehensive evaluation of ten docking programs on a diverse set of protein-ligand complexes: The prediction accuracy of sampling power and scoring power. Phys. Chem. Chem. Phys. 2016, 18, 12964–12975. [Google Scholar] [CrossRef] [PubMed]
  23. Kolb, P.; Irwin, J.J. Docking screens: Right for the right reasons? Curr. Top. Med. Chem. 2009, 9, 755–770. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Banerjee, S.; Arnst, K.E.; Wang, Y.; Kumar, G.; Deng, S.; Yang, L.; Li, G.B.; Yang, J.; White, S.W.; Li, W.; et al. Heterocyclic-Fused Pyrimidines as Novel Tubulin Polymerization Inhibitors Targeting the Colchicine Binding Site: Structural Basis and Antitumor Efficacy. J. Med. Chem. 2018, 61, 1704–1718. [Google Scholar] [CrossRef] [PubMed]
  25. La Regina, G.; Bai, R.; Coluccia, A.; Famiglini, V.; Pelliccia, S.; Passacantilli, S.; Mazzoccoli, C.; Ruggieri, V.; Sisinni, L.; Bolognesi, A.; et al. New pyrrole derivatives with potent tubulin polymerization inhibiting activity as anticancer agents including hedgehog-dependent cancer. J. Med. Chem. 2014, 57, 6531–6552. [Google Scholar] [CrossRef]
  26. Boichuk, S.; Galembikova, A.; Syuzov, K.; Dunaev, P.; Bikinieva, F.; Aukhadieva, A.; Zykova, S.; Igidov, N.; Gankova, K.; Novikova, M.; et al. The Design, Synthesis, and Biological Activities of Pyrrole-Based Carboxamides: The Novel Tubulin Inhibitors Targeting the Colchicine-Binding Site. Molecules 2021, 26, 5780. [Google Scholar] [CrossRef]
  27. Chen, Y.C. Beware of docking! Trends Pharmacol. Sci. 2015, 36, 78–95. [Google Scholar] [CrossRef]
  28. Ferreira, L.G.; Dos Santos, R.N.; Oliva, G.; Andricopulo, A.D. Molecular docking and structure-based drug design strategies. Molecules 2015, 20, 13384–13421. [Google Scholar] [CrossRef] [Green Version]
  29. Alonso, H.; Bliznyuk, A.A.; Gready, J.E. Combining docking and molecular dynamic simulations in drug design. Med. Res. Rev. 2006, 26, 531–568. [Google Scholar] [CrossRef]
  30. Sakano, T.; Mahamood, M.I.; Yamashita, T.; Fujitani, H. Molecular dynamics analysis to evaluate docking pose prediction. Biophys. Physicobiol. 2016, 13, 181–194. [Google Scholar] [CrossRef] [Green Version]
  31. Wang, T.C.; Cheng, L.P.; Huang, X.Y.; Zhao, L.; Pang, W. Identification of potential tubulin polymerization inhibitors by 3D-QSAR, molecular docking and molecular dynamics. RSC Adv. 2017, 7, 38479–38489. [Google Scholar] [CrossRef] [Green Version]
  32. Rao, C.; Naidu, N.; Priya, J.; Rao, K.P.C.; Ranjith, K.; Shobha, S.; Chowdary, B.S.; Siddiraju, S.; Yadam, S. Molecular docking and dynamic simulations of benzimidazoles with beta-tubulins. Bioinformation 2021, 17, 404–412. [Google Scholar] [PubMed]
  33. Kumbhar, B.V.; Borogaon, A.; Panda, D.; Kunwar, A. Exploring the Origin of Differential Binding Affinities of Human Tubulin Isotypes αβII, αβIII and αβIV for DAMA-Colchicine Using Homology Modelling, Molecular Docking and Molecular Dynamics Simulations. PLoS ONE 2016, 11, e0156048. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Ghinet, A.; Abuhaie, C.M.; Gautret, P.; Rigo, B.; Dubois, J.; Farce, A.; Belei, D.; Bicu, E. Studies on indolizines. Evaluation of their biological properties as microtubule-interacting agents and as melanoma targeting compounds. Eur. J. Med. Chem. 2015, 89, 115–127. [Google Scholar] [CrossRef] [PubMed]
  35. Moise, I.-M.; Bicu, E.; Farce, A.; Dubois, J.; Ghinet, A. Indolizine-phenothiazine hybrids as the first dual inhibitors of tubulin polymerization and farnesyltransferase with synergistic antitumor activity. Bioorg. Chem. 2020, 103, 104184. [Google Scholar] [CrossRef]
  36. Mangalagiu, G.; Ungureanu, M.; Grosu, G.; Mangalagiu, I.; Petrovanu, M. New pyrrolo-pyrimidine derivatives with antifungal or antibacterial properties in vitro. Ann. Pharm. Fr. 2001, 59, 139–140. [Google Scholar]
  37. Bahner, C.T.; Norton, L.L. Some Quaternary Salts of Pyrazine. J. Am. Chem. Soc. 1950, 72, 2881–2882. [Google Scholar] [CrossRef]
  38. Antoci, V.; Oniciuc, L.; Amariucai-Mantu, D.; Moldoveanu, C.; Mangalagiu, V.; Amarandei, A.M.; Lungu, C.N.; Dunca, S.; Mangalagiu, I.I.; Zbancioc, G. Benzoquinoline Derivatives: A Straightforward and Efficient Route to Antibacterial and Antifungal Agents. Pharmaceuticals 2021, 14, 335. [Google Scholar] [CrossRef]
  39. Lucescu, L.; Bicu, E.; Belei, D.; Dubois, J.; Ghinet, A. Synthesis and Biological Evaluation of Some New Indolizine Derivatives as Antitumoral Agents. Lett. Drug Des. Discov. 2016, 13, 479–488. [Google Scholar] [CrossRef]
  40. Georgescu, E.; Draghici, C.; Iuhas, P.C.; Georgescu, F. A new approach for the synthesis of benzo[f]pyrrolo[1,2-a]-quinolines. Arkivoc 2005, 10, 95–104. [Google Scholar] [CrossRef] [Green Version]
  41. Shoemaker, R.H. The NCI60 human tumour cell line anticancer drug screen. Nat. Rev. Cancer 2006, 6, 813–823. [Google Scholar] [CrossRef] [PubMed]
  42. Skehan, P.; Storeng, R.; Scudiero, D.; Monks, A.; McMahon, J.; Vistica, D.; Warren, J.T.; Bokesch, H.; Kenney, S.; Boyd, M.R. New colorimetric cytotoxicity assay for anticancer-drug screening. J. Natl. Cancer Inst. 1990, 82, 1107–1112. [Google Scholar] [CrossRef]
  43. Boyd, M.R. The NCI Human Tumor Cell Line (60-Cell) Screen. In Anticancer Drug Development Guide: Preclinical Screening, Clinical Trials, and Approval; Teicher, B.A., Andrews, P.A., Eds.; Humana Press: Totowa, NJ, USA, 2004; pp. 41–61. [Google Scholar]
  44. Sharma, A.; Tarbox, L.; Kurc, T.; Bona, J.; Smith, K.; Kathiravelu, P.; Bremer, E.; Saltz, J.H.; Prior, F. PRISM: A Platform for Imaging in Precision Medicine. JCO Clin. Cancer Inform. 2020, 4, 491–499. [Google Scholar] [CrossRef]
  45. Sisco, E.; Barnes, K.L. Design, Synthesis, and Biological Evaluation of Novel 1,3-Oxazole Sulfonamides as Tubulin Polymerization Inhibitors. ACS Med. Chem. Lett. 2021, 12, 1030–1037. [Google Scholar] [CrossRef]
  46. Kamal, A.; Kumar, G.B.; Vishnuvardhan, M.V.; Shaik, A.B.; Reddy, V.S.; Mahesh, R.; Sayeeda, I.B.; Kapure, J.S. Synthesis of phenstatin/isocombretastatin-chalcone conjugates as potent tubulin polymerization inhibitors and mitochondrial apoptotic inducers. Org. Biomol. Chem. 2015, 13, 3963–3981. [Google Scholar] [CrossRef]
  47. Dohle, W.; Jourdan, F.L.; Menchon, G.; Prota, A.E.; Foster, P.A.; Mannion, P.; Hamel, E.; Thomas, M.P.; Kasprzyk, P.G.; Ferrandis, E.; et al. Quinazolinone-Based Anticancer Agents: Synthesis, Antiproliferative SAR, Antitubulin Activity, and Tubulin Co-crystal Structure. J. Med. Chem. 2018, 61, 1031–1044. [Google Scholar] [CrossRef] [PubMed]
  48. Dohle, W.; Prota, A.E.; Menchon, G.; Hamel, E.; Steinmetz, M.O.; Potter, B.V.L. Tetrahydroisoquinoline Sulfamates as Potent Microtubule Disruptors: Synthesis, Antiproliferative and Antitubulin Activity of Dichlorobenzyl-Based Derivatives, and a Tubulin Cocrystal Structure. ACS Omega 2019, 4, 755–764. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Steinmetz, M.O.; Prota, A.E. Microtubule-Targeting Agents: Strategies to Hijack the Cytoskeleton. Trends Cell. Biol. 2018, 28, 776–792. [Google Scholar] [CrossRef]
  50. Muhlethaler, T.; Gioia, D.; Prota, A.E.; Sharpe, M.E.; Cavalli, A.; Steinmetz, M.O. Comprehensive Analysis of Binding Sites in Tubulin. Angew. Chem. Int. Ed. Engl. 2021, 60, 13331–13342. [Google Scholar] [CrossRef]
  51. Prota, A.E.; Bargsten, K.; Diaz, J.F.; Marsh, M.; Cuevas, C.; Liniger, M.; Neuhaus, C.; Andreu, J.M.; Altmann, K.H.; Steinmetz, M.O. A new tubulin-binding site and pharmacophore for microtubule-destabilizing anticancer drugs. Proc. Natl. Acad. Sci. USA 2014, 111, 13817–13821. [Google Scholar] [CrossRef] [Green Version]
  52. Zhu, L.; Zhang, C.; Lu, X.; Song, C.; Wang, C.; Zhang, M.; Xie, Y.; Schaefer, H.F., 3rd. Binding modes of cabazitaxel with the different human beta-tubulin isotypes: DFT and MD studies. J. Mol. Model. 2020, 26, 162. [Google Scholar] [CrossRef] [PubMed]
  53. Trott, O.; Olson, A.J. AutoDock Vina: Improving the speed and accuracy of docking with a new scoring function, efficient optimization, and multithreading. J. Comput. Chem. 2010, 31, 455–461. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Makeneni, S.; Thieker, D.F.; Woods, R.J. Applying Pose Clustering and MD Simulations To Eliminate False Positives in Molecular Docking. J. Chem. Inf. Model. 2018, 58, 605–614. [Google Scholar] [CrossRef] [PubMed]
  55. Wei, L.; Chi, B.; Ren, Y.; Rao, L.; Wu, J.; Shang, H.; Liu, J.; Xiao, Y.; Ma, M.; Xu, X.; et al. Conformation Search Across Multiple-Level Potential-Energy Surfaces (CSAMP): A Strategy for Accurate Prediction of Protein-Ligand Binding Structures. J. Chem. Theory Comput. 2019, 15, 4264–4279. [Google Scholar] [CrossRef]
  56. Thomas, E.; Gopalakrishnan, V.; Hegde, M.; Kumar, S.; Karki, S.S.; Raghavan, S.C.; Choudhary, B. A Novel Resveratrol Based Tubulin Inhibitor Induces Mitotic Arrest and Activates Apoptosis in Cancer Cells. Sci. Rep. 2016, 6, 34653. [Google Scholar] [CrossRef] [Green Version]
  57. Samad, A.; Naffaa, M.M.; Bakht, M.A.; Malhotra, M.; Ganaie, M.A. Target Based Designing of Anthracenone Derivatives as Tubulin Polymerization Inhibiting Agents: 3D QSAR and Docking Approach. Int. J. Med. Chem. 2014, 2014, 658016. [Google Scholar] [CrossRef]
  58. Zhang, Y.L.; Li, B.Y.; Yang, R.; Xia, L.Y.; Fan, A.L.; Chu, Y.C.; Wang, L.J.; Wang, Z.C.; Jiang, A.Q.; Zhu, H.L. A class of novel tubulin polymerization inhibitors exert effective anti-tumor activity via mitotic catastrophe. Eur. J. Med. Chem. 2019, 163, 896–910. [Google Scholar] [CrossRef]
  59. Lowe, J.; Li, H.; Downing, K.H.; Nogales, E. Refined structure of alpha beta-tubulin at 3.5 A resolution. J. Mol. Biol. 2001, 313, 1045–1057. [Google Scholar] [CrossRef]
  60. Reijo, R.A.; Cooper, E.M.; Beagle, G.J.; Huffaker, T.C. Systematic mutational analysis of the yeast beta-tubulin gene. Mol. Biol. Cell. 1994, 5, 29–43. [Google Scholar] [CrossRef] [Green Version]
  61. Chattopadhyaya, S.; Chakravorty, D.; Basu, G. A collective motion description of tubulin betaT7 loop dynamics. Biophys. Physicobiol. 2019, 16, 264–273. [Google Scholar] [CrossRef] [Green Version]
  62. Cury, N.M.; Muhlethaler, T.; Laranjeira, A.B.A.; Canevarolo, R.R.; Zenatti, P.P.; Lucena-Agell, D.; Barasoain, I.; Song, C.; Sun, D.; Dovat, S.; et al. Structural Basis of Colchicine-Site targeting Acylhydrazones active against Multidrug-Resistant Acute Lymphoblastic Leukemia. iScience 2019, 21, 95–109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Prota, A.E.; Danel, F.; Bachmann, F.; Bargsten, K.; Buey, R.M.; Pohlmann, J.; Reinelt, S.; Lane, H.; Steinmetz, M.O. The novel microtubule-destabilizing drug BAL27862 binds to the colchicine site of tubulin with distinct effects on microtubule organization. J. Mol. Biol. 2014, 426, 1848–1860. [Google Scholar] [CrossRef] [PubMed]
  64. Niu, L.; Yang, J.; Yan, W.; Yu, Y.; Zheng, Y.; Ye, H.; Chen, Q.; Chen, L. Reversible binding of the anticancer drug KXO1 (tirbanibulin) to the colchicine-binding site of beta-tubulin explains KXO1’s low clinical toxicity. J. Biol. Chem. 2019, 294, 18099–18108. [Google Scholar] [CrossRef] [PubMed]
  65. Li, Y.; Yang, J.; Niu, L.; Hu, D.; Li, H.; Chen, L.; Yu, Y.; Chen, Q. Structural insights into the design of indole derivatives as tubulin polymerization inhibitors. FEBS Lett. 2020, 594, 199–204. [Google Scholar] [CrossRef] [PubMed]
  66. Majcher, U.; Klejborowska, G.; Moshari, M.; Maj, E.; Wietrzyk, J.; Bartl, F.; Tuszynski, J.A.; Huczynski, A. Antiproliferative Activity and Molecular Docking of Novel Double-Modified Colchicine Derivatives. Cells 2018, 7, 192. [Google Scholar] [CrossRef] [Green Version]
  67. Gutierrez, E.; Benites, J.; Valderrama, J.A.; Calderon, P.B.; Verrax, J.; Nova, E.; Villanelo, F.; Maturana, D.; Escobar, C.; Lagos, R.; et al. Binding of dihydroxynaphthyl aryl ketones to tubulin colchicine site inhibits microtubule assembly. Biochem. Biophys. Res. Commun. 2015, 466, 418–425. [Google Scholar] [CrossRef]
  68. Torres, P.H.M.; Sodero, A.C.R.; Jofily, P.; Silva-Jr, F.P. Key Topics in Molecular Docking for Drug Design. Int. J. Mol. Sci. 2019, 20, 4574. [Google Scholar] [CrossRef] [Green Version]
  69. de Campos, L.J.; Palermo, N.Y.; Conda-Sheridan, M. Targeting SARS-CoV-2 Receptor Binding Domain with Stapled Peptides: An In Silico Study. J. Phys. Chem. B 2021, 125, 6572–6586. [Google Scholar] [CrossRef]
  70. Wang, Q.; Arnst, K.E.; Wang, Y.; Kumar, G.; Ma, D.; White, S.W.; Miller, D.D.; Li, W.; Li, W. Structure-Guided Design, Synthesis, and Biological Evaluation of (2-(1H-Indol-3-yl)-1H-imidazol-4-yl)(3,4,5-trimethoxyphenyl) Methanone (ABI-231) Analogues Targeting the Colchicine Binding Site in Tubulin. J. Med. Chem. 2019, 62, 6734–6750. [Google Scholar] [CrossRef]
  71. Usui, T.; Watanabe, H.; Nakayama, H.; Tada, Y.; Kanoh, N.; Kondoh, M.; Asao, T.; Takio, K.; Watanabe, H.; Nishikawa, K.; et al. The anticancer natural product pironetin selectively targets Lys352 of alpha-tubulin. Chem. Biol. 2004, 11, 799–806. [Google Scholar] [CrossRef] [Green Version]
  72. Shi, J.; Shen, Q.; Cho, J.H.; Hwang, W. Entropy Hotspots for the Binding of Intrinsically Disordered Ligands to a Receptor Domain. Biophys. J. 2020, 118, 2502–2512. [Google Scholar] [CrossRef] [PubMed]
  73. Dolenc, J.; Baron, R.; Oostenbrink, C.; Koller, J.; van Gunsteren, W.F. Configurational entropy change of netropsin and distamycin upon DNA minor-groove binding. Biophys. J. 2006, 91, 1460–1470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Genheden, S.; Ryde, U. The MM/PBSA and MM/GBSA methods to estimate ligand-binding affinities. Expert Opin. Drug Discov. 2015, 10, 449–461. [Google Scholar] [CrossRef]
  75. Schiff, P.B.; Fant, J.; Horwitz, S.B. Promotion of microtubule assembly in vitro by taxol. Nature 1979, 277, 665–667. [Google Scholar] [CrossRef] [PubMed]
  76. Schofield, A.V.; Gamell, C.; Suryadinata, R.; Sarcevic, B.; Bernard, O. Tubulin polymerization promoting protein 1 (Tppp1) phosphorylation by Rho-associated coiled-coil kinase (rock) and cyclin-dependent kinase 1 (Cdk1) inhibits microtubule dynamics to increase cell proliferation. J. Biol. Chem. 2013, 288, 7907–7917. [Google Scholar] [CrossRef] [Green Version]
  77. Morris, G.M.; Huey, R.; Lindstrom, W.; Sanner, M.F.; Belew, R.K.; Goodsell, D.S.; Olson, A.J. AutoDock4 and AutoDockTools4: Automated docking with selective receptor flexibility. J. Comput. Chem. 2009, 30, 2785–2791. [Google Scholar] [CrossRef] [Green Version]
  78. Hanwell, M.D.; Curtis, D.E.; Lonie, D.C.; Vandermeersch, T.; Zurek, E.; Hutchison, G.R. Avogadro: An advanced semantic chemical editor, visualization, and analysis platform. J. Cheminform. 2012, 4, 17. [Google Scholar] [CrossRef] [Green Version]
  79. Zhu, K.; Day, T.; Warshaviak, D.; Murrett, C.; Friesner, R.; Pearlman, D. Antibody structure determination using a combination of homology modeling, energy-based refinement, and loop prediction. Proteins 2014, 82, 1646–1655. [Google Scholar] [CrossRef] [Green Version]
  80. Shaw, D.E. A fast, scalable method for the parallel evaluation of distance-limited pairwise particle interactions. J. Comput. Chem. 2005, 26, 1318–1328. [Google Scholar] [CrossRef] [Green Version]
  81. Webb, B.; Sali, A. Protein structure modeling with MODELLER. Methods Mol. Biol. 2014, 1137, 1–15. [Google Scholar]
  82. Pettersen, E.F.; Goddard, T.D.; Huang, C.C.; Couch, G.S.; Greenblatt, D.M.; Meng, E.C.; Ferrin, T.E. UCSF Chimera--a visualization system for exploratory research and analysis. J. Comput. Chem. 2004, 25, 1605–1612. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Kaminski, G.A.; Friesner, R.A.; Tirado-Rives, J.; Jorgensen, W.L. Evaluation and Reparametrization of the OPLS-AA Force Field for Proteins via Comparison with Accurate Quantum Chemical Calculations on Peptides. J. Phys. Chem. B 2001, 105, 6474–6487. [Google Scholar] [CrossRef]
  84. Jensen, K.P.; Jorgensen, W.L. Halide, Ammonium, and Alkali Metal Ion Parameters for Modeling Aqueous Solutions. J. Chem. Theory Comput. 2006, 2, 1499–1509. [Google Scholar] [CrossRef] [PubMed]
  85. Jorgensen, W.L.; Chandrasekhar, J.; Madura, J.D.; Impey, R.W.; Klein, M.L. Comparison of simple potential functions for simulating liquid water. J. Chem. Phys. 1983, 79, 926–935. [Google Scholar] [CrossRef]
  86. Martyna, G.J.; Klein, M.L.; Tuckerman, M. Nosé–Hoover chains: The canonical ensemble via continuous dynamics. J. Chem. Phys. 1992, 97, 2635–2643. [Google Scholar] [CrossRef]
  87. Martyna, G.J.; Tobias, D.J.; Klein, M.L. Constant pressure molecular dynamics algorithms. J. Chem. Phys. 1994, 101, 4177–4189. [Google Scholar] [CrossRef]
  88. Wickham, H. ggplot2: Elegant Graphics for Data Analysis; Springer: New York, NY, USA, 2016. [Google Scholar]
  89. Schlitter, J. Estimation of Absolute and Relative Entropies of Macromolecules Using the Covariance Matrix. Chem. Phys. Lett. 1993, 215, 617–621. [Google Scholar] [CrossRef]
  90. Abraham, M.J.; Murtola, T.; Schulz, R.; Páll, S.; Smith, J.C.; Hess, B.; Lindahl, E. GROMACS: High Performance Molecular Simulations through Multi-Level Parallelism from Laptops to Supercomputers. SoftwareX 2015, 1–2, 19–25. [Google Scholar] [CrossRef] [Green Version]
  91. Lagunin, A.A.; Dubovskaja, V.I.; Rudik, A.V.; Pogodin, P.V.; Druzhilovskiy, D.S.; Gloriozova, T.A.; Filimonov, D.A.; Sastry, N.G.; Poroikov, V.V. CLC-Pred: A freely available web-service for in silico prediction of human cell line cytotoxicity for drug-like compounds. PLoS ONE 2018, 12, e0191838. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Rational design strategy of pyrrolo-heterocyclic target compounds [6,11,14,35].
Figure 1. Rational design strategy of pyrrolo-heterocyclic target compounds [6,11,14,35].
Pharmaceuticals 16 00865 g001
Scheme 1. Reaction pathway for pyrrolopyridazines 4ad.
Scheme 1. Reaction pathway for pyrrolopyridazines 4ad.
Pharmaceuticals 16 00865 sch001
Scheme 2. Reaction pathway for pyrrolopyrimidines 7ad.
Scheme 2. Reaction pathway for pyrrolopyrimidines 7ad.
Pharmaceuticals 16 00865 sch002
Scheme 3. Reaction pathway for pyrroloquinolines 10ad.
Scheme 3. Reaction pathway for pyrroloquinolines 10ad.
Pharmaceuticals 16 00865 sch003
Scheme 4. Reaction pathway for benzo[f]pyrroloquinoline 13ad.
Scheme 4. Reaction pathway for benzo[f]pyrroloquinoline 13ad.
Pharmaceuticals 16 00865 sch004
Scheme 5. Reaction pathway for pyrroloisoquinolines 16ad.
Scheme 5. Reaction pathway for pyrroloisoquinolines 16ad.
Pharmaceuticals 16 00865 sch005
Figure 2. Effect of compound 10a, phenstatin, and paclitaxel (10−5 M) on microtubule dynamics in an in vitro tubulin polymerization assay.
Figure 2. Effect of compound 10a, phenstatin, and paclitaxel (10−5 M) on microtubule dynamics in an in vitro tubulin polymerization assay.
Pharmaceuticals 16 00865 g002
Figure 3. A 2D interaction diagram of tubulin and compound 10a in (a) BM I, (b) BM II, and (c) BM III. Residues are annotated with a three-letter amino acid code and corresponding chain and colored as follows: hydrophobic—green; polar uncharged—light blue; negatively charged—red; and positively charged—dark blue; hydrogen bonds are indicated as pink arrows.
Figure 3. A 2D interaction diagram of tubulin and compound 10a in (a) BM I, (b) BM II, and (c) BM III. Residues are annotated with a three-letter amino acid code and corresponding chain and colored as follows: hydrophobic—green; polar uncharged—light blue; negatively charged—red; and positively charged—dark blue; hydrogen bonds are indicated as pink arrows.
Pharmaceuticals 16 00865 g003
Figure 4. RMSD of 10a-tubulin complexes and colchicine-tubulin complex from 10 ns MD simulations; (a) protein backbone RMSD; (b) ligand RMSD with respect to first frame—Lig fit Lig; (c) ligand RMSD with respect to protein—Lig fit Prot; investigated binding modes of 10a are BM I—green, BM II—black, BM III—orange, and colchicine—blue.
Figure 4. RMSD of 10a-tubulin complexes and colchicine-tubulin complex from 10 ns MD simulations; (a) protein backbone RMSD; (b) ligand RMSD with respect to first frame—Lig fit Lig; (c) ligand RMSD with respect to protein—Lig fit Prot; investigated binding modes of 10a are BM I—green, BM II—black, BM III—orange, and colchicine—blue.
Pharmaceuticals 16 00865 g004
Figure 5. Amino acid interactions throughout the MD simulation for (a) BM I, (b) BM II, and (c) BM III; Hbond—H-bond interactions; Hydroph—Hydrophobic contacts; WaterBridge—hydrogen-bonded protein-ligand interactions mediated by a water molecule; amino acids are abbreviated as chain: 3-letter code:number.
Figure 5. Amino acid interactions throughout the MD simulation for (a) BM I, (b) BM II, and (c) BM III; Hbond—H-bond interactions; Hydroph—Hydrophobic contacts; WaterBridge—hydrogen-bonded protein-ligand interactions mediated by a water molecule; amino acids are abbreviated as chain: 3-letter code:number.
Pharmaceuticals 16 00865 g005
Figure 6. Swiss ADME bioavailability chart of compound 10a.
Figure 6. Swiss ADME bioavailability chart of compound 10a.
Pharmaceuticals 16 00865 g006
Table 1. Selected results of the in vitro growth inhibition (GI%) of tested compounds against human cancer cell lines in the single-dose assay a.
Table 1. Selected results of the in vitro growth inhibition (GI%) of tested compounds against human cancer cell lines in the single-dose assay a.
Cell TypeCompound
Cell Line
GI (%) (10–5 M) a
10a10b10c12a12b12c12d13b16bPhenstatin
LeukemiaCCRF-CEM925651631732094
K-5628831133366326528291
SR9036140931571093
HL-60(TB)100 b,c5083883890100 b,h
MOLT-4660111155248085
RPMI-822689320-----087
Non-small Cell Lung CancerA549/ATCC65021018416262382
HOP-926115172736214912048
HOP-62791515513224343077
NCI-H460860192061748093
NCI-H52298361416261119221788
Colon CancerCOLO205100 b,d000160195058
HCT-1167511126154249491696
HCT-15751216000039096
HT-299150175775017085
SW-6208214722242336078
KM1272427081220091
CNS CancerSF-2957843835632196100 b,i
SF-539100 b,c700147162026100 b,j
SNB-75--521015163015100 b,d100 b,k
U2515144129105371979
MelanomaLOX IMVI60915217318301285
M1490100117416270100 b,l
MDA-MB-435100 b,e4311-----0100 b,m
UACC-6268181691781325955
SK-MEL-2502000000040
SK-MEL-59150424252050100 b,n
Ovarian CancerOVCAR-3100 b,f0012302231170100 b,d
NCI/ADR-RES831380000338100 b,o
SK-OV-366170160130253
IGROV169144121401914086
Renal cancerA498100 b,g330702031125
RXF393100 b,h7502403203999
UO-3168174210161441125-
Breast cancerMCF7826202230283128094
MDA-MB-4689263426053451826100 b,p
HS 578T8907112911171871
BT-54934100411202688
Prostate cancerPC-37820121225123137080
DU-1457800060819090
a Data obtained from NCI’s in vitro sixty-cell one-dose screening at 10−5M; b cytotoxic effect; cell growth percent: c—13; d—7; e—55; f—17; g—11; h—10; i—29; j—22; k—1; l—4; m—41; n—60; o—32; and p—14. The best values in terms of growth inhibition caused by the tested compounds are highlighted in bold.
Table 2. Selected results of the five-dose in vitro human cancer cell growth inhibition a for compound 10a and phenstatin.
Table 2. Selected results of the five-dose in vitro human cancer cell growth inhibition a for compound 10a and phenstatin.
Cell TypeCompound→
Cell Line ↓
10aPhenstatin
GI50
(μM)
TGI
(μM)
GI50
(μM)
TGI
(μM)
LeukemiaK-5620.052>100<0.010>100
HL-60(TB)0.247n.d.0.0110.084
SR0.070>100<0.010>100
CCRF-CEM0.309>1000.309>100
MOLT-40.436>1000.436>100
RPMI-82260.392>1000.392>100
Non-small
Cell Lung Cancer
NCI-H4600.341>1000.033>100
NCI-H5220.0500.8520.027>100
A549/ATCC0.246>1000.057>100
Colon
Cancer
HT290.2081.062.95>100
SW-6200.096>100<0.010>100
KM120.395>100<0.010>100
HCT-1160.219>1000.038>100
COLO 2050.2160.8613.0525.7
CNS
Cancer
SF-2950.059n.d.0.367>100
SF-5390.2160.7230.0110.056
SNB-750.1020.911<0.010>100
U2510.428>1000.043>100
MelanomaLOX IMVI0.751>1000.013>100
M140.137>100<0.010>100
MDA-MB-4350.0350.137<0.010>100
UACC-620.209>1000.448>100
SK-MEL-20.411>100n.d.>100
SK-MEL-50.446>1000.0400.378
Ovarian
Cancer
OVCAR-30.0770.4180.0210.053
NCI/ADR-RES0.118>1000.012>100
Renal
Cancer
786-00.701>1000.905>100
A4980.0270.2152.288.22
RXF 3930.2060.7230.016>100
Breast
cancer
MCF70.075>1000.033>100
HS 578T0.280>1000.031>100
BT-5490.677>1000.034>100
T-47D0.697>10030.4>100
MDA-MB-4680.0630.4730.271n.d.
Prostate
cancer
PC-30.260>1000.045>100
DU-1450.328>1000.039>100
a Data obtained from NCI’s in vitro sixty-cell five-dose screening [41,42,43]. GI50—the molar concentration of the tested compound causing 50% growth inhibition of tumor cells. TGI—the molar concentration of tested compound causing total growth inhibition of tumor cells. n.d.—not determined. The best values showed by compound 10a are highlighted in bold.
Table 3. In silico prediction of ADME parameters for compound 10a.
Table 3. In silico prediction of ADME parameters for compound 10a.
ADME Parameter10a
Physicochemical Properties
Molecular weight433.45 g/mol
Log Po/w (MLOGP)2.30
Number of H-bond acceptors6
Number of H-bond donors0
Number of rotatable bonds8
TPSA75.47 Å2
Pharmacokinetics
Gastrointestinal (GI) absorptionhigh
Blood brain barrier (BBB) permeantno
P-gp substrateno
Drug likeness
Log S (ESOL)−6.00
Water solubility classmoderately soluble
Lipinski ruleno violation
Veber ruleno violation
Bioavailability0.55
Medicinal Chemistry
PAINS alerts0
Brenk alerts0
Synthetic accessibility3.27
Table 4. Results of the prediction for cytotoxicity of compound 10a.
Table 4. Results of the prediction for cytotoxicity of compound 10a.
PaPiCell LineCell Type
0.7510.005T-47DBreast carcinoma
0.7020.002SNU-398Hepatocellular carcinoma
0.3800.009RajiB-lymphoblastic cells
0.3890.062HT-29Colon adenocarcinoma
0.3270.065DU-145Prostate carcinoma
0.3190.127MCF7Breast adenocarcinoma
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

Amărandi, R.-M.; Al-Matarneh, M.-C.; Popovici, L.; Ciobanu, C.I.; Neamțu, A.; Mangalagiu, I.I.; Danac, R. Exploring Pyrrolo-Fused Heterocycles as Promising Anticancer Agents: An Integrated Synthetic, Biological, and Computational Approach. Pharmaceuticals 2023, 16, 865. https://doi.org/10.3390/ph16060865

AMA Style

Amărandi R-M, Al-Matarneh M-C, Popovici L, Ciobanu CI, Neamțu A, Mangalagiu II, Danac R. Exploring Pyrrolo-Fused Heterocycles as Promising Anticancer Agents: An Integrated Synthetic, Biological, and Computational Approach. Pharmaceuticals. 2023; 16(6):865. https://doi.org/10.3390/ph16060865

Chicago/Turabian Style

Amărandi, Roxana-Maria, Maria-Cristina Al-Matarneh, Lăcrămioara Popovici, Catalina Ionica Ciobanu, Andrei Neamțu, Ionel I. Mangalagiu, and Ramona Danac. 2023. "Exploring Pyrrolo-Fused Heterocycles as Promising Anticancer Agents: An Integrated Synthetic, Biological, and Computational Approach" Pharmaceuticals 16, no. 6: 865. https://doi.org/10.3390/ph16060865

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