Next Article in Journal
Synthesis, Surface Activity, Emulsifiability and Bactericidal Performance of Zwitterionic Tetrameric Surfactants
Previous Article in Journal
Effective Detoxification of Olive Mill Wastewater Using Multi-Step Surfactant-Based Treatment: Assessment of Environmental and Health Impact
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Novel Autotaxin Inhibitor ATX-1d Significantly Enhances Potency of Paclitaxel—An In Silico and In Vitro Study

by
Prateek Rai
1,2,†,
Christopher J. Clark
1,2,†,
Carl B. Womack
3,
Curtis Dearing
2,
Joshua Thammathong
2,
Derek D. Norman
4,
Gábor J. Tigyi
4,
Subhabrata Sen
5,
Kevin Bicker
1,2,
April M. Weissmiller
1,3 and
Souvik Banerjee
1,2,*
1
Molecular Biosciences, Middle Tennessee State University, Murfreesboro, TN 37132, USA
2
Department of Chemistry, Middle Tennessee State University, Murfreesboro, TN 37132, USA
3
Department of Biology, Middle Tennessee State University, Murfreesboro, TN 37132, USA
4
Department of Physiology, University of Tennessee Health Science Center, Memphis, TN 38163, USA
5
Department of Chemistry, School of Natural Sciences, Shiv Nadar Institution of Eminence Deemed to be University, Dadri 201314, UP, India
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2024, 29(18), 4285; https://doi.org/10.3390/molecules29184285
Submission received: 7 August 2024 / Revised: 5 September 2024 / Accepted: 6 September 2024 / Published: 10 September 2024

Abstract

:
The development of drug resistance in cancer cells poses a significant challenge for treatment, with nearly 90% of cancer-related deaths attributed to it. Over 50% of ovarian cancer patients and 30–40% of breast cancer patients exhibit resistance to therapies such as Taxol. Previous literature has shown that cytotoxic cancer therapies and ionizing radiation damage tumors, prompting cancer cells to exploit the autotaxin (ATX)–lysophosphatidic acid (LPA)–lysophosphatidic acid receptor (LPAR) signaling axis to enhance survival pathways, thus reducing treatment efficacy. Therefore, targeting this signaling axis has become a crucial strategy to overcome some forms of cancer resistance. Addressing this challenge, we identified and assessed ATX-1d, a novel compound targeting ATX, through computational methods and in vitro assays. ATX-1d exhibited an IC50 of 1.8 ± 0.3 μM for ATX inhibition and demonstrated a significant binding affinity for ATX, as confirmed by MM-GBSA, QM/MM-GBSA, and SAPT in silico methods. ATX-1d significantly amplified the potency of paclitaxel, increasing its effectiveness tenfold in 4T1 murine breast carcinoma cells and fourfold in A375 human melanoma cells without inducing cytotoxic effects as a single agent.

1. Introduction

The autotaxin–lysophosphatidic acid–lysophosphatidic acid receptors (ATX-LPA-LPAR) signaling axis is pivotal in numerous physiological and pathological processes within the body. Autotaxin (ATX) is central to this complex system, as it converts lysophosphatidylcholine (LPC) into lysophosphatidic acid (LPA), thereby precisely regulating LPA levels within the body. ATX was first identified in conditioned media from A2058 melanoma cells as an autocrine motility factor. It generates LPA, activating six bona fide G protein-coupled receptors (GPCR) designated as LPAR1-6. This interface initiates a signaling process that leads to cell proliferation, migration, survival, and inflammation [1,2,3,4]. Dysregulation of the ATX-LPA-LPAR axis has been linked to several diseases, including cancer, fibrosis, autoimmune disorders, neurodegenerative diseases, and cardiovascular disorders [5,6]. Notably, in cancer biology, this axis regulates vital processes such as malignant transformation, invasion, and metastasis [7]. The biggest challenges for effective cancer treatment have been identified as (i) resistance to conventional chemo- and radiation therapy, (ii) the intricate connections within the tumor microenvironment (TME), and evasion of tumor immunity [8,9,10,11,12]. Cytotoxic cancer therapies and ionizing radiation damage the tumor, initiating a self-repair mechanism that is profoundly affected by the ATX-LPA-LPAR axis. Consequently, cancer cells leverage this mechanism to upregulate survival pathways, thereby diminishing the efficacy of subsequent therapy [13,14,15]. The ATX-LPA-LPAR signaling axis has been implicated in chemotherapeutic and radiotherapeutic resistance through multiple mechanisms (Figure 1). These include (i) facilitation of cell survival and proliferation via the PI3K-AKT and RAS-MAPK pathways [16], (ii) augmentation of DNA repair mechanisms [15], (iii) induction of anti-apoptotic processes mediated by the NF-κB pathway, thereby impeding apoptosis [17], (iv) promotion of angiogenesis [18], and (v) modulation of the tumor microenvironment by fostering inflammation and immune suppression [19]. Emerging evidence suggests that the anticancer efficacy of taxanes, doxorubicin, tamoxifen, and sunitinib is intricately linked to the ATX-LPA-LPAR axis [13,20,21,22]. This relationship is partly due to the signaling ability of LPA to reinforce the stability of the transcription factor NRF2. The overexpression of ATX, alongside LPAR1 and LPAR2, has been documented across various tumor types, encompassing including ovarian cancer, osteocarcinoma, metastatic melanoma, neuroblastoma, breast cancer, pancreatic cancer, prostate cancer, and hepatocellular carcinoma [23]. Targeting the ATX-LPA-LPAR signaling axis with inhibitors holds promise for alleviating treatment resistance in patients whose cancer cells resist conventional anticancer therapies. Previous research has highlighted the role of the ATX-LPA-LPAR signaling axis in overcoming cancer resistance when inhibited in combination with chemotherapeutic agents. Li et al. (2011) revealed an intrinsic HDAC inhibitor (HDACi)-resistant mechanism in cancer cells, suggesting that inhibiting the ATX-LPA axis could enhance the efficacy of HDACi-based therapies [24]. Similarly, Long et al. (2023) demonstrated that the combined inhibition of EZH2 and the ATX-LPA-LPAR2 axis exerts synergistic antitumor effects in colon cancer cells [25]. In a significant advancement, the FDA has approved the ATX inhibitor IOA-289 as a potential therapeutic option for patients with pancreatic cancer. IOA-289 was evaluated for metastatic pancreatic cancer in combination with the standard care regimen of nab-paclitaxel and gemcitabine [26].
ATX features a tripartite binding site comprising a catalytic site with two zinc ions, a hydrophobic pocket, and a tunnel. Researchers have explored lipid-like and non-lipid ATX inhibitors, leveraging different mechanisms and binding modes to achieve effective inhibition [27]. Early efforts focused on lipid-like ATX inhibitors that mimic natural substrates like LPC. However, these inhibitors faced limitations in preclinical development due to high partition coefficients (logP > 5), impacting their pharmacokinetic profiles [28]. Despite these challenges, structural insights into ATX, mainly through X-ray cocrystal structures, have facilitated the design of more potent inhibitors [29]. A significant advancement came with the discovery of potent non-acidic inhibitors, such as compound 3, which exhibited an ATX IC50 of 2.2 nM [30]. These inhibitors demonstrated the potential for high efficacy without relying on acidic moieties, typically required to interact with the Zn2+ ions at the active site of the enzyme. Recent research has also explored hybrid inhibitors that combine features of multiple binding modes. These compounds, like GLPG1690, aim to enhance potency and specificity by targeting both the hydrophobic pocket and the tunnel of ATX. Zhang et al. reviewed the recent development of ATX inhibitors [31].
In this study, we report the discovery of a novel hit compound, ATX-1d, identified through in silico and in vitro methods, which enhances the potency of paclitaxel (PTX) in breast cancer and melanoma cell lines. Our approach began with a small library of 60 unique compounds subjected to molecular docking against ATX (PDB ID: 6W35), followed by an ATX enzyme inhibition assay. The library contained compounds from six distinct chemical families: itaconimides, imide-fused pyrroles, 4,6-dioxo-hexahydro-1H-furo [3,4-c]pyrrole derivatives, phenyloxazoli-din-3-yl)pentane-1,5-dione derivatives, spiro[cyclopropane-1,3′-indolin]-2′-one derivatives, and glycine betaine derivatives. To our knowledge, none of the compounds from these families were tested before for their biological activity against autotaxin. This process identified one compound, ATX-1d, that inhibited the ATX hydrolysis of the LPC-like fluorogenic substrate FS-3 by over 50% at a concentration of 10 μM. We further explored the potential interactions of ATX-1d using computational analyses, including molecular dynamics (MD) simulations, molecular mechanics/generalized Born surface area (MM-GBSA), and QM/MM-GBSA. Quantum chemical calculations employing symmetry-adapted perturbation theory (SAPT) and conceptual DFT were conducted to evaluate the strength of intermolecular interactions between the protein and the ligand and to explore the reactive properties of the compound. The computational analyses demonstrated that ATX-1d exhibits an effective binding affinity to ATX. Subsequently, ATX-1d was tested in combination with PTX through an in vitro synergistic assay against A375 melanoma cells and 4T1 breast carcinoma cells, demonstrating significant enhancement of the antitumor activity of PTX.

2. Methods

The methodology followed in this study is demonstrated in Figure 2.

3. Computational Details

3.1. Molecular Docking

The three-dimensional structures of 60 compounds were generated using LigandScout (version 4.4.3) modeling software through MMFF94 energy minimization [32,33]. The 3D structure of the protein target ATX (PDB ID: 6W35) was obtained from the RCSB Protein Data Bank (PDB) [34,35]. The energy minimization of the protein structures was done using the GROMOS96 force field within the Swiss PDB-Viewer 4.1 [36] program to remove bad contacts and unfavorable torsion angles. Molecular docking was conducted using PLANTS software, version 1.2 [37], which integrates an ant colony optimization algorithm with an empirical scoring function for predicting and scoring binding poses of ligands within protein structures. For each compound, 50 binding poses were calculated and scored using the ChemPLP scoring function at a speed setting of 1. All other PLANTS settings were kept at their default values. The results, represented by the best binding affinity, were visualized using Biovia Discovery Studio (https://www.3ds.com/about, accessed on 5 September 2024) [38] to illustrate the interactions between the ligands and the targets. The docking procedure was validated by re-docking the native ligands to ensure alignment with their crystal structure poses. The root-mean-square deviations (RMSDs) between the docked and crystal poses were calculated using the DockRMSD (https://zhanggroup.org/DockRMSD/, accessed on 5 September 2024) web server to confirm the accuracy [39].

3.2. Molecular Dynamics (MD) Simulations

The initial coordinates of the protein–ligand complexes were obtained from molecular docking. MD simulations were performed using the Groningen Machine for Chemical Simulations (GROMACS) [40] software with the CHARMM36 force field [41] for the protein and the CHARMM general force field (CGenFF) for the ligand [42]. The protein–ligand complex was placed in a dodecahedron simulation box with periodic boundary conditions and filled with TIP3P water molecules. Na+ and Cl ions were added to neutralize the system. The system was parametrized, and the geometry was minimized through 5000 steps of the steepest descent method followed by 10,000 steps of the conjugate gradient method to remove any undesirable interactions and relax the system. The system was gradually annealed under an NVT ensemble for 50 ps to reach an optimal temperature of 300 K. This was followed by a 1 ns density equilibration under NPT ensemble conditions, maintaining a constant temperature of 300 K and constant pressure of 1.0 atm. A Langevin thermostat with a collision frequency of 2 ps and a Berendsen barostat with a pressure relaxation time of 1 ps were used to maintain these conditions. The MD simulation employed the particle mesh Ewald method to describe long-range electrostatic interactions with a van der Waals cutoff of 1.2 nm. Production dynamics were then carried out for 100 ns to investigate the behavior of the complex. The average protein–ligand pose over the simulation period was determined using the gmx_cluster module in GROMACS. This method clustered the MD simulation trajectories, employing the GROMOS clustering algorithm with a cutoff of 0.15 nm, to identify the most representative pose of the protein–ligand complex throughout the simulation.

3.3. Binding Free Energy Calculations

The binding free energy calculations were conducted on the most converged trajectories identified during root–mean–square deviation (RMSD) analysis. The MM-GBSA (Molecular Mechanics Generalized Born Surface Area) method was used for these calculations, utilizing the gmx_MMPBSA tool [43]. The calculations included contributions from van der Waals interactions, electrostatic energy, and the electrostatic and non-polar components of the solvation-free energy, with the electrostatic contribution computed using the Generalized Born equation and the non-polar contribution estimated by an empirical model. The exterior dielectric constant was set to 78.5 for the solvation energy calculations, and the solute dielectric constant (εin) was set to 1. The polar component of the solvation energy (ΔEpolar) was calculated using the modified Generalized Born model 2 (GB-OBC2) [44], as developed by Onufriev et al. (referred to as igb = 5 in gmx_MMPBSA). The non-polar component of the solvation energy (ΔEnon-polar) was determined based on the solvent-accessible surface area (SASA), computed using the Linear Combination of Pairwise Overlaps (LCPO) algorithm [45], with the formula ΔEnon-polar = 0.0072 * ΔSASA, with a probe radius of 1.4 Å. Water molecules and added ions were removed from the system prior to the MM-GBSA analysis. The binding free energies (ΔGtotal) of the ligands with the target proteins were calculated using the following formula:
ΔGTotal = ΔGgas + ΔGsolv
where, ΔGgas = ΔEvdw + ΔEele
and, ΔGsolv = ΔEpolar + ΔEnon-polar
ΔGTotal represents the sum of the gas-phase molecular mechanics energy (ΔGgas) and the solvation-free energy (ΔGsolv). ΔGgas is further divided into van der Waals (ΔEvdw) and electrostatic (ΔEele) contributions, while ΔGsolv includes polar (ΔEpolar) and non-polar (ΔEnon-polar) contributions to the solvation energy. The detailed protocol for this method is available in the previous literature [46]. For each complex, the interaction spectrum between the ligand and the protein target on a per-residue basis was calculated using MM/GBSA decomposition analysis supported by the gmx_MMPBSA tool.

3.4. QM/MM-GBSA Calculations

QM/MM-GBSA calculations were performed similarly to MM-GBSA, with the distinction that the quantum mechanical (QM) part specifically employed the semiempirical Hamiltonian PM6-DH+ method [47]. The remainder of the system was treated at the molecular mechanics (MM) level using the ff99SB force field [48]. The QM region encompassed a radius of 5 Å around the binding pocket, comprising both the ligand and amino acid residues from the target protein. This approach enabled the quantum mechanical modeling of interactions within the binding site while efficiently describing the surrounding environment with molecular mechanics.

3.5. Symmetry-Adapted Perturbation Theory (SAPT) Calculations

The interaction energies between protein targets and the ligand were analyzed using SAPT calculations performed with the PSI4 quantum chemistry package [49,50,51]. The methodology involved decomposing these energies into electrostatic (Eele), exchange (Eex), induction (Eind), and dispersion (Edisp) components. The calculations utilized the jun-cc-pVDZ basis set [52] and employed density fitting for improved computational speed. The initial guess for wavefunction optimization was based on the superposition of atomic densities (SAD), while frozen core approximation was implemented to treat inner electron shells as inert during calculations. To enhance computational efficiency, residues involved in interactions with the ligand were partitioned into smaller blocks, effectively reducing the complexity of SAPT calculations [53]. Each partitioned system underwent separate SAPT analysis, with hydrogen atoms capped to preserve chemical valency and ensure accuracy in quantum mechanical calculations. The interaction energies between each ligand and the protein residues were subsequently obtained by summing the energies calculated from these partitioned systems. We used zeroth-order symmetry-adapted perturbation theory (SAPT0) to balance reasonable accuracy with modest computational cost. This methodological approach facilitated a comprehensive assessment of protein–ligand interactions, offering insights into the specific molecular forces that govern binding affinity.

3.6. Conceptual DFT Parameters

We employed Conceptual Density Functional Theory (DFT) to determine the global reactivity parameters of molecular systems using Gaussian 09 software [54]. The parameters calculated include ionization potential, electron affinity, energy gap, electronegativity, chemical potential, global hardness, global softness, electrophilicity index, nucleophilicity index, and maximum charge transfer index [55,56]. The electronic structure calculations were performed with optimized geometries obtained at the M06-2X/6-311 + G(d,p) level of theory [57].
Molecular interactions and structures were visualized using the Biovia Discovery Studio visualizer [38]. Graphs were created with OriginPro 8.5, OriginLab Corporation, Northampton, MA, USA, and GraphPad Prism 5.0a, Boston, MA, USA, www.graphpad.com, while figures were designed with BioRender.com.

4. Experimental Details

4.1. Chemical Samples

Chemical samples of the compounds were received in powder form. The synthesis and characterization of these compounds have been previously reported [58,59,60,61,62,63]. The stock solutions were prepared by dissolving the compounds in DMSO to a final concentration of 25 mM. Further dilutions were performed manually using DMSO or according to the specific assay protocols.

4.2. In Vitro ATX Enzyme Inhibition Assay

Human recombinant ATX (hATX) was purified as previously described [64]. The enzymatic activity of ATX was assessed by measuring the hydrolysis of the synthetic lipid-like FRET-based substrate FS-3 from Echelon Biosciences [64]. Compounds predicted to have potential inhibitory effects through molecular docking were initially screened at a concentration of 10 μM. The reaction mixtures were prepared in 60 μL volumes and loaded into triplicate wells of black-wall 96-well plates. The assay buffer consisted of 50 mM Tris, 140 mM NaCl, 5 mM KCl, 1 mM CaCl2, 1 mM MgCl2, and 10 μM BSA, with the pH adjusted to 8.0. Each assay contained 4 nM ATX and 1 μM FS-3. The test compounds were diluted in an assay buffer and tested in technical triplicates. For dose–response and IC50 determinations, reaction mixtures included 1 μM FS-3, 0 or 4 nM ATX, and varying concentrations of the test compound ranging from 0 to 30 μM. Fluorescence readings were taken at 2 min intervals over 1 h using excitation and emission wavelengths of 485 and 538 nm. Fluorescence data were collected as relative fluorescence units (RFUs) and averaged for the triplicate samples. A compound was considered an experimentally validated hit if it caused more than 50% inhibition of FS-3 hydrolysis at 10 µM concentration. For dose–response studies, nonlinear regression analysis was performed using GraphPad Prism (GraphPad Software, version 10.0.2 (232), San Diego, CA, USA). A variable four-parameter slope model was applied to fit the data, allowing for determining the Hill slope and interpolation of the IC50 ± SD. Fluorescence measurements were conducted using the CLARIOstar Plus plate reader from BMG Labtech (BMG Labtech, Ortenberg, Baden-Württemberg, Germany).

4.3. Cell Viability Assay

To further characterize the compound, we assessed its effects on the viability of 4T1 murine breast cancer cells and A375 melanoma cells. The 4T1 cells were cultured in RPMI-1640 medium (with L-glutamine (Corning, Durham, NC, USA)), supplemented with 10% charcoal-stripped fetal bovine serum (FBS) (Gibco, ThermoFisher Scientific, Waltham, MA, USA) and 1% penicillin/streptomycin. The cells were maintained at 37 °C in a humidified atmosphere with 5% CO2 and passaged before reaching confluence to ensure they remained in the logarithmic growth phase. Similarly, human A375 melanoma cells were purchased from the American Type Culture Collection (ATCC, Manassas, VA, USA) and were cultured in DMEM medium (Corning) supplemented with 10% charcoal-stripped FBS and 1% penicillin/streptomycin. The cells were maintained at 37 °C in a humidified atmosphere containing 5% CO2. To measure the antiproliferative effects of the ATX-1d and PTX alone, 750 4T1 or 1500 A375 cells were plated in triplicate in a 96-well plate with either DMSO or nine doses of each test compound. After 72 h of incubation at 37 °C in a humidified 5% CO2 atmosphere, cell viability was measured using the CellTiter-Glo assay (Promega, Fitchburg, WI, USA). Luminescence values were measured using a CLARIOstar plate reader (BMG Labtech, Ortenberg, Baden-Württemberg, Germany), and data for each dose were normalized to averaged DMSO-treatment values.
To determine if the potential inhibitor could enhance the efficacy of PTX, 4T1 or A375 cells were plated in triplicate on a 96-well plate with 3 μM ATX inhibitor or matched DMSO for 24 h. The next day, nine doses of PTX were added directly to the pre-treated cells and allowed to incubate for an additional 48 h at 37 °C in a humidified 5% CO2 atmosphere. PTX was also added alone for comparison. Cell viability was measured as described in the single compound assay. All viability data were analyzed and plotted using GraphPad Prism (GraphPad Software, version 10.0.2 (232), San Diego, CA, USA). Nonlinear regression analysis was performed using an (inhibitor) vs. normalized response dose–response inhibition model to determine the GI50 concentrations for the ATX inhibitor alone or in combination with PTX.

5. Results and Discussion

5.1. Molecular Docking

We screened a small in-house library of 60 compounds to evaluate their potential to disrupt the ATX-LPA-LPAR signaling axis by targeting ATX (PDB ID: 6W35). Molecular docking, a crucial tool in structure-based drug design, was employed to predict binding affinity between these small molecules and ATX. We validated our docking procedure by docking the native ligand (PDB ID: 6W35) into its binding site, achieving an RMSD of 0.328 between the docking pose and the crystal pose using the DockRMSD server [39], which confirms the reliability of our docking protocol. The superimposition of the crystal structure and the best docking pose obtained from PLANTS docking [37] is shown in the Supplementary Information, Figure S1. The native ligand was identified as a Type IV hybrid tunnel–pocket [27] inhibitor at the ATX binding site. Docking analysis, conducted with the PLANTS software (version 1.2) [37], assessed various ligand poses for their non-bonding interactions within the ATX binding pocket. Many ligands bound similarly to the native ligand (PDB ID: 6W35), interacting with the crucial hydrophobic pocket and tunnel residues such as ALA218, LEU214, PHE274, PHE211, TRP261, TRP255, PHE275, and LEU79. Compounds with docking scores of −90 or below, calculated using the ChemPLP scoring function [37], were considered hits for the in vitro single-dose ATX enzyme inhibition assay. Given the small size of our library, the threshold was set leniently. Of the 60 compounds, 45 met this criterion (Supplementary Information, Table S1). However, despite not being among the top predicted compounds by docking, ATX-1d was the only compound to exhibit more than 50% inhibition in the enzyme inhibition assay. This discrepancy likely stems from the inherent limitations of molecular docking in accurately estimating binding affinities [65]. Detailed analysis revealed that ATX-1d forms bifurcated hydrogen bonds with PHE274 and PHE275 via its carboxylate group (Figure 3). The pyridine group of ATX-1d positioned itself along the tunnel side, while one N-ethyl succinamide group extended towards the side tunnel region and the other towards the hydrophobic pocket. Recognizing the limitations of molecular docking, we further explored the binding mode and affinity of ATX-1d through MD simulations.

5.2. In Vitro Screening through ATX Enzyme Inhibition Assay

Inhibition assays targeting the ATX-mediated hydrolysis of FS-3 were performed with 45 selected compounds, each at a fixed concentration of 10 μM. Compounds were classified as experimentally validated hits if they achieved more than 50% inhibition of the hydrolysis of the model lysophospholipase D substrate FS-3. From this initial screening, only one compound, ATX-1d, met this criterion (Supplementary Information, Table S1). ATX-1d was then subjected to a dose–response assay to determine its IC50 value (Figure 4). This assay, conducted in biological duplicates, produced an average IC50 of 1.8 ± 0.3 μM. BMP-22, a previously validated inhibitor of ATX, was used as a control. The second replicate is shown in the Supplementary Information, Figure S2.

5.3. MD Simulations

MD simulations provide comprehensive insights into how a drug interacts with its biological target, offering a holistic understanding crucial for guiding the drug discovery process. This approach enables informed decisions regarding lead optimization and rational drug design strategies. Unlike molecular docking, which predicts the spatial orientation of a ligand within the active pocket of a protein, MD simulations account for the dynamic nature of proteins under physiological conditions, considering factors such as conformational stability and flexibility. Our study employed MD simulations to examine the equilibration of binding dynamics of ATX-1d over time. In addition to the widely used MM-GBSA end-point binding free energy calculations, we further employed hybrid QM/MM-GBSA calculations to assess the binding affinity of ATX-1d against the target. The QM/MM-GBSA approach integrates quantum mechanics for the crucial amino acids and the ligand with molecular mechanics for the remainder of the system, potentially offering improved accuracy over MM-GBSA. The enhanced performance of QM/MM-GBSA in predicting binding affinities has been documented in several studies. Mishra and Koca (2018) [66], Wichapong et al. (2014) [67], Pu et al. (2017) [68], and Su et al. (2015) have demonstrated the accuracy of the method when evaluating ligand–target interactions. Our utilization of QM/MM-GBSA thus aims to leverage these benefits for a more reliable prediction of the binding affinity of ATX-1d.
We conducted a 100 ns MD simulation of our hit compound, ATX-1d, in a complex with ATX. The average root–mean–square deviations (RMSDs) were observed to be 2.71 Å for the protein backbone, 1.98 Å for the ligand, and 3.87 Å for the protein–ligand complex (Figure 5B). In comparison, the native ligand (PDB ID: 6W35) exhibited RMSDs of 2.99 Å, 0.86 Å, and 1.48 Å for the protein backbone, ligand, and protein–ligand complex, respectively (Figure 5A). The higher RMSD for the ATX-1d protein–ligand complex suggests some movement of ATX-1d at the binding site compared to its initial docking pose. Nonetheless, ATX-1d remained stably bound at the tripartite binding site of ATX throughout the simulation. Analysis of the average pose over the 100 ns simulation revealed that the native ligand (PBD ID: 6W35) forms hydrogen bonds with tunnel residues TRP255 and SER82 and a halogen bond with the hydrophobic pocket residue PHE274. The residues such as LEU214, PHE250, and TRP261 are also involved in π-π and hydrophobic interactions. Superimposition of the average poses from the MD simulations indicated that ATX-1d does not penetrate deeply into the hydrophobic pocket as the native ligand (PBD ID: 6W35) does (Figure 6A). Instead, it remains positioned between the hydrophobic pocket and the tunnel. For ATX-1d, the average pose showed a bifurcated hydrogen bond with tunnel residue LYS249 and a carbon–hydrogen bond with SER82. The interaction profile also included π-π and hydrophobic interactions with residues TRP255, PHE250, TRP261, PHE275, and PHE211 (Figure 6B). To gain deeper insights into the binding affinity of ATX-1d at the ATX binding site, we extended the production MD simulation by an additional 100 ns, resulting in a cumulative simulation time of 200 ns for the ATX-1d–ATX complex. The RMSD plots for the entire 200 ns simulation of ATX-1d are provided in the Supplementary Information, Figure S3. Analysis of the MD simulation trajectories demonstrated that ATX-1d consistently remained within the binding site of protein for the entire 200 ns of the simulation. There were no occurrences of the ligand dissociating from the binding pocket, suggesting that ATX-1d forms a stable and persistent interaction with the target protein. Figure 7 presents the root–mean–square fluctuation (RMSF) analysis, which measures the deviation of Cα atoms from their reference positions over time. Notably, the RMSF values for ATX-1d exhibit minimal variation when compared to the native ligand, suggesting that ATX-1d maintains the structural stability of the protein similarly to the native ligand (Figure 7). Additionally, hydrogen bond analysis indicates that ATX-1d forms between zero and one hydrogen bonds with ATX, whereas the native ligand forms between one and two hydrogen bonds (Figure 8A,B). Binding free energy calculations on the last 55 ns using the MM-GBSA method [43] yielded a value of −26.85 kcal/mol for ATX-1d, predominantly driven by van der Waals interactions (Figure 9). In contrast, the native ligand (PBD ID: 6W35) exhibited a more stable binding free energy of −50.49 kcal/mol (Figure 9). This indicates that ATX-1d has a lower binding affinity than the native ligand (PDB ID: 6W35). The binding free energy of ATX-1d, calculated over the final 100 ns of the 200 ns simulation, is presented in the Supplementary Information, Table S2. Per residue decomposition of the binding free energy for ATX-1d identified PHE275, TRP255, LYS249, PHE250, PHE211, TRP261, and LEU79 as significant contributors (Table 1). The electrostatic contribution from LYS249 was the highest at −5.32 kcal/mol, reflecting its role in hydrogen bond formation. The residue TRP255 contributed the most through van der Waals interactions at −2.62 kcal/mol, followed by PHE275, located between the pocket and the tunnel. To further validate our findings, we performed QM/MM-GBSA calculations on snapshots from the MD simulation. The binding free energies were −20.02 kcal/mol for the native ligand (PDB ID: 6W35) and −8.37 kcal/mol for ATX-1d (Table 2), consistent with the MM-GBSA results and confirming the moderate binding affinity of ATX-1d towards ATX. The moderate binding affinity of ATX-1d observed in silico was also observed in the in vitro ATX enzyme inhibition assay, which produced the IC50 value of 1.8 ± 0.3 μM.

5.4. SAPT Calculations

SAPT is a perturbative method known for accurately calculating interaction energies. However, SAPT is computationally intensive, particularly for large systems. We performed a one-step SAPT0 calculation that included all significant residues of the target as one fragment and the ligand as another, which was not feasible due to the substantial computational time required. To address this, we divided the critical residues predicted by the MD simulations of each target into several blocks. We then separately calculated the interaction energy of each block with ATX-1d (Supplementary Information, Table S4). The total interaction energy of ATX-1d with the significant residues was obtained by summing the interaction energies of the individual blocks. This approach relies on the pairwise additivity assumption to estimate the total interaction energy, simplifying the computational process without significantly compromising accuracy. This methodology aligns with the strategy employed by Bamdad et al. (2022) [69] in their study on the decomposition of interaction energies between various flavonoids and Escherichia coli DNA gyrase using the SAPT method. Similarly, Stasyuk et al. applied this approach to calculate interactions in protein–DNA complexes [70]. Recently, Ding et al. performed SAPT calculations to quantify the strength of intermolecular interactions of GPCRs [71]. These previous studies have demonstrated that SAPT can predict interaction energies with an accuracy comparable to the coupled-cluster method, CCSD(T) [72,73,74].
We evaluated the interactions of ATX-1d with the amino acid residues at the ATX binding site. The analysis revealed that LYS249 exhibited the most stable total SAPT0 interaction energy at −16.83 kcal/mol (Supplementary Information, Table S4), aligning well with the per-residue binding free energy decomposition calculations. The stability is attributed to the bifurcated hydrogen bond formation between ATX-1d and LYS249, resulting in a significant contribution from the electrostatic component. Following LYS249, PHE275 and PHE211 were identified with total SAPT0 interaction energies of −7.74 kcal/mol and −7.60 kcal/mol, respectively. These residues showed the highest contributions of the dispersion component, which can be explained by the presence of π-π and hydrophobic interactions. Based on their total SAPT0 interaction energies, other significant residues included TRP255, LEU244, SER82, and TYR215. The overall interaction energy for ATX-1d was determined to be −49.78 kcal/mol, with the dispersion component contributing the most (Table 3, Figure 10). For comparison, we performed SAPT0 calculations for the native ligand (PDB ID: 6W35), which showed a total interaction energy of −65.80 kcal/mol (Table 3) (Supplementary Information, Table S3). This comparison underscores the strength of the binding affinity of the native ligand (PDB ID: 6W35), reflected in its lower (more negative) interaction energy.
Our comprehensive computational analysis, encompassing molecular docking, MD simulations, MM-GBSA, QM/MM-GBSA, and SAPT0 calculations, indicates that ATX-1d exhibits significant activity against ATX. The pharmacokinetic properties and drug-likeness of ATX-1d were predicted using the SwissADME [75] and pkCSM web servers [76]. SwissADME provided insights into key drug-like properties such as lipophilicity, solubility, and absorption, while pkCSM predicted various pharmacokinetic parameters, including ADMET (Absorption, Distribution, Metabolism, Excretion, and Toxicity) profiles. The results from these computational predictions are summarized in the Supplementary Information, Tables S5 and S6. To evaluate the potential off-target effects of ATX-1d, we utilized the SwissTargetPrediction webserver [77]. The SMILES code of ATX-1d was submitted to this online tool, which predicts possible biological targets based on chemical similarity to known ligands. The analysis yielded low probability scores for other potential targets, suggesting that ATX-1d likely exhibits minimal off-target effects (Supplementary Information, Table S7). Consequently, the observed biological effects are predominantly attributable to the action of ATX-1d on its primary targets.

5.5. Cellular Activity of ATX-1d

To further evaluate the anticancer potential of our hit compound, ATX-1d, we investigated its cellular activity against 4T1 murine breast cancer cells and A375 human melanoma cells using cell viability assays. As expected, each cell line shows a potent response to PTX treatment over three days (Table 4). In contrast, treatment with ATX-1d as a single agent does not reduce cell viability below 50% in either cell line, even at the highest dose, 20 μM. These data indicate that ATX-1d treatment alone is ineffective at impairing cell viability, and aligns with previous findings that ATX inhibitors are non-cytotoxic [78,79].

5.6. ATX Inhibition by ATX-1d Improves Cellular Response to PTX

Given the role of ATX and LPA in pro-survival signaling, we next investigated whether ATX inhibition by ATX-1d could be used in combination with PTX to enhance the cellular response to PTX treatment [80]. To determine this, A375 cells were exposed to increasing concentrations of PTX for 48 h, with or without a 24 h pretreatment with 3 μM ATX-1d, a concentration that does not have any observable cytotoxicity by itself (Table 4). The combination of PTX and ATX-1d resulted in a fourfold increase in the potency of PTX, with the GI50 improving from 38 ± 12 nM (PTX alone) to 10 ± 0 nM (combination, Figure 11A,C). Remarkably, in 4T1 cells, ATX-1d also increases the potency of PTX tenfold, reducing the GI50 from 258 ± 78 nM (PTX alone) to 25 ± 11 nM (combination treatment, Figure 11B,C).

5.7. Conceptual DFT Parameters

The global reactive parameters of ATX-1d, displayed in Table 5, provide critical insights into its chemical behavior. By examining frontier molecular orbitals (FMOs), specifically the highest occupied (HOMO) and lowest unoccupied (LUMO) orbitals, chemists can understand the reactivity and regioselectivity of ATX-1d. Important descriptors such as ionization potential (IP), electron affinity (EA), HOMO-LUMO energy gap (Eg), electronegativity (χ), chemical potential (μ), global hardness (η), global softness (S), electrophilicity index (ω), nucleophilicity index (ε), and maximum charge transfer index (ΔNmax) were calculated.

6. Conclusions

Cancer therapeutic resistance remains a significant obstacle in oncology, compromising treatment efficacy and patient outcomes. Recent studies indicate that the inhibition of ATX can significantly enhance the efficacy of chemotherapeutic agents and prevent the development of resistance, providing a potential pathway for more effective cancer treatments. This study utilized computational methods alongside in vitro assays to identify and evaluate ATX-1d, a novel compound targeting ATX. ATX-1d demonstrated an IC50 of 1.8 (±0.3) μM in the ATX enzyme inhibition assay. Computational studies revealed that ATX-1d exhibits significant activity against ATX. The binding affinities were evaluated using MM-GBSA, QM/MM-GBSA, and SAPT methods, consistently indicating effective interactions with ATX. Molecular dynamics simulations highlighted that ATX-1d binds stably within the tunnel and hydrophobic regions of the ATX tripartite binding site, with LYS249 forming a critical bifurcated hydrogen bond contributing to the binding stability. Importantly, ATX-1d exhibited no cytotoxicity as a single agent in the cell viability assay, but significantly enhanced the efficacy of paclitaxel, increasing its potency fourfold in A375 melanoma cells and tenfold in 4T1 breast carcinoma cells. These results establish ATX-1d as a promising hit compound for further optimization and in vivo studies. This discovery also validates a novel pharmacophore model for developing next-generation ATX inhibitors. The insights gained from these computational and experimental analyses will guide the refinement of the inhibitory mechanism of ATX-1d and inform the future development of therapeutic strategies.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29184285/s1, Figure S1: Superimposition of the crystal pose (yellow) and the docking pose (pink) demonstrates the accuracy of the docking procedure. The evaluation was conducted to validate this process, confirming that the docked pose of the native ligand aligns well with its crystal structure. The root–mean–square deviation (RMSD) between the crystal and docking poses was calculated to be 0.328 Å using the DockRMSD webserver, as detailed in the main text. Table S1: SMILES notation, docking scores, and % inhibition of 60 compounds tested at 10 μM concentration. Figure S2: A second replicate of the in vitro autotaxin enzyme inhibition assay was conducted. The average IC50 values obtained from the biological duplicates were 1.8 ± 0.3 μM for ATX-1d and 0.2 ± 0.1 μM for BMP-22, as mentioned in the main text. Figure S3: RMSD plot for the ATX-1d–ATX complex obtained through the 200 ns MD simulation. Table S2: Binding free energy of ATX-1d calculated using the MM-GBSA method for the last 100 ns of the 200 ns MD simulation. Table S3: Results of the SAPT0 calculations for the SKV (native ligand, PDB ID: 6W35)–ATX complex. Table S4: Results of the SAPT0 calculations for the ATX-1d-ATX complex. Table S5: Evaluation of pharmacokinetics, drug-likeness, and medicinal chemistry friendliness predictions of ATX-1d through SwissADME webserver. Table S6: Prediction of ADMET properties of ATX-1d through pkCSM webserver based on graph-based signatures. Table S7: Results of the SwissTargetPrediction webserver.

Author Contributions

Conceptualization, P.R., C.J.C., D.D.N., G.J.T., S.S., A.M.W. and S.B.; data curation, C.D. and J.T.; funding acquisition, S.B.; Methodology, P.R., C.J.C., C.B.W. and S.B.; project administration, S.B.; software, P.R. and S.B.; supervision, S.B.; Validation, S.B.; visualization, P.R.; writing—original draft, P.R., C.J.C., K.B., A.M.W. and S.B.; writing—review and editing, P.R., C.J.C., C.B.W., D.D.N., G.J.T., S.S., K.B., A.M.W. and S.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Middle Tennessee State University startup funding [210268] and the MT Internal Grant Opportunities [MT-IGO; 220000]. The APC was funded by Middle Tennessee State University startup funding [210268].

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data supporting this article have been included in the main text or as part of the Supplementary Information. The computational chemistry files, including molecular docking output files, molecular dynamics simulation trajectories and output files, SAPT files, and conceptual DFT files, can be made available on request.

Acknowledgments

S.B. would like to thank Middle Tennessee State University (MTSU) and its Molecular Biosciences Ph.D. program for funding from a generous startup package. S.B. would also like to thank MTSU MT-IGO for its funding and crucial assistance with necessary supplies and chemicals. The authors thank Anatoliy Volkov for assisting with MD simulations with the VOLTRON server.

Conflicts of Interest

Authors declare no conflicts of interest.

Abbreviations

ATX, autotaxin; LPAR, lysophosphatidic acid receptor; LPA, lysophosphatidic acid; GPCR, G protein-coupled receptor; TME, tumor microenvironment; PI3K, phosphoinositide 3-kinase; AKT, protein kinase B; RAS, rat sarcoma; MAPK, mitogen-activated protein kinase; DNA, deoxyribonucleic acid; NF-κB, nuclear factor kappa B; NRF2, nuclear factor erythroid 2-related factor 2; HDAC, histone deacetylase; EZH2, enhancer of zeste homolog 2; FDA, food and drug administration; MMFF94, Merck Molecular Force Field 1994; MM-GBSA, molecular mechanics/generalized Born surface area; QM/MM-GBSA, quantum mechanics/molecular mechanics/generalized Born surface area; SAPT, symmetry-adapted perturbation theory; DFT, density functional theory; PTX, paclitaxel; PDB, protein databank; RMSD, root–mean–square deviation; GROMOS96, GROningen MOlecular Simulation 1996; GROMACS, GROningen MAchine for Chemical Simulations; CHARMM36, Chemistry at HARvard Macromolecular Mechanics (version 36); MD, molecular dynamics; TIP3P, Three Point Charge Model (TIP3P); NVT, number of particles (N)–volume (V)–temperature (T); NPT, number of particles–pressure–temperature; SASA, solvent-accessible surface area; LCPO, Linear Combination of Pairwise Overlaps; PM6-DH+, Parametric Method 6 with Dispersion and Hydrogen bonding corrections; ff99SB, Force Field 1999 Side-Chain and Backbone; QM, quantum mechanics; MM, molecular mechanical; SAD, superposition of atomic densities; hATX, human recombinant ATX; FBS, fetal bovine serum; CCSD(T), Coupled Cluster with Single and Double Excitations and Perturbative Triple Excitations.

References

  1. Perrakis, A.; Moolenaar, W.H. Autotaxin: Structure-Function and Signaling. J. Lipid Res. 2014, 55, 1010–1018. [Google Scholar] [CrossRef] [PubMed]
  2. van Meeteren, L.A.; Moolenaar, W.H. Regulation and Biological Activities of the Autotaxin–LPA Axis. Prog. Lipid Res. 2007, 46, 145–160. [Google Scholar] [CrossRef] [PubMed]
  3. Nakanaga, K.; Hama, K.; Aoki, J. Autotaxin—An LPA Producing Enzyme with Diverse Functions. J. Biochem. 2010, 148, 13–24. [Google Scholar] [CrossRef] [PubMed]
  4. Zhang, X.; Li, M.; Yin, N.; Zhang, J. The Expression Regulation and Biological Function of Autotaxin. Cells 2021, 10, 939. [Google Scholar] [CrossRef] [PubMed]
  5. Liu, S.; Murph, M.; Panupinthu, N.; Mills, G.B. ATX-LPA Receptor Axis in Inflammation and Cancer. Cell Cycle 2009, 8, 3695–3701. [Google Scholar] [CrossRef]
  6. Hemmings, D.G.; Brindley, D.N. Signalling by Lysophosphatidate and Its Health Implications. Essays Biochem. 2020, 64, 547–563. [Google Scholar] [CrossRef]
  7. Samadi, N.; Bekele, R.; Capatos, D.; Venkatraman, G.; Sariahmetoglu, M.; Brindley, D.N. Regulation of Lysophosphatidate Signaling by Autotaxin and Lipid Phosphate Phosphatases with Respect to Tumor Progression, Angiogenesis, Metastasis and Chemo-Resistance. Biochimie 2011, 93, 61–70. [Google Scholar] [CrossRef]
  8. Meng, L.; Zheng, Y.; Liu, H.; Fan, D. The Tumor Microenvironment: A Key Player in Multidrug Resistance in Cancer. Oncologie 2024, 26, 41–58. [Google Scholar] [CrossRef]
  9. Morin, P.J. The Microenvironment and Drug Resistance. In Cancer Drug Resistance; Humana Press: Totowa, NJ, USA, 2006; pp. 201–210. [Google Scholar]
  10. Dzobo, K.; Senthebane, D.A.; Dandara, C. The Tumor Microenvironment in Tumorigenesis and Therapy Resistance Revisited. Cancers 2023, 15, 376. [Google Scholar] [CrossRef]
  11. Lin, K.; Lee, S.C.; Dacheux, M.A.; Norman, D.D.; Balogh, A.; Bavaria, M.; Lee, H.; Tigyi, G. E2F7 Drives Autotaxin/Enpp2 Transcription via Chromosome Looping: Repression by P53 in Murine but Not in Human Carcinomas. FASEB J. 2023, 37, e23058. [Google Scholar] [CrossRef]
  12. Koirala, M.; DiPaola, M. Overcoming Cancer Resistance: Strategies and Modalities for Effective Treatment. Biomedicines 2024, 12, 1801. [Google Scholar] [CrossRef] [PubMed]
  13. Samadi, N.; Gaetano, C.; Goping, I.S.; Brindley, D.N. Autotaxin Protects MCF-7 Breast Cancer and MDA-MB-435 Melanoma Cells against Taxol-Induced Apoptosis. Oncogene 2009, 28, 1028–1039. [Google Scholar] [CrossRef] [PubMed]
  14. Tang, X.; Wuest, M.; Benesch, M.G.K.; Dufour, J.; Zhao, Y.; Curtis, J.M.; Monjardet, A.; Heckmann, B.; Murray, D.; Wuest, F.; et al. Inhibition of Autotaxin with GLPG1690 Increases the Efficacy of Radiotherapy and Chemotherapy in a Mouse Model of Breast Cancer. Mol. Cancer Ther. 2020, 19, 63–74. [Google Scholar] [CrossRef] [PubMed]
  15. Deng, W.; Shuyu, E.; Tsukahara, R.; Valentine, W.J.; Durgam, G.; Gududuru, V.; Balazs, L.; Manickam, V.; Arsura, M.; Vanmiddlesworth, L.; et al. The Lysophosphatidic Acid Type 2 Receptor Is Required for Protection against Radiation-Induced Intestinal Injury. Gastroenterology 2007, 132, 1834–1851. [Google Scholar] [CrossRef]
  16. Weiner, J.A.; Chun, J. Schwann Cell Survival Mediated by the Signaling Phospholipid Lysophosphatidic Acid. Proc. Natl. Acad. Sci. USA 1999, 96, 5233–5238. [Google Scholar] [CrossRef]
  17. Liu, S.; Umezu-Goto, M.; Murph, M.; Lu, Y.; Liu, W.; Zhang, F.; Yu, S.; Stephens, L.C.; Cui, X.; Murrow, G.; et al. Expression of Autotaxin and Lysophosphatidic Acid Receptors Increases Mammary Tumorigenesis, Invasion, and Metastases. Cancer Cell 2009, 15, 539–550. [Google Scholar] [CrossRef]
  18. Yasuda, D.; Kobayashi, D.; Akahoshi, N.; Ohto-Nakanishi, T.; Yoshioka, K.; Takuwa, Y.; Mizuno, S.; Takahashi, S.; Ishii, S. Lysophosphatidic Acid–Induced YAP/TAZ Activation Promotes Developmental Angiogenesis by Repressing Notch Ligand Dll4. J. Clin. Investig. 2019, 129, 4332–4349. [Google Scholar] [CrossRef]
  19. Tigyi, G.J.; Yue, J.; Norman, D.D.; Szabo, E.; Balogh, A.; Balazs, L.; Zhao, G.; Lee, S.C. Regulation of Tumor Cell—Microenvironment Interaction by the Autotaxin-Lysophosphatidic Acid Receptor. Adv. Biol. Regul. 2019, 71, 183–193. [Google Scholar] [CrossRef]
  20. Venkatraman, G.; Benesch, M.G.K.; Tang, X.; Dewald, J.; McMullen, T.P.W.; Brindley, D.N. Lysophosphatidate Signaling Stabilizes Nrf2 and Increases the Expression of Genes Involved in Drug Resistance and Oxidative Stress Responses: Implications for Cancer Treatment. FASEB J. 2015, 29, 772–785. [Google Scholar] [CrossRef]
  21. Bekele, R.T.; Venkatraman, G.; Liu, R.-Z.; Tang, X.; Mi, S.; Benesch, M.G.K.; Mackey, J.R.; Godbout, R.; Curtis, J.M.; McMullen, T.P.W.; et al. Oxidative Stress Contributes to the Tamoxifen-Induced Killing of Breast Cancer Cells: Implications for Tamoxifen Therapy and Resistance. Sci. Rep. 2016, 6, 21164. [Google Scholar] [CrossRef]
  22. Su, S.-C.; Hu, X.; Kenney, P.A.; Merrill, M.M.; Babaian, K.N.; Zhang, X.-Y.; Maity, T.; Yang, S.-F.; Lin, X.; Wood, C.G. Autotaxin–Lysophosphatidic Acid Signaling Axis Mediates Tumorigenesis and Development of Acquired Resistance to Sunitinib in Renal Cell Carcinoma. Clin. Cancer Res. 2013, 19, 6461–6472. [Google Scholar] [CrossRef]
  23. Banerjee, S.; Lee, S.; Norman, D.D.; Tigyi, G.J. Designing Dual Inhibitors of Autotaxin-LPAR GPCR Axis. Molecules 2022, 27, 5487. [Google Scholar] [CrossRef]
  24. Li, S.; Wang, B.; Xu, Y.; Zhang, J. Autotaxin Is Induced by TSA through HDAC3 and HDAC7 Inhibition and Antagonizes the TSA-Induced Cell Apoptosis. Mol. Cancer 2011, 10, 18. [Google Scholar] [CrossRef] [PubMed]
  25. Long, Y.; Wang, Y.; Qu, M.; Zhang, D.; Zhang, X.; Zhang, J. Combined Inhibition of EZH2 and the Autotaxin-LPA-LPA2 Axis Exerts Synergistic Antitumor Effects on Colon Cancer Cells. Cancer Lett. 2023, 566, 216226. [Google Scholar] [CrossRef]
  26. Centonze, M.; Di Conza, G.; Lahn, M.; Fabregat, I.; Dituri, F.; Gigante, I.; Serino, G.; Scialpi, R.; Carrieri, L.; Negro, R.; et al. Autotaxin Inhibitor IOA-289 Reduces Gastrointestinal Cancer Progression in Preclinical Models. J. Exp. Clin. Cancer Res. 2023, 42, 197. [Google Scholar] [CrossRef]
  27. Salgado-Polo, F.; Perrakis, A. The Structural Binding Mode of the Four Autotaxin Inhibitor Types That Differentially Affect Catalytic and Non-Catalytic Functions. Cancers 2019, 11, 1577. [Google Scholar] [CrossRef]
  28. Castagna, D.; Budd, D.C.; Macdonald, S.J.F.; Jamieson, C.; Watson, A.J.B. Development of Autotaxin Inhibitors: An Overview of the Patent and Primary Literature. J. Med. Chem. 2016, 59, 5604–5621. [Google Scholar] [CrossRef] [PubMed]
  29. Nishimasu, H.; Okudaira, S.; Hama, K.; Mihara, E.; Dohmae, N.; Inoue, A.; Ishitani, R.; Takagi, J.; Aoki, J.; Nureki, O. Crystal Structure of Autotaxin and Insight into GPCR Activation by Lipid Mediators. Nat. Struct. Mol. Biol. 2011, 18, 205–212. [Google Scholar] [CrossRef]
  30. Jones, S.B.; Pfeifer, L.A.; Bleisch, T.J.; Beauchamp, T.J.; Durbin, J.D.; Klimkowski, V.J.; Hughes, N.E.; Rito, C.J.; Dao, Y.; Gruber, J.M.; et al. Novel Autotaxin Inhibitors for the Treatment of Osteoarthritis Pain: Lead Optimization via Structure-Based Drug Design. ACS Med. Chem. Lett. 2016, 7, 857–861. [Google Scholar] [CrossRef]
  31. Zhang, C.; Liu, Y.; Zhou, Q.; Fan, H.; Liu, X.; Hu, J. Recent Research Advances in ATX Inhibitors: An Overview of Primary Literature. Bioorg. Med. Chem. 2023, 90, 117374. [Google Scholar] [CrossRef]
  32. Halgren, T.A.; Nachbar, R.B. Merck Molecular Force Field. IV. Conformational Energies and Geometries for MMFF94. J. Comput. Chem. 1996, 17, 587–615. [Google Scholar] [CrossRef]
  33. Wolber, G.; Langer, T. LigandScout: 3-D Pharmacophores Derived from Protein-Bound Ligands and Their Use as Virtual Screening Filters. J. Chem. Inf. Model 2005, 45, 160–169. [Google Scholar] [CrossRef] [PubMed]
  34. Cuozzo, J.W. A New Autotaxin Inhibitor for the Treatment of Idiopathic Pulmonary Fibrosis: A Clinical Candidate Discovered Using DNA-Encoded Chemistry. Available online: https://www.wwpdb.org/pdb?id=pdb_00006w35 (accessed on 12 June 2024).
  35. Chrencik, J.E.; Roth, C.B.; Terakado, M.; Kurata, H.; Omi, R.; Kihara, Y.; Warshaviak, D.; Nakade, S.; Asmar-Rovira, G.; Mileni, M.; et al. Crystal Structure of Human Lysophosphatidic Acid Receptor 1 in Complex with ONO-9910539. Cell 2015, 161, 1633–1643. [Google Scholar] [CrossRef] [PubMed]
  36. Guex, N.; Peitsch, M.C. SWISS-MODEL and the Swiss-Pdb Viewer: An Environment for Comparative Protein Modeling. Electrophoresis 1997, 18, 2714–2723. [Google Scholar] [CrossRef] [PubMed]
  37. Korb, O.; Stützle, T.; Exner, T.E. Empirical Scoring Functions for Advanced Protein–Ligand Docking with PLANTS. J. Chem. Inf. Model. 2009, 49, 84–96. [Google Scholar] [CrossRef]
  38. Biovia, D.S.; Berman, H.M.; Westbrook, J.; Feng, Z.; Gilliland, G.; Bhat, T.N.; Richmond, T.J. Discovery Studio Visualizer; Dassault Systèmes BIOVIA: San Diego, CA, USA, 2020. [Google Scholar]
  39. Bell, E.W.; Zhang, Y. DockRMSD: An Open-Source Tool for Atom Mapping and RMSD Calculation of Symmetric Molecules through Graph Isomorphism. J. Cheminform. 2019, 11, 40. [Google Scholar] [CrossRef]
  40. 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, 19–25. [Google Scholar] [CrossRef]
  41. Huang, J.; MacKerell, A.D. CHARMM36 All-Atom Additive Protein Force Field: Validation Based on Comparison to NMR Data. J. Comput. Chem. 2013, 34, 2135–2145. [Google Scholar] [CrossRef]
  42. Vanommeslaeghe, K.; Hatcher, E.; Acharya, C.; Kundu, S.; Zhong, S.; Shim, J.; Darian, E.; Guvench, O.; Lopes, P.; Vorobyov, I.; et al. CHARMM General Force Field: A Force Field for Drug-like Molecules Compatible with the CHARMM All-atom Additive Biological Force Fields. J. Comput. Chem. 2010, 31, 671–690. [Google Scholar] [CrossRef]
  43. Valdés-Tresanco, M.S.; Valdés-Tresanco, M.E.; Valiente, P.A.; Moreno, E. Gmx_MMPBSA: A New Tool to Perform End-State Free Energy Calculations with GROMACS. J. Chem. Theory Comput. 2021, 17, 6281–6291. [Google Scholar] [CrossRef]
  44. Onufriev, A.; Bashford, D.; Case, D.A. Exploring Protein Native States and Large-scale Conformational Changes with a Modified Generalized Born Model. Proteins Struct. Funct. Bioinform. 2004, 55, 383–394. [Google Scholar] [CrossRef] [PubMed]
  45. Weiser, J.; Shenkin, P.S.; Still, W.C. Approximate Atomic Surfaces from Linear Combinations of Pairwise Overlaps (LCPO). J. Comput. Chem. 1999, 20, 217–230. [Google Scholar] [CrossRef]
  46. Fogolari, F.; Brigo, A.; Molinari, H. Protocol for MM/PBSA Molecular Dynamics Simulations of Proteins. Biophys. J. 2003, 85, 159–166. [Google Scholar] [CrossRef]
  47. Korth, M. Third-Generation Hydrogen-Bonding Corrections for Semiempirical QM Methods and Force Fields. J. Chem. Theory Comput. 2010, 6, 3808–3816. [Google Scholar] [CrossRef]
  48. Hornak, V.; Abel, R.; Okur, A.; Strockbine, B.; Roitberg, A.; Simmerling, C. Comparison of Multiple Amber Force Fields and Development of Improved Protein Backbone Parameters. Proteins Struct. Funct. Bioinform. 2006, 65, 712–725. [Google Scholar] [CrossRef]
  49. Parrish, R.M.; Sherrill, C.D. Spatial Assignment of Symmetry Adapted Perturbation Theory Interaction Energy Components: The Atomic SAPT Partition. J. Chem. Phys. 2014, 141, 044115. [Google Scholar] [CrossRef] [PubMed]
  50. Parrish, R.M.; Parker, T.M.; Sherrill, C.D. Chemical Assignment of Symmetry-Adapted Perturbation Theory Interaction Energy Components: The Functional-Group SAPT Partition. J. Chem. Theory Comput. 2014, 10, 4417–4431. [Google Scholar] [CrossRef]
  51. Smith, D.G.A.; Burns, L.A.; Simmonett, A.C.; Parrish, R.M.; Schieber, M.C.; Galvelis, R.; Kraus, P.; Kruse, H.; Di Remigio, R.; Alenaizan, A.; et al. PSI4 1.4: Open-source software for high-throughput quantum chemistry. J. Chem. Phys. 2020, 152, 184108. [Google Scholar] [CrossRef]
  52. Dunning, T.H. Gaussian Basis Sets for Use in Correlated Molecular Calculations. I. The Atoms Boron through Neon and Hydrogen. J. Chem. Phys. 1989, 90, 1007–1023. [Google Scholar] [CrossRef]
  53. Gonthier, J.F.; Sherrill, C.D. Density-Fitted Open-Shell Symmetry-Adapted Perturbation Theory and Application to π -Stacking in Benzene Dimer Cation and Ionized DNA Base Pair Steps. J. Chem. Phys. 2016, 145, 134106. [Google Scholar] [CrossRef]
  54. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 09; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  55. Chattaraj, P.K.; Sarkar, U.; Roy, D.R. Electrophilicity Index. Chem. Rev. 2006, 106, 2065–2091. [Google Scholar] [CrossRef] [PubMed]
  56. Geerlings, P.; De Proft, F.; Langenaeker, W. Conceptual Density Functional Theory. Chem. Rev. 2003, 103, 1793–1874. [Google Scholar] [CrossRef] [PubMed]
  57. Zhao, Y.; Truhlar, D.G. The M06 Suite of Density Functionals for Main Group Thermochemistry, Thermochemical Kinetics, Noncovalent Interactions, Excited States, and Transition Elements: Two New Functionals and Systematic Testing of Four M06-Class Functionals and 12 Other Functionals. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar] [CrossRef]
  58. Maiti, D.; Das, R.; Prabakar, T.; Sen, S. Blue LED Induced Solvent-Free Multicomponent Reactions among Aryl Diazoacetates, Pyridine Derivatives and Maleimides: Direct Eco-Friendly Synthesis of Densely Functionalized Itaconimides. Green Chem. 2022, 24, 3001–3008. [Google Scholar] [CrossRef]
  59. Maiti, D.; Saha, A.; Guin, S.; Maiti, D.; Sen, S. Unveiling Catalyst-Free Electro-Photochemical Reactivity of Aryl Diazoesters and Facile Synthesis of Oxazoles, Imide-Fused Pyrroles and Tetrahydro-Epoxy-Pyridines via Carbene Radical Anions. Chem. Sci. 2023, 14, 6216–6225. [Google Scholar] [CrossRef]
  60. Khan, H.; Guha, S.; Baruah, M.; Yadav, S.; Maheshwari, S.; Sainani, S.; Maiti, D.; Sen, S. Blue LED Induced Three Component Reactions for the Generation of 4,6-Dioxo-hexahydro-1H-furo [3, 4-c] Pyrrole: Their Evaluation as Anticancer Agents through PARP-1 Inhibition. Chem. Asian J. 2023, 18, e202300420. [Google Scholar] [CrossRef]
  61. Singh, S.; Gremaud, L.; Sen, S.; Maiti, D. A Combination of Computational and Experimental Studies to Correlate Electronic Structure and Reactivity of Donor–Acceptor Singlet Carbenes. Synlett 2024, 35, 319–324. [Google Scholar] [CrossRef]
  62. Maurya, R.A.; Reddy, C.N.; Mani, G.S.; Kapure, J.S.; Adiyala, P.R.; Nanubolu, J.B.; Singarapu, K.K.; Kamal, A. Highly Diastereoselective Synthesis of Spiro[Cyclopropane-1,3′-Indolin]-2′-Ones via Catalyst-Free Cyclopropanation Using Ethyl Diazoacetate. Tetrahedron 2014, 70, 4709–4717. [Google Scholar] [CrossRef]
  63. Rath, S.; Maiti, D.; Modi, M.; Pal, P.; Munan, S.; Mohanty, B.; Bhatia, A.; Bhowal, R.; Priyadarshini, R.; Samanta, A.; et al. Metal-Free Synthesis and Study of Glycine Betaine Derivatives in Water for Antimicrobial and Anticancer Applications. iScience 2023, 26, 107285. [Google Scholar] [CrossRef]
  64. North, E.J.; Howard, A.L.; Wanjala, I.W.; Pham, T.C.T.; Baker, D.L.; Parrill, A.L. Pharmacophore Development and Application Toward the Identification of Novel, Small-Molecule Autotaxin Inhibitors. J. Med. Chem. 2010, 53, 3095–3105. [Google Scholar] [CrossRef]
  65. Chen, Y.-C. Beware of Docking! Trends Pharmacol. Sci. 2015, 36, 78–95. [Google Scholar] [CrossRef] [PubMed]
  66. Mishra, S.K.; Koča, J. Assessing the Performance of MM/PBSA, MM/GBSA, and QM–MM/GBSA Approaches on Protein/Carbohydrate Complexes: Effect of Implicit Solvent Models, QM Methods, and Entropic Contributions. J. Phys. Chem. B 2018, 122, 8113–8121. [Google Scholar] [CrossRef] [PubMed]
  67. Wichapong, K.; Rohe, A.; Platzer, C.; Slynko, I.; Erdmann, F.; Schmidt, M.; Sippl, W. Application of Docking and QM/MM-GBSA Rescoring to Screen for Novel Myt1 Kinase Inhibitors. J. Chem. Inf. Model 2014, 54, 881–893. [Google Scholar] [CrossRef] [PubMed]
  68. Pu, C.; Yan, G.; Shi, J.; Li, R. Assessing the Performance of Docking Scoring Function, FEP, MM-GBSA, and QM/MM-GBSA Approaches on a Series of PLK1 Inhibitors. Medchemcomm 2017, 8, 1452–1458. [Google Scholar] [CrossRef]
  69. Bamdad, F.; Farrokhpour, H.; Ashrafizaadeh, M.; Najafi, B. Decomposition of the Interaction Energy of Several Flavonoids with Escherichia Coli DNA Gyr Using the SAPT (DFT) Method: The Relation between the Interaction Energy Components, Ligand Structure, and Biological Activity. Biochim. Biophys. Acta Gen. Subj. 2022, 1866, 130111. [Google Scholar] [CrossRef]
  70. Stasyuk, O.A.; Jakubec, D.; Vondrášek, J.; Hobza, P. Noncovalent Interactions in Specific Recognition Motifs of Protein–DNA Complexes. J. Chem. Theory Comput. 2017, 13, 877–885. [Google Scholar] [CrossRef] [PubMed]
  71. Ding, T.; Karlov, D.S.; Pino-Angeles, A.; Tikhonova, I.G. Intermolecular Interactions in G Protein-Coupled Receptor Allosteric Sites at the Membrane Interface from Molecular Dynamics Simulations and Quantum Chemical Calculations. J. Chem. Inf. Model 2022, 62, 4736–4747. [Google Scholar] [CrossRef]
  72. Matczak, P. Description of Weak Halogen Bonding Using Various Levels of Symmetry-Adapted Perturbation Theory Combined with Effective Core Potentials. J. Chem. 2017, 2017, 9031494. [Google Scholar] [CrossRef]
  73. Pitoňák, M.; Riley, K.E.; Neogrády, P.; Hobza, P. Highly Accurate CCSD(T) and DFT–SAPT Stabilization Energies of H-Bonded and Stacked Structures of the Uracil Dimer. Chem. Phys. Chem. 2008, 9, 1636–1644. [Google Scholar] [CrossRef]
  74. Jansen, G. Symmetry-adapted Perturbation Theory Based on Density Functional Theory for Noncovalent Interactions. WIREs Comput. Mol. Sci. 2014, 4, 127–144. [Google Scholar] [CrossRef]
  75. Daina, A.; Michielin, O.; Zoete, V. SwissADME: A Free Web Tool to Evaluate Pharmacokinetics, Drug-Likeness and Medicinal Chemistry Friendliness of Small Molecules. Sci. Rep. 2017, 7, 42717. [Google Scholar] [CrossRef]
  76. Pires, D.E.V.; Blundell, T.L.; Ascher, D.B. PkCSM: Predicting Small-Molecule Pharmacokinetic and Toxicity Properties Using Graph-Based Signatures. J. Med. Chem. 2015, 58, 4066–4072. [Google Scholar] [CrossRef]
  77. Gfeller, D.; Grosdidier, A.; Wirth, M.; Daina, A.; Michielin, O.; Zoete, V. SwissTargetPrediction: A Web Server for Target Prediction of Bioactive Small Molecules. Nucleic Acids Res. 2014, 42, W32–W38. [Google Scholar] [CrossRef] [PubMed]
  78. Banerjee, S.; Norman, D.D.; Lee, S.C.; Parrill, A.L.; Pham, T.C.T.; Baker, D.L.; Tigyi, G.J.; Miller, D.D. Highly Potent Non-Carboxylic Acid Autotaxin Inhibitors Reduce Melanoma Metastasis and Chemotherapeutic Resistance of Breast Cancer Stem Cells. J. Med. Chem. 2017, 60, 1309–1324. [Google Scholar] [CrossRef] [PubMed]
  79. Iwaki, Y.; Ohhata, A.; Nakatani, S.; Hisaichi, K.; Okabe, Y.; Hiramatsu, A.; Watanabe, T.; Yamamoto, S.; Nishiyama, T.; Kobayashi, J.; et al. ONO-8430506: A Novel Autotaxin Inhibitor That Enhances the Antitumor Effect of Paclitaxel in a Breast Cancer Model. ACS Med. Chem. Lett. 2020, 11, 1335–1341. [Google Scholar] [CrossRef]
  80. Brindley, D.N.; Lin, F.-T.; Tigyi, G.J. Role of the Autotaxin–Lysophosphatidate Axis in Cancer Resistance to Chemotherapy and Radiotherapy. Biochim. Biophys. Acta Mol. Cell Biol. Lipids 2013, 1831, 74–85. [Google Scholar] [CrossRef]
Figure 1. ATX-LPA-LPAR signaling axis in cancer therapeutic resistance.
Figure 1. ATX-LPA-LPAR signaling axis in cancer therapeutic resistance.
Molecules 29 04285 g001
Figure 2. Overview of this study.
Figure 2. Overview of this study.
Molecules 29 04285 g002
Figure 3. Non-bonding interactions within the binding site residues of ATX (PDB ID: 6W35) for (A) native ligand (PDB ID: 6W35) and (B) ATX-1d. Zinc ions are depicted using the CPK model. Hydrogen bonds are shown as green dotted lines and halogen bonds as cyan dotted lines, and other non-bonding interactions such as π-π and π-alkyl interactions are represented by pink dotted lines.
Figure 3. Non-bonding interactions within the binding site residues of ATX (PDB ID: 6W35) for (A) native ligand (PDB ID: 6W35) and (B) ATX-1d. Zinc ions are depicted using the CPK model. Hydrogen bonds are shown as green dotted lines and halogen bonds as cyan dotted lines, and other non-bonding interactions such as π-π and π-alkyl interactions are represented by pink dotted lines.
Molecules 29 04285 g003
Figure 4. Representative dose–response curve of the ATX enzyme inhibition assay. The hit compound, ATX-1d (blue), and BMP-22 (brown), a previously reported ATX inhibitor, were evaluated for their inhibitory effects on ATX activity. The assay was conducted in duplicate (n = 2), yielding an average IC50 of 1.8 ± 0.3 μM and 0.2 ± 0.1 μM for ATX-1d and BMP-22, respectively.
Figure 4. Representative dose–response curve of the ATX enzyme inhibition assay. The hit compound, ATX-1d (blue), and BMP-22 (brown), a previously reported ATX inhibitor, were evaluated for their inhibitory effects on ATX activity. The assay was conducted in duplicate (n = 2), yielding an average IC50 of 1.8 ± 0.3 μM and 0.2 ± 0.1 μM for ATX-1d and BMP-22, respectively.
Molecules 29 04285 g004
Figure 5. RMSDs obtained through 100 ns MD simulations for (A) native ligand (PDB ID: 6W35) and (B) ATX-1d against ATX.
Figure 5. RMSDs obtained through 100 ns MD simulations for (A) native ligand (PDB ID: 6W35) and (B) ATX-1d against ATX.
Molecules 29 04285 g005
Figure 6. (A) Superimposition of the average poses for the native ligand (PDB ID: 6W35) (yellow stick) and ATX-1d (orange stick). The wire mesh represents the protein surface based on hydrophobicity. Blue color represents the hydrophilic regions while brown color represents the hydrophobic regions. (B) Non-bonding interactions of ATX-1d in its average pose obtained through MD simulation against ATX. Hydrogen bonds are shown as green dotted lines, halogen bonds as cyan dotted lines, and other non-bonding interactions such as π-π and π-alkyl interactions are represented by pink dotted lines.
Figure 6. (A) Superimposition of the average poses for the native ligand (PDB ID: 6W35) (yellow stick) and ATX-1d (orange stick). The wire mesh represents the protein surface based on hydrophobicity. Blue color represents the hydrophilic regions while brown color represents the hydrophobic regions. (B) Non-bonding interactions of ATX-1d in its average pose obtained through MD simulation against ATX. Hydrogen bonds are shown as green dotted lines, halogen bonds as cyan dotted lines, and other non-bonding interactions such as π-π and π-alkyl interactions are represented by pink dotted lines.
Molecules 29 04285 g006
Figure 7. Root–mean–square fluctuation (RMSF) of each residue from ATX in complex with SKV (native ligand) (orange) and ATX-1d (blue).
Figure 7. Root–mean–square fluctuation (RMSF) of each residue from ATX in complex with SKV (native ligand) (orange) and ATX-1d (blue).
Molecules 29 04285 g007
Figure 8. Number of hydrogen bonds determined between the protein and the ligand for (A) SKV (native ligand) and (B) ATX-1d throughout the 100 ns MD simulations.
Figure 8. Number of hydrogen bonds determined between the protein and the ligand for (A) SKV (native ligand) and (B) ATX-1d throughout the 100 ns MD simulations.
Molecules 29 04285 g008
Figure 9. Binding free energies for native ligand (PDB ID: 6W35) and ATX-1d in complex with ATX obtained through the MM-GBSA method [43].
Figure 9. Binding free energies for native ligand (PDB ID: 6W35) and ATX-1d in complex with ATX obtained through the MM-GBSA method [43].
Molecules 29 04285 g009
Figure 10. Decomposition of total protein–ligand interaction energy into electrostatic, exchange, induction, and dispersion components [49,50,51].
Figure 10. Decomposition of total protein–ligand interaction energy into electrostatic, exchange, induction, and dispersion components [49,50,51].
Molecules 29 04285 g010
Figure 11. Synergistic effects of ATX-1d with PTX in multiple cancer cell lines. (A) A375 cells were treated for 24 h with 3 μM ATX-1d or DMSO before increasing concentrations of PTX were added for an additional 48 h. Cell viability was determined using CellTiter-Glo and all values were normalized to DMSO-treated samples to obtain the final percent cell viability. The red line indicates the percent cell viability obtained at each dose with paclitaxel treatment in combination with DMSO, and the blue line indicates the same for paclitaxel in combination with ATX-1d (n = 3 biological replicates). (B) Cell viability data for 4T1 cells performed as in (A) (n = 3 biological replicates). (C) Bar graph displays the enhancement in PTX potency by showing extracted GI50 values for each replicate and condition. The error bars represent the standard deviation (SD) obtained from biological triplicates. The GraphPad Prism was used to conduct two-tailed Student’s t-tests for analyzing independent groups in the synergistic assay. Significance levels are denoted as follows: * p < 0.05, ** p < 0.01.
Figure 11. Synergistic effects of ATX-1d with PTX in multiple cancer cell lines. (A) A375 cells were treated for 24 h with 3 μM ATX-1d or DMSO before increasing concentrations of PTX were added for an additional 48 h. Cell viability was determined using CellTiter-Glo and all values were normalized to DMSO-treated samples to obtain the final percent cell viability. The red line indicates the percent cell viability obtained at each dose with paclitaxel treatment in combination with DMSO, and the blue line indicates the same for paclitaxel in combination with ATX-1d (n = 3 biological replicates). (B) Cell viability data for 4T1 cells performed as in (A) (n = 3 biological replicates). (C) Bar graph displays the enhancement in PTX potency by showing extracted GI50 values for each replicate and condition. The error bars represent the standard deviation (SD) obtained from biological triplicates. The GraphPad Prism was used to conduct two-tailed Student’s t-tests for analyzing independent groups in the synergistic assay. Significance levels are denoted as follows: * p < 0.05, ** p < 0.01.
Molecules 29 04285 g011
Table 1. Per-residue decomposition of the binding free energy for ATX-1d in complex with ATX, obtained using the MM-GBSA method. The energy values are given in kcal/mol.
Table 1. Per-residue decomposition of the binding free energy for ATX-1d in complex with ATX, obtained using the MM-GBSA method. The energy values are given in kcal/mol.
Residuesvan der WaalsElectrostaticPolar Solv.Non-Polar Solv.TOTAL
Avg.Std. Err. of MeanAvg.Std. Err. of MeanAvg.Std. Err. of MeanAvg.Std. Err. of MeanAvg.Std. Err. of Mean
GLN:67−0.090.00−0.460.000.450.00−0.040.00−0.140.00
LEU:79−1.290.010.060.00−0.210.00−1.220.00−2.660.01
SER:82−0.470.00−0.060.00−0.080.00−0.430.00−1.030.01
TYR:83−0.780.01−0.550.010.130.00−0.650.01−1.850.02
PHE:211−1.520.01−0.390.000.110.00−1.330.01−3.130.01
TYR:215−0.930.01−0.200.000.020.00−0.720.00−1.830.01
LEU:244−0.740.000.210.00−0.260.00−0.700.00−1.480.01
ARG:245−0.090.000.200.00−0.200.000.000.00−0.100.01
LYS:249−1.830.00−5.320.034.150.02−1.540.00−4.530.03
PHE:250−1.780.00−0.270.000.170.00−1.450.00−3.340.01
TRP:255−2.630.01−0.600.00−0.130.00−1.920.01−5.280.01
TRP:261−1.430.01−0.100.00−0.090.00−1.250.00−2.860.01
ILE:262−0.260.000.000.000.010.00−0.220.00−0.460.00
THR:273−0.180.00−0.020.000.100.00−0.150.00−0.250.00
PHE:274−0.410.00−0.080.000.060.00−0.180.00−0.600.00
PHE:275−2.440.01−0.660.000.060.00−2.020.00−5.060.01
Table 2. Results of QM/MM-GBSA calculations for the ligand and amino acid residues within the 5 Å radius of the ligand. The energy values are given in kcal/mol.
Table 2. Results of QM/MM-GBSA calculations for the ligand and amino acid residues within the 5 Å radius of the ligand. The energy values are given in kcal/mol.
Compoundsvan der WaalsElectrostatics (SCF)Polar Solv.Non-Polar Solv.TOTAL
Avg.Std. Err. of MeanAvg.Std. Err. of MeanAvg.Std. Err. of MeanAvg.Std. Err. of MeanAvg.Std. Err. of Mean
SKV (native ligand, PDB ID: 6W35) vs. ATX−5.850.01−68.090.1158.780.11−7.790.01−22.950.06
ATX−1d vs. ATX−7.060.06−65.580.4369.860.46−5.60.02−8.370.11
Table 3. Results of SAPT0 calculations for the ligand and interacting amino acid residues. The energy values are given in kcal/mol.
Table 3. Results of SAPT0 calculations for the ligand and interacting amino acid residues. The energy values are given in kcal/mol.
CompoundsTotal SAPT0 Protein–Ligand Interaction Energy (in kcal/mol)
SKV (native ligand, PDB ID: 6W35) vs. ATX−65.80
ATX-1d vs. ATX−49.78
Table 4. Cell growth assay for cytotoxicity. Indicated cell lines were plated in triplicate with nine serial dilutions of either ATX-1d or PTX for three days, and cell viability was determined using the CellTiter-Glo assay. Calculated GI50 values are shown below for each agent with the standard error of the mean. The “>20” indicates that the highest dose used in the assay did not reduce cell viability below 50% (n = 3 biological replicates for ATX-d treatment and two biological replicates for PTX treatment).
Table 4. Cell growth assay for cytotoxicity. Indicated cell lines were plated in triplicate with nine serial dilutions of either ATX-1d or PTX for three days, and cell viability was determined using the CellTiter-Glo assay. Calculated GI50 values are shown below for each agent with the standard error of the mean. The “>20” indicates that the highest dose used in the assay did not reduce cell viability below 50% (n = 3 biological replicates for ATX-d treatment and two biological replicates for PTX treatment).
4T1 Breast Cancer CellsA375 Melanoma Cells
CompoundsGI50 (in nM)CompoundsGI50 (in nM)
Paclitaxel62 ± 18Paclitaxel4 ± 1
ATX-1d>20,000ATX-1d>20,000
Table 5. Conceptual DFT parameters calculated at M06-2X/6-311 + G(d,p) level of theory.
Table 5. Conceptual DFT parameters calculated at M06-2X/6-311 + G(d,p) level of theory.
CompoundATX−1d
ELUMO−0.041
EHOMO−0.299
Energy Gap (Eg) {Eg = ELUMO − EHOMO}0.258
Ionization Potential (IP) {IP = −EHOMO}0.299
Electron Affinity (EA) {EA = −ELUMO}0.041
Electronegativity (χ) {χ = ( E LUMO + E HOMO ) 2 } 0.170
Chemical Potential (μ) {μ = ( E LUMO + E HOMO ) 2 } −0.170
Global Hardness (η) {η = E g 2 }0.129
Global Softness (S) {S = 1 2 η }3.875
Electrophilicity Index (ω) {ω = μ 2 2 η }0.112
Nucleophilicity Index (ε) {ε = 1 ω }8.964
Maximum Charge Transfer Index (ΔNmax) {ΔNmax = μ η }1.315
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

Rai, P.; Clark, C.J.; Womack, C.B.; Dearing, C.; Thammathong, J.; Norman, D.D.; Tigyi, G.J.; Sen, S.; Bicker, K.; Weissmiller, A.M.; et al. Novel Autotaxin Inhibitor ATX-1d Significantly Enhances Potency of Paclitaxel—An In Silico and In Vitro Study. Molecules 2024, 29, 4285. https://doi.org/10.3390/molecules29184285

AMA Style

Rai P, Clark CJ, Womack CB, Dearing C, Thammathong J, Norman DD, Tigyi GJ, Sen S, Bicker K, Weissmiller AM, et al. Novel Autotaxin Inhibitor ATX-1d Significantly Enhances Potency of Paclitaxel—An In Silico and In Vitro Study. Molecules. 2024; 29(18):4285. https://doi.org/10.3390/molecules29184285

Chicago/Turabian Style

Rai, Prateek, Christopher J. Clark, Carl B. Womack, Curtis Dearing, Joshua Thammathong, Derek D. Norman, Gábor J. Tigyi, Subhabrata Sen, Kevin Bicker, April M. Weissmiller, and et al. 2024. "Novel Autotaxin Inhibitor ATX-1d Significantly Enhances Potency of Paclitaxel—An In Silico and In Vitro Study" Molecules 29, no. 18: 4285. https://doi.org/10.3390/molecules29184285

Article Metrics

Back to TopTop