Next Article in Journal
Effect of Polyethylene Glycol Additive on the Structure and Performance of Fabric-Reinforced Thin Film Composite
Next Article in Special Issue
Synthesis, Characterization, and Anticancer Activity of New N,N′-Diarylthiourea Derivative against Breast Cancer Cells
Previous Article in Journal
The Study of Optimal Adsorption Conditions of Phosphate on Fe-Modified Biochar by Response Surface Methodology
Previous Article in Special Issue
Synthesis and the In Vitro Evaluation of Antitumor Activity of Novel Thiobenzanilides
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Design and Synthesis of New Acyl Urea Analogs as Potential σ1R Ligands

1
Pharmaceutical Sciences Department, Feik School of Pharmacy, University of the Incarnate Word, San Antonio, TX 78209, USA
2
Department of Physics and Astronomy and Neuroscience Program, Trinity University, San Antonio, TX 78212, USA
3
Department of Chemistry, Florida A&M University, Tallahassee, FL 32307, USA
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(5), 2319; https://doi.org/10.3390/molecules28052319
Submission received: 7 February 2023 / Revised: 21 February 2023 / Accepted: 27 February 2023 / Published: 2 March 2023
(This article belongs to the Special Issue Research of Organic Chemicals for Biological Applications)

Abstract

:
In search of synthetically accessible open-ring analogs of PD144418 or 5-(1-propyl-1,2,5,6-tetrahydropyridin-3-yl)-3-(p-tolyl)isoxazole, a highly potent sigma-1 receptor (σ1R) ligand, we herein report the design and synthesis of sixteen arylated acyl urea derivatives. Design aspects included modeling the target compounds for drug-likeness, docking at σ1R crystal structure 5HK1, and contrasting the lower energy molecular conformers with that of the receptor-embedded PD144418—a molecule we opined that our compounds could mimic pharmacologically. Synthesis of our acyl urea target compounds was achieved in two facile steps which involved first generating the N-(phenoxycarbonyl) benzamide intermediate and then coupling it with the appropriate amines weakly to strongly nucleophilic amines. Two potential leads (compounds 10 and 12, with respective in vitro σ1R binding affinities of 2.18 and 9.54 μM) emerged from this series. These leads will undergo further structure optimization with the ultimate goal of developing novel σ1R ligands for testing in neurodegeneration models of Alzheimer’s disease (AD).

1. Introduction

Inspired by the highly potent σ1R ligand [PD144418 or 5-(1-propyl-1,2,5,6-tetrahydropyridin-3-yl)-3-(p-tolyl)isoxazole] which requires multi-step synthesis, we opted to deconstruct its isoxazole ring and convert it into a more readily accessible acyl urea scaffold. This structural change facilitates quick synthetic access to a broad range of open-ring acyl-urea analogs, with patentable space, as potential σ1R ligands—evidenced by compound SW-43 [1]. The acyl urea scaffold also allows access to analogs with additional rotatable bonds, conformations, and aqueous solubility due to H-donor/acceptor capabilities [2]. Interestingly, small molecules containing urea pharmacophores have adaptable clinical applications, therefore medicinal chemists are also exploring the utility of this functionality in potential anti-neurodegenerative agents [2,3,4]. In fact, literature not only indicates that σ1R is a promising druggable target in neurodegenerative pathologies like AD, but it also suggests additional roles in other disease pharmacologies (addiction, pain, cancer proliferation, HIV, COVID-19, etc.) [1,5,6,7,8]. Now, the unsatisfactory level of the urea’s blood-brain barrier (BBB) penetration due to H-bonding/ionization/polarity/molecular flexibility [2] suggests making structural elaborations to endow derivatives with “drug-like” pharmacokinetics (absorption, distribution, metabolism, elimination, toxicity or ADMET). Drug-like pharmacokinetics are determined by the molecule’s hydrophobicity, flexibility, size, and electronic nature [9,10,11,12], and computational predictions can help identify such in early drug design. To that end, our sixteen compounds were computationally profiled for pharmacokinetics and docked at σ1R prior to their synthesis. Structural design aspects of our arylated acyl-ureas are illustrated in Figure 1.
Online Molinspiration [13] and Osiris property explorer [14] were used to calculate/predict drug-like molecular properties for the target compounds. These studies involved retrieving the σ1R crystal structure (5HK1) with embedded ligand (PD144418), preparation (e.g., H2O and ligand removal), and docking the target molecules via the PyRx and AutoDock Vina suite [15,16]. All sixteen compounds comparably docked to the 5HK1 ligand binding site, suggesting that acyl ureas could also bind to σ1R. Amino acid (AA) interactions in the receptor binding site were also revealed. Synthetically, a variety of approaches for obtaining acyl ureas were reviewed [17,18,19,20,21,22,23,24,25,26,27,28,29] and the most feasible conditions were attempted with limited success—our goal was to obtain our compounds using the shortest and most cost-effective approach. In our hands, even the reported convenient Stokes’ method [23] which required obtaining the same key N-(phenoxycabonyl)benzamide intermediate from diphenyl carbonate, proved futile. Instead, 1H NMR analysis of reaction mixtures revealed phenol, methyl benzoate, and remaining starting materials. We therefore slightly modified Stokes’ conditions [23] by preparing the N-(phenoxycabonyl)benzamide intermediate from phenol and benzoyl isocynate. Coupling the intermediate with amines or amides or thioamides via simple nucleophilic displacement afforded our series of N-acyl ureas plus N-acyl thioureas (Table 1) in good yields (>64%). To confirm our pharmacological assumptions regarding these molecules behaving like PD144418, the compounds were sent to the National Institute of Mental Health’s Psychoactive Drug Screening Program and evaluated in radioligand displacement assays [30] to determine their in vitro σ1R affinities or Ki values.

2. Results and Discussion

2.1. Chemistry

Literature is replete with synthetic options for obtaining acyl ureas. For this project, we reviewed and tried several established approaches [17,18,19,20,21,22,23,24,25,26,27,28,29]. However, our attempts at synthesizing the target compounds were besieged by complex reaction mixtures and poor reaction yields after tedious purifications. For instance, classical conditions (e.g., Scheme 1) which involve initially reacting benzamides and reactive oxalyl chloride followed by the addition of corresponding amines, lead us to dark reaction mixtures with multiple product spots by thin layer chromatography (TLC). Perhaps the acidic conditions were too reactive or unstable and therefore thwarted desired product formation.
We eventually settled on the reported convenient Stokes’ method [23] and slightly modified it by first generating the N-(phenoxycarbonyl)benzamide intermediate from benzoyl isocyanate and phenol (rather than from diphenylcarbonate). Obtaining this key intermediate, in this manner, facilitated the synthesis of our target acyl urea or N-benzoyl urea compounds in Table 1 conveniently and cost-effectively, in two steps (Scheme 2) and in good yields (64–94%). Essentially, our compounds were synthesized by directly coupling amines or amides, or thioamides with a readily scalable N-(phenoxycarbonyl)benzamide intermediate, under reflux, via simple nucleophilic displacements.
This synthetic approach tolerated both weakly and strongly nucleophilic amines, under mild reaction conditions, and is amenable to both parallel and high throughput access to diverse arylated acyl ureas for different applications. All compounds were purified (via filtration and/or combi-Flash) and structurally confirmed by spectroscopic methods [1H and 13C-NMR, high-resolution MS (ES)]. 1H NMRs of this family of compounds showed the presence of an exchangeable signal characterizing the -NH proton at the region δ 9 to 15 ppm DMSO-d6. 13C NMRs ascertained the presence of two to three C=O groups.

2.2. Drug-Likeness Predictions

At a minimum, drug-likeness implies ligands exhibiting Lipinski’s rule of five (Ro5), topological polar surface area (TPSA), and the number of rotatable bond (NRB) limit considerations [9,10,11,12,31]. Table 2 itemizes partial molecular descriptors and drug-likeness properties obtained via Molinspiration /Osiris [13,14] predictions on our target compounds (entries 1–16). Ro5 upper limits are listed in the bottom row.
All target compounds were of low molecular weight (MW < 500), their range was 231.21–396.44. Low molecular weight drug molecules are readily absorbed, diffuse, and are easier to transport versus heavier molecules. Lipophilicity (logP) and TPSA values are predictors of drug oral bioavailability. Calculated log p values of 1.53–3.53 were in the acceptable range of drug molecules capable of penetrating biological membranes including the blood-brain barrier (BBB). The numbers of H-bond acceptors (O and N atoms) and H-bond donors (OH and NH) in the target compounds were within the Ro5 guidelines. TPSA values for all molecules were within the acceptable range. TPSA values are used to characterize potential drug absorption (including intestinal absorption), bioavailability, Caco-2 permeability, and BBB penetration. These values are calculated from the surface areas occupied by oxygen, nitrogen, and the hydrogen atoms attached to them. Thus, the TPSA is closely related to the hydrogen bonding potential of a compound. Interestingly, all compounds exhibited 52.65–119.7 Å TPSA values, indicating potential good oral bioavailabilities. Good bioavailability is more likely for compounds with ≤10 rotatable bonds and TPSA of ≤140 Å. As the NRBs increase, the molecule becomes more flexible and more adaptable for efficient interactions in particular binding pockets. Interestingly all compounds have 1- 4 rotatable bonds and are flexible. Positive drug-likeness scores suggest that the designed molecules retain predominant structural features of known/common commercial drugs. All but two of the sixteen target molecules (compounds 2 and 9) exhibited positive drug-likeness scores.

2.3. Molecular Docking

Docking was conducted to evaluate protein-ligand interactions and predict binding affinities of the target compounds versus the comparator potent σ1R ligand PD144418. Our target molecules were docked using the PyRx and AutoDock Vina platforms [15,16]. Analysis of the docking scores and prediction of binding modes between the ligand and σ1R was conducted using BIOVIA Discovery Studio 4.5 visualizer [32]. Figure 2 contrasts the before (A) and after (B) preparation visualizations of PD14418 in σ1R or 5HK1 binding pocket.
Docking yielded the most stable conformer free-binding energy scores (i.e., binding affinities in kcal/mol) and H-bonding interactions within the active site of σ1R (Table 3).
Additionally, docking revealed relevant amino acids for bonding interactions in the active site of σ1R (5HK1). Example interactions are illustrated in Figure 3.
The binding affinity values of the most stable/best-docked 5HK1 bound conformers ranged from −11.40 to −8.6 kcal/mol. The binding energies of all compounds were significant and comparable to the reference ligand PD14418 (−10.2 kcal/mol) and were commonly due to conventional H-bonding interactions between GLU 172 and the amide bond hydrogen atom. Representative compound 11 exhibited the strongest (three conventional hydrogen bonds via -OH, carbonyl, and -NH) polar interactions with ASP126, THR 181, and TYR 120 residues. In addition to the H-bond interactions illustrated in Figure 3, the 5HK1-compound 11 complexes were also stabilized by π-sigma (Tyr 120) and π-anion (Glu 172, and Asp 126) interactions which also contributed to the calculated lowest root mean square deviation (RMSD) value and provided the best-fit result. On the other hand, the comparator (PD14418) formed: π-sigma with His 194; π-alkyl with Val 84, Leu 105, Phe 133, Val 162, and Ala 185; π-anion with Asp 126; π-donor with Phe 107, Ser 117, Tyr 120, Ile 124, Trp 164, Glu 172 and Thr 181residues.
Computational (kcal/mol) versus in vitro (μM, from the National Institute of Mental Health Psychoactive Drug Screening Program or NIMH-PDSP [https://pdsp.unc.edu/pdspweb/ (accessed 28 February 2023)] σ1R binding affinity data are illustrated in Table 4.
It is worth noting that although we used modeling/docking to guide our synthetic target prioritization, we did not observe a good correlation between predicted versus the in vitro affinity data. This discrepancy is perhaps a result of quantitative limitations of the free and open-source software (FOSS) platforms (e.g., Open Babel) we utilized. For example, ligand protonation states can affect virtual binding affinities [34], ergo, multiple ionization states needed to be included in the prediction data. Moving forward, we endeavor to utilize more robust computational platforms (either through collaborations or by subscription) to optimize the binding affinities of the two leads before synthesizing any follow-up compounds for the SAR studies.

3. Experimental

3.1. Materials and General Methods

Reagents and solvents were purchased from vendors [Fischer Scientific and Acros Organics (Waltham, MA, USA)/Sigma Aldrich (St. Louis, MO, USA] and used as received unless otherwise indicated. All reactions were performed using established procedures. Melting points were taken on the Mel-temp capillary apparatus and reported uncorrected. Teledyne Combiflash Rf flash chromatography fitted with Redisep Rf silica gel cartridges was used for compound purifications. Analytical TLC plates from EM Science (silica Gel 60 F254) were used. 1H and 13C spectra (Supplementary Materials) were recorded using CDCl3 or DMSO-d6 as solvents on a Bruker 300 MHz spectrometer at ambient probe temperature unless otherwise indicated. 1H and 13C chemical shifts are reported versus SiMe4 and were determined by reference to the residual 1H and 13C solvent peaks. Coupling constants (J) are reported in hertz (Hz). Characterization data were reported as follows: chemical shift, multiplicity (s = singlet, d = doublet, t = triplet, q = quartet, br = broad, m = multiplet), coupling constants, number of protons, and mass-to-charge (m/z) ratios. High-Resolution Mass Spectrometry (HRMS) was carried out using an Agilent 6230 electrospray ionization time-of-flight instrument. Where necessary, the purity/mass of the target compounds was determined by analytical reversed-phase high-performance liquid chromatography/mass spectrometer (HPLC/MS) tandem on a Dionex Ultimate 3000 HPLC system. HPLC was conducted using Nova-Pak C18 Column, 60Å, 4 µm, (4.6 mm × 150 mm) at ambient temperature, and a flow rate of 1.0 mL/min., CH3CN & H2O eluent (containing 0.1% acetic acid); gradient, 5% CH3CN to 100% CH3CN; 8 min.; UV detection at 254 nm. Mass was obtained in positive ion mode using heated electrospray ionization (ISQ-HESI) source. MS conditions: capillary voltage, 3.0000 V; drying gas flow, 0.2mL/min; vaporizer gas temperature, 350 °C; and gas pressure, 28.8 psig. MS data were acquired with Chromeleon 7.

3.2. Synthetic Procedures and Compound Data

Compounds 1–15. Individual amines/amides/thioamides were added to N-(Phenoxycarbonyl) benzamide—a versatile common precursor in toluene (20 mL), and resulting mixtures were stirred at room temperature (25 °C) for 10 min, followed by 30–60 min of reflux. Suspended solids that formed upon cooling were collected by filtration, triple rinsed (with 0 °C toluene), and dried under vacuum to afford the desired products. In cases where suspensions did not form after cooling, reaction mixtures were concentrated by rotavapor and then cooled to room temperature to promote crystallization or solidification. Flash chromatography was used as needed to further purify the products. This procedure provided acyl urea and acyl thiourea compounds (1–15) in 64–94% yields. Compound 16 was prepared in 65% yield by directly condensing benzoyl isothiocyanate with 2-aminoisoquinoline in THF for 2 h.

3.2.1. N-((5-Methylisoxazol-3-yl)carbamoyl)benzamide (1)

N-(phenoxycarbonyl) benzamide (632 mg, 2.62 mmol) was reacted with 3-amino-5-methylisoxazole (290 mg, 2.95 mmol) in toluene (60 min reflux) and the pure product was collected as a white crystalline solid, yield: (490 mg, 76%), m.p. 212–215 °C. 1H NMR (300 MHz, CDCl3) δ 11.20 (s, 1H, NH), 9.04 (s, 1H, NH), 7.95 (dd, J = 7.3, 1.8 Hz, 2H, arom.CH), 7.72–7.60 (m, 3H, arom.CH), 6.59 (s, 1H, arom.CH), 2.43 (d, J = 0.9 Hz, 3H, CH3). 13C NMR (75 MHz, CDCl3) δ 170.30, 168.13, 157.71, 151.18, 134.06, 132.05, 129.43, 128.09, 97.01, 13.01. HRMS (ES) m/z calcd. for (C12H11N3O3; M + H)+ 246.0878 found 246.0874.

3.2.2. N-((1H-Benzo[D]imidazol-2-yl)carbamothioyl)benzamide (2)

N-(phenoxycarbonyl) benzamide (560 mg, 2.32 mmol) was condensed with 5-aminoindole (331 mg, 2.49 mmol) in toluene (30 min reflux). The product was obtained as a white crystalline solid, yield: (615 mg, 94%), m.p. 220–223 °C. 1H NMR (300 MHz, DMSO-d6) δ 11.10 (s, 1H, NH), 10.99 (s, 1H, NH), 10.77 (s, 1H, NH), 8.10–8.00 (m, 2H, arom.CH), 7.85 (d, J = 2.0 Hz, 1H, arom.CH), 7.72–7.60 (m, 1H, arom.CH), 7.61–7.45 (m, 2H, arom.CH), 7.42–7.32 (m, 2H, arom.CH), 7.18 (dd, J = 8.7, 2.1 Hz, 1H, arom.CH), 6.42 (ddd, J = 3.0, 1.9, 0.9 Hz, 1H, arom.CH). 13C NMR (75 MHz, DMSO-d6) δ 169.34, 151.78, 133.53, 132.81, 129.91, 129.59, 129.13, 128.59, 128.10, 127.87, 126.71, 115.97, 115.67, 112.11, 112.08, 101.70, 79.38. HRMS (ES) m/z calcd. for (C16H13N3O2; M + H)+ 280.1086 found 280.1080.

3.2.3. N-((Pyridine-2-carbonothioyl)carbamoyl)benzamide (3)

N-(phenoxycarbonyl) benzamide (336 mg, 1.39 mmol) and 2-pyridine thioamide (210 mg, 1.52 mmol) were reacted in toluene (60 min reflux). The desired product was obtained as a purple solid, yield: (304 mg, 75%), m.p. 180–182 °C. 1H NMR (300 MHz, DMSO-d6) δ 14.63 (s, 1H, NH), 11.51 (s, 1H, NH), 8.72 (ddd, J = 4.7, 1.8, 0.9 Hz, 1H, arom.CH), 8.46 (dt, J = 8.1, 1.1 Hz, 1H, arom.CH), 8.12–8.01 (m, 3H, arom.CH), 7.79–7.64 (m, 2H, arom.CH), 7.61–7.52 (m, 2H, arom.CH). 13C NMR (75 MHz, DMSO-d6) δ 194.24, 168.41, 151.60, 149.75, 147.68, 138.12, 133.32, 132.10, 128.59, 127.47, 124.87. HRMS (ES) m/z calcd. for (C14H11N3O2S; M + H)+ 286.0650 found 286.0642.

3.2.4. N-((2-Methyl-1H-indol-5-yl)carbamoyl)benzamide (4)

N-(phenoxycarbonyl) benzamide (740 mg, 3.07 mmol) was reacted with 2-amino-5-methylindole (510 mg, 3.48 mmol) in toluene (40 min reflux). The purified product was a dark brown solid, yield: (635 mg, 71%), m.p. 209–211 °C. 1H NMR (300 MHz, DMSO-d6) δ 10.93 (d, J = 15.5 Hz, 2H, NH), 10.73 (s, 1H, NH), 8.09–7.99 (m, 2H, arom.CH), 7.72–7.62 (m, 2H, arom.CH), 7.55 (dd, J = 8.2, 6.8 Hz, 2H, arom.CH), 7.23 (d, J = 8.6 Hz, 1H, arom.CH), 7.08 (dd, J = 8.6, 2.1 Hz, 1H, arom.CH), 6.11 (dt, J = 2.0, 1.0 Hz, 1H, arom.CH), 2.41–2.34 (m, 3H, CH3). 13C NMR (75 MHz, DMSO-d6) δ 169.32, 151.71, 137.22, 133.72, 133.49, 132.80, 129.90, 129.39, 129.11, 128.57, 115.66, 114.68, 111.17, 111.08, 99.73, 79.37, 13.78. HRMS (ES) m/z calcd. for (C17H15N3O2; M + H)+ 294.1242 found 294.1240.

3.2.5. N-Benzoyl-4-(pyrimidin-2-yl)piperazine-1-carboxamide (5)

N-(phenoxycarbonyl) benzamide (525 mg, 5.05 mmol) was condensed with 1-(2-Pyrimidyl) (315.8 mL, 2.22 mmol) in toluene (60 min reflux). The target compound was collected as solid, yield: (470 mg, 70%), m.p. 145–147 °C. 1H NMR (300 MHz, CDCl3) δ 8.75 (s, 1H, NH), 8.32 (d, J = 4.8 Hz, 2H, arom.CH), 7.63–7.51 (m, 1H, arom.CH), 7.47 (dd, J = 8.2, 6.7 Hz, 2H, arom.CH), 4.03–3.83 (m, 4H, CH2), 3.63 (t, J = 5.2 Hz, 4H, CH2). 13C NMR (75 MHz, CDCl3) δ 166.07, 161.65, 157.92, 153.78, 132.96, 132.85, 128.87, 128.03, 110.55, 43.57. HRMS (ES) m/z calcd. for (C16H17N5O2; M + H)+ 312.1460 found 312.1456.

3.2.6. N-Benzoyl-4-(pyridin-4-yl)piperazine-1-carboxamide (6)

N-(phenoxycarbonyl) benzamide (590 mg, 2.44 mmol) and 1-(4-pyridyl)piperazine (411.51 mg, 2.51 mmol) were reacted in toluene (60 min reflux). This reaction afforded a brown solid, yield: (700 mg, 92%), m.p. 161–163 °C. 1H NMR (300 MHz, CDCl3) δ 9.05 (bs, 1H, NH), 8.33–8.25 (m, 2H, arom.CH), 7.95–7.86 (m, 2H, arom.CH), 7.64–7.53 (m, 1H, arom.CH), 7.48 (dd, J = 8.2, 6.8 Hz, 2H, arom.CH), 6.70–6.61 (m, 2H, arom.CH), 3.75–3.66 (m, 4H, CH2), 3.50–3.40 (m, 4H, CH2). 13C NMR (75 MHz, CDCl3) δ 166.22, 154.57, 153.75, 150.38, 133.00, 132.60, 129.49, 128.80, 127.96, 115.70, 108.58, 45.72. HRMS (ES) m/z calcd. for (C17H18N4O2; M + H)+ 311.1508 found 311.1505.

3.2.7. N-Benzoyl-4-hydroxypiperazine-1-carboxamide (7)

N-(phenoxycarbonyl) benzamide (620 mg, 2.67 mmol) and 4-hydroxypiperidine (270 mg, 3.48 mmol) were stirred in toluene (60 min reflux). The pure product was a white solid, yield: (585 mg, 91%), m.p. 170–172 °C. 1H NMR (300 MHz, DMSO-d6) δ 10.11 (s, 1H, NH), 7.92–7.81 (m, 2H, arom.CH), 7.65–7.54 (m, 1H, arom.CH), 7.55–7.44 (m, 2H, arom.CH), 4.79 (d, J = 4.1 Hz, 1H, OH), 3.367–3.70 (m, 1H, CH), 3.21–3.12 (m, 2H, CH2), 2.52 (m, 2H, CH2), 1.75 (m, 2H,), 1.38 (m, 2H, CH2). 13C NMR (75 MHz, DMSO-d6) δ 166.60, 152.95, 133.78, 132.62, 128.86, 128.44, 65.80, 34.52. HRMS (ES) m/z calcd. for (C13H16N2O3; M + H)+ 249.1234 found 249.1235.

3.2.8. N-Benzoyl-4-(p-tolyl)piperazine-1-carboxamide (8)

N-(phenoxycarbonyl) benzamide (510 mg, 2.11 mmol) dissolve (20 mL) and 1-(4-methylphenyl) piperazine (400 mg, 2.22 mmol) were coupled in toluene (60 min reflux). The filtered pure product was a crystalline white solid, yield: (490 mg, 72 %), m.p. 158–160 °C. 1H NMR (300 MHz, CDCl3) δ 8.53 (s, 1H, NH), 7.93–7.81 (m, 2H, arom.CH), 7.65–7.52 (m, 1H, arom.CH), 7.54–7.41 (m, 2H, arom.CH), 7.14–7.05 (m, 2H, arom.CH), 6.92–6.80 (m, 2H, arom.CH), 3.72 (s, 4H, CH2), 3.26–3.15 (m, 4H, CH2), 2.28 (s, 3H, CH3) 13C NMR (75 MHz, CDCl3) δ 166.01, 153.30, 149.00, 132.99, 132.92, 130.30, 129.90, 129.00, 128.94, 127.95, 117.27, 50.11, 20.60. HPLC/MS tR 5.02 min.; m/z calcd. for (C19H21N3O2; M+) 323.1634 found 323.9460.

3.2.9. Tert-butyl 4-(3-benzoylureido)piperidine-1-carboxylate (9)

N-(phenoxycarbonyl) benzamide (740 mg, 3.07 mmol) and 4-amino-1-Boc-piperidine (510 mg, 3.48 mmol) were stirred in toluene (40 min reflux). The target product was obtained as a white solid, yield: (563 mg, 71%), m.p. 164–166 °C. 1H NMR (300 MHz, CDCl3) δ 8.95 (s, 1H, NH), 8.73 (d, J = 7.5 Hz, 1H, NH), 7.95–7.85 (m, 2H, arom.CH), 7.66–7.55 (m, 1H, arom.CH), 7.49 (dd, J = 8.3, 6.9 Hz, 2H, arom.CH), 4.13–3.83 (m, 2H, CH2), 2.96 (t, J = 12.3 Hz, 2H, CH2), 1.98 (dd, J = 13.0, 3.8 Hz, 2H, CH2), 1.47 (s, 11H, CH). 13C NMR (75 MHz, CDCl3) δ 168.00, 154.74, 153.19, 133.23, 132.33, 129.64, 128.91, 127.56, 115.33, 79.73, 47.38, 42.02, 31.83, 28.45. HRMS-ES) m/z calcd. for (C18H25N3O4; M + H)+ 348.1923 found 348.1880.

3.2.10. N-(Benzo[d]oxazol-2-ylcarbamoyl)benzamide (10)

N-(phenoxycarbonyl) benzamide (571 mg, 2.36 mmol) and 2-amino benzoxazole (340 mg, 2.43 mmol) were condensed in at 25 °C in toluene (60 min reflux). The purified suspension afforded a white solid, yield: (551 mg, 83%), m.p. 180–182 °C. 1H NMR (300 MHz, DMSO-d6) δ 11.74 (s, 1H, NH), 8.03 (dd, J = 7.1, 1.8 Hz, 2H, arom.CH), 7.75–7.52 (m, 5H, arom.CH), 7.40–7.22 (m, 2H, arom.CH). 13C NMR (75 MHz, DMSO-d6) δ 167.59, 155.06, 148.41, 147.68, 140.00, 139.32, 133.37, 132.30, 128.82, 128.25, 124.83, 123.83, 118.11, 110.26. HRMS (ES) m/z calcd. for (C15H11N3O3; M + H)+ 282.0873 found 282.0866.

3.2.11. N-(Benzoylcarbamoyl)-6-hydroxy-2,5,7,8-tetramethylchromane-2-carboxamide (11)

N-(phenoxycarbonyl) benzamide (510 mg, 2.11 mmol) was reacted with 6-hydroxy-2,5,7,8-tetramethylchromane-2-carboxamide (554 mg, 2.22 mmol) in toluene (60 min reflux). The reaction mixture yielded a white solid, yield: (605 mg, 72 %), m.p. 155–157 °C. 1H NMR (300 MHz, CDCl3) δ 8.50 (s, 1H, NH), 7.96–7.86 (m, 2H, arom.CH), 7.69–7.57 (m, 1H, arom.CH), 7.52 (dd, J = 8.2, 6.7 Hz, 2H, arom.CH), 6.36 (s, 1H, OH), 2.63 (q, J = 6.1 Hz, 2H, CH2), 2.37 (dt, J = 12.3, 5.9 Hz, 1H, CH2), 2.15 (d, J = 12.7 Hz, 3H, CH3), 2.06 (s, 3H, CH3), 1.89 (ddd, J = 13.9, 8.1, 6.4 Hz, 1H, CH2), 1.57 (s, 3H, CH3). 13C NMR (75 MHz, CDCl3) δ 176.96, 164.91, 149.64, 148.57, 141.00, 133.39, 132.89, 129.11, 127.91, 126.27, 122.73, 118.42, 78.88, 29.16, 24.55, 20.47, 13.18, 12.32, 12.15. HRMS (ES) m/z calcd. for (C22H24N2O5; M + H)+ 397.1763 found 397.1759.

3.2.12. N-((4-Hydroxy-3-methoxyphenylcarbonothioyl)carbamoyl) benzamide (12)

N-(phenoxycarbonyl) benzamide (478 mg, 1.98 mmol) and 4-hydroxy-3- methoxythiobenzamide (381 mg, 2.10 mmol) were reacted in toluene (30 min reflux). The pure product was collected as a yellow solid, yield: (420 mg, 64%), m.p. 176–177 °C. 1H NMR (300 MHz, DMSO-d6) δ 12.90–12.83 (m, 1H, arom.CH), 11.55 (s, 1H, arom.CH), 10.15 (s, 1H, OH), 8.02 (d, J = 8.0 Hz, 2H, arom.CH), 7.80–7.33 (m, 5H, arom.CH), 6.87 (d, J = 8.4 Hz, 1H, arom.CH), 3.85 (s, 3H, OCH3). 13C NMR (75 MHz, DMSO-d6) δ 169.23, 152.00, 147.64, 133.94, 133.68, 132.52, 129.21, 128.79, 121.90, 115.29, 112.68, 79.36, 56.13. HRMS (ES) m/z calcd. for (C16H14N2O4S; M + H)+ 331.0752 found 331.0745.

3.2.13. N-((4,6-Dimethylpyrimidin-2-yl)carbamoyl)benzamide (13)

N-(phenoxycarbonyl) benzamide (1.22 g, 5.05 mmol) and 2-amino-4,6-dimethylpyrimidine (650 mg, 5.27 mmol) were condensed in toluene (30 min reflux). The suspension obtained after being cooled down the reaction mixture was collected by filtration and dried. The pure product was collected by filtration and dried under a vacuum. White crystalline solid, yield: (913 mg, 68%), m.p. 183–185 °C. 1H NMR (300 MHz, CDCl3) δ 8.09–7.98 (m, 2H, arom.CH), 7.67–7.55 (m, 1H, arom.CH), 7.58 –7.45 (m, 2H, arom.CH), 6.77 (s, 1H, arom.CH), 2.48 (s, 6H, 2CH3). 13C NMR (75 MHz, CDCl3) δ 156.87, 149.58, 133.00, 128.79, 127.89, 23.85. HRMS (ES) m/z calcd. for (C14H14N4O2; M + H)+ 271.1195 found 271.1188.

3.2.14. N-(Quinolin-3-ylcarbamoyl)benzamide (14)

N-(phenoxycarbonyl) benzamide (460 mg, 1.90 mmol) dissolved in toluene (20 mL) and 3-aminoquinoline (291 mg, 2.0 mmol) were coupled in toluene (60 min reflux). The resulting pure product was a white fluffy solid, yield: (505 mg, 94 %), m.p. 233–236 °C. 1H NMR (300 MHz, DMSO-d6) δ 11.27 (s, 1H, NH), 11.15 (s, 1H, NH), 9.00 (d, J = 2.5 Hz, 1H, arom.CH), 8.69 (d, J = 2.6 Hz, 1H, arom.CH), 8.14–7.91 (m, 4H, arom.CH), 7.75–7.49 (m, 5H, arom.CH). 13C NMR (75 MHz, DMSO-d6) δ 169.19, 152.13, 145.68, 144.93, 133.65, 132.68, 132.03, 129.10, 129.07, 128.84, 128.58, 128.24, 128.22, 127.65, 123.62. HRMS (ES) m/z calcd. for (C17H13N3O2; M + H)+ 292.1086 found 292.1082.

3.2.15. N-(Isoquinolin-3-ylcarbamoyl)benzamide (15)

N-(phenoxycarbonyl) benzamide (608 mg, 2.52 mmol) and 3-aminoisoquinoline (430 mg, 2.98 mmol) were coupled in toluene (30 min reflux). The purified desired compound was a white solid, yield: (691 mg, 94 %), m.p. 237–238 °C. 1H NMR (300 MHz, DMSO-d6) δ 11.45 (s, 1H, NH), 11.28 (s, 1H, NH), 9.18 (d, J = 1.1 Hz, 1H, arom.CH), 8.42 (s, 1H, arom.CH), 8.14–8.01 (m, 3H, arom.CH), 7.98 (dd, J = 8.5, 1.1 Hz, 1H, arom.CH), 7.81–7.62 (m, 2H, arom.CH), 7.62–7.51 (m, 3H, arom.CH). 13C NMR (75 MHz, DMSO-d6) δ 169.39, 152.26, 151.31, 146.56, 137.69, 133.61, 132.77, 131.61, 129.05, 128.88, 128.19, 126.89, 126.46, 126.38, 107.06. HRMS (ES) m/z calcd. for (C17H13N3O2; M + H)+ 292.1086 found 292.1083.

3.2.16. N-(Isoquinolin-3-ylcarbamothioyl)benzamide (16)

Benzoyl isothiocyanate (1.0 mL, 8.37 mmol) was added dropwise to 3-aminoisoquinoline (1.2 gm, 8.33 mmol) dissolved in THF (20 mL) at room temperature. The reaction mixture was refluxed for 2 h, cooled, and filtered. The yellow solid obtained was further purified by flash chromatography DCM/MeOH (9:1) to afford a pale yellow solid, yield: (1.5 gm, 65%), m.p. 162–165 °C. 1H NMR (300 MHz, CDCl3) δ 13.35 (s, 1H, NH), 9.29 (s, 1H, NH), 9.10 (s, 2H, arom.CH), 8.00–7.85 (m, 4H, arom.CH), 7.75–7.48 (m, 6H, arom.CH). 13C NMR (75 MHz, CDCl3) δ 176.24, 166.54, 151.55, 146.10, 137.23, 133.83, 131.82, 131.10, 129.32, 127.68, 127.65, 127.40, 127.03, 110.89. HPLC/MS tR 5.76 min.; m/z calcd for (C17H13N3OS; M+) 307.0779 found 307.8182.

3.3. Molecular Docking

The docking protocol included: deriving two-dimension (2D) structures of target compounds by ChemDraw Ultra 19.1 and saving them as structural data files (SDFs) prior to converting them into three-dimension (3D) images with openBabel 3.1.1 program [35] in PyRx [15]. Energy minimized conformers were then converted into Protein Data Bank, Partial Charge (Q), and Atom Type (T) or pdbqt format. The σ1R/PD-14418 complex structure was obtained from the Protein Data Bank database (http://www.rcsb.org/pdb, PDB, ID: 5HK1), and imported into Discovery Studio software for visualizing the binding domain of the complex and identification of the amino acid (AA) residues in the binding pocket. Hydrogen atoms were added to the protein to correct the ionization and tautomeric states of amino acid residues. Furthermore, H2O molecules and complexes bound to σ1R were removed before the docking. In addition, the protein was subjected to energy minimization by applying the UFF force field, with a maximum number of 200 steps at an RMS gradient of 0.02. Amino acids within 8Å of the binding sites were used for docking studies in Autodock vina [16]. The optimized protein was saved in pdbqt format and imported to the PyRx workstation for molecular docking via Autodock Vina virtual screening tool.
The auto grid generated by vina wizard in PyRx was utilized for arrangement affiliations of grid maps and contained 1.00 Å spacing with box dimensions of 11.78X × 37.22Y × −34.86Z Å and centers of 76.160X × 71.016Y × 71.999Z Å. The docking grid box was centered to cover residues within the active site in the protein structure. The validation of the docking study was performed by re-docking PD14418 into an appropriate protein cavity. Re-docking was accepted if the root mean square deviation (RMSD) was <2.0 Å. Hydrophobic and hydrogen bond interactions between the docked compounds and protein were investigated utilizing the discovery studio software.

3.4. In Vitro σ1R Binding Assay

Sigma-1 binding affinity experiments (Ki determinations) were conducted at the National Institute of Mental Health’s Psychoactive Drug Screening Program (NIMH PDSP). Sigma-1 receptor binding assay protocol details (radioligand displacement conditions, etc.) are available at https://pdsp.unc.edu/pdspweb/content/PDSP%20Protocols%20II%202013-03-28.pdf (accessed 28 February 2023) [30]. The Ki binding affinity values of our target compounds are also displayed in Table 4.

4. Conclusions

Sixteen acyl ureas (i.e., benzoyl ureas and benzoyl thioureas) were rationally designed, and readily synthesized in good to high yields (64–94%). Per the SciFinder database, fourteen of these compounds are new. This compound series exhibited drug-like properties (complied with Ro5 and TPSA/NRB parameters), possessed modeled σ1R affinities (−8.6 to −11.4 kcal/mol) in the range comparable to the reference PD14418 (−10.1 kcal/mol), and the in vitro σ1R binding data revealed two compounds (10 and 12, with respective Ki values of 2.18 and 9.54 μM) as promising leads. The two new leads have now opened the door for optimization in our follow-up structure-activity relationship (SAR) studies in search of potent σ1R acyl urea ligands with anti-neurodegenerative applications.

Supplementary Materials

Compound 1HNMR and 13CNMR spectra are available online at https://www.mdpi.com/article/10.3390/molecules28052319/s1.

Author Contributions

R.T., performed synthesis and structural analysis. R.F. assisted with synthesis and molecular modeling. K.H.C. and B.M. contributed to the docking studies. D.S. wrote and edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This project is funded by the NIH NGMS grant #1SC3GM140980-01 (Donald Sikazwe, PI), and the Welch Foundation grant W-2057-20210327 (Kwan H. Cheng, PI).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Thanks to Adam Urbach (Trinity University, San Antonio, TX, USA) for conducting the HRMS analysis. Thanks to Bryan L. Roth, (Director of the NIMH PDSP, Contract # HHSN-271-2018-00023-C. https://pdsp.unc.edu/pdspweb/ (accessed 28 February 2023), North University Carolina at Chapel Hill, Chapel Hill, NC, USA), and Jamie Driscoll (Project Officer, NIMH, Bethesda MD, USA) for providing the in vitro σ1R binding affinity or Ki data. Thanks to Osbert N. Sikazwe (formerly of University of Zambia, Lusaka, Zambia) for inspirational guidance on this project.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fallica, A.N.; Pittalà, V.; Modica, M.N.; Salerno, L.; Romeo, G.; Marrazzo, A.; Helal, M.A.; Intagliata, S. Recent Advances in the Development of Sigma Receptor Ligands as Cytotoxic Agents: A medicinal Chemistry Perspective. J. Med. Chem. 2021, 64, 7926–7962. [Google Scholar] [CrossRef]
  2. Ghosh, A.K.; Brindisi, M. Urea derivatives in modern drug discovery and medicinal chemistry. J. Med. Chem. 2020, 63, 2751–2788. [Google Scholar] [CrossRef]
  3. Bialer, M.; Yagen, B.; Shimshoni, J.A. Acyl-Urea Derivatives and Uses Thereof. U.S. Patent US 8,846,903, 30 September 2014. [Google Scholar]
  4. Ujan, R.; Channar, P.A.; Ali Bahadur, A.; Abbas, Q.; Shah, M.; Rashid, S.G.; Iqbal, S.; Saeed, A.; Abd-Rabboh, H.S.M.; Raza, H.; et al. Synthesis, kinetics and biological assay of some novel aryl bis-thioureas: A potential drug candidates for Alzheimer’s disease. J. Mol. Struct. 2021, 1246, 131136. [Google Scholar] [CrossRef]
  5. Ryskamp, D.A.; Korban, S.; Zhemkov, V.; Kraskovskaya, N.; Bezprozvanny, I. Neuronal Sigma-1 Receptors: Signaling Functions and Protective Roles in Neurodegenerative Diseases. Front. Neurosci. 2019, 13, 862. [Google Scholar] [CrossRef] [Green Version]
  6. Nguyen, L.; Lucke-Wold, B.P.; Mookerjee, S.; Kaushal, N.; Matsumoto, R.R. Sigma-1 Receptors and Neurodegenerative Diseases: Towards a Hypothesis of Sigma-1 Receptors as Amplifiers of Neurodegeneration and Neuroprotection. Adv. Exp. Med. Biol. 2017, 964, 133–152. [Google Scholar]
  7. Maurice, T.; Su, T.-P. The Pharmacology of sigma-1 receptors. Pharmacol. Ther. 2009, 124, 195–206. [Google Scholar] [CrossRef] [Green Version]
  8. Vela, J.M. Repurposing sigma-1 receptor ligands for COVID-19 therapy? Front. Pharmacol. 2020, 11, 582310. [Google Scholar] [CrossRef]
  9. Lipinski, C.A.; Lombardo, F.; Dominy, B.W.; Feeney, P.J. Experimental and computational approaches to estimate solubility and permeability in drug discovery and development settings. Adv. Drug Deliv. Rev. 2012, 64, 4–17. [Google Scholar] [CrossRef]
  10. Lipinski, C.A. Lead- and drug-like compounds: The rule-of-five revolution. Drug Discov. Today Technol. 2004, 1, 337–341. [Google Scholar] [CrossRef]
  11. Topliss, J.G. Chance Correlations in Structure-Activity Studies Using Multiple Regression Analysis. J. Med. Chem. 1972, 15, 1006–1009. [Google Scholar] [CrossRef]
  12. Bhowmik, R.; Roy, S.; Sengupta, S.; Ravi, L. Computer aided drug design of florfenicol to target chloramphenicol acetyltransferase of vibriosis causing pathogens. J. Appl. Biol. Biotech. 2022, 10, 76–84. [Google Scholar]
  13. Molinspiration Cheminformatics. Available online: http://www.molinspiration.com/cgi-bin/properties (accessed on 25 January 2023).
  14. OSIRIS Property Explorer. Available online: https://www.organic-chemistry.org/prog/peo/ (accessed on 20 January 2023).
  15. Dallakyan, S.; Olson, A.J. Small-molecule library screening by docking with PyRx. Methods Mol. Biol. 2015, 1263, 243–250. [Google Scholar] [PubMed]
  16. 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]
  17. Hurst, D.T.; Stacey, A.D.; Nethercleft, M.; Rahim, A.; Harnden, M.R. The synthesis of some pyrimidinyl and thiazolyl ureas and thioureas and some related compounds. Aust. J. Chem. 1988, 41, 1221–1229. [Google Scholar] [CrossRef]
  18. Ravn, J.; Ankersen, M.; Begtrup, M.; Lau, J.F. A novel solid-phase synthesis of di- and trisubstituted N-acyl ureas. Tetrahedron Lett. 2003, 44, 6931–6935. [Google Scholar] [CrossRef]
  19. Verbrugge, P.A.; De Waal, J.J. Process for Preparing Carbamates and Intermediates Therein. European Patent EP0387946A1, 19 September 1990. [Google Scholar]
  20. Voronkov, M.G.; Vlasova, N.N.; Grigor’eva, O.Y.; Belousova, L.I.; Vlasov, A.V. Acyl Iodides in Organic Synthesis. Reactions of Acetyl Iodide with Urea, Thiourea, and Their N,N′-Disubstituted Derivatives. Russ. J. Org. Chem. 2009, 45, 486–490. [Google Scholar] [CrossRef]
  21. Singh, A.K.; Chawla, R.; Yadav, L.D.S. In situ slow release of isocyanates: Synthesis and organocatalytic application of N-acylureas. Tetrahedron Lett. 2013, 54, 5099–5102. [Google Scholar] [CrossRef]
  22. Sun, C.; Zhang, X.; Huang, H.; Zhou, P. Synthesis and evaluation of a new series of substituted acyl(thio)urea and thiadiazolo [2,3-a] pyrimidine derivatives as potent inhibitors of influenza virus neuraminidase. Bioorg. Med. Chem. 2006, 14, 8574–8581. [Google Scholar] [CrossRef]
  23. Stokes, S.; Martin, N.G. A simple and efficient synthesis of N-benzoyl ureas. Tetrahedron Lett. 2012, 53, 4802–4804. [Google Scholar] [CrossRef]
  24. Hernandez, A.G.; Grooms, G.M.; El-Alfy, A.T.; Stec, J. Convenient one-pot two step synthesis of symmetrical and unsymmetrical diacyl ureas, acyl urea/carbamate/thiocarbamate derivatives, and related compounds. Synthesis 2017, 49, 2163–2176. [Google Scholar]
  25. Wang, H.; Zhai, Z.-W.; Shi, Y.-X.; Tan, C.-X.; Weng, J.-Q.; Han, L.; Li, B.-J.; Liu, X.-H. Novel trifluoromethylpyrazole acyl urea derivatives: Synthesis, crystal structure, fungicidal activity and docking study. J. Mol. Struct. 2018, 1171, 631–638. [Google Scholar] [CrossRef]
  26. Sroor, F.M.; Abdelmoniem, A.M.; Abdelhamid, I.A. Facile Synthesis, Structural Activity Relationship, Molecular Modeling and In vitro Biological Evaluation of New Urea Derivatives with Incorporated Isoxazole and Thiazole Moieties as Anticancer Agents. ChemistrySelect 2019, 4, 10113–10121. [Google Scholar] [CrossRef]
  27. Sroor, F.M.; Othman, A.M.; Tantawy, M.A.; Mahrous, K.F.; El-Naggar, M.E. Synthesis, antimicrobial, anti-cancer and in silico studies of new urea derivatives. Bioorg. Chem. 2021, 112, 104953. [Google Scholar] [CrossRef]
  28. Dovlatyan, V.V.; Eliazyan, K.A.; Kazaryan, E.A.; Pivazyan, V.P. N-pyrimidinyl-N-acylureas and thioureas. Zek. Hay. Gitut’yunneri Azg. Akad. 2004, 104, 119–122. [Google Scholar]
  29. Mohammadpoor-Baltork, I.; Sadeghi, M.M.; Esmayilpour, K. A Convenient and Inexpensive Method for Conversion of Thiocarbonyl Compounds to Their Oxo Derivatives Using Oxone Under Solvent-Free Conditions. Synth. Commun. 2003, 33, 953–959. [Google Scholar] [CrossRef]
  30. Roth, B. National Institute of Mental Health’s Psychoactive Drug Screening Program. In Assay Protocol Book; Contract # HHSN-271-2018-00023-C (NIMH PDSP); National Institute of Mental Health: Bethesda, MD, USA, 2018; p. 359. Available online: https://pdsp.unc.edu/pdspweb/content/PDSP%20Protocols%20II%202013-03-28.pdf (accessed on 28 February 2023).
  31. Veber, D.F.; Johnson, S.R.; Cheng, H.Y.; Smith, B.R.; Ward, K.W.; Kopple, K.D. Molecular properties that influence the oral bioavailability of drug candidates. J. Med. Chem. 2002, 45, 2615–2623. [Google Scholar] [CrossRef]
  32. Biovia, D.S.; Berman, H.M.; Westbrook, J.; Feng, Z.; Gilliland, G.; Bhat, T.N.; Richmond, T.J. Dassault systèmes BIOVIA, discovery studio visualizer, v. 17.2, San Diego: Dassault Systèmes, 2016. J. Chem. Phys. 2000, 10, 21–9991. [Google Scholar]
  33. Akunne, H.C.; Whetzel, S.Z.; Wiley, J.N.; Corbin, A.E.; Ninteman, F.W.; Tecle, H.; Pei, Y.; Pugsley, T.A.; Heffner, T.G. The pharmacology of the novel and selective sigma ligand, PD 144418. Neuropharmacology 1997, 36, 51–62. [Google Scholar] [CrossRef]
  34. Ropp, P.J.; Kaminsky, J.C.; Yablonski, S.; Durrant, J.D. Dimorphite-DL: An open-source program for enumerating the ionization states of drug-like small molecules. J. Cheminform. 2019, 11, 14. [Google Scholar] [CrossRef]
  35. O’Boyle, N.M.; Banck, M.; James, C.A.; Morley, C.; Vandermeersch, T.; Hutchinson, G.J. Open Babel: An open chemical toolbox. J. Cheminform. 2011, 3, 33. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Acyl urea scaffold with alkyl (R), aromatic/heteroaromatic features (Ar1 and Ar2).
Figure 1. Acyl urea scaffold with alkyl (R), aromatic/heteroaromatic features (Ar1 and Ar2).
Molecules 28 02319 g001
Scheme 1. Classical synthesis of acyl ureas from benzamides.
Scheme 1. Classical synthesis of acyl ureas from benzamides.
Molecules 28 02319 sch001
Scheme 2. Two-step synthesis of acyl ureas from isocynates and phenol via the N-(phenoxycarbonyl)benzamide intermediate.
Scheme 2. Two-step synthesis of acyl ureas from isocynates and phenol via the N-(phenoxycarbonyl)benzamide intermediate.
Molecules 28 02319 sch002
Figure 2. Schematic structure of the (A) X-ray crystallographic structure of σ1R co-crystallized with PD14418 (PDB:5HK1) (B) Discovery Studio prepared 5HK1.
Figure 2. Schematic structure of the (A) X-ray crystallographic structure of σ1R co-crystallized with PD14418 (PDB:5HK1) (B) Discovery Studio prepared 5HK1.
Molecules 28 02319 g002
Figure 3. 2D Interactions of representative compound 11 and PD14418 with 5HK1. Legend: (Dark green, Conventional H-bonding, Light green (π-donor hydrogen), Purple (π-sigma), Pink (π-alkyl), and Orange (π-anion) polar and non-polar interactions.
Figure 3. 2D Interactions of representative compound 11 and PD14418 with 5HK1. Legend: (Dark green, Conventional H-bonding, Light green (π-donor hydrogen), Purple (π-sigma), Pink (π-alkyl), and Orange (π-anion) polar and non-polar interactions.
Molecules 28 02319 g003
Table 1. Synthesized acyl ureas plus starting amines/amide/thioamides. a Characterized by 1H, 13C NMR, HRMS. b Yields of pure isolated products. ** Known compounds.
Table 1. Synthesized acyl ureas plus starting amines/amide/thioamides. a Characterized by 1H, 13C NMR, HRMS. b Yields of pure isolated products. ** Known compounds.
Compds.AmineAcyl Urea a% Yield bMP (°C)
1Molecules 28 02319 i001** [25,26]
Molecules 28 02319 i002
76212–215
(Not reported)
2Molecules 28 02319 i003Molecules 28 02319 i00494220–223
3Molecules 28 02319 i005Molecules 28 02319 i00675180–182
4Molecules 28 02319 i007Molecules 28 02319 i00871209–211
5Molecules 28 02319 i009Molecules 28 02319 i01070145–147
6Molecules 28 02319 i011Molecules 28 02319 i01292161–163
7Molecules 28 02319 i013Molecules 28 02319 i01491170–172
8Molecules 28 02319 i015Molecules 28 02319 i01672158–160
9Molecules 28 02319 i017Molecules 28 02319 i01871164–166
10Molecules 28 02319 i019Molecules 28 02319 i02083180–182
11Molecules 28 02319 i021Molecules 28 02319 i02272155–157
12Molecules 28 02319 i023Molecules 28 02319 i02464176–177
13Molecules 28 02319 i025** [27]
Molecules 28 02319 i026
68183–185
(Reported 184–186)
14Molecules 28 02319 i027Molecules 28 02319 i02894233–236
15Molecules 28 02319 i029Molecules 28 02319 i03094237–238
16Molecules 28 02319 i031Molecules 28 02319 i03265162–165
Table 2. Some Molinspiration/OSIRIS predicted pharmacokinetic properties. Calculated partition coefficients (cLogP values) were Octanol/H2O based.
Table 2. Some Molinspiration/OSIRIS predicted pharmacokinetic properties. Calculated partition coefficients (cLogP values) were Octanol/H2O based.
EntryMWcLogPTPSAONOHNHNRBDrug-Likeness
1231.21.7584.236221.02
2279.32.6373.99532−0.05
3285.32.14103.15231.75
4293.33.0373.995320.59
5311.31.5378.437128.16
6310.31.7265.546126.83
7249.20.3272.896215.88
8323.43.0752.655120.81
9332.42.6478.95614−27.5
10281.22.9784.236222.54
11396.43.6087.667321.80
12330.32.67119.76342.42
13270.22.1983.986221.62
14291.32.9171.095221.66
15291.33.1471.095221.66
16
Ro5
307.3
≤500
3.53
≤5
86.114
≤10
2
≤5
40.73
Table 3. PyRx generated binding affinity values, and the number of H-bonds and residue(s) involved in H-bonding.
Table 3. PyRx generated binding affinity values, and the number of H-bonds and residue(s) involved in H-bonding.
EntriesBinding
Affinity
#H-Bond
Interactions
Residues Involved in 5HK1
H-Bond Interactions
1−10.02TYR 120, THR 181
2−9.61GLU 172
3−9.61GLU 172
4−10.41GLU 172
5−10.01TYR 120
6−9.61GLU 172
7−8.61GLU 172
8−9.71GLU 172
9−9.01GLU 172
10−10.63THR 181, TYR 108, GLU 172
11−9.53TYR 120, THR 181, ASP 126
12−9.92ASP 126, GLU 172
13−10.33GLU 213, SER 99, SER 33
14−10.61GLU 172
15−11.40-
16−10.52GLU 172, TYR 103
PD-14418−10.20-
Table 4. Modeled (kcal/mol) versus in vitro (μM) binding affinity values. * PD144418 is a selective and high affinity (Ki = 0.08 nM) σ1R ligand [33].
Table 4. Modeled (kcal/mol) versus in vitro (μM) binding affinity values. * PD144418 is a selective and high affinity (Ki = 0.08 nM) σ1R ligand [33].
EntriesCalculated σ1R Affinity (kcal/mol)In Vitro σ1R Affinity (Ki, μM)
1−10.0>10
2−9.6>10
3−9.6>10
4−10.4>10
5−10.0>10
6−9.6>10
7−8.6>10
8−9.7>10
9−9.0>10
10−10.62.18
11−9.5>10
12−9.99.54
13−10.3>10
14−10.6>10
15−11.4>10
16−10.5>10
PD-14418−10.2*
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

Thapa, R.; Flores, R.; Cheng, K.H.; Mochona, B.; Sikazwe, D. Design and Synthesis of New Acyl Urea Analogs as Potential σ1R Ligands. Molecules 2023, 28, 2319. https://doi.org/10.3390/molecules28052319

AMA Style

Thapa R, Flores R, Cheng KH, Mochona B, Sikazwe D. Design and Synthesis of New Acyl Urea Analogs as Potential σ1R Ligands. Molecules. 2023; 28(5):2319. https://doi.org/10.3390/molecules28052319

Chicago/Turabian Style

Thapa, Rajesh, Rafael Flores, Kwan H. Cheng, Bereket Mochona, and Donald Sikazwe. 2023. "Design and Synthesis of New Acyl Urea Analogs as Potential σ1R Ligands" Molecules 28, no. 5: 2319. https://doi.org/10.3390/molecules28052319

Article Metrics

Back to TopTop