Next Article in Journal / Special Issue
Nucleic Acids Analytical Methods for Viral Infection Diagnosis: State-of-the-Art and Future Perspectives
Previous Article in Journal
The Essential Role of Stathmin in Myoblast C2C12 for Vertical Vibration-Induced Myotube Formation
Previous Article in Special Issue
SARS-CoV-2 Mpro: A Potential Target for Peptidomimetics and Small-Molecule Inhibitors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Pseudo-Symmetric N-benzyl Hydroxyethylamine Core in a New Series of Heteroarylcarboxyamide HIV-1 Pr Inhibitors: Synthesis, Molecular Modeling and Biological Evaluation

1
Dipartimento di Scienze, Università della Basilicata, Via Ateneo Lucano 10, 85100 Potenza, Italy
2
Dipartimento di Scienze Chimiche e Farmaceutiche, Università di Trieste, Via Licio Giorgieri 1, 34127 Trieste, Italy
*
Authors to whom correspondence should be addressed.
Biomolecules 2021, 11(11), 1584; https://doi.org/10.3390/biom11111584
Submission received: 20 September 2021 / Revised: 21 October 2021 / Accepted: 23 October 2021 / Published: 26 October 2021

Abstract

:
Here, we report the synthesis, enzyme inhibition and structure–activity relationship studies of a new potent class of HIV-1 protease inhibitors, which contain a pseudo-symmetric hydroxyethylamine core and heteroarylcarboxyamide moieties. The simple synthetic pathway furnished nine compounds in a few steps with high yields. The compounds were designed taking into account our previous results on other series of inhibitors with different substituents at P’ and P’’ and different ways of linking them to the inhibitor core. Potent inhibitory activity was obtained with nanomolar IC50 values measured with a standard fluorimetric test in 100 mM MES buffer, pH 5.5, containing 400 mM NaCl, 1 mM EDTA, 1 mM DTT and 1 mg/ml BSA. Compounds 9ac, containing the indole ring in P1, exhibited an HIV-1 protease inhibitory activity more powerful than darunavir in the same assay. To obtain molecular insight into the binding properties of these compounds, docking analysis was performed, and their binding properties were also compared.

1. Introduction

The AIDS epidemic is still one of the most challenging problems [1], although great efforts are made for the discovery of new drugs for its treatment. Among many strategies to combat the disease, antiretroviral therapy (ART) containing at least one HIV-1 protease inhibitor (PIs) is considered the most effective treatment [2,3,4,5]. When a protease inhibitor binds to the active site, it prevents the cleavage of nascent viral proteins, thereby halting viral replication [6]. The synthesis of compounds able to block the action of the HIV protease, the enzyme which plays a key role in maintaining infectivity, is currently a huge aim.
Nowadays, nine FDA-approved PIs are available on the market, but due to the rapid genomic evolution of HIV, an inevitable consequence in the treatment of the infection has been the rise of drug resistance, and therefore, the dramatic reduction of the marketed inhibitors efficiency [7,8].
Thus, the emergence of highly mutated viral strains cross-resistant to antivirals, the occurrence of various side effects and the high cost of ART prompted scientists to seek novel PIs, preferably with alternative frameworks.
Notably, the introduction of heterocyclic moieties in a bioactive molecule can have important effects on physicochemical and pharmacological properties [9]. This strategy has been widely adopted in medicinal chemistry for the design of new drugs, because of their chemical stability and structural rigidity, less entropic energy was lost upon binding. In our experience of the synthesis of highly functionalized small molecules with aryl and heteroaryl structures [10,11,12], we evidenced the crucial effect of the presence of heterocyclic moiety in PIs structure, either in the core of the inhibitors [13,14,15] or in the P2 position [16,17,18].
Inspired by the success of darunavir, an FDA-approved drug for the treatment of HIV, and its analog TMC-126, containing a bis-tetrahydrofuran heterocyclic system as P2 ligand (Figure 1), our aim was focused on designing inhibitors containing heteroaryl moieties that specifically target and maximize interactions with the backbone. Both extensive hydrogen bonding and hydrophobic interactions with enzyme subsites can limit the protease’s ability to acquire drug resistance as the geometry of the catalytic site must be conserved to maintain functionality [19,20,21].
Recently the preparation and the activity, in vitro and in mammalian cells, of new HIV protease inhibitors, compounds 1 and 2, were reported by our group (Figure 1) [22,23].
They were designed having heterocycle as the P2 ligand linked by a carboxyamidic or carbammic moiety to the core, with or without the benzyl group, and a 3,4-dimethoxyphenylsulfonyl-N-isobutylamide [22] or a 4-methoxyphenylsulfonyl-N-isobutylamide [23] as the P2’ ligand. Compounds with benzyl in the core showed in vitro activity against native protease with IC50 values in the range of <0.6–13 nM.
As HIV protease has been shown to exist as a C2-symmetric homodimer in its active form, several dipeptide isosteres, such as diaminoalcohol, diaminodiol and hydroxyethylhydrazine, have also been employed in the development of pseudo-symmetric inhibitors (that is, inhibitors that are lacking the same C2 symmetry of the enzyme, but bear the same group at P1 and P1’) [24] (Figure 2).
In order to obtain a pseudo-symmetric hydroxyethylamine core, the isobutyl portion present in the structure of compounds 1 and 2 (Figure 1) was replaced with a benzyl group. Therefore, a library of compounds containing different heteroarenes and sulfonamide portions was prepared. The general structure of these newly synthesized compounds is reported as A in Figure 2. The structure takes into account the outcomes of our previous evaluations and the indications obtained in the models developed and described in refs [13,14,16,17,18,22,23].
The effects on the inhibitory activity of the heteroatom (S, O, N) in the heteroarylcarboxamidic portion and the electronic properties of the substituents on the sulfonamidic moiety were also evaluated. The novel inhibitors were tested in vitro on HIV-1 protease with a standard fluorometric assay, and their activities were compared with that of darunavir.

2. Materials and Methods

2.1. Chemistry

Preparative chromatography was carried out on Merck silica gel (0.063–0.200 mm particle size) by progressive elution with opportune solvent mixtures. 1H and 13C NMR spectra were normally carried out in CDCl3 solutions on a VARIAN INOVA 500 MHz or Bruker 400 MHz and referenced to CDCl3. Mass spectra were obtained with a Hewlett-Packard 5971 mass-selective detector on a Hewlett–Packard 5890 gas chromatograph ((OV-1 capillary column between 70 and 250 °C (20 °C min−1)). The optical purity was evaluated by using a polarimeter JASCO Mod Dip-370. CH2Cl2 was dried by distillation over anhydrous CaCl2 in an inert atmosphere. Dry THF and DMF were commercially available.
tert-butyl (2S,3R)-4-(benzylamino)-3-hydroxy-1-phenylbutan-2-ylcarbamate (4). Compound 4 was prepared from a solution of (2S,3S)-1,2-epoxy-3-(Boc-amino)-4-phenylbutane (1.5 mmol) and benzylamine (1.5 mmol) in i-PrOH (10 mL) that was stirred under reflux for 16 h. The reaction mixture was rotary evaporated, and the crude product was purified by recrystallization in methanol/water (7:3). Compound 4 was obtained as a white solid, yield 88%. 1H and 13C NMR spectra were consistent with the data in the literature [25].
General procedure for the preparation of tert-butyl ((2S,3R)-n-[4-(n-benzyl-4-R-phenylsulfonamido)-3-hydroxy-1-phenylbutan-2-yl)carbamates (5). To a stirred solution of aminoalcohol 4 (0.78 mmol) in anhydrous CH2Cl2 (20 mL), Et3N (2.02 mmol) and arylsulfonyl chloride (0.93 mmol) were added at room temperature and under an Ar atmosphere. After 24 h, the reaction was quenched with 5% aqueous H2SO4 solution. The organic layer was washed, adding saturated aqueous NaHCO3 solution and brine. The organic phases collected were dried over Na2SO4, filtered and concentrated in vacuo. The crude product was purified by column chromatography on silica gel.
(1S,2R)-{1-Benzyl-2-hydroxy-3-[n-benzyl-(4-nitrobenzenesulfonyl)-amino]-propyl}-carbamic acid tert-butyl ester (5a). Compound 5a was isolated as a white solid (CH2Cl2/EtOAc 98/2), yield 94%. [ α ]   D 20 = +6.3° (c: 1.0, CHCl3). 1H NMR (400 MHz, DMSO-d6): δ 8.37 (d, J = 9.0 Hz, 2H), 8.10 (d, J = 9.0, 2H), 7.33 (m, 5H), 7.21 (m, 2H), 7.13 (m, 3H), 6.63 (d, J = 9.0 Hz, 1H), 5.00 (d, J = 6.5, 1H), 4.68 (d, J = 15.5 Hz, 1H), 4.44 (d, J = 15.5 Hz, 1H), 3.46 (m, 2H), 3.35 (m, 1H), 3.10 (dd, J1 = 15.0 Hz, J2 = 9.0 Hz, 1H), 2.89 (dd, J1 = 13.7 Hz, J2 = 3.0 Hz, 1H), 2,42 (dd, J1 = 13.5 Hz, J2 = 11.0 Hz, 1H), 1.20 (s, 9H). 13C{1H} NMR (100 MHz, DMSO-d6): δ 155.2, 149.5, 146.1, 139.3, 136.1, 129.0, 128.5, 128.4, 128.1, 127.8, 127.5, 125.6, 124.3, 77.5, 71.2, 54.9, 50.9, 50.2, 35.2, 28.1. Anal. Calcd for C28H33N3O7S: C, 60.52; H, 5.99; N, 7.56; S, 5.77. Found: C, 60.54; H, 5.93; N, 7.50; S,5.75.
(1S,2R)-{1-Benzyl-2-hydroxy-3-[n-benzyl-(4-methoxybenzenesulfonyl)-amino]-propyl}-carbamic acid tert-butyl ester (5b). Compound 5b was obtained as a white solid (CH2Cl2/EtOAc 95/5), yield 85 %. [ α ]   D 20 = +5.4° (c: 1.0, CHCl3). 1H NMR (400 MHz, DMSO-d6): δ 7.78 (d, J = 8.5 Hz, 2H), 7.28 (m, 8H), 7.16–7.08 (m, 4H), 6.60 (d, J = 8.8 Hz, 1H), 4.94 (d, J = 6.0 Hz, 1H), 4.50 (d, J = 15.6 Hz, 1H), 4.37 (d, J = 15.6, 1H), 3.85 (s, 3H), 3.48 (m, 2H), 3.35 (m, 1H), 2.93 (m, 2H), 2.45 (dd, J = 13.8 Hz, J = 10.4 Hz, 1H), 1.22 (s, 9H). 13C{1H} NMR (100 MHz, DMSO-d6): δ 155.2, 139.5, 136.6, 131.6, 129.2, 129.1, 128.2, 128.0, 127.8, 127.2, 125.6, 114.3, 77.4, 72.0, 54.9, 55.6, 54.8, 51.4, 50.7, 35.2, 28.1. Anal. Calcd for C29H36N2O6S: C, 64.42; H, 6.71; N, 5.18; S, 5.93. Found: C, 64.49; H, 6.76; N, 5.23; S, 5.88.
(1S,2R)-{1-Benzyl-2-hydroxy-3-[n-Benzyl-(3,4-dimethoxybenzenesulfonyl)-amino]-propyl}-carbamic acid tert-butyl ester (5c). Compound 5c was obtained as a white solid (CH2Cl2/EtOAc 95/5), yield 90 %. [ α ]   D 20 = +6.7° (c: 1.0, CHCl3). 1H NMR (400 MHz, DMSO-d6): δ 7.45 (dd, J = 2.0 Hz, J = 6.8 MHz, 1H), 7.54 (m, 10H), 6.94 (d, J = 8 Hz, 1H), 4.45 (d, J = 14, 1H), 4.16 (d, J = 27.2 Hz, 1H), 3.95 (s, 3H), 3.89 (s, 3H), 3.64 (d, 1H), 3.34 (d, 1H), 3.18 (d, 1H), 2.8 (m, 2H), 1.32 (s, 9H). 13C{1H} NMR (100 MHz, DMSO-d6): δ 155.2, 149.5, 146.1, 139.3, 136.1, 129.0, 128.5, 128.4, 128.1, 127.8, 127.5, 125.6, 124.3, 77.5, 71.2, 54.9, 50.9, 50.2, 35.2, 28.1. Anal. Calcd for C30H38N2O7S: C, 63.14; H, 6.71; N, 4.91; S, 5.62. Found: C, 63.17; H, 4.97; N, 4.87; S, 5.69.
General procedure for the preparation of n-((2R,3S)-3-amino-2-hydroxy-4-phenylbutyl)-n-benzyl-R-benzenesulfonamide (6). To a stirred solution of 5ac (0.78 mmol) in anhydrous CH2Cl2 (30 mL), trifluoroacetic acid (13 mL) was added at room temperature. After 1 hour, the reaction mixture was concentrated and treated with toluene (3 × 20 mL), evaporated in vacuum. The crude product was treated with Et3N and purified by chromatography on silica gel (CHCl3/AcOEt 9/1).
N-((2R,3S)-3-Amino-2-hydroxy-4-phenylbutyl)-n-benzyl-4-nitrobenzenesulfonamide (6a). Compound 6a was obtained as a white solid (CH2Cl2/EtOAc 9/1), yield 41%. [ α ]   D 20 = +6.4° (c: 1.0, CHCl3). 1H NMR (400 MHz, DMSO-d6): δ 8.34 (d, J = 8.6 Hz, 2H), 8,11 (bs, 2H,), 8.06 (d, J = 8.6 Hz, 2H), 7.26 (m, 10H), 5.67 (d, J = 5.6, 1H), 4.56 (d, J = 16.0 Hz, 1H), 4.49 (d, J = 16.0 Hz, 1H), 3.96 (bs, 1H), 3.38 (m, 2H), 3.16 (dd, J = 14.8 Hz, J = 8.8 Hz, 1H), 2.87 (dd, J = 14.4 Hz, J = 7.2 Hz, 1H), 2.82 (dd, J = 14.2 Hz, J = 7.6 Hz, 1H). 13C{1H} NMR (100 MHz, DMSO-d6): δ 149.6, 145.3, 136.4, 135.9, 129.3, 128.6, 128.4, 128.1, 127.6, 126.8, 124.4, 67.8, 55.2, 51.5, 49.1, 32.8. Anal. Calcd for C23H25N3O5S: C, 60.64; H, 5.53; N, 9.22; S, 7.04. Found: C, 60.68; H, 5.59; N, 9.30; S, 7.10.
N-((2R,3S)-3-Amino-2-hydroxy-4-phenyl-butyl)-n-benzyl-4-methoxybenzenesulfonamide (6b). Compound 6b was obtained as a white solid (CH2Cl2/EtOAc 9/1), yield 35%. [ α ]   D 20 = +7.5° (c: 1.0, CHCl3). 1H NMR (400 MHz, DMSO-d6): δ 7.75 (d, J = 8.4 Hz, 2H), 7.20 (m, 10H), 6.99 (d, J = 8.8 Hz, 2H), 4.33 (d, J = 14.4 Hz, 1H), 4.17 (d, J = 14.4Hz, 1H), 4.00 (m, 1H), 3.88 (s, 3H), 3.50 (m, 2H), 3.45 (m, 2H), 2.79 (m, 2H). 13C{1H} NMR (100 MHz, DMSO-d6): δ 136.5, 130.8, 129.3, 128.5, 128.3, 128.0, 127.3, 126.7, 114.4, 68.3, 56.0, 55.7, 51.8, 49.5, 32.5. Anal. Calcd for C24H28N2O4S: C, 65.43; H, 6.41; N, 6.36; S, 7.28. Found: C, 65.45; H, 6.46; N, 6.40; S, 7.25.
N-((2R,3S)-3-Amino-2-hydroxy-4-phenyl-butyl)-n-benzyl-3,4-methoxybenzenesulfonamide (6c). Compound 6c was obtained as a white solid (CH2Cl2/EtOAc 9/1), yield 52%. [ α ]   D 20 = +8.3° (c: 1.0, CHCl3). 1H NMR (400 MHz, DMSO-d6): δ 7.41 (dd, J = 8 Hz, J = 2.1 Hz, 1H), 7.15 (m, 10H), 6.9 (d, J =8.4 Hz, 2H), 4.35 (d, J = 14.8 Hz, 2H), 4.18 (d, J = 14.8 Hz, 2H), 3.98 (s, 1H), 3.92 (s, 3H), 3.82 (s, 3H), 3.56 (s, 1H), 3.25 (m, 2H), 2.82 (m, 2H). 13C{1H} NMR (100 MHz, DMSO-d6): δ 135.9, 130.5, 128.7, 128.2, 127.9, 127.6, 126.9, 126.5, 113.5, 67.9, 56.2, 55.8, 55.3, 51.1, 48.9, 31.8. Anal. Calcd for C25H30N2O5S: C, 63.81; H, 6.43; N, 5.95; S, 6.81. Found: C, 63.85; H, 6.41; N, 5.89; S, 6.75.
General procedure for the preparation of n-((2S,3R)-3-hydroxy-4-(n-benzyl-arylsulfonamido)-1-phenylbutan-2-yl)heteroarene-5-carboxamide (7, 8, 9). To a solution of 5-heterobenzoic acid (0.13 mmol), EDCI (0.20mmol), HOBt (0.20 mmol) in anhydrous CH2Cl2, a solution of amine 6ac (0.13 mmol) and diisopropylethylamine (0.78 mmol) in anhydrous CH2Cl2 was added at 0 °C under an argon atmosphere and it was stirred for 16 h at room temperature. The reaction mixture was quenched with water and extracted with CH2Cl2. The organic layers were dried on Na2SO4 and filtered and concentrated under reduced pressure. The residue was purified by silica gel column chromatography (CH2Cl2/AcOEt 9/1) to furnish inhibitors 7ac, 8ac, 9ac.
N-((2S,3R)-3-hydroxy-4-(n-benzyl-4-nitrophenylsulfonamido)-1-phenylbutan-2-yl)benzo[b]thiophene-5-carboxamide (7a). Following the general procedure, compound 7a was obtained as a white solid, yield 50%. [ α ]   D 20 = +14.5° (c: 1, CHCl3). 1H NMR (400 MHz, CDCl3): δ 8.29 (d, J = 8.8 Hz, 1H), 7.95 (m, 4H), 7.54 (d, J = 6.4 Hz, 1H), 7.44 (d, J = 8.4 Hz, 1H), 7.75 (m, 12H), 6.07 (d, J = 8.0 Hz, 1H), 4.43 (m, 2H), 4.20 (m, 1H), 3.66 (m, 1H), 3.36 (m, 2H), 3.98 (m, 2H). 13C{1H} NMR (100 MHz, CDCl3): δ 168.6, 150.0, 144.9, 143.0, 139.4, 137.1, 135.0, 129.8, 129.3, 129.0, 128.8, 128.7, 128.4, 123.3, 126.8, 124.3, 124.1, 122.7, 122.1, 71.7, 54.9, 53.6, 51.5, 35.3. Calcd for C32H29N3O6S2: C, 62.42; H, 4.75; N, 6.82; S, 10.42. Found: C, 62.49; H, 4.81; N, 6.74; S, 10.35.
N-((2S,3R)-3-hydroxy-4-(n-benzyl-4-methoxyphenylsulfonamido)-1-phenylbutan-2-yl)benzo[b]thiophene-5-carboxamide (7b). Following the general procedure, compound 7b was obtained as a white solid, yield 57%. [ α ]   D 20 = +1.5° (c: 0.32, CHCl3). 1H NMR (400 MHz, CDCl3): δ 7.98 (s, 1H), 7.88 (d, J = 8.4 Hz, 1H), 7.71 (d, J = 9.2 Hz, 2H), 7.53 (d, J = 5.6 Hz, 1H), 7.46 (d, J = 8.4 Hz, 1H), 7.37 (d, J = 5.2 Hz, 1H), 7.20 (m, 10H), 6.94 (d, J = 9.2 Hz, 2H), 6.03 (d, J = 8.0 Hz, 1H), 4.25 (m, 3H), 3.86 (s, 3H), 3.50 (m, 1H), 3.34 (m, 1H), 3.08 (m, 2H), 3.98 (m, 1H). 13C{1H} NMR (100 MHz, CDCl3): δ 168.1, 163.1, 142.8, 139.3, 137.5, 135.9, 130.1, 129.8, 129.4, 129.4, 128.8, 128.7, 128.5, 128.1, 127.9, 126.6, 124.1, 121.6, 122.5, 122.2, 114.4, 72.0, 55.6, 54.5, 54.2, 52.5, 35.2. Calcd for C33H32N2O5S2: C, 65.98; H, 5.37; N, 4.66; S, 10.68. Found: C, 65.95; H, 5.41; N, 4.62; S, 10.72.
N-((2S,3R)-3-hydroxy-4-(n-benzyl-3,4-dimethoxyphenylsulfonamido)-1-phenylbutan-2-yl)benzo[b]thiophene-5-carboxamide (7c). Following the general procedure, compound 7c was obtained as a white solid, yield 55%. [ α ]   D 20 = +4.6° (c: 0.1, CHCl3). 1H NMR (400 MHz, CDCl3): δ 7.99 (d, J = 1.6 Hz, 1H), 7.87 (d, J = 8.8 Hz, 1H), 7.45 (m, 15H), 6.88 (d, J = 8.4 Hz, 1H), 6.09 (d, J = 8.4 Hz, 1H), 4.13 (m, 3H), 3.98 (s, 3H), 3.79 (s, 3H), 3.58 (m, 1H), 3.38 (m, 1H), 3.06 (m, 2H), 2.98 (m, 1H). 13C{1H} NMR (100 MHz, CDCl3): δ 168.1, 152.8, 149.2, 142.8, 139.3, 137.5, 135.9, 130.0, 129.4, 128.8, 128.7, 128.6, 128.1, 128.0, 126.6, 124.2, 122.6, 122.2, 121.2, 110.7, 109.6, 72.0, 56.2, 56.1, 54.4, 54.3, 52.3, 35.2. Anal. Calcd for C34H34N2O6S2: C, 64.74; H, 5.43; N, 4.44; S, 10.17. Found: C, 64.68; H, 5.41; N, 4.40; S, 10.11.
N-((2S,3R)-3-hydroxy-4-(n-benzyl-4-nitrophenylsulfonamido)-1-phenylbutan-2-yl)benzofuran-5-carboxamide (8a). Following the general procedure, compound 8a was obtained as a white solid, yield 53%. [ α ]   D 20 = +3.1° (c: 0.22, CHCl3). 1H NMR (400 MHz, CDCl3): δ 8.18 (d, J = 8.8 Hz, 2H), 7.98 (d, J = 8.8 Hz, 2H), 8.80 (s, 1H), 8.70 (s, 1H), 7.45 (m, 12H), 6.82 (s, 1H), 6.03 (d, J = 8.0 Hz, 1H), 4.25 (m, 2H), 4.16 (m, 1H), 3.67 (m, 1H), 3.17 (m, 2H), 3.01 (m, 2H).13C{1H} NMR (100 MHz, CDCl3): δ 168.6, 156.7, 150.0, 146.5, 137.1, 135.0, 129.2, 128.8, 128.7, 128.4, 128.3, 127.6, 126.8, 124.3, 123.1, 120.7, 111.5, 71.7, 55.0, 53.5, 51.5, 35.3. Calcd for C32H29N3O7S: C, 64.09; H, 4.87; N, 7.01; S, 5.35. Found: C, 64.12; H, 4.81; N, 7.10; S, 5.39.
N-((2S,3R)-3-hydroxy-4-(n-benzyl-4-methoxyphenylsulfonamido)-1-phenylbutan-2-yl)benzofuran-5-carboxamide (8b). Following the general procedure, compound 8b was obtained as a white solid, yield 54%. [ α ]   D 20 = +3.6° (c: 1, CHCl3). 1H NMR (400 MHz, CDCl3): δ 7.79 (s, 1H), 7.70 (m, 1H), 7.47 (m, 2H), 7.18 (m, 10H); 6.95 (d, J = 8.4 Hz, 2H), 6.81 (m, 1H), 6.00 (d, J = 8 Hz, 1H), 4.25 (m, 3H), 3.87 (s, 3H), 3.6 (m, 1H), 3.55 (m,1H), 3.35 (m, 1H), 3.18 (m, 2H), 2.98 (m, 1H). 13C{1H} NMR (100 MHz, CDCl3): δ 168.2, 163.1, 156.6, 146.4, 137.5, 135.9, 129.4, 129.4, 129.3, 128.8, 128.6, 128.5, 128.0, 127.5, 126.5, 123.2, 120.6, 114.6, 114.4, 111.4, 106.9, 72.6, 55.6, 54.5, 54.2, 52.5, 35.2, 31.9. Calcd for C33H32N2O6S: C, 67.79; H, 5.52; N, 4.79; S, 5.48. Found: C, 67.75; H, 5.60; N, 4.73; S, 5.54.
N-((2S,3R)-3-hydroxy-4-(n-benzyl-3,4-dimethoxyphenylsulfonamido)-1-phenylbutan-2-yl)benzofuran-5-carboxamide (8c). Following the general procedure, compound 8c was obtained as a white solid, yield 56%. [ α ]   D 20 = +19.5° (c: 1, CHCl3). 1H NMR (400 MHz, CDCl3): δ 7.80 (s, 1H), 7.6 (d, J = 2.0 Hz, 1H), 7.43 (m, 3H), 7.20 (m, 12H), 6.89 (d, J = 8.4 Hz, 1H), 6.8 (d, J = 2.1 Hz, 1H), 6.1 (d, J = 7.6Hz, 1H), 3.93 (s, 3H), 3.80 (s, 3H), 3.52 (m, 1H), 3.37 (m, 1H), 3.06 (m, 2H), 2.92 (m, 1H). 13C{1H} NMR (100 MHz, CDCl3): δ 168.2, 156.6, 152.8, 130.1, 129.4, 128.9, 128.8, 128.7, 128.5, 128.1, 127.5, 126.6, 123.2, 121.1, 120.7, 111.6, 110.7, 109.6, 106.9, 72.0, 56.2, 56.1, 54.3, 52.4, 35.2, 30.9. Calcd for C34H34N2O7S: C, 66.43; H, 5.57; N, 4.56; S, 5.22. Found: C, 66.45; H, 5.61; N, 4.50; S, 5.25.
N-((2S,3R)-3-hydroxy-4-(n-benzyl-4-nitrophenylsulfonamido)-1-phenylbutan-2-yl)1H-indol-5-carboxamide (9a). Following the general procedure, compound 9a was obtained as a white solid, yield 33%. [ α ]   D 20 = +25.8° (c: 1.3, CHCl3). 1H NMR (400 MHz, CDCl3): δ 8.64 (s, 1H), 7.84 (s, 1H), 7.72 (d, J = 8.8 Hz, 2H), 7.40-7.15 (m, 14H), 6.93 (d, J = 8.8 Hz, 2H), 6.59 (s, 1H), 6.03 (d, J =6 Hz, 1H), 4.25 (m, 2H), 3.84 (s, 3H), 3.50 (m, 1H), 3.35 (dd, J = 15 Hz, J = 4.4 Hz, 1H), 3.17 (m, 2H), 2.90 (m, 2H). 13C{1H} NMR (100 MHz, CDCl3): δ 169.2, 163.0, 137.7, 136.6, 136.0, 129.5, 129.4, 128.8, 128.7, 128.5, 128.0, 127.4, 126.5, 125.7, 120.8, 120.4, 114.4, 111.0, 103.6, 72.1, 55.6, 54.5, 54.2, 52.5, 35.3. Calcd for C32H30N4O6S: C, 64.20; H, 5.05; N, 9.36; S, 5.36. Found: C, 64.25; H, 4.81; N, 9.30; S, 5.39.
N-((2S,3R)-3-hydroxy-4-(n-benzyl-4-methoxyphenylsulfonamido)-1-phenylbutan-2-yl)1H-indol-5-carboxamide (9b). Following the general procedure, compound 9b was obtained as a white solid, yield 43%. [ α ]   D 20 = +10.2° (c: 0.4, CHCl3). 1H NMR (400 MHz, CDCl3): δ 8.47 (s, 1H) 8.24 (d, J = 8.8 Hz, 2H), 7.90(d, J = 8.8 Hz, 2H), 7.81 (s, 1H), 7.37-7.17 (m, 14H), 6.59 (s, 1H), 6.04 (d, J = 7.2 Hz, 1H), 4.52 (d, J = 14.4 Hz, 1H), 4.37 (d, J =14.4 Hz, 1H), 4.18 (m, 2H), 3.81-3.24 (m, 2H), 2.95 (m, 2H). 13C{1H} NMR (100 MHz, CDCl3): 169.7, 149.9, 145.0, 137.7, 137.2, 135.1, 129.3, 128.8, 128.7, 128.6, 128.4, 128.3, 127.5, 126.8, 125.8, 125.3, 124.3, 120.9, 120.4, 111.1, 103.7, 71.8, 55.1, 53.4, 51.3, 35.6. Calcd for C33H33N3O5S: C, 67.90; H, 5.70; N, 7.20; S, 5.49. Found: C, 67.96; H, 5.76; N, 7.22; S, 5.56.
N-((2S,3R)-3-hydroxy-4-(n-benzyl-3,4-dimethoxyphenylsulfonamido)-1-phenylbutan-2-yl)1H-indol-5-carboxamide (9c). Following the general procedure, compound 9c was obtained as a white solid, yield 44%. [ α ]   D 20 = +18.7° (c: 1, CHCl3). 1H NMR (400 MHz, CDCl3): δ 8.76 (bs, 1H), 7.85 (s, 1H), 7.37 (m, 3H), 7.23 (m, 11H), 6.85 (d, J = 8.4Hz, 1H), 6.56 (s, 1H), 6.12 (d, J = 8Hz, 1H), 4.37 (d, J = 14.8Hz, 1H), 3.90 (s, 3H), 3.76 (s, 3H), 3.55 (m, 1H), 3.40 (dd, J = 12Hz, J = 4Hz, 1H), 3.10 (m, 2H), 2.92 (m, 1H). 13C{1H} NMR (100 MHz, CDCl3): δ 169.2, 152.6, 149.1, 137.6, 135.8, 130.1, 129.3, 128.7, 128.6, 128.4, 127.9, 127.7, 127.4, 126.4, 125.7, 125.3, 121.1, 120.7, 120.2, 110.9, 110.6, 109.6, 103.4, 72.0, 60.3, 56.1, 55.9, 54.4, 54.1, 52.3, 35.2, 31.4, 29.5, 22.5, 20.9, 14.0. Calcd for C34H35N3O6S: C, 66.54; H, 5.75; N, 6.85; S, 5.22. Found: C, 66.62; H, 5.81; N, 6.80; S, 5.27.
(2R,3S)-3-amino-1-(benzylamino)-4-phenylbutan-2-ol (10). A solution of (2S,3S)-1,2-epoxy-3-(Boc-amino)-4-phenylbutane 3 (1.6 mmol) and benzylamine (1.5 mmol) in i-PrOH (10 mL) was stirred under reflux for 16 h. The reaction mixture was rotary evaporated, and the crude product was purified by recrystallization in methanol/water (7:3) to afford compound 4 as a white solid. Then product 4 (1 mmol) was dissolved in MeCN (10 ml) and tosic acid monohydrate was added (3 mmol); the resulting mixture was stirred at room temperature for 5 h. The precipitate formed was filtered off and washed with Et2O to give compound 10 as a white solid, 60% yield. 1H NMR (400 MHz, DMSO-d6): δ 9.00 (bs, 1H), 8.89 (bs, 1H), 7.97 (bs, 3H), 7.49 (d, J = 7.6 Hz, 4H), 7.45 (m, 5H), 7.30 (m, 5H), 7.13 (d, J = 7.6 Hz, 4H), 6.11 (bs, 1H), 4.15 (m, 2H), 4.06 (d, J = 10.4 Hz, 1H), 3.53 (m, 1H), 3.13 (m, 1H), 2.86 (m, 3H), 2.29 (s, 6H). 13C{1H} NMR (100 MHz, DMSO-d6): δ 145.3, 137.9, 135.9, 131.3, 130.2, 129.3, 129.1, 128.9, 128.7, 128.1, 127.0, 125.5, 65.7, 54.9, 50.2, 47.3, 33.1, 20.8. Calcd for C17H22N2O: C, 75.52; H, 8.20; N, 10.36. Found: C, 75.59; H, 8.28; N, 10.28.

2.2. In Vitro Activity Test

IC50 values were determined at pH 5.5 using recombinant wild-type HIV-1 PR from Bachem and the fluorogenic substrate Abz-Thr-Ile-Nle-Phe(p-NO2)-Gln-Arg-NH2 (Abz-NF⁄-6; Bachem AG, Bubendorf, CH). Darunavir was used in this assay as a reference inhibitor for titration of the active enzyme. The assay was performed as follows: Dilution Buffer: 100 mM MES buffer, pH 5.5, containing 400 mM NaCl, 1 mM EDTA, 1 mM DTT and 1 mg/mL BSA. Solution A: 10 μL of a substrate stock solution in DMSO (10 mg/ml, 10.6 mM) were diluted in 1.99 ml of dilution buffer to a final concentration of 53mM. Solution B: 10 μL of a protease stock solution (0.4 mg/mL) in 10 mM sodium phosphate buffer, pH 6.5, containing 1 mM EDTA, 10% glycerol, 0.05% mercaptoethanol, 50 mM NaCl, were diluted 100 times with the dilution buffer, pH 5.5, to a final concentration of 0.004 mg/ml.
Assay: 114 μL of solution A, 11 μL of solution B and 75 μL of the dilution buffer were pre-incubated in a cuvette at 25 °C, and the fluorescence was recorded at 325 nm excitation and 420 nm emission for 10 min. A total of 2 μL of the inhibitor in DMSO was then added, and the fluorescence was recorded for a further 10 min. Final concentrations in the assay were 1.2 nM protease, 30 μM substrate and 0.1–10 μM inhibitor. IC50 was obtained by measuring the relative residual enzyme activity (ratio of the increase of fluorescence velocities before and after the addition of inhibitor) and by fitting the residual activity vs. inhibitor concentration semilog plots to a tetraparametric logistic function (Sigma plot 2001, SPSS Inc., Chicago, IL, USA). All the measures were triplicated.

2.3. Molecular Modeling

A crystallographic structure of the wt-HIV-Pr complex with darunavir (PDB id. 4LL3) was used as the starting geometry of the model complexes. The structure was prepared by adding hydrogen atoms, removing water crystallization molecules but keeping the essential one inside the catalytic site and always choosing the most symmetrical option for amino acid side chains, allowing more solutions. The structure was then optimized with the Amber* force field as implemented in the Schrödinger suite [26,27]. After docking, the complexes were thermalized by a MD run carried out with Yasara (NTV, 300 °K, 500 ps) and finally optimized as previously described. The models of the heterocycle–methane complexes were obtained at the MO62X/6-311++G(d,p) level with Gaussian 09 [28].

3. Results and Discussion

3.1. Chemistry

The preparation of aromatic sulfonamides (general structure A, Figure 2) started from homochiral N-Boc-protected amino epoxide 3, keeping the established stereochemistry during the synthesis [29,30]. The epoxide was firstly opened with benzylamine to afford the monoprotected diaminoalcohol 4. Then, the substituted benzenesulfonyl groups were introduced, and the N-Boc group was efficiently displaced by treatment with trifluoroacetic acid in dichloromethane. The crude ammonium trifluoroacetate derivatives were treated with NEt3, affording the free amines 6ac. The amines were reacted with 5-heteroarylcarboxylic acids, previously activated with 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide and hydroxybenzotriazole. Thus, the final products 7ac, 8ac and 9ac were obtained in four steps and with excellent overall yield (Scheme 1).
This synthetic pathway appears very solid, high-yielding and general, irrespective of the N-group, the sulfonamide or the type of heteroaryl moiety chosen. The easy access of substrates represents an open door to molecules with synergic biological activity, as anticancer activity, especially because there has been growing interest in repurposing PIs for the treatment of cancer [31].
Despite the fact that this pathway proved to be solid, diversity-oriented synthesis was studied to introduce different functionalities according to needs. In particular, the removal of the Boc group immediately after opening the commercial epoxide 3 with benzylamine allowed to diaminoalcohol 10. In this way, it should be possible to firstly introduce the desired heteroaryl moiety on the primary amine and then the different aromatic sulfonyls on the sterically hindering secondary amine. Unfortunately, this strategy proved to not be applicable because, under these conditions, diamine 10 did not react to afford the desired heterocarboxyamide derivative (Scheme 2).

3.2. In Vitro Activity

IC50 values were obtained on recombinant wild-type HIV protease by measuring the initial rates of hydrolysis of the fluorogenic substrate Abz-Thr-Ile-Nle-Phe(NO2)-Gln-Arg (Table 1) [13,14,15,16,17,18]. The initial rates of the enzyme-catalyzed reactions were measured at different inhibitors concentrations, while the reference rates for the not-inhibited reactions were measured each time before the addition of the inhibitors. IC50 was obtained by measuring the relative residual enzyme activity (ratio of the increase of fluorescence velocities before and after the addition of inhibitor) and by fitting the residual activity vs. inhibitor concentration semilog plots to a tetraparametric logistic function. The results are the mean of three independent experiments and are reported in Table 1, while the plots are reported in Supplementary Figure S1.
All the inhibitors proved to be active and capable of reducing the enzyme activity to less than 10% of that of the free enzyme within the tested range of concentrations (0.5–10 μM) with similar efficacy. Their power is clearly related to the nature of the heterocyclic system at P2. The indole derivatives 9ac are the most powerful inhibitors and perform better than darunavir under our experimental conditions. When tested at 0.5 nM concentrations, they are able to inhibit most of the enzyme activity. We do not report the IC50s for the indole compounds as the residual activities measured at 0.5 nM are already less than 50%, and we are unable to measure the IC50 as the nominal concentration of the enzyme in the test is 1.2 nM, as estimated from the amount of enzyme declared by the producer in the sample. Even if this concentration is most likely overestimated, we prefer not to evaluate the IC50 for such compounds, as the values could be physically meaningless. One order of magnitude in affinity is lost on changing the indole to benzothiophene, and compounds 7ac show nanomolar IC50s. A further decrease is observed with benzofuran at P2, as in inhibitors 8ac, with tenth nanomolar IC50s. Most of the IC50s of compounds 7 and 8 are in the same order of magnitude as that of darunavir, although statistically distinguishable from it (see the t-test plot in Supplementary Figure S2.
A minor effect is given by the substituents at the arylsulfonamide group at P2’, where, at least in series 7 and 8, the 4-methoxyphenyl moiety seems slightly better than 4-nitrophenyl and 3,4-dimethoxyphenyl. This effect, if present, cannot be evaluated in the more active compounds, 9ac.
The beneficial effect of indole in comparison with benzothiophene and benzofuran was already observed in our previous studies on compounds 1 and 2, which are different at P1’ (an isobutyl group is present), and in 2 also as to the length of the chain connecting P2 with the core (a one-atom longer carbamate linker).

3.3. Molecular Modeling

The models were, therefore, built to gain insight into the structural effects at the origin of the observed activities.
The optimized complexes of HIV-Pr with all the indole derivatives 9ac, 1c and 2c were obtained and compared with the experimental crystallographic structure of the complex of darunavir with the enzyme. The model complexes of the benzothiophene and benzofuran derivatives were then also obtained.
All the heterocyclic systems are hosted by the S2 site of the protein in a very similar way. An overlay of the structures of darunavir and 9b is reported in Figure 3a, while the overlay of the structures of 7b, 8b and 9b is reported in Figure 3b.
The heteroatoms (S, O, N) are closely superimposed, while the heterocyclic systems are quite more exposed to the solvent than the dioxabicyclo octane side chain of darunavir. Nevertheless, the indole derivatives inhibit the enzyme better than darunavir. Details of the interactions established by darunavir and 9b are reported in Figure 3c,d.
A clear difference is given by the ability of the indole NH group to act as a hydrogen bond donor towards the carboxylate group of Asp30. This interaction cannot be established by darunavir nor by our benzofuran–benzothiophene compounds, which can only accept hydrogen bonds. However, the heteroatoms in compounds 7, 8 and 9 point outside the binding site and are largely exposed to the solvent. Thus, the interaction with Asp30 is expected to be rather weak, and other effects are most likely operating. The aromatic rings of our inhibitors can clearly establish significant interactions mediated by their π systems. A recent study has compared, at different levels of theory, the ability of indole, thiophene and benzofuran as partners in the formation of π–π stacking interactions with DNA bases [32]. Very interestingly, indole was capable of establishing the strongest π–π stacking interactions, followed by benzothiophene and then by benzofuran. This order resembles that of the inhibitory activity of our compounds. By the way, aromatic side chains are not present in subsite S2 of HIV-Pr, rather there is a number of methyl groups wallpapering the surface of S2, and those from Ala28 and Ile47 (to a minor extent) are found to interact with the heterocyclic side chains of our compounds. We have, therefore, carried out a preliminary evaluation on the ability of indole, thiophene and benzofuran in CH3/π interactions by modeling their complexes with methane. We have followed one of the approaches reported by Toupkanloo and Rahmani, optimizing the structures at the MO62X/6-311++G(d,p) level, and we have actually found that the strength of the CH3/π interaction follows the same order found for π–π stacking. The superior performance of indole over benzofuran and thiophene is thus probably due to this effect, which is likely very general when comparing the interactions of such compounds with biomolecules. Moreover, the indole system is also capable of acting as an acceptor in π-acceptor hydrogen bonding, and we find a couple of interactions where the donors are the backbone NH of Asp29 and Asp30. These interactions replace the hydrogen bonding ones given by the backbone NH of Ala28 and Asp29 towards the oxygens in darunavir (Figure 3a,b).
A further point in favor of 9b in comparison with darunavir may be given by the benzyl side chain that replaces the alkyl chain of the drug at P1’. The aromatic side chain seems able to actually establish more favorable hydrophobic interactions (see the Supplementary Figures S4 and S5 with the maps of the recognized interactions). This may also explain the better performance of the set of inhibitors reported in the present paper in comparison with other sets previously described by us (namely, the difference between 9c and 1c).
As to the minor effect given by the substituents at P2’, a very simple explanation is found in the relatively small size of the S2’ subsite, which can fit the aromatic ring with one methoxy group well, but is unable to host both the 3,4-dimethoxyphenly and the 4-nitrophenyl groups without suffering from conformational distortions of the ligands (Supplementary Figure S3).

4. Conclusions

In conclusion, all the newly synthesized molecules with pseudo-symmetric hydroxyethylamine cores proved to be active, with excellent IC50 values and with several interactions with the enzyme site. Thus, we can highlight that the presence of a bis benzyl in the core can give rigidity to the molecules and maximize the interaction. Furthermore, the indole ring is apparently the heterocycle, which confers greater inhibitory activity; this makes compounds 9ac very promising molecules, regardless of the nature of the substituent present on the sulfonamide.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/biom11111584/s1. Copy of 1H and 13C NMR spectra of compounds 5c, 6c, 7a–c, 8a–c and 9a–c inhibition assays and model structures of complex of compound 9b and darunavir with the protease are available.

Author Contributions

Conceptualization, L.C.; Data curation, F.B.; Funding acquisition, M.F.; Investigation, M.F., P.L.; Methodology, R.D.; Methodology, T.L.; Software, F.B.; Writing—original draft, R.D.; Writing—review and editing, L.C. All authors have read and agreed to the published version of the manuscript.

Funding

Financial support has been provided by the MIUR (Italian Ministry of University) PON Ricerca e Innovazione 2014–2020-Area SALUTE-ARS01 01081,“Prodotti INnovativi ad alto contenuto biotecnologico per il settore BIOMEDicale” (INBIOMED) and the University of Basilicata.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Global Report: UNAIDS Report on the Global AIDS Epidemic 2021. Available online: https://www.unaids.org/sites/default/files/media_asset/2021-global-aids-update_en.pdf. (accessed on 17 September 2021).
  2. Wensing, A.M.J.; Van Maarseveen, N.M.; Nijhuis, M. Fifteen years of HIV Protease Inhibitors: Raising the barrier to resistance. Antiviral Res. 2010, 85, 59–74. [Google Scholar] [CrossRef] [PubMed]
  3. World Health Organization, Geneva, Switzerland. Available online: https://www.who.int/hiv/pub/arv/chapter4.pdf. (accessed on 17 September 2021).
  4. Ghosh, A.K.; Chapsal, B.D. Aspartic acid proteases as therapeutic targets. In Methods and Principles in Medicinal Chemistry; Ghosh, A.K., Ed.; Wiley-VCH: Weinheim, Germany, 2010; Volume 45, pp. 169–204. [Google Scholar]
  5. Tomasselli, A.G.; Heinrikson, R.L. Targeting the HIV-protease in AIDS therapy: A current clinical perspective. Biochim. et Biophys. Acta 2000, 1477, 189–214. [Google Scholar] [CrossRef]
  6. Brik, A.; Wong, C.H. HIV-1 protease: Mechanism and drug discovery. Org. Biomol. Chem. 2003, 1, 5–14. [Google Scholar] [CrossRef] [PubMed]
  7. Ghosh, A.K.; Rao, K.V.; Nyalapatla, P.R.; Kovela, S.; Brindisi, M.; Osswald, H.L.; Reddy, B.S.; Agniswamy, J.; Wang, Y.-F.; Aoki, M.; et al. Design of Highly Potent, Dual-Acting and Central-Nervous-System-Penetrating HIV-1 Protease Inhibitors with Excellent Potency against Multidrug-Resistant HIV-1 Variants. Chem. Med. Chem. 2018, 13, 803–815. [Google Scholar] [CrossRef] [Green Version]
  8. Ghosh, A.K.; Williams, J.N.; Ho, R.Y.; Simpson, H.M.; Hatton, S.-I.; Hayashi, H.; Agniswamy, J.; Wang, Y.-F.; Weber, I.T. Design and Synthesis of Potent HIV-1 Protease Inhibitors Containing Bicyclic Oxazolidinone Scaffold as the P2 Ligands: Structure-Activity Studies and Biological and X-ray Structural Studies. J. Med. Chem. 2018, 61, 9722–9737. [Google Scholar] [CrossRef] [PubMed]
  9. Sun, S.; Jia, Q.; Zhang, Z. Applications of amide isosteres in medicinal chemistry. Bioorg. Med. Chem. Lett. 2019, 29, 2535–2550. [Google Scholar] [CrossRef] [PubMed]
  10. Bonini, C.; Chiummiento, L.; De Bonis, M.; Funicello, M.; Lupattelli, P.; Pandolfo, R. Application of Sharpless asymmetric dihydroxylation to thienyl- and benzothienyl acrylates and crotonates. Tetrahedron Asymmetry 2006, 17, 2919–2924. [Google Scholar] [CrossRef]
  11. Bochicchio, A.; Cefola, R.; Choppin, S.; Colobert, F.; Di Noia, M.A.; Funicello, M.; Hanquet, G.; Pisano, I.; Todisco, S. Chiummiento, L. Selective Claisen rearrangement and iodination for the synthesis of polyoxygenated allyl phenol derivatives. Tetrahedron Lett. 2016, 57, 4053–4055. [Google Scholar] [CrossRef]
  12. Cerminara, I.; Chiummiento, L.; Funicello, M.; Guarnaccio, A.; Lupattelli, P. Heterocycles in Peptidomimetics and Pseudopeptides: Designand Synthesis. Pharmaceuticals 2012, 5, 297–316. [Google Scholar] [CrossRef] [PubMed]
  13. Bonini, C.; Chiummiento, L.; De Bonis, M.; Di Blasio, N.; Funicello, M.; Lupattelli, P.; Pandolfo, R.; Tramutola, F.; Berti, F. Synthesis of New Thienyl Ring Containing HIV-1 Protease Inhibitors: Promising Preliminary Pharmacological Evaluation against Recombinant HIV-1 Proteases. J. Med. Chem. 2010, 53, 1451–1457. [Google Scholar] [CrossRef]
  14. Bonini, C.; Chiummiento, L.; De Bonis, M.; Funicello, M.; Lupattelli, P.; Suanno, G.; Berti, F.; Campaner, P. Synthesis, biological activity and modelling studies of two novel anti HIV PR inhibitors with a thiophene containing hydroxyethylamino core. Tetrahedron 2005, 61, 6580–6589. [Google Scholar] [CrossRef]
  15. Bonini, C.; Chiummiento, L.; De Bonis, M.; Funicello, M.; Lupattelli, P. Synthesis of a First Thiophene Containing Analog of the HIV Protease Inhibitor Nelfinavir. Tetrahedron Lett. 2004, 45, 2797–2799. [Google Scholar] [CrossRef]
  16. Chiummiento, L.; Funicello, M.; Lupattelli, P.; Tramutola, F.; Campaner, P. New indolic non-peptidic HIV protease inhibitors from (S)-glycidol: Synthesis and preliminary biological activity. Tetrahedron 2009, 65, 5984–5989. [Google Scholar] [CrossRef]
  17. Chiummiento, L.; Funicello, M.; Lupattelli, P.; Tramutola, F.; Berti, F.; Marino-Merlo, F. Synthesis and biological evaluation of novel small non-peptidic HIV-1 PIs: The benzothiophene ring as an effective moiety. Bioorg. Med. Chem. Lett. 2012, 22, 2948–2950. [Google Scholar] [CrossRef]
  18. Bonini, C.; Chiummiento, L.; Di Blasio, N.; Funicello, M.; Lupattelli, P.; Tramutola, F.; Berti, F.; Ostric, A.; Miertus, S.; Frecer, V.; et al. Synthesis and biological evaluation of new simple indolic non peptidic HIV Protease inhibitors: The effect of different substitution patterns. Bioorg. Med. Chem. 2014, 22, 4792–4802. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Ghosh, A.K.; Xu, C.X.; Rao, K.V.; Baldridge, A.; Agniswamy, J.; Wang, Y.F.; Weber, I.T.; Aoki, M.; Miguel, S.G.P.; Amano, M.; et al. Probing multidrug-resistance/protein-ligand interaction with oxatricyclic designed ligands in HIV-1 Protease inhibitors. ChemMedChem 2010, 5, 1850–1854. [Google Scholar] [CrossRef] [Green Version]
  20. Ghosh, A.K.; Anderson, D.D.; Weber, I.T.; Mitsuya, H. Enhancing Protein Backbone Binding-A Fruitful Concept for Combating Drug-Resistant HIV. Angew. Chem. Int. Ed. Engl. 2012, 51, 1778–1802. [Google Scholar] [CrossRef] [PubMed]
  21. Zhang, H.; Wang, Y.F.; Shen, C.H.; Agniswamy, J.; Rao, K.V.; Xu, X.; Ghosh, A.K.; Harrison, R.W.; Weber, I.T. Novel P2 Tris-tetrahydrofuran Group in Antiviral Compound 1 (GRL-0519) Fills the S2 Binding Pocket of Selected Mutants of HIV-1 Protease. J. Med. Chem. 2013, 56, 1074–1083. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Funicello, M.; Chiummiento, L.; Tramutola, F.; Armentano, M.F.; Bisaccia, F.; Miglionico, R.; Milella, L.; Benedetti, F.; Berti, F.; Lupattelli, P. Synthesis and biological evaluation in vitro and in mammalian cells of new heteroaryl carboxyamides as HIV-protease inhibitors. Bioorg. Med. Chem. 2017, 25, 4715–4722. [Google Scholar] [CrossRef] [PubMed]
  23. Tramutola, F.; Armentano, M.F.; Berti, F.; Chiummiento, L.; Lupattelli, P.; D’Orsi, R.; Miglionico, R.; Milella, L.; Bisaccia, F.; Funicello, M. New heteroaryl carbamates: Synthesis and biological screening in vitro and in mammalian cells of wild-type and mutant HIV-protease inhibitors. Bioorg. Med. Chem. 2019, 27, 1863–1870. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Hidaka, K.; Kimura, T.; Hayashi, Y.; McDaniel, K.F.; Dekhtyar, T.; Colletti, L.; Kiso, Y. Design and Synthesis of Pseudo-Symmetric HIV Protease Inhibitors Containing a Novel Hydroxymethylcarbonyl (HMC)-Hydrazide Isostere. Bioorg. Med. Chem. Lett. 2003, 13, 93–96. [Google Scholar] [CrossRef]
  25. Facchinetti, V.; Moreth, M.; Gomes, C.R.B.; do Ó Pessoa, C.; Rodrigues, F.A.R.; Cavalcanti, B.C.; Oliveira, A.C.A.; Carneiro, T.R.; Gama, I.L.; de Souza, M.V.N. Evaluation of (2S,3R)-2-(amino)-[4-(N-benzylarenesulfonamido)-3-hydroxy-1-phenylbutane derivatives: A promising class of anticancer agents. Med. Chem. Res. 2015, 24, 533–542. [Google Scholar] [CrossRef]
  26. Schrödinger Release 2021-3: MacroModel. Schrödinger, LLC.: New York, NY, USA, 2021; The inhibitors were then docked inside the catalytic site of the enzyme with the Schrödinger Glide protocol.
  27. Friesner, R.A.; Murphy, R.B.; Repasky, M.P.; Frye, L.L.; Greenwood, J.R.; Halgren, T.A.; Sanschagrin, P.C.; Mainz, D.T. Extra Precision Glide: Docking and Scoring Incorporating a Model of Hydrophobic Enclosure for Protein-Ligand Complexes. J. Med. Chem. 2006, 49, 6177–6196. [Google Scholar] [CrossRef] [Green Version]
  28. Trucks, G.W.; Frisch, M.J.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. (2009) Gaussian 09; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  29. Bai, X.; Yang, Z.; Zhu, M.; Dong, B.; Zhu, L.; Zhang, G.; Wang, J.; Wang, Y. Design and synthesis of potent HIV-1 protease inhibitors with (S)-tetrahydrofuran-tertiary amine-acetamide as P2ligand: Structure activity studies and biological evaluation. Eur. J. Med. Chem. 2017, 137, 30–44. [Google Scholar] [CrossRef] [PubMed]
  30. Yang, Z.-H.; Bai, X.-G.; Zhou, L.; Wang, J.-X.; Liu, H.-T.; Wang, Y.-C. Synthesis and biological evaluation of novel HIV-1 protease inhibitors using tertiary amine as P2-ligands. Bioorg. Med. Chem. Lett. 2015, 25, 1880–1883. [Google Scholar] [CrossRef] [PubMed]
  31. Chow, W.A.; Jiang, C.; Guan, M. Anti-HIV drugs for cancer therapeutics: Back to the future? Lancet Oncol. 2009, 10, 61–71. [Google Scholar] [CrossRef]
  32. Toupkanloo, H.A.; Rahmani, Z. An in-depth study on non-covalent stacking interactions between DNA bases and aromatic drug fragments using DFT method and AIM analysis: Conformers, binding energies, and charge transfer. App. Biol. Chem. 2018, 61, 209–226. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Commercial and not HIV-protease inhibitors.
Figure 1. Commercial and not HIV-protease inhibitors.
Biomolecules 11 01584 g001
Figure 2. Examples of pseudo-symmetrical core of HIV-protease inhibitors.
Figure 2. Examples of pseudo-symmetrical core of HIV-protease inhibitors.
Biomolecules 11 01584 g002
Scheme 1. Synthesis of inhibitors 7ac, 8ac and 9ac: (a) BnNH2, i-PrOH, 60°C, 4h (88% yield); (b) arylsulfonyl chloride, Et3N, CH2Cl2, rt, 24h (5a 94%; 5b 85%; 5c 90%); (c) TFA/CH2Cl2 30%, rt, 1 h; then Et3N; (d) 5-heteroarylcarboxylic acid, EDC, HOBt, then 6a–c, iPr2NEt, CH2Cl2, 24h, rt (7a, 50%; 7b 57%; 7c 55%; 8a 53%; 8b 54%; 8c 56%; 9a 33%; 9b 43%; 9c 44%).
Scheme 1. Synthesis of inhibitors 7ac, 8ac and 9ac: (a) BnNH2, i-PrOH, 60°C, 4h (88% yield); (b) arylsulfonyl chloride, Et3N, CH2Cl2, rt, 24h (5a 94%; 5b 85%; 5c 90%); (c) TFA/CH2Cl2 30%, rt, 1 h; then Et3N; (d) 5-heteroarylcarboxylic acid, EDC, HOBt, then 6a–c, iPr2NEt, CH2Cl2, 24h, rt (7a, 50%; 7b 57%; 7c 55%; 8a 53%; 8b 54%; 8c 56%; 9a 33%; 9b 43%; 9c 44%).
Biomolecules 11 01584 sch001
Scheme 2. (a) BnNH2, i-PrOH, 60 °C, 4 h; (b) p-TsOH, MeCN, 5 h, rt (60% yield from 3); (c) 5-heteroarylcarboxylic acid, EDC, HOBt, then 10, iPr2NEt, CH2Cl2, 24 h, rt; (d) p-methoxybenzene sulfonyl chloride, Et3N, CH2Cl2.
Scheme 2. (a) BnNH2, i-PrOH, 60 °C, 4 h; (b) p-TsOH, MeCN, 5 h, rt (60% yield from 3); (c) 5-heteroarylcarboxylic acid, EDC, HOBt, then 10, iPr2NEt, CH2Cl2, 24 h, rt; (d) p-methoxybenzene sulfonyl chloride, Et3N, CH2Cl2.
Biomolecules 11 01584 sch002
Figure 3. (a): Overlay of the crystallographic structure of the Hiv-pr complex with darunavir and the model structure of the complex of 9b; (b): overlay of the model structures of the complexes of 7b, 8b and 9b with the protease; (c): interactions of the dioxabicyclo octane group of darunavir with the protein; (d): interactions of the indole system of 9b with the protein.
Figure 3. (a): Overlay of the crystallographic structure of the Hiv-pr complex with darunavir and the model structure of the complex of 9b; (b): overlay of the model structures of the complexes of 7b, 8b and 9b with the protease; (c): interactions of the dioxabicyclo octane group of darunavir with the protein; (d): interactions of the indole system of 9b with the protein.
Biomolecules 11 01584 g003
Table 1. In vitro inhibition activity of compounds 7.
Table 1. In vitro inhibition activity of compounds 7.
Entry Biomolecules 11 01584 i001CompoundXRIC50 (nM)Std. Error (nM)
17aS4-NO24.20.6
27bS4-OMe2.30.4
37cS3,4-diOMe6.22.1
48aO4-NO247.06.0
58bO4-OMe9.60.1
68cO3,4-diOMe82.07.0
79aNH4-NO227.7% *3.5%
89bNH4-OMe36.3% *5.1%
99cNH3,4-diOMe10.2% *2.4%
10darunavir 1.80.3
* residual activity at the 0.5 nM inhibitor.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

D’Orsi, R.; Funicello, M.; Laurita, T.; Lupattelli, P.; Berti, F.; Chiummiento, L. The Pseudo-Symmetric N-benzyl Hydroxyethylamine Core in a New Series of Heteroarylcarboxyamide HIV-1 Pr Inhibitors: Synthesis, Molecular Modeling and Biological Evaluation. Biomolecules 2021, 11, 1584. https://doi.org/10.3390/biom11111584

AMA Style

D’Orsi R, Funicello M, Laurita T, Lupattelli P, Berti F, Chiummiento L. The Pseudo-Symmetric N-benzyl Hydroxyethylamine Core in a New Series of Heteroarylcarboxyamide HIV-1 Pr Inhibitors: Synthesis, Molecular Modeling and Biological Evaluation. Biomolecules. 2021; 11(11):1584. https://doi.org/10.3390/biom11111584

Chicago/Turabian Style

D’Orsi, Rosarita, Maria Funicello, Teresa Laurita, Paolo Lupattelli, Federico Berti, and Lucia Chiummiento. 2021. "The Pseudo-Symmetric N-benzyl Hydroxyethylamine Core in a New Series of Heteroarylcarboxyamide HIV-1 Pr Inhibitors: Synthesis, Molecular Modeling and Biological Evaluation" Biomolecules 11, no. 11: 1584. https://doi.org/10.3390/biom11111584

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

Article Metrics

Back to TopTop