Next Article in Journal
Fructose Might Contribute to the Hypoglycemic Effect of Honey
Next Article in Special Issue
Synthesis and Antimicrobial Evaluation of Some Pyrazole Derivatives
Previous Article in Journal
A New Triterpene From Uncaria macrophylla and Its Antitumor Activity
Previous Article in Special Issue
Benzoxetes and Benzothietes ¾ Heterocyclic Analogues of Benzocyclobutene
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A New Class of Heterocycles: 1,4,3,5-Oxathiadiazepane 4,4-dioxides

by
Amel Bendjeddou
1,*,
Tahar Abbaz
2,
Zine Regainia
1 and
Nour-Eddine Aouf
1
1
Laboratoire de Chimie Organique Appliquée, Groupe de chimie Hétérocyclique, Département de Chimie, Faculté des Sciences, Université d’Annaba, BP 12, 23000, Algérie
2
Laboratoire de Chimie des Matériaux Organiques, Université de Tébessa, Route de Constantine, Tébessa, 12000, Algérie
*
Author to whom correspondence should be addressed.
Molecules 2012, 17(2), 1890-1899; https://doi.org/10.3390/molecules17021890
Submission received: 19 January 2012 / Revised: 6 February 2012 / Accepted: 8 February 2012 / Published: 14 February 2012
(This article belongs to the Special Issue Heterocycles)

Abstract

:
This work reports the synthesis of novel 1,4,3,5-oxathiadiazepanes 4,4-dioxides from the reaction of N’-benzyl-N-(2-hydroxyethyl)-sarcosine or proline sulfamide with aromatic aldehydes under acid catalysis. To prepare the starting materials N-Boc-sulfamide derivatives of sarcosine or proline were alkylated with benzyl alcohol under Mitsunobu reaction conditions, the Boc group was removed chemoselectively by acidolysis, and the resulting product reduced to the corresponding alcohol in good yields.

1. Introduction

Sulfamides and their analogs have a rich chemical and biological history and have emerged as a promising class of compounds in drug discovery [1,2]. Sultams (cyclic sulfamides), although not found in Nature, have also shown potent biological activity, including several displaying a wide spectrum of activities [3,4]. The more prominent include a number of benzodithiazine dioxides and benzoxathiazepine 1,1-dioxides displaying anti-HIV-1 activity [5] and the ability to activate glucokinase [6] (type II diabetes), respectively.
In addition, pyrrolo[2,1-c]benzodiazepine antibiotics and some of their heterocyclic analogs show anticancer activity [7,8,9,10], it has been proposed that the cytotoxicity and antitumour activity of these compounds results from the formation of a covalent bond between the azomethine unit of the diazepine ring and the C(2)-amino group guanine nucleus in the minor groove of the DNA double spiral [11,12,13,14].
In a continuous of previous work [15,16,17,18,19], we have described a convenient access to a series of n-membered cyclic sulfamides A and heterocyclic constrained peptides containing sulfamide groups B, starting from natural amino acids, chloroethylamine and chlorosulfonylisocynanate (CSI). In continuation of our efforts to design and synthesize new cyclic sulfamides, we have extended our studies to a series of seven membered heterocyclic compounds C containing sulfamide groups (Figure 1), we describe for the first time the first example of a new heterocyclic class 1,4,3,5-oxathiadiazepane 4,4-dioxides which can be described as structural analogs of the anti-HIV-1 compounds mentioned above. This compound is an interesting candidate for pharmaceutical purposes.
Figure 1. Cyclosulfamide stuctures.
Figure 1. Cyclosulfamide stuctures.
Molecules 17 01890 g001

2. Results and Discussion

In recent years, considerable attention was paid to compounds resulting from the condensation of substituted amino alcohols with aldehydes [20,21,22,23,24]. As outlined in Scheme 1, the substituted amino alcohols (N’-benzyl-N-(2-hydroxyethyl)-proline or sarcosine sulfamides) 1c and 2c were prepared in a three-step reaction sequence starting from (tert-butyloxycarbonylsulfonyl) L-amino acid methyl esters 1 and 2. These compounds were synthesized by sulfamoylation of aminoester derivatives (Pro, Sar) as previously described [18].
Scheme 1. General synthesis of substituted aminoalcohol sulfamides.
Scheme 1. General synthesis of substituted aminoalcohol sulfamides.
Molecules 17 01890 g002
In these tert-butyloxycarbonylsulfamides 1 and 2, the Boc (tert-butyloxycarbonyl) group increases the acidity of the adjacent NH group and allows an expedient regiospecific alkylation under Mitsunobu conditions [25,26,27,28] using benzylic alcohol, which provides the N-substituted Boc-sulfamides 1a and 2a in 80% and 85% yield, respectively. Selective cleavage of the tert-butyloxycarbonyl protective group with trifluoroacetic acid gives compounds 1b and 2b in 95% and 98% yield, respectively. The substituted amino alcohols sulfamides 1c and 2c were obtained from the deprotected sulfamides by NaBH4 reduction in 87% and 85% yield, respectively.
Substituted 1,4,3,5-oxathiadiazepane 4,4-dioxides were obtained in accordance with the methodology shown in the literature [20]. The substituted amino alcohols 1c and 2c were allowed to react with aromatic aldehydes in dichloromethane in a cycolodehydration reaction to obtain the corresponding compounds 1d4d, 1e4e. The yields are listed in Table 1. These compounds can furnish after debenzylation new ring opened products by nucleophilic attack by organometallic reagents [27,28].
Table 1. General synthesis of substituted oxathiadiazepane 4,4-dioxides. Molecules 17 01890 i001
Table 1. General synthesis of substituted oxathiadiazepane 4,4-dioxides. Molecules 17 01890 i001
CompoundsYield (%)
1d50
2d50
3d43
4d41
1e55
2e45
3e40
4e42
In the 1H-NMR (CDCl3) spectra of proline derivatives 1e4e, the asymmetric carbon protons of compounds 3e and 4e resonate around δ 6.52–6.35 ppm as a singlet peak. Compounds 1e and 2e exhibit a doublet of doublets at δ ~ 5.48 ppm and 7.90 ppm, with coupling constants of 3.43, 3.49 Hz and 1.53, 1.57 Hz, respectively. Thus, the asymmetric carbon proton of sarcosine derivatives, showed for compounds 3d and 4d a doublet at δ ~ 7.90 ppm and δ ~ 7.50 ppm with coupling constants of 1.47 Hz and 1.49 Hz, respectively, while compound 2d exhibit a doublet of doublets at δ ~ 7.80 ppm with coupling constants of 1.37, 1.55 Hz, and compound 1d presents a triplet at δ ~ 5.10 ppm.

3. Experimental

3.1. General

All commercial chemicals and solvents were used as received. Melting points were determined in open tubes on a Büchi apparatus and are uncorrected. IR spectra were recorded on a Perkin-Elmer Spectrum 1000 spectrophotometer. Microanalyses were performed in the Microanalysis Laboratory of ENSCM (Montpellier). 1H and 13C-Nuclear Magnetic Resonance spectra were determined on a Brüker AC 250 spectrometer. Chemical shifts are recorded in ppm (δ) and coupling constants (J) in Hertz, relative to tetramethylsilane used as internal standard. Multiplicity is indicated as s (singlet), d (doublet), q (quadruplet), m (multiplet) and combinations of these signals. Fast-atom bombardment mass spectra (FAB) were recorded in positive or negative mode with glycerol (G), thioglycerol (GT), or 3-nitrobenzylalcohol (NOBA) as matrix. Optical rotations for solutions in CHCl3 were measured with a POLAX model 2L digital polarimeter. All reactions were monitored by Thin Layer Chromatography (TLC) on silica gel Merck 60 F254 precoated aluminium plates, developed by spraying with ninhydrin solution. Column chromatography was performed using silica gel 60 (203–400 mesh).

3.2. General Synthetic Procedure for Carbamoylation-Sulfamoylation: Preparation of 1 and 2

A solution of N-chlorosulfonyl tert-butylcarbamate (0.05 mol) was prepared by addition of tert-butanol (408 mL in 50 mL of dichloromethane) to a solution of CSI (7.1 g in the same solvent). The resulting Boc-sulfamoyl chloride solution (25 mL) and triethylamine (17.40 g, 17.1 mL, 0.085 mol) in dichloromethane (100 mL) was added into a suspension of aminoester (0.05 mol) in the same solvent (120 mL) at 0 °C. The reaction was complete in 45 minutes. The reaction mixture was then diluted with dichloromethane (100 mL) and washed with two portions of 0.1 N HCl solution. The organic layer was dried with (Na2SO4) and concentrated in vacuo to give the crude product, which was purified by column chromatography eluting with dichloromethane to give compounds 1 and 2.
(S)-Methyl [N-methyl(N’-tert-butyloxycarbonyl)-sulfamoyl]-glycinate (1). Yield = 80%; TLC: Rf = 0.76 (CH2Cl2-MeOH 9:1); cristallizable oil; IR (KBr) ν cm−1: 3200 (NH), 1770, 1763 (C=O), 1360 and 1150 (SO2); 1H-NMR (CDCl3) δ ppm: 7.62 (s, 1H, NH), 4.14 (s, 2H, CH2), 3.75 (s, 3H, OCH3), 3.04 (s, 3H, NCH3), 1.49 (s, 9H, tBu); M.S: (NOBA, FAB > 0): 283 [M+H]+, 565. M = 282; Anal. Calcd for C9H18N2O6S: C, 38.29; H, 6.38; N, 9.93; S, 11.34; found: C, 38.24; H, 6.43; N, 9.87; S, 11.28.
(S)-Methyl [N-(N’-tert-butyloxycarbonyl)-sulfamoyl]-prolinate (2). Yield = 80%; TLC: Rf = 0.58 (CH2Cl2-MeOH 95:5); m.p. = 132–133 °C; [α]D = −9.5 (c = 1; MeOH); IR (KBr) ν cm−1: 1730, 1712 (C=O), 1340 and 1150 (SO2); 1H-NMR (CDCl3) δ ppm: 7.52 (s, 1H, NH); M.S: (NOBA, FAB > 0): 309 [M+H]+, 208 ([M+H]+-Boc). M = 308; Anal. Calcd for C11H20N2O6S: C, 42.85; H, 6.69; N, 9.09; S, 10.38; found: C, 42.27; H, 6.53; N, 9.06; S, 10.41.

3.3. General Procedure for the Synthesis of N-Boc, N′-(benzyl)sulfamides 1a and 2a

A solution of N-alkyl carboxylsulfamide (0.0065 mol, 2 g), triphenylphosphine (2.56 g) and benzylic alcohol (0.7 g) in THF (15 mL) was added dropwise (20 min, 5 °C) to a solution of equimolar quantities of diethyl(diisopropyl)azodicarboxylate (1.97 g) in THF (5 mL), the reaction medium was stirred under an atmosphere of dry nitrogen for about 45 min. TLC reveals the formation of substituted compound (UV, ninhydrin) less polar than its precursor. Oxidoreduction compounds were removed by filtration after precipitation into diethylether. The filtrate was concentred and the crude residue was purified by column chromatography eluted with dichloromethane.
(S)-Methyl [N-(N′-tert-butyloxycarbonyl, N′-benzyl)-sulfamoyl]-glycinate (1a). Yield = 80%; TLC: Rf = 0.76 (CH2Cl2-MeOH 95:5); m.p. = 72 °C; 1H-NMR (CDCl3) δ ppm: 7.34 (m, 5H, ArH), 4.86 (s, 2H, CH2-ph), 4.04 (s, 2H, CH2-N), 3.72 (s, 3H, OCH3), 2.84 (s, 3H, CH3), 1.49 (s, 9H, tBu); M.S: (NOBA, FAB > 0): 373 [M+H]+. M = 372; Anal. Calcd for C16H24N2O6S: C, 51.61; H, 6.45; N, 7.52; S, 8.60; found: C, 51.43; H, 6.38; N, 7.49; S, 8.62.
(S)-Methyl [N-(N′-tert-butyloxycarbonyl, N′-benzyl)-sulfamoyl]-prolinate (2a). Yield = 85%; TLC: Rf = 0.73 (CH2Cl2-MeOH 95:5); m.p. = 73 °C; 1H-NMR (CDCl3) δ ppm: 7.36 (m, 5H, ArH), 4.86 (m, 1H, CH), 4.66 (2dd, 2H, J = 3.14, 5.45 Hz, CH2-Ph), 3.72 (s, 3H, OCH3), 3.48 (m, 1H, CH*), 3.15 (m, 1H, CH), 1.88–2.08 (m, 4H, 2CH2), 1.48 (s, 9H, tBu); M.S: (NOBA, FAB > 0): 399 [M+H]+. M = 398; Anal. Calcd for C18H26N2O6S: C, 54.71; H, 6.53; N, 7.038; S,8.04; found: C, 54.65; H, 6.62; N, 7.05; S, 8.13.

3.4. General Deprotection Procedure: Preparation of 1b and 2b

A solution of trifluoroacetic acid (50% in dichloromethane; 3 equiv.) was dropwise added into a stirred solution of substituted N-carboxylsulfamide(0.92 g, 2.6 mmol) in dichloromethane (10 mL) at 0 °C. The reaction midium was stirred during two hours, concentrated under reduced pressure and coevaporated with diethylether. The residue was purified by column chromatography eluted with dichloromethane (or recrystallized from an AcOEt-hexane mixture) to afford deprotected sulfamides 1b and 2b.
N[N′-Benzyl]sulfamoylglcinate de methyl (1b). Yield = 95%; TLC: Rf = 0.55 (CH2Cl2-MeOH 95:5); m.p. = 100–102 °C; IR (KBr) ν cm−1: 3312 (NH), 1360 and 1160 (SO2); 1H-NMR (CDCl3) δ ppm: 7.34 (m, 5H, ArH), 5.70 (t, 1H, NH), 4.30 (d, 2H, J = 6.0 Hz, CH2-Ph), 4.0 (s, 2H, CH2), 3.75 (s, 3H, OCH3), 2.90 (s, 3H, CH3); M.S: (NOBA, FAB > 0): 273 [M+H]+, 546; M = 272; Anal. Calcd for C11H16N2O4S: C, 48.52; H ,5.88; N ,10.29; S ,11.76; found: C, 48.39; H, 5.80; N, 10.25; S, 11.74.
N[N′-Benzyl]sulfamoylprolinate de methyl (2b). Yield = 98%; TLC: Rf = 0.60 (CH2Cl2-MeOH 95:5); oil; IR (KBr) ν cm−1: 3320 (NH), 1355 and 1159 (SO2); 1H-NMR (CDCl3) δ ppm: 7.34 (m, 5H, ArH), 5.96 (m, 1H, NH), 4.41 (dd, 1H, J = 3.81, 4.70Hz, C*H), 4.30 (2dd, 2H, J = 3.11, 4.70 Hz, CH2-Ph), 3.70 (s, 3H, OCH3), 3.46 (m, 2H, CH2N), 2.20–2.0 (m, 4H, CH2β and CH2γ); M.S: (NOBA, FAB > 0): 299 [M+H]+. M = 298; Anal. Calcd for C13H18N2O4S: C, 52.34; H, 6.04; N, 9.39; S, 10.73; found: C, 52.29; H, 6.00; N, 9.37; S, 10.68.

3.5. General Reduction Procedure; Preparation of 1c and 2c

Deprotected product 1b or 2b (3.30 mmol) in THF (15 mL) was added dropwise to a suspension of NaBH4 (0.006 mol) in THF-Water (4:1, v/v, 20 mL) at 0 °C. When the addition was complete, the reaction mixture was acidified slowly with HCl 5% and concentrated in vacuo. The aqueous layer was extracted with ethyl acetate (3 × 150 mL). The combined organic layers were dried (Na2SO4) and concentrated in vacuo. The crude product was purified by column chromatography eluted with dichlomethane-methanol (90:10).
N-Methyl(N′-benzylsulfamoyl) glycinol (1c). Yield = 87%; TLC: Rf = 0.50 (CH2Cl2-MeOH 95:5); m.p. = 98 °C; IR (KBr) ν cm−1: 3315 (NH), 1330 (OH); 1H-NMR (CDCl3) δ ppm: 7.35 (m, 5H, ArH), 4.72 (t, 1H, NH), 4.39 (d, 2H, J = 5.99 Hz, CH2-Ph), 3.78 (t, 2H, CH2OH), 3.35 (t, 2H, CH2N), 2.88 (s, 3H, CH3), 1.98 (s, 1H, OH); M.S: (ESI+): 267 [M+Na]+, 511 [2M+Na]+; M = 244; Anal. Calcd for C10H16N2O3S: C, 49.18; H, 6.55; N, 11.47; S, 13.11; found: C, 49.14; H, 6.52; N, 11.45; S, 13.20.
N-Methyl(N′-benzylsulfamoyl) prolinolnol (2c). Yield = 85%; TLC: Rf = 0.48 (CH2Cl2-MeOH 95:5); m.p. = 109–112 °C; IR (KBr) ν cm−1: 3320 (NH), 1370 (OH); 1H-NMR (CDCl3) δ ppm: 7.30 (m, 5H, ArH), 5.0 (t, 1H, NH), 4.25 (d, 2H, J = 5.5.75 Hz, CH2-Ph), 3.81 (m, 1H, CH*), 3.55 (m, 2H, CH2OH), 3.30 (t, 2H, CH2N), 2.31 (band large, s, 1H, OH), 1.85 (m, 4H, CH2β and CH2γ); M.S: (ESI+): 293 [M+Na]+, 563 [2M+Na]+; M = 270; Anal. Calcd for C12H18N2O3S: C, 53.33; H, 6.66; N, 10.37; S, 11.85; found: C, 53.31; H, 6.60; N, 10.39; S, 11.89.

3.6. General Procedure for the Preparation of 1,4,3,5-Oxathiadiazepanes 4,4-dioxides 1d4d, 1e4e

Compounds 1c and 2c (0.01 mol) were dissolved separately in dichloromethane (25 mL), and the aromatic aldehyde (0.01 mol) was added. A drop of concentrated sulfuric acid was also added, and the reaction mixture was stirred for 3h at room temperature. The reaction mixture was washed with a 5% solution of sodium bicarbonate, water and then with brine. The organic layer was dried over anhydrous sodium sulfate, and evaporated under reduced pressure on a rotary evaporator. The residue was purified by column chromatography eluting with dichloromethane to give the 1,4,3,5-oxathiadiazepanes 4,4-dioxide.
(N3,2)-Dibenzyl, N5-methyl 1,4,3,5-oxathidiazepane 4,4-dioxyde (1d). Yield = 59%; TLC: Rf = 0.75 (CH2Cl2); m.p. = 82–85 °C; IR (KBr); 1H-NMR (CDCl3) δ ppm: 7.30 (m, 10H, 2ArH), 7.09 (m, 2H, C*-CH2-ph), 5.1 (t, J = 5.37, 1H, CH*), 4.55 (s, 2H, CH2-N), 3.90 (m, 2H, CH2-O), 2.97 (S, 3H, CH3-N), 2.85 (m, 2H, CH2-N); 13C-NMR (CDCl3) δ ppm: 37.80, 41.34, 49.74, 52.80, 68, 89.31, 128; M.S: (ESI+): 369 [M+Na]+; M = 346; Anal. Calcd for C18H21N2O3S: C, 62.42; H, 6.06; N, 8.09; S, 9.24; found: C, 62.44; H, 6.02; N, 8.03; S, 9.20.
N3-Benzyl, N5-methyl, 2-phenyl 1,4,3,5-oxathidiazepane 4,4-dioxyde (2d). Yield = 50%; TLC: Rf = 0.75 (CH2Cl2); oil; 1H-NMR (CDCl3) δ ppm: 7.80 (dd, J = 1.37, 1.55 Hz, 1H, CH*), 7.30 (m, 10H, 2Ar), 4.35 (t, 2H, CH2-O), 4.25 (dd, J = 27.57, 27.52 Hz, 2H, CH2-Ph), 3.30 (t, 2H, CH2-N), 2.90 (S, 3H, CH3-N); 13C-NMR (CDCl3) δ ppm: 37.80, 49.70, 52.80, 68, 87.50, 128.50; (ESI+): 355 [M+Na]+; M = 332; Anal. Calcd for C17H20N2O3S: C, 61.44; H, 6.02; N, 8.43; S ,9.63; found: C, 61.40; H, 6.10; N, 8.50; S, 9.67.
N3-Benzyl, N5-methyl, 2-(2-chlorophenyl) 1,4,3,5-oxathidiazepane 4,4-dioxyde (3d). Yield = 43%; TLC: Rf = 0.80 (CH2Cl2); m.p. = 80–82 °C; 1H-NMR (CDCl3) δ ppm: 7.90 (d, J = 1.47 Hz, 1H, CH*), 7.40 (m, 9H, 2Ar), 4.50 (t, J = 5.49 Hz, 2H, CH2-O), 4.20 (d, J = 5.98 Hz, 2H, CH2-Ph), 3.57 (t, J = 5.49 Hz , 2H, CH2-N), 2.92 (S, 3H, CH3-N); 13C-NMR (CDCl3) δ ppm: 37.80, 49.50, 53, 68.4, 84.60, 129, 132; M.S: (ESI+): 389 [M+Na]+; M = 366; Anal. Calcd for C17H19N2O3SCl: C,55.73; H, 5.19; N, 7.65; S, 8.74; found: C, 55.68; H, 5.14; N, 7.59; S, 8.70.
N3-Benzyl, N5-methyl, 2-(4-chlorophenyl) 1,4,3,5-oxathidiazepane 4,4-dioxyde (4d). Yield = 41%; TLC: Rf = 0.80 (CH2Cl2); m.p. = 80–82 °C; 1H-NMR (CDCl3) δ ppm: 7.50 (d, J = 1.49 Hz, 1H, CH*), 7.20 (m, 9H, 2Ar), 4.30 (t, J = 1.63 Hz, 2H, CH2-O), 4.1 and 4.4 (2d, J = 15.89 Hz, 1H and J = 15.91 Hz, 1H, CH2-Ph), 3.5 (t, J = 1.49, 2H, CH2-N), 3.10 (s, 3H , CH3-N); 13C-NMR (CDCl3) δ ppm: 37.80, 49.50, 55.90, 68.3, 85, 129, 131.5; M.S: (ESI+): 389 [M+Na]+; M = 366; Anal. Calcd for C17H19N2O3SCl: C, 55.73; H, 5.19; N, 7.65; S, 8.74; found: C, 55.76; H, 5.23; N, 7.72; S, 8.70.
(N3,2)-Dibenzyl, (N5,6)-trimethylene 1,4,3,5-oxathiadiazepane 4,4-dioxide (1e). Yield = 55%; TLC: Rf = 0.82 (CH2Cl2); m.p. = 107–108 °C; 1H-NMR (CDCl3) δ ppm: 7.30 (m, 10H, 2ArH), 5.48 (dd, J = 3.43, J = 3.49 Hz, 1H, C1H*), 5.31 (s, 2H, CH2-N), 4.81 and 4.55 (2d, J = 16.77 Hz, 1H and J = 16.68 Hz, 1H, CH2-Ph), 3.81 (m, 2H, CH2-O), 3.50 (m, 3H, CH2-N and C2H*-N), 1.35 (m, 4H, CH2β and CH2γ); 13C-NMR (CDCl3) δ ppm: 24.80, 28.97, 29.71, 41.41, 48.19, 59.15, 73.06, 88.98, 129; M.S: (ESI+): 395 [M+Na]+; M = 372; Anal. Calcd for C20H24N2O3S: C, 64.5; H, 6.45; N, 7.52; S, 8.60; found: C, 64.49; H, 6.46; N, 7.47; S, 8.65.
2-Phenyl, N3-benzyl, (N5,6)-trimethylene 1,4,3,5-oxathiadiazepane 4,4-dioxide (2e). Yield = 45%; TLC: Rf = 0.79 (CH2Cl2); m.p. = 97–99 °C; 1H-NMR (CDCl3) δ ppm: 7.90 (dd, J = 1.53, J = 1.57 Hz, 1H, C1H*), 7.45 (m, 10H, 2ArH), 4.30 (s, 2H, CH2-Ph), 3.90 (m, 1H, C2H*), 3.70–3.60 (m, 2H, CH2-O), 3.50 (m, 2H, CH2-N), 1.80 (m, 4H, CH2β and CH2γ); 13C-NMR (CDCl3) δ ppm: 24.80, 28.97, 41, 48.20, 59.15, 72.80, 89.80, 129; M.S: (ESI+): 381 [M+Na]+; M = 358; Anal. Calcd for C19H22N2O3S: C, 63.68; H, 6.14; N, 7.82; S, 8.93; found: C, 63.72; H, 6.16; N, 7.79; S, 8.89.
2-(2-Chlorophenyl), N3-benzyl, (N5,6)-trimethylene 1,4,3,5-oxathiadiazepane 4,4-dioxide (3e). Yield = 40%; TLC: Rf = 0.80 (CH2Cl2); oil; 1H-NMR (CDCl3) δ ppm: 6.94 (m, 9H, 2ArH), 6.52 (s, 1H, C1H*), 4.50 (dd, J = 15.81, J = 16.24 Hz, 2H, CH2-Ph), 4.10–3.90 (m, 3H, CH2-O and C2H*), 3.60 (t, 2H, CH2-N), 2.08 (m, 2H, CH2β,), 1.60 (m, 2H, CH2γ); 13C-NMR (CDCl3) δ ppm: 24.80, 28.97, 41, 48.20, 59.15, 72.80, 89.80, 129, 131.5; M.S: (ESI+): 415 [M+Na]+; M = 392.5; Anal. Calcd for C19H21N2O3SCl: C, 58.08; H, 5.35; N, 7.13; S, 8.15; found: C, 58.00; H, 5.29; N, 7.09; S, 8.20.
2-(4-Chlorophenyl), N3-benzyl, (N5,6)-trimethylene 1,4,3,5-oxathiadiazepane 4,4-dioxide (4e). Yield = 42%; TLC: Rf = 0.78 (CH2Cl2); yellow powder, m.p. = 154–157 °C; 1H-NMR (CDCl3) δ ppm: 7.25 (m, 9H, 2ArH), 6.35 (s, 1H, C1H*), 4.48 (dd, J = 15.85, J = 16.21 Hz, 2H, CH2-Ph), 4.11 (m, 1H, C2H*), 3.85 (m, 2H, CH2-O), 3.55 (m, 2H, CH2-N), 2.15 (m, 4H, CH2β and CH2γ,); 13C-NMR (CDCl3) δ ppm: 24.80, 28.97, 41, 48.20, 59.15, 72.80, 89.80, 129, 131; M.S: (ESI+): 415 [M+Na]+, 807 [2M+Na]+; M = 392.5; Anal. Calcd for C19H21N2O3SCl: C, 58.08; H, 5.35; N, 7.13; S, 8.15; found: C, 58.10; H, 5.39; N, 7.16; S, 8.19.

4. Conclusions

In conclusion, we have successfully prepared a new class of seven-membered heterocyclic substituted 1,4,3,5-oxathiadiazepanes 4,4-dioxides via a simple strategy. The biological evaluation of the resulting compounds, their stereochemical study, reopening after debenzylation by nucleophilic attack by organometallic reagents and their incorporation into biomolecule analogues are currently underway and will be reported in due course.

Acknowledgments

This work was partially supported by Algerian Research Ministry, MERS.

References and Notes

  1. Drews, J. Drug Discovery: A Historical Perspective. Science 2000, 287, 1960–1964. [Google Scholar] [CrossRef]
  2. Scozzafava, A.; Owa, T.; Mastrolorenzo, A.; Supuran, C.T. Update on the Development of HIV Entry Inhibitors. Curr. Med. Chem. 2003, 10, 925–953. [Google Scholar] [CrossRef]
  3. Silvestri, R.; Marfe, G.; Artico, M.; la Regina, G.; Lavecchia, A.; Novellino, E.; Morgante, M.; di Stefano, C.; Catalano, G.; Filomeni, G.; Abruzzese, E.; Ciriolo, M.R.; Russo, M.A.; Amadori, S.; Cirilli, R.; la Torre, F.; Salimei, P.S. Pyrrolo[1,2-b][1,2,5]benzothiadiazepines (PBTDs): A New Class of Agents with High Apoptotic Activity in Chronic Myelogenous Leukemia K562 Cells and in Cells from Patients at Onset and Who Were Imatinib-Resistant. J. Med. Chem. 2006, 49, 5840–5844. [Google Scholar]
  4. Lebegue, N.; Gallet, S.; Flouquet, N.; Carato, P.; Pfeiffer, B.; Renard, P.; Le´once, S.; Pierre, A.; Chavatte, P.; Berthelot, P. Novel Benzopyridothiadiazepines as Potential Active Antitumor Agents. J. Med. Chem. 2005, 48, 7363–7373. [Google Scholar] [CrossRef]
  5. Brzozowski, F.; Saczewski, F.; Neamati, N. Synthesis and anti-HIV-1 activity of a novel series of 1,4,2-benzodithiazine-dioxides. Bioorg.Med. Chem. Lett. 2006, 16, 5298–5302. [Google Scholar] [CrossRef]
  6. Iyenger, B.S.; Dorr, R.T.; Alberts, D.S.; Solyom, A.M.; Krutzsch, M.; Remers, W.A. 1,4-Disubstituted Anthracene Antitumor Agents. J. Med. Chem. 1997, 40, 3734–3738. [Google Scholar] [CrossRef]
  7. Thurston, D.E.; Subhas, B.D.; Howard, P.W.; Jen-kins, T.C.; Leoni, A.; Baraldi, P.G.; Guiotto, A.; Cacciari, B.; Kelland, L.R.; Follope, M.; Rault, S.J. Effect of A-Ring Modifications on the DNA-Binding Behavior and Cytotoxicity of Pyrrolo[2,1-c][1,4]benzodiazepines. J. Med. Chem. 1999, 42, 1951–1964. [Google Scholar] [CrossRef]
  8. Basolo, L.; Beccalli, E.M.; Borsini, E.; Broggini, G.; Khansaa, M.; Rigamonti, M. Access to a Novel Class of Tetracyclic 1,4-Benzodiazepin-5-ones Starting from α-Amino Acids by Pd-Catalyzed Amination/1,3-Dipolar Cycloaddition as the Key Steps. Eur. J. Org. Chem. 2010, 2010, 1694–1703. [Google Scholar]
  9. Mowbray, C.E.; Burt, C.; Corbau, R.; Gayton, S.; Hawes, M.; Perros, I.; Tran, D.A.; Price, F.J.; Quinton, M.D.; Selby, P.A.; Stupple, R.; Webster, A.W. Pyrazole NNRTIs 4: Selection of UK-453,061 (lersivirine) as a Development Candidate. Bioorg. Med. Chem. Lett. 2009, 19, 5857–5860. [Google Scholar]
  10. Broggini, G.; Marchi, I.; Martinelli, M.; Paladino, G.; Pilati, T.; Terraneo, A. Effective Synthesis of Enantiopure [1,2,3]Triazolo[1,5-a] and Pyrazolo[1,5a]pyrrolo[2,1c][1,4] benzodiazepines by Diastereoselective Intramolecular Azide and Nitrilimine Cycloadditions. Synthesis 2005, 2246–2252. [Google Scholar]
  11. Hurley, L.H.; Petrusek, R.L. Proposed structure of the anthramycin-DNA adduct. Nature 1979, 282, 529–531. [Google Scholar] [CrossRef]
  12. Cheatham, S.; Kook, A.; Hurley, L.H.; Barkley, M.D.; Remers, W.J. One and two dimensional proton NMR, fluorescence and molecular modeling studies on the tomaymycin-d(ATGCAT)2 adduct. Evidence for two covalent adducts with opposite orientations and stereochemistries at the covalent linkage site. J. Med. Chem. 1988, 31, 583–590. [Google Scholar] [CrossRef]
  13. Wang, J.J.; Hill, G.C.; Hurley, L.H. Template-directed design of a DNA-DNA crosslinker based upon a bis-tomaymycin-duplex adduct. J. Med. Chem. 1992, 35, 2995–3002. [Google Scholar] [CrossRef]
  14. Montzouris, J.A.; Wang, J.J.; Thurston, D.E.; Hurley, L.H. Comparison of a DSB-120 DNA Interstrand Cross-Linked Adduct with the Corresponding Bis-tomaymycin Adduct: An Example of a Successful Template-Directed Approach to Drug Design Based upon the Monoalkylating Compound Tomaymycin. J. Med. Chem. 1994, 37, 3132–3140. [Google Scholar] [CrossRef]
  15. Boudjabi, S.; Dewynter, G.; Voyer, N.; Toupet, L.; Montero, J.L. Sulfahydantoins as Tripeptide Constraints: Synthesis and Structure of Chiral Substituted 3-Oxo-1,2,5-thiadiazolidine 1,1-Dioxides. Eur. J. Org. Chem. 1999, 2275–2283. [Google Scholar]
  16. Regainia, Z.; Abdaoui, M.; Aouf, N.E.; Dewynter, G.; Montero, J.L. Synthesis of 1,2,5-Thiadiazolidines 1,1-dioxides (Cyclosulfamides) Starting from Amino Acids and Chlorosulfonyl Isocyanate. Tetrahedron 2000, 56, 381–387. [Google Scholar] [CrossRef]
  17. Berredjem, M.; Djebbar, H.; Regainia, Z.; Aouf, N.E.; Dewynter, G.; Winum, J.Y.; Montero, J.L. Synthese et N-Acylation Regiospecifique de 1,2,5-Thiadiazolidines 1,1-Dioxydes Chirales. Phosphorus Sulfur 2003, 178, 693–705. [Google Scholar] [CrossRef]
  18. Regainia, Z.; Winum, J.Y.; Smain, F.Z.; Toupet, L.; Aouf, N.E.; Montero, J.L. General synthesis of n-membered cyclic sulfamides. Tetrahedron 2003, 59, 6051–6056. [Google Scholar] [CrossRef]
  19. Bendjeddou, A.; Djeribi, R.; Regainia, Z.; Aouf, N. N,N’-Substituted 1,2,5 Thiadiazolidine 1,1-Dioxides: Synthesis, Selected Chemical and Spectral Proprieties and Antimicrobial Evaluation. Molecules 2005, 10, 1387–1398. [Google Scholar] [CrossRef]
  20. Kumar, G.B.; Patel, H.V.; Shah, A.C.; Trenkle, M.; Cardin, C.J. Diastereoselective formation of 2-aryl-3-arenesulfonyl 4-ethyl-1,3-oxazolidines: An X-ray crystallographic and 1H NMR study. Tetrahedron-Asymmetry 1996, 7, 3391–3396. [Google Scholar]
  21. Wilken, J.; Wallbaum, S.; Saak, W.; Haase, D.; Pohl, S.; Laxmikant, N.; Patkar, N.; Arun, N.; Chittari, P.; Rajappa, S.; Martens, J. Utilization of Industrial Waste Materials, 6. Utilization of Derivatives of the Bicyclic Proline Analog (all-R)-Octahydrocyclopenta[b]pyrrol-2-carboxylic Acid in the Stereoselective Synthesis. Liebigs Ann. 1996, 927–934. [Google Scholar]
  22. Heydenreich, M.; Koch, A.; Lazar, L.; Szatmari, I.; Sillanpaa, R.; Kleinpeter, E.; Fulop, F. Synthesis and stereochemical studies of 1- and 2-phenyl-substituted 1,3-oxazino[4,3-a]isoquinoline derivatives. Tetrahedron 2003, 59, 1951–1959. [Google Scholar] [CrossRef]
  23. Perlmutter, P.; Tabone, M. A Simple Route to .alpha.-Substituted-.beta.-Amino Ester Precursors of Carbapenem Antibiotics. J. Org. Chem. 1995, 60, 6515–6522. [Google Scholar]
  24. Arnecke, R.; Bohmer, V.; Friebe, S.; Gebauer, S.; Krauss, G.J.; Thondorf, I.; Vgt, W. Regio- and diastereoselective condensation of resorcarenes with primary amines and formaldehyde. Tetrahedron Lett. 1995, 36, 6221–6224. [Google Scholar]
  25. Mitsunobu, O. The Use of Diethyl Azodicarboxylate and Triphenylphosphine in Synthesis and Transformation of Natural Products. Synthesis 1981, 1, 1–29. [Google Scholar] [CrossRef]
  26. Hughes, D.L. The Mitsunobu Reaction. Org. Reactions 1992, 42, 335–380. [Google Scholar]
  27. Takahashi, H.; Morimoto, I.; Higashiyama, K. Highly Diastereoselective Reaction of (3aS,6R)-5-Oxa-7,8a-diazaperhydroazulen-8-ones with Diethylzinc. Chem. Pharm. Bull. 1990, 38, 2627–2631. [Google Scholar] [CrossRef]
  28. Takahashi, H.; Morimoto, I.; Higashiyama, K. Asymmetirc Induction by Chiral Heterocyclic Compounds: Highly Diastereoselective Reaction of 5-Oxa-7,8a-diazaperhydroazulen-8-ones with Organometallic Reagents. Heterocycles 1990, 30, 287–290. [Google Scholar] [CrossRef]
  • Sample Availability: Samples of the compounds 1d4d and 1e4e are available from the authors.

Share and Cite

MDPI and ACS Style

Bendjeddou, A.; Abbaz, T.; Regainia, Z.; Aouf, N.-E. A New Class of Heterocycles: 1,4,3,5-Oxathiadiazepane 4,4-dioxides. Molecules 2012, 17, 1890-1899. https://doi.org/10.3390/molecules17021890

AMA Style

Bendjeddou A, Abbaz T, Regainia Z, Aouf N-E. A New Class of Heterocycles: 1,4,3,5-Oxathiadiazepane 4,4-dioxides. Molecules. 2012; 17(2):1890-1899. https://doi.org/10.3390/molecules17021890

Chicago/Turabian Style

Bendjeddou, Amel, Tahar Abbaz, Zine Regainia, and Nour-Eddine Aouf. 2012. "A New Class of Heterocycles: 1,4,3,5-Oxathiadiazepane 4,4-dioxides" Molecules 17, no. 2: 1890-1899. https://doi.org/10.3390/molecules17021890

Article Metrics

Back to TopTop