797
Views
11
CrossRef citations to date
0
Altmetric
Research Article

Synthesis and biological evaluation of fused oxepinocoumarins as free radicals scavengers

, , &
Pages 805-812 | Received 14 May 2010, Accepted 11 Jan 2011, Published online: 07 Mar 2011

Abstract

Some fused dihydrooxepino[f]-, [g]-, and [h]coumarins were obtained from the ring-closing metathesis of the corresponding o-allyl-allyloxycoumarins under the treatment with the first generation Grubbs’ catalyst. These compounds were tested in vitro for their antioxidant activity, and they present significant scavenging activity. They were also showed to inhibit in vitro soybean lipoxygenase.

Introduction

Coumarin derivatives are an interesting class of heterocyclic system, since the coumarin ring is an essential core moiety for a variety of natural and synthetic biologically active compoundsCitation1–3. In particular, coumarins fused with a ring containing an O-atom such as furocoumarins and pyranocoumarins are important as photochemotherapeuticCitation4–10 agents and exhibit antitumorialCitation11, antifungalCitation12, insecticidalCitation12, anticancerCitation12, anti-HIVCitation6,Citation13, anti-inflammatoryCitation3,Citation14, and antioxidantCitation3,Citation14 activities. The synthesis of those coumarins has been achieved mainly by formation of furan or pyran ring starting from hydroxycoumarins and using the tandem Claisen rearrangement-cyclization reactionCitation15,Citation16 of the intermediate propargyloxy- or allyloxycoumarinsCitation17–21. The Ru-catalyzed ring-closing metathesis (RCM)Citation22–26 has been applied in the synthesis of furan and pyran ring during the last decadeCitation25,Citation27–30 and especially in the synthesis of fused furo- or pyranocoumarinsCitation25,Citation27,Citation29. With this method, oxepinesCitation25,Citation27,Citation31–34, oxocinesCitation25,Citation27,Citation34,Citation35, or larger O-containing ringsCitation25,Citation36–38 have also been prepared. In the course of our interest on the synthesisCitation21,Citation39 of coumarin derivatives and the studyCitation3,Citation14,Citation40,Citation41 of their biological activities and in continuation to our previous work on RCM37,42, we wish to report here the synthesis of [6,5]-, [7,6]-, [7,8]- and [8,7]-fused oxepinocoumarins through the combination of Claisen rearrangement, allylation, and RCM starting from allyloxycoumarins.

There is an increasing interest in antioxidants, particularly in those intended to prevent the presumed deleterious effects of free radicals in the human body. Free radicals are molecules produced when human body breaks down foods, or by environmental exposures such as tobacco smoke and radiation and have been implicated in the pathology of more than 50 human diseases. Oxidative stress, occurring when antioxidant defenses are inadequate, can damage lipids, proteins, carbohydrates, and DNA. Several antioxidants are available for therapeutic use. They include molecules naturally present in the body as well as synthetic antioxidantsCitation43. Thus, we found interesting the biological screening of the resulted compounds as possible free-radicals scavengers and lipoxygenase (LOX) inhibitors. It is well known that free radicals play an important role in inflammatory processCitation44. Consequently compounds with antioxidant properties could be expected to offer protection in rheumatoid arthritis and inflammation and to lead to potentially effective drugs. The reactions studied and the products received are depicted in Schemes 1 to 3.

Materials and methods

Melting points were determined on a Kofler hot-stage apparatus and are uncorrected. Infrared (IR) spectra were obtained with a Perkin–Elmer 1310 spectrophotometer as KBr pellets or Nujol mulls. Nuclear magnetic resonance (NMR) spectra were recorded on a Bruker AM 300 (300 and 75 MHz for 1H and 13C, respectively) using CDCl3 as solvent and tetramethylsilane (TMS) as an internal standard. J values are reported in Hertz. Mass spectra were determined on a LCMS-2010 EV Instrument (Shimadzu) under electrospray ionization (ESI) conditions or on a VG-250 spectrometer at 70 eV under electron impact (EI) conditions. Microanalyses were performed on a Perkin–Elmer 2400-II Element analyzer. Silica gel (no. 60, Merck A.G.) was used for column chromatography. All the reagents used were commercially available. 1,1-Diphenyl-2-picrylhydrazyl (DPPH) and nordihydroguaiaretic acid (NDGA) were purchased from the Aldrich Chemical Co. (Milwaukee, WI). Soybean LOX, linoleic acid sodium salt, and 2,2-azobis(2-amidinopropane) dihydrochloride (AAPH) were obtained from Sigma Chemical, Co. (St. Louis, MO).

Synthesis

General procedure for the Claisen rearrangement of allyloxycoumarins

A solution of 2.133 g (10.5 mmol) of 7-allyloxycoumarinCitation45 in ethyleneglycol (90 mL) was refluxed under stirring for 9 h. Cold water (100 mL) was added and it was refrigerated overnight. The precipitated solid was filtered and separated by column chromatography [silica gel No. 60, hexane/ethyl acetate (2:1)] to give unreacted starting material (214 mg, 10%), 6-allyl-7-hydroxy-2H-chromen-2-one (4) (363 mg, 17%), m.p. 140–141°C (DCM–CH3OH) (lit45. 137–139°C) and 8-allyl-7-hydroxy-2H-chromene-2-one (3a) (1.45 g, 68%), m.p. 164–166°C (DCM–CH3OH) (lit46. 165–166°C).

General procedure for the allylation of o-hydroxy-allylcoumarins

To a solution of compound 3a (340 mg, 1.68 mmol) in dry acetone (30 mL) anhydrous K2CO3 (1.12 g, 8.1 mmol) was added, followed by allyl bromide (0.87 mL, 1.25 g, 10.3 mmol). The mixture was heated under reflux and stirring for 2 h and filtered while hot. The filtrate was concentrated and the residue was left for crystallization in the freezer to give 8-allyl-7-(allyloxy)-2H-chromene-2-one (5a) (325 mg, 80%), m.p. 82–84°C (DCM–hexane); IR(Nujol) 3060, 1710, 1600 cm−1; 1H NMR (CDCl3, 300 MHz) δ 3.64 (d, J =  6.4 Hz, 2H), 4.65 (d, J =  5.1 Hz, 2H), 4.99 (d, J =  11.5 Hz, 1H), 5.09 (d, J =  19.1 Hz, 1H), 5.31 (d, J =  8.9 Hz, 1H), 5.44 (d, J =  19.1 Hz, 1H), 5.92–6.10 (m, 2H), 6.25 (d, J =  8.9 Hz, 1H), 6.83 (d, J =  8.9 Hz, 1H), 7.30 (d, J =  8.9 Hz, 1H), 7.62 (d, J =  8.9 Hz, 1H); 13C NMR (CDCl3, 75 MHz) δ: 27.0, 69.3, 108.6, 110.0, 113.2, 115.5, 117.7, 120.9, 126.5, 132.5, 139.2, 143.7, 148.9, 159.3, 161.9; MS (EI) 242 (M+·, 20%), 214 (9), 201 (63), 187 (58), 173 (44), 145 (30), 117 (17), 115 (100). Anal. Calcd for C15H14O3: C, 74.36; H, 5.83. Found: C, 74.70; H, 6.14.

General procedure for the RCM reaction of o-allyl-allyloxycoumarins

The catalyst 6 (12.2 mg, 0.015 mmol) was added to a solution of derivative 5a (105 mg, 0.43 mmol) in dry dichloromethane (DCM) (50 mL) after removing of the air by a pump and introducing Argon (repeating in three cycles). The solution was stirred (the air was removed at the beginning and Argon was passed in the solution for three times again) for 4 h, a new amount of the catalyst 6 (3.3 mg, 0.004 mmol, 4.3 mol% totally) was added and the stirring was continued for 20 h more (totally 24 h stirring). After the evaporation of the solvent, the residue was separated by column chromatography (silica gel No. 60, DCM) to give 8,11-dihydro-2H-oxepino[2,3-h]chromen-2-one (7a), (83 mg, 90%), m.p. 119–121°C (DCM–hexane); IR(Nujol) 3070, 1695, 1595 cm−1; 1H NMR (CDCl3, 300 MHz) δ 3.78 (d, J =  3.8 Hz, 2H), 4.65 (d, J =  3.8 Hz, 2H), 5.58 (dt, J1 = 11.5 Hz, J2 = 3.8 Hz,1H), 5.87–5.97 (m, 1H), 6.31 (d, J =  8.9 Hz, 1H), 6.99 (d, J =  8.9 Hz, 1H), 7.30 (d, J =  8.9 Hz, 1H), 7.67 (d, J =  8.9 Hz, 1H); 13C NMR (CDCl3, 75 MHz) δ: 22.2, 70.8, 114.2, 115.2, 118.2, 123.4, 125.9, 126.6, 127.4, 143.7, 151.4, 160.8, 162.1; MS (EI) 214 (M+·, 87%), 186 (20), 185 (43), 170 (100), 160 (16), 158 (25), 157 (30), 142 (55). Anal. Calcd for C13H10O3: C, 72.89; H, 4.71. Found: C, 72.97; H, 4.55.

8-Allyl-7-hydroxy-4-methyl-2H-chromen-2-one (3b)

(80% yield), m.p. 197–199°C (EtOH) (lit46. 198–199°C)

8-Allyl-7-(allyloxy)-4-methyl-2H-chromene-2-one (5b) (90% yield), 92–93°C (acetone) (lit27. 94°C).

4-Methyl-8,11-dihydro-2H-oxepino[2,3-h]chromen-2-one (7b) (90% yield− 0.9 mol% of 6 was added at once), m.p. 109–111°C (DCM); IR(Nujol) 3060, 1690, 1590 cm−1; 1H NMR (CDCl3, 300 MHz) δ 2.41 (s, 3H), 3.78 (s, 2H), 4.65 (s, 2H), 5.56 (d, J1 = 11.4 Hz, 1H), 5.87–5.98 (m, 1H), 6.19 (s, 1H), 7.01 (d, J =  7.6 Hz, 1H), 7.43 (d, J =  7.6 Hz, 1H); 13C NMR (CDCl3, 75 MHz) δ: 18.8, 22.3, 70.8, 113.0, 116.3, 117.8, 123.2, 123.4, 126.0, 127.3, 150.8, 152.6, 160.8, 161.9; MS (EI) 228 (M+·, 94%), 213 (92), 200 (29), 199 (74), 186 (97), 185 (78), 184 (14), 156 (10), 128 (100). Anal. Calcd for C14H12O3: C, 73.67; H, 5.30. Found: C, 73.61; H, 5.12.

6-Allyl-7-(allyloxy)-2H-chromene-2-one (8) (91% yield), m.p. 94–95°C (acetone); IR(Nujol) 3060, 1700, 1600 cm−1; 1H NMR (CDCl3, 300 MHz) δ 3.42 (d, J =  5.9 Hz, 2H), 4.61 (d, J =  4.9 Hz, 2H), 5.09 (d, J =  12.8 Hz, 1H), 5.10 (d, J =  11.8 Hz, 1H), 5.34 (d, J =  11.8 Hz, 1H), 5.46 (d, J =  18.7 Hz, 1H), 5.85–6.14 (m, 2H), 6.24 (d, J =  8.9 Hz, 1H), 6.77 (s, 1H), 7.22 (s, 1H), 7.64 (d, J =  8.9 Hz, 1H); 13C NMR (CDCl3, 75 MHz) δ: 33.7, 69.1, 99.7, 112.0, 112.9, 116.3, 117.9, 126.3, 128.0, 132.1, 135.8, 143.4, 154.5, 159.3, 161.3; MS (EI) 242 (M+·, 64%), 214 (8), 201 (61), 173 (39), 145 (30), 117 (44), 115 (100), 91 (49). Anal. Calcd for C15H14O3: C, 74.36; H, 5.83. Found: C, 74.52; H, 5.59.

6,9-Dihydro-2H-oxepino[3,2-g]chromen-2-one (9) (97% yield− 3.9 mol% of 6 was added in four portions during 15 h), m.p. 118–120°C (DCM–hexane); IR (KBr) 3080, 2927, 1732, 1622, 1561 cm−1; 1H NMR (CDCl3, 300 MHz) δ 3.53 (d, J =  3.0, 2H), 4.67 (dd, J1 = 3.0, J2 = 2.0, 2H), 5.52 (dt, J1 = 11.8 Hz, J2 = 2.0 Hz,1H), 5.86–5.97 (m, 1H), 6.34 (d, J =  9.9 Hz, 1H), 7.04 (s, 1H), 7.21 (s, 1H), 7.43 (d, J =  9.9 Hz, 1H); 13C NMR (CDCl3, 75 MHz) δ: 31.1, 71.6, 110.2, 115.0, 115.1, 125.8, 127.2, 127.7, 132.9, 143.0, 155.0, 160.9, 161.8; MS (ESI) 215 [M+H]+, 237 [M+Na]+. Anal. Calcd for C13H10O3: C, 72.89; H, 4.71. Found: C, 73.01; H, 4.97.

5-Allyl-6-hydroxy-4-methyl-2H-chromen-2-one (12) (78% yield, after the reflux of compound 11 in ethyleneglycol for 16 h), m.p. 176–178°C (EtOH) (lit47. 176–177°C).

5-Allyl-6-(allyloxy)-4-methyl-2H-chromene-2-one (13) (83% yield, after 4 h of refluxing), m.p. 62–64°C (acetone); IR (KBr) 3097, 2978, 2934, 1695, 1598, 1563 cm−1; 1H NMR (CDCl3, 300 MHz) δ 2.67 (s, 3H), 3.87 (m, 2H), 4.57 (dd, J1 = 4.9 Hz, J2 = 2.0 Hz, 2H), 4.75 (dq, J1 = 18.7 Hz, J2 = 2.0 Hz, 1H), 5.08 (dq, J1 = 10.8 Hz, J2 = 2.0 Hz, 1H), 5.28 (dt, J1 = 11.8 Hz, J2 = 3.0 Hz, 1H), 5.41 (dt, J1 = 18.7 Hz, J2 = 2.0 Hz, 1H), 5.96–6.13 (m, 2H), 6.25 (s, 1H), 7.12 (d, J =  8.9 Hz, 1H), 7.23 (d, J =  8.9 Hz, 1H); 13C NMR (CDCl3, 75 MHz) δ: 24.2, 30.6, 70.2, 115.5, 116.3, 116.4, 117.3, 117.7, 120.0, 126.1, 133.0, 136.8, 149.1, 153.4, 153.6, 160.3; MS (EI) 256 (M+·, 73%), 215 (27), 201 (21), 200 (42), 187 (16), 171 (18), 115 (48), 91 (40), 41 (100). Anal. Calcd for C16H16O3: C, 74.97; H, 6.30. Found: C, 75.12; H, 6.27.

4-Methyl-8,11-dihydro-3H-oxepino[3,2-f]chromen-3-one (14) (98% yield− 2.2 mol% of 6 was added in four portions during 15 h), m.p. 126–128°C (DCM–hexane); IR (KBr) 3079, 2939, 2843, 1732, 1593, 1568 cm−1; 1H NMR (CDCl3, 300 MHz) δ 2.63 (s, 3H), 3.90 (d, J =  5.9 Hz, 2H), 4.64 (t, J =  2.0 Hz, 2H), 5.46 (dt, J1 = 10.8 Hz, J2 = 2.0 Hz, 1H), 5.88–5.98 (m, 1H), 6.26 (s, 1H), 7.19 (d, J =  8.9 Hz, 1H), 7.28 (d, J =  8.9 Hz, 1H); 13C NMR (CDCl3, 75 MHz) δ: 25.0, 26.1, 72.0, 116.4, 117.5, 120.8, 124.6, 124.9, 128.4, 136.2, 150.9, 152.7, 155.5, 159.1; MS (ESI) 229 [M+H]+, 251 [M+Na]+. Anal. Calcd for C14H12O3: C, 73.67; H, 5.30. Found: C, 73.85; H, 5.51.

7-Allyl-8-hydroxy-2H-chromen-2-one (17) (89% yield, after the reflux of compound 16 in ethyleneglycol for 20 h), m.p. 151–153°C (DCM–hexane) (lit48. 154°C).

7-Allyl-8-(allyloxy)-2H-chromene-2-one (18) (82% yield), m.p. 51–53°C (acetone); IR (Nujol) 3040, 1715, 1590 cm−1; 1H NMR (CDCl3, 300 MHz) δ 3.50 (d, J =  6.9 Hz, 2H), 4.70 (d, J =  5.9 Hz, 2H), 5.07 (d, J =  17.7 Hz, 1H), 5.09 (d, J =  8.9 Hz, 1H), 5.25 (d, J =  10.8 Hz, 1H), 5.40 (d, J =  15.8 Hz, 1H), 5.88–6.0 (m, 1H), 6.05–6.18 (m, 1H), 6.36 (d, J =  9.8 Hz, 1H), 7.08 (d, J =  7.9 Hz, 1H), 7.15 (d, J =  7.9 Hz, 1H), 7.67 (d, J =  9.8 Hz, 1H); 13C NMR (CDCl3, 75 MHz) δ: 34.3, 74.6, 115.6, 116.5, 118.2, 118.4, 122.3, 125.4, 128.4, 133.5, 135.9, 137.5, 143.7, 147.1, 160.1; MS (ESI) 265 [M+Na]+. Anal. Calcd for C15H14O3: C, 74.36; H, 5.83. Found: C, 74.41; H, 5.96.

7,10-Dihydro-2H-oxepino[3,2-h]chromen-2-one (19) (83% yield− 6.9 mol% of 6 was added in three portions over 12 h), m.p. 109–111°C (ethyl acetate); IR (Nujol) 3050, 1710, 1590 cm−1; 1H NMR (CDCl3, 300 MHz) δ 3.58 (d, J =  2.6, 2H), 4.71 (d, J =  2.6, 2H), 5.53 (dt, J1 = 12.0 Hz, J2 = 2.6 Hz,1H), 5.80–5.91 (m, 1H), 6.39 (d, J =  9.5 Hz, 1H), 7.02 (d, J =  7.7 Hz, 1H), 7.15 (d, J =  7.7 Hz, 1H), 7.69 (d, J =  9.5 Hz, 1H); 13C NMR (CDCl3, 75 MHz) δ: 31.7, 70.8, 114.7, 117.0, 121.5, 123.7, 125.2, 126.6, 128.6, 142.5, 144.5, 155.3, 160.5; MS (ESI) 237 [M+Na]+. Anal. Calcd for C13H10O3: C, 72.89; H, 4.71. Found: C, 73.08; H, 4.60.

Biological assay

In vitro experiments

In the in vitro assays, each experiment was performed at least in triplicate and the standard deviation of absorbance was <10% of the mean.

Determination of the reducing activity of the stable radical DPPHCitation41

An equal volume of the compounds dissolved in dimethylsulfoxide (DMSO) was added to a solution of DPPH (0.1 mM) in absolute ethanol. Ethanol was used as the control solution. The concentration of the solutions of the compounds was 0.1 mM. After 20 and 60 min at room temperature, the absorbance was recorded at 517 nm (). NDGA was used as a standard.

Table 1.  Interaction % with DPPH at 0.1 mM; competition % of compounds with DMSO for hydroxyl radical (HO %); inhibition of lipid peroxidation at 0.1 mM (LP %); in vitro inhibition of soybean LOX IC50 μM.

Competition of the tested compounds with DMSO for hydroxyl radicalsCitation49

The hydroxyl radicals generated by the Fe3+/ascorbic acid system, were detected according to Nash, by the determination of formaldehyde produced from the oxidation of DMSO. The reaction mixture contained EDTA (0.1 mM), Fe3+ (167 mM), DMSO (33 mM) in phosphate buffer (50 mM, pH 7.4), the tested compounds (concentration 0.1 mM), and ascorbic acid (10 mM). After 30 min of incubation (37°C), the reaction was stopped with CCl3COOH (17% w/v) (). Trolox was used as a standard.

Inhibition of linoleic acid lipid peroxidationCitation41

Production of conjugated diene hydroperoxide by oxidation of linoleic acid sodium salt in an aqueous solution was monitored at 234 nm. AAPH was used as a free-radical initiator. Ten microliters of the 16 mM linoleic acid sodium salt solution was added to the UV cuvette containing 0.93 mL of 0.05 M phosphate buffer, pH 7.4 prethermostated at 37°C. The oxidation reaction was initiated at 37°C under air by the addition of 50 μL of 40 mM AAPH solution. Oxidation was carried out in the presence of aliquots (10 μL) of oxepincoumarins. In the assay without antioxidant, lipid oxidation was measured in the presence of the same level of DMSO. The rate of oxidation at 37°C was monitored by recording the increase in absorption at 234 nm caused by conjugated diene hydroperoxides.

Soybean LOX inhibition study in vitro

In vitro study was evaluated as reported previouslyCitation41. The tested compounds dissolved in ethanol were incubated at room temperature with sodium linoleate (0.1 mM) and 0.2 mL of enzyme solution (1/9 × 10−4 w/v in saline). The conversion of sodium linoleate to 13-hydroperoxylinoleic acid at 234 nm was recorded and compared with the appropriate standard inhibitor caffeic acid (IC50 600 mM) ().

Results and discussion

Synthesis

The treatment of 8-allyl-7-hydroxycoumarin (3a) [preparedCitation45 by the Claisen rearrangement under reflux in ethyleneglycol of 7-allyloxycoumarin (2a), received from umbelliferone (1a)] with allyl bromide and K2CO3 in dry acetone resulted in the 8-allyl-7-(allyloxy) coumarin (5a) in 80% yield (). The RCM of 5a with the first generation Grubbs’ catalyst 6 (4.3 mol%, added in two portions during 4 h) in dichloromethane solution under stirring at room temperature furnished the dihydrooxepin derivative 7a in 90% yield.

Scheme 1.  Reagents and conditions: (i) allyl bromide, K2CO3, acetone (dry), reflux, 2 h; (ii) ethyleneglycol, reflux, 9 h; and (iii) 6, DCM (dry), room temperature, 24 h.

Scheme 1.  Reagents and conditions: (i) allyl bromide, K2CO3, acetone (dry), reflux, 2 h; (ii) ethyleneglycol, reflux, 9 h; and (iii) 6, DCM (dry), room temperature, 24 h.

The Claisen rearrangement of 7-allyloxy-4-methylcoumarin (2b)50 in refluxing ethyleneglycol for 9 h, in analogous way to 2a, gave 8-allyl-7-hydroxy-4-methylcoumarin (3b) (80% yield). Allylation of 3b provided 8-allyl-7-(allyloxy)-4-methylcoumarin (5b)27 (90%). From the RCM of 5b with the catalyst 6 (0.9 mol%, added at once) in dichloromethane solution under stirring at room temperature, the dihydrooxepin derivative 7b was obtained in 90% yield. The same reaction was performed in refluxing dichloromethane (10 mol% of catalyst), and the product 7b was received in 33% yield along with a pyrano[7,8]coumarin derivative (24%)Citation27.

The 6-allyl-7-hydroxycoumarin (4) (isolatedCitation45 also from the mixture of the Claisen rearrangement reaction of 2a) allylated with allyl bromide and gave the 6-allyl-7-(allyloxy) coumarin (8) (91%). The RCM reaction of derivative 8 with the catalyst 6 (3.9 mol%) (added in four portions during 15 h) in dichloromethane solution at room temperature resulted in the dihydrooxepin compound 9 in 97% yield.

The allylation of 5-allyl-6-hydroxy-4-methylcoumarin (12)47 (prepared in 78% yield by the heating under reflux of allyloxy derivative 11 in ethyleneglycol) provided the 5-allyl-6-allyloxy-4-methylcoumarin (13) (83%). The RCM reaction of 5-allyl-6-(allyloxy)coumarin derivative 13 with the catalyst 6 (2.2 mol%, added in four portions during 15 h) led to the dihydrooxepin derivative 14 in 98% yield ().

Scheme 2.  Reagents and conditions: (i) allyl bromide, K2CO3, acetone (dry), reflux, 2 h; (ii) ethyleneglycol, reflux, 9 h; and (iii) 6, DCM (dry), room temperature, 24 h.

Scheme 2.  Reagents and conditions: (i) allyl bromide, K2CO3, acetone (dry), reflux, 2 h; (ii) ethyleneglycol, reflux, 9 h; and (iii) 6, DCM (dry), room temperature, 24 h.

The Claisen rearrangement of 8-allyloxycoumarin (16)48 in refluxing ethyleneglycol resulted in 7-allyl-8-hydroxycoumarin (17) (89%)Citation48. The later allylated with allyl bromide and gave 7-allyl-8-(allyloxy)coumarin (18) (82%), which by RCM reaction with the catalyst 6 (6.9 mol%, added in three portions over 12 h) furnished the dihydrooxepin derivative 19 in 83% yield ().

Scheme 3.  Reagents and conditions: (i) allyl bromide, K2CO3, acetone (dry), reflux, 2 h; (ii) ethyleneglycol, reflux, 9 h; and (iii) 6, DCM (dry), room temperature, 24 h.

Scheme 3.  Reagents and conditions: (i) allyl bromide, K2CO3, acetone (dry), reflux, 2 h; (ii) ethyleneglycol, reflux, 9 h; and (iii) 6, DCM (dry), room temperature, 24 h.

In all the above cases, the RCM reaction product received as a sole product in excellent yield. The loading of the catalyst usually is more than one portion.

Biological studies

Herein the antioxidant activity was evaluated in several in vitro tests. In view of the differences among the test systems available, the results of a single assay can give only a suggestion on the protective potential of tested compounds. Thus, we have used three different types of assays to measure in vitro antioxidant activity of fused oxepino coumarins: (a) interaction with the stable free-radical DPPH, (b) competition with DMSO for hydroxyl radicals, and (c) interaction with the water soluble azo compound AAPH. All the assays require a spectrophotometric measurement and a certain reaction time to obtain reproducible resultsCitation51.

The DPPH test is very useful in the micromolar range demanding minutes to hours for both lipophilic and hydrophilic samples and indicates their reducing ability in an iron-free system. The interaction of the examined compounds with the stable free-radical DPPH was studied by the use of the stable 1,1-diphenyl-2-picrylhydrazyl radical DPPHCitation41 at 0.1 mM after 20 and 60 min (). The results showed that this interaction was very low, if any, compared with the reference compound NDGA. Small changes were observed with the time and only for compounds 7a and 7b, the rest remain unchanged (). Compound 14 does not present any interaction under the reported conditions. The low interaction values compared with NDGA should be mainly attributed to the absence of easily oxidized functionalities like the ones present in NDGA (two catechol subunits). Lipophilicity is not well correlated with the results. There is no evidence for any structural characteristic of the tested compounds that is correlated with the antioxidant activity. The presence of the coumarin nucleus is implicated by itself in the reducing procedure.

It has been claimed that hydroxyl radical scavengers could serve as protectors, thus decreasing prostaglandin synthesis. During the inflammatory process, phagocytes generate the superoxide anion radical at the inflamed site and this is connected to other oxidizing species such as ·OH that are among the most reactive oxygen species and are considered to be responsible for some of the tissue damage occurring during inflammationCitation52. The competition of compounds with DMSO for ·OH radicalsCitation49, generated by the Fe3+/ascorbic acid system, expressed as the inhibition of formaldehyde production was used for the evaluation of their hydroxyl radical scavenging activity. From the tested derivatives, only compounds 7a and 7b highly compete with DMSO (33 mM) at 0.1 mM in comparison with trolox (). Lower lipophilicity is well correlated with the results [clog P 7a (2.56) < clog P 7b (3.06); ]. Antioxidants of hydrophilic or lipophilic character are both needed to act as radical scavengers in the aqueous phase or as chain-breaking antioxidants in biological membranes.

AAPH-induced linoleic acid oxidation has been developed as a quick and a reliable method for measuring the antioxidant activity and provides a measure of how efficiently antioxidants protect against lipid oxidation in vitro. Oxidation of exogenous linoleic acid by a thermal free radical producer (AAPH) is followed by UV spectrophotometry in a highly diluted sampleCitation41. Compounds 9, 14, and 19 effectively inhibit AAPH-induced lipid peroxidation, showing higher activity than the reference compound trolox (14 and 19, ). Higher lipophilicity value (14 > 9) is correlated with higher lipid peroxidation inhibition (87% > 65%; ). It also seems that angular analogues (14 and 19) are more potent than the linear 9 (). The standard inhibitor trolox obviously exerts its inhibitory effect on lipid peroxidation mainly through the ability of its 6-hydroxy-5,7,8-trimethylchromane moiety to break the radical chain. Although no phenol moieties are present in oxepinocoumarins 9, 14, and 19, they could break the radical chain through the initial abstraction of hydrogen from the methyl group, e.g., compound 14, or through other mechanisms. The new radicals thus created could be efficiently stabilized through resonance.

LOXs play a role in membrane lipid peroxidation by forming hydroperoxides in the lipid bilayerCitation53. Inhibitors of LOX have attracted attention initially as potential agents for the treatment of inflammatory diseases, e.g. cancer and atheromatosisCitation54,Citation55. Our compounds were further evaluated for inhibition of soybean LOX by the UV absorbance-based enzyme assayCitation41. Compound 7b (IC50 180 μM) is the most active within the set compared with the reference compound caffeic acid, whereas compounds 9, 14, and 19 do not present any inhibition. The majority of the LOX inhibitors act as: (a) antioxidants or free radical scavengersCitation56, (b) inhibitors to reduce Fe3+ at the active site to the catalytically inactive Fe2+ (LOXs contain a “non-heme” iron per molecule in the enzyme active site, and (c) excellent ligands for Fe3+. On comparing our results, it seems that there is no correlation between their antioxidant ability and their LOX inhibitory activity (compounds 9, 14, and 19). This is in accordance with the finding of Curini et al.Citation57 who have studied the antioxidant and LOX inhibitory activity of five natural prenyloxycarboxylic acids and showed that the most efficient LOX inhibitor (boropinic acid) is not the most active DPPH radical scavenger. Lipophilicity is referredCitation58 as an important physicochemical property for LOX inhibitors, and the above tested derivatives 7a and 7b seem to follow this concept (clog P 7b > clog P 7a, IC50 7b > IC50 7a).

Conclusion

In this study, fused dihydrooxepinocoumarins are prepared in excellent yields as a sole product of the RCM reactions. The loading of the catalyst is more than one portion. The newly synthesized compounds present interesting biological activities. Our study indicates that LOX or lipid peroxidation inhibitory activity is not always accompanied by DPPH radical scavenging activity. Thus, although compounds such as 9, 14, and 19 inhibit lipid peroxidation potently they present low, if any, DPPH and hydroxyl radical scavenging activity. However, a better correlation exists between LOX and hydroxyl radical scavenging activity for compounds 7a and 7b.

Acknowledgement

We thank Dr C. Hansch and Biobyte Corp. 201West 4th Str., Suite 204, Claremont CA California, 91711, for free access to the C-QSAR program.

Declaration of interest

The authors report no conflict of interest.

References

  • Murray DH, J. Mendez, J, Brown SA. The Natural Coumarins: Occurrence, Chemistry and Biochemistry. New York: Wiley, 1982.
  • O’Kennedy R, Thornes RD. Coumarins: Biology, Applications and Mode of Action. Chichester: Wiley, 1997.
  • Fylaktakidou KC, Hadjipavlou-Litina DJ, Litinas KE, Nicolaides DN. Natural and synthetic coumarin derivatives with anti-inflammatory/antioxidant activities. Curr Pharm Des 2004;10:3813–3833.
  • Dall’Acqua F, Vedaldi D. Molecular basis of psoralen photochemotherapy. In: Horspool WM, Song PS, eds., CRC Handbook of Organic Photochemistry and Photobiology. Boca Raton, FL: CRC Press, 1995, pp. 1341–1350.
  • Santana L, Uriarte E, Roleira F, Milhazes N, Borges F. Furocoumarins in medicinal chemistry. Synthesis, natural occurrence and biological activity. Curr Med Chem 2004;11:3239–3261.
  • Kulkarni MV, Kulkarni GM, Lin CH, Sun CM. Recent advances in coumarins and 1-azacoumarins as versatile biodynamic agents. Curr Med Chem 2006;13:2795–2818.
  • Wulff H, Rauer H, Düring T, Hanselmann C, Ruff K, Wrisch A et al. Alkoxypsoralens, novel nonpeptide blockers of Shaker-type K+ channels: synthesis and photoreactivity. J Med Chem 1998;41:4542–4549.
  • Chimichi S, Boccalini M, Cosimelli B, Viola G, Vedaldi D, Dall’ Acqua F. A convenient synthesis of psoralens. Tetrahedron 2002;58:4859–4863.
  • Kitamura N, Kohtani S, Nakagaki R. Molecular aspects of furocoumarin reactions: Photophysics, photochemistry, photobiology, and structural analysis. J Photochem Photobiol C Photochem Rev 2005;6:168–185.
  • Mali RS, Pandhare NA, Sindkhedkar MD. Convenient two-step syntheses of seselin and angelicin derivatives. Tetrahedron Lett 1995;36:7109–7110.
  • Santana L, Uriarte E, Dalla Via L, Gia O. A new benzoangelicin with strong photobiological activity. Bioorg Med Chem Lett 2000;10:135–137.
  • Mali RS, Joshi PP, Sandhu PK, Manekar-Tilve A. Efficient syntheses of 6-prenylcoumarins and linear pyranocoumarins: Total synthesis of suberosin, toddaculin, O-methylapigravin (O-methylbrosiperin), O-methylbalsamiferone, dihydroxanthyletin, xanthyletin and luvangetin J Chem Soc Perkin Trans 2002;1:371–376.
  • Xie L, Takeuchi Y, Cosentino LM, McPhail AT, Lee KH. Anti-AIDS agents. 42. Synthesis and anti-HIV activity of disubstituted (3′R,4′R)-3′,4′-di-O-(S)-camphanoyl-(+)-cis-khellactone analogues. J Med Chem 2001;44:664–671.
  • Nicolaides DN, Gautam DR, Litinas KE, Hadjipavlou-Litina DJ, Fylaktakidou KC. Synthesis and evaluation of the antioxidant and antiinflammatory activities of some benzo[l]khellactone derivatives and analogues. Eur J Med Chem 2004;39:323–332.
  • Moody CJ. Claisen Rearrangement in heteroaromatic systems. In: Katritzky AR, ed., Advances in Heterocyclic Chemistry, Vol. 42. Orlando, FL: Academic Press, 1987, p. 231.
  • Lutz RP. Catalysis of cope and claisen rearrangements. Chem Rev 1984;84:205–247.
  • Saidi MR, Bigdeli K. Microwave promoted and improved thermal synthesis of pyranocoumarins and furocoumarins. J Chem Res 1998;800–801.
  • Saidi MR, Rajabi F. Microwave and BF3 promoted rearrangement of allyloxycoumarins to allylcoumarins and dihydrofurocoumarins. Heterocycles 2001;55:1805–1812.
  • Zhang Q, Chen Y, Xia Y, Yang Z, Xia P. Thermal ring closure reaction of 4-methyl-7-(1,1-disubstituted propyn-2-yloxy)chromen-2-ones: the effects of the substituents at propargylic position on reactivity and products. Synth Commun 2004;34:4507–4515.
  • Garazd MM, Garazd YL, Khilya VP. Neoflavones. 2. Methods for synthesizing and modifying 4-arylcoumarins. Chem Nat Comp 2005;41:245–271.
  • Litinas KE, Symeonidis T. Convenient synthesis of fused pyran[3,2-h]- and furo[3,2-h]coumarins from naphthalene-2,3-diol. Tetrahedron 2010;66:1289–1293.
  • Grubbs RH, Chang S. Recent advances in olefin metathesis and its application in organic synthesis. Tetrahedron 1998;54:4413–4450.
  • Armstrong SK. Ring closing diene metathesis in organic synthesis. J Chem Soc Perkin Trans 1998;1:371–388.
  • Fürstner A. Olefin Metathesis and Beyond. Angew Chem Int Ed Engl 2000;39:3012–3043.
  • Deiters A, Martin SF. Synthesis of oxygen- and nitrogen-containing heterocycles by ring-closing metathesis. Chem Rev 2004;104:2199–2238.
  • Vougioukalakis GC, Grubbs RH. Ruthenium-based heterocyclic carbene-coordinated olefin metathesis catalysts. Chem Rev 2010;110:1746–1787.
  • Chattopadhyay SK, Maity S, Panja S. Combined Claisen rearrangement and ring-closing metathesis as a route to oxepin- and oxocin-annulated coumarins. Tetrahedron Lett 2002;43:7781–7783.
  • van Otterlo WAL, Ngidi EL, de Koning CB. Sequential isomerisation and ring-closing metathesis: Masked styryl and vinyloxyaryl groups for the synthesis of benzo-fused heterocycles. Tetrahedron Lett 2003;44:6483–6486.
  • Tsai T-W, Wang E-C. A new synthesis of angelicin from 7-hydroxycoumarin via C-propenation-O-vinylation and ring-closing metathesis. J Chin Chem Soc 2004;51:1019–1023.
  • Virolleaud M-A, Piva O. Selective formation of dihydropyran derivatives by a tandem domino ring-closing metathesis/cross-metathesis. Tetrahedron Lett 2007;48:1417–1420.
  • Snyder NL, Haines HM, Peczuh MW. Recent developments in the synthesis of oxepins. Tetrahedron 2006;62:9301–9320.
  • Majumdar KC, Rahaman H, Islam R, Roy B. Tandem Claisen rearrangement and ruthenium catalyzed enyne bond reorganization as a route to the synthesis of tricyclic 1,8-naphthyridinones. Tetrahedron Lett 2006;47:2111–2113.
  • Chattopadhyay SK, Roy SP, Ghosh D, Biswas G. Synthesis of oxepine-, oxocine- and azepine-annulated carbazole derivatives by combined Claisen rearrangement and diene/enyne metathesis. Tetrahedron Lett 2006;47:6895–6898.
  • Chattopadhyay SK, Dey R, Biswas S. Regioselective synthesis of oxepin- and oxocin-annulated 2-quinolones. Synthesis 2005;403–406.
  • Buszek KR, Sato N, Jeong Y. Total synthesis of octalactin A via ring-closing metathesis reaction. Tetrahedron Lett 2002;43:181–184.
  • Fürstner A, Langemann K. Conformationally Unbiased Macrocyclization Reactions by Ring Closing Metathesis. J Org Chem 1996;61:3942–3943.
  • Litinas KE, Salteris BE. Unsaturated macrocyclic lactone synthesis via catalytic ring-closing metathesis. J Chem Soc Perkin Trans 1997;1:2869–2872.
  • Mohapatra DK, Ramesh DK, Giardello MA, Chorghade MS, Gurjar MK, Grubbs RH. Protecting group directed ring-closing metathesis (RCM): the first total synthesis of an anti-malarial nonenolide. Tetrahedron Lett 2007;48:2621–2625.
  • Baldoumi V, Gautam DR, Litinas KE, Nicolaides DN. Convenient synthesis of linear pyrano[3,2-g]-, [2,3-g]- and angular pyrano[3,2-f]coumarins from 4[(1,1-dimethyl-2-propynyl)oxy]phenol. Tetrahedron 2006;62:8016–8020.
  • Kontogiorgis C, Litinas KE, Makri A, Nicolaides DN, Vronteli A, Hadjipavlou-Litina DJ et al. Synthesis and biological evaluation of novel angular fused Pyrrolocoumarins. J Enzyme Inhib Med Chem 2008;23:43–49.
  • Symeonidis T, Chamilos M, Hadjipavlou-Litina DJ, Kallitsakis M, Litinas KE. Synthesis of hydroxycoumarins and hydroxybenzo[f]- or [h]coumarins as lipid peroxidation inhibitors. Bioorg Med Chem Lett 2009;19:1139–1142.
  • Gallos JK, Koftis TV, Sarli VC, Litinas KE. A straightforward synthesis of perbenzylated conduritols from alditols by ring closing olefin metathesis. J Chem Soc Perkin Trans 1999; 1:3075–3077.
  • Halliwell B. Drug antioxidant effects. A basis for drug selection? Drugs 1991;42:569–605.
  • Saldanha LA, Elias G, Rao MN. Oxygen radical scavenging activity of phenylbutenones and their correlation with antiinflammatory activity. Arzneimittelforschung 1990;40:89–91.
  • Loutfy MA. The Claisen rearrangement of umbelliferone allyl ether. Pharmazie 1979;34:672.
  • Kaufman KD. Synthetic Furocoumarins. I. A new synthesis of methyl-substituted psoralenes and isopsoralenes. J Org Chem 1961;26:117–121.
  • Kaufman KD, Keana JFW, Kelly RC, McBride DW, Slomp G. Synthetic furocoumarins. VI Analogs of psoralene derived from hydroquinone. J Org Chem 1962;27:2567–2572.
  • Guiotto A, Manzini P, Chilin A, Pastorini G, Rodighiero P. 13C-NMR spectra and carbon–proton coupling constants of variously annulated furocoumarins. J Heterocyclic Chem 1985;22:649–656.
  • Symeonidis T, Fylaktakidou KC, Hadjipavlou-Litina DJ, Litinas KE. Synthesis and anti-inflammatory evaluation of novel angularly or linearly fused coumarins. Eur J Med Chem 2009;44:5012–5017.
  • Baker W, Lothian OM. Studies in chelation. Part II. The stabilisation of Kekulé forms in o-hydroxyacetophenones. J Chem Soc 1935;628–633.
  • Kulisic T, Radonic A, Katalinic V, Milos M. Use of different methods for testing antioxidative activity of oregano essential oil. Food Chem 2004;85:633–640.
  • Halliwell B, Gutterridge JMC. Free Radicals in Biology and Medicine. Oxford: Clarendon, 1989.
  • Kuhn H, Belkner J, Wiesner R, Brash AR. Oxygenation of biological membranes by the pure reticulocyte lipoxygenase. J Biol Chem 1990;265:18351–18361.
  • Shureiqi I, Lippman SM. Lipoxygenase modulation to reverse carcinogenesis. Cancer Res 2001;61:6307–6312.
  • Zhao L, Funk CD. Lipoxygenase pathways in atherogenesis. Trends Cardiovasc Med 2004;14:191–195.
  • Müller K. 5-Lipoxygenase and 12-lipoxygenase: attractive targets for the development of novel antipsoriatic drugs. Arch Pharm (Weinheim) 1994;327:3–19.
  • Curini M, Epifano F, Genovese S, Menghini L, Ricci D, Fraternale D, Giamperi L, Bucchini A, Bellacchio E. Lipoxygenase inhibitory activity of boropinic acid, active principle from Boronia pinnata. Nat Prod Commun 2006:1141–1145.
  • Pontiki E, Hadjipavlou-Litina D. Lipoxygenase inhibitors: a comparative QSAR study review and evaluation of new QSARs. Med Res Rev 2008;28:39–117.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.