1,578
Views
26
CrossRef citations to date
0
Altmetric
Research Paper

Discovery of curcumin inspired sulfonamide derivatives as a new class of carbonic anhydrase isoforms I, II, IX, and XII inhibitors

, , , , , , & show all
Pages 1274-1281 | Received 26 Aug 2017, Accepted 13 Sep 2017, Published online: 02 Oct 2017

Abstract

A series of curcumin inspired sulfonamide derivatives was prepared from various chalcones and 4-sulfamoyl benzaldehyde via Claisen–Schmidt condensation. All new compounds were assayed as inhibitors of four human isoforms of the metalloenzyme carbonic anhydrase (hCA, EC 4.2.1.1) isoforms hCA I, II, IX and XII. Interesting inhibitory activities were observed against all these isoforms. hCA I, an isoform involved in several eye diseases was inhibited moderately with KIs in the range of 191.8–904.2 nM, hCA II, an antiglaucoma drug target was highly inhibited by the new sulfonamides, with KIs in the range of 0.75–8.8 nM. hCA IX, a tumor-associated isoform involved in cancer progression and metastatic spread was potently inhibited by the new sulfonamides, with KIs in the range of 2.3–87.3 nM, whereas hCA XII, and antiglaucoma and anticancer drug target, was inhibited with KIs in the range of 6.1–71.8 nM. It is noteworthy that one of the new compounds, 5d, was found to be almost 9 times more selective against hCA II (KI = 0.89 nM) over hCA IX and hCA XII, whereas 5e was 3 and 70 times more selective against hCA II (KI = 0.75 nM) over hCA IX and hCA XII, respectively.

Introduction

Carbonic anhydrase (CAs, EC 4.2.1.1) enzymes are ubiquitous metalloproteins present in prokaryotes and eukaryotes, and they catalyze the fundamental biochemical process of carbon dioxide hydration (a reversible reaction producing a bicarbonate anion and a proton), being therefore one of the principal regulators of cellular pH homeostasisCitation1. These enzymes also take part in several vital biological processes such as carbon dioxide and bicarbonate ion transport, respiration, electrolyte secretion, bone resorption, gluconeogenesis, lipogenesis and ureagenesis among othersCitation2–6. Indeed, seven (α-, β-, γ-, δ-, ζ-, η-, and θ-) genetically different CA families are known so far, which incorporate different metal ions at their active site including, Zn(II) (in all classes), Fe(II) (in the γ-CAs), Co(II) (in δ-CAs), and Cd(II) (in ζ-CAs)Citation7,Citation8. To date, in mammals, 15 α-CA isozymes have been identified, that differ in subcellular localization, tissue distribution, and catalytic activityCitation1,Citation2. Some of these isozymes are cytosolic (CA I, CA II, CA III, CA VII, and CA XIII), others are membrane bound (CA IV, CA IX, CA XII, and CA XIV), two are mitochondrial (CA VA and CA VB), and one is secreted in saliva and milk (CA VI)Citation2,Citation9–12.

The potential of this enzyme family as an important class of biological targets for pharmacologic intervention was recognized several decades agoCitation13. An atypical expression level or dysregulated activities of these enzymes were proven to be linked with several human diseases, such as glaucoma (hCAs I, II, IV, and XII), cancer (hCAs IX and XII), some central-nervous system syndromes, including epilepsy, neuropathic pain, and idiopathic intracranial hypertension (hCAs I, II, and VII), edema (hCA II, IV, XII, and XIV), obesity (hCAs VA and VB), and osteoporosis (hCA IV and XIV)Citation9,Citation11,Citation14,Citation15. Currently, the known carbonic anhydrase inhibitors (CAIs) can be divided into various classes: those that coordinate to the active site metal ionCitation9 and those that do not interact with itCitation16. Primary sulfonamides are the central and historically most prominent class of CAIs, being discovered in 1940, with many representatives in clinical use for decadesCitation17. The sulfonamide functionality, denoted as a zinc-binding group (ZBG), coordinates in deprotonated form to the Zn(II) ion within the hCA active sites and establishes hydrogen bonds with a residues nearby (e.g. Thr199 in α-CAs)Citation3,Citation4. Such binding features are common among the active site architectures of all the 15 human isozymes, that all belong to the α-classCitation9,Citation14. Furthermore, mercaptophenols, ureates/hydroxamates, metal complexing anion inhibitors, and the bioisosteres of sulfonamides (such as sulfamates and sulfamides) exhibit CA inhibitory activity following a similar mechanismCitation9,Citation16. Some of the structures of CAIs [acetazolamide, methazolamide, brinzolamide, dorzolamide (antiglaucoma drugs), celecoxib (COX-2 selective nonsteroidal anti-inflammatory drug), topiramate (anticonvulsant drug), and indisulam (in clinical development as an anticancer agent)] in clinical use/development are depicted in . The main drawback associated with the use of CAIs is their lack of selectivity in inhibiting various isoforms, considering the fact that many of these isoforms are rather similar from the structural view point and even subcellular localizationCitation3,Citation4,Citation9, resulting thus in undesired side effectsCitation3,Citation4. Thus, it is still challenging to design selective/specific agents with distinct inhibition profiles (inhibitors or activators) for any of these isoforms. In the last decades, many efforts have been carried out to design isoform-selective sulfonamide-inhibitors by employing two principal methods: the ring and the tail approachesCitation9,Citation14,Citation18,Citation19. The first resides in modulating the ring (chiefly its chemical nature) directly connected to the sulfonamide ZBG, whereas the latter consists in attaching different tails to the aromatic/heterocyclic ring carrying various ZBGs, of the sulfonamide, sulfamide, sulfamate, carboxylate, hydroxamate, or dithiocarbamate typeCitation3,Citation4. This enabled modulation of the interactions that the ligand establishes with the middle and outer parts of the active site cavity, which are the most variable regions among the 15 hCA isoforms mentioned above, and led to a variety of isoform-selective CAIsCitation3,Citation4,Citation9.

Figure 1. Structures of some drugs and drug candidates possessing sulfonamide/sulfamate moiety.

Figure 1. Structures of some drugs and drug candidates possessing sulfonamide/sulfamate moiety.

Among the CA isoforms, CA I and CA II are the two major isozymes present at high concentrations in the cytosol in erythrocytes, and CA II (together with CA IX) is the most active among all the α-CAsCitation2,Citation9–12. Sulfonamides which are known as strong inhibitors of CA II, have been utilized as commercial drugs (, structures A, B, C, and D) to treat glaucoma or diuretics for a long periodCitation20. In view of lack of selectivity and side effects associated with the use of the existing drugs, exhaustive search for novel CAIs is ongoing either through synthesis of new derivatives of known drugs or from new molecular bases.

Systematic search for new chemotypes may be tackled by investigating compounds from natural productsCitation21,Citation22. Natural compounds such as resveratrol, catechin, silymarin, dobutamin, and curcumin (KI = 7.44 µM against hCA II isozyme) exhibit CA inhibitory activityCitation23. Curcumin, the main curcuminoid of popular Indian spice turmeric (Curcuma longa), was reported to possess neuroprotective properties, which may be effective in the prevention and treatment of glaucomaCitation24. Curcumin as well as its analogues and 4′-(phenylurenyl)chalcones () were proven to act as CAIsCitation25,Citation26. Intrigued by these findings and in continuation of our research into the synthesis and biological evaluation of curcumin inspired analoguesCitation27, herein, we wish to report curcumin inspired sulfonamide derivatives as a new class of CAIs against isoforms I, II, IX, and XII.

Figure 2. Design of curcumin inspired sulfonamide derivatives as hCA inhibitors.

Figure 2. Design of curcumin inspired sulfonamide derivatives as hCA inhibitors.

Materials and methods

Chemistry

All reagents and solvents were obtained from commercial suppliers and used without further purification. The reactions were monitored by thin layer chromatography (TLC), using MERCK pre-coated silica gel 60-F254 aluminum plates. Column chromatography with 60–120 mesh silica gel was used as separation and purification method. Ethyl acetate and hexane were used as eluents. Melting points were obtained on Stuart digital melting-point apparatus/SMP 30 and were uncorrected. All IR spectra were recorded on a Perkin Elmer, FT-IR spectrometer using KBr discs. 1H NMR spectra were recorded on an Avance NMR instrument operated at 500 MHz. 13C NMR spectra were recorded on an Avance NMR instrument operated at 125 MHz. Chemical shift values were reported in ppm with TMS as an internal reference and J values were given in Hertz. The following abbreviations were used for 1H NMR spectra to indicate the signal multiplicity: s (singlet), d (doublet), dd (doublet of doublet), and m (multiplet). HRMS were determined with Agilent QTOF mass spectrometer 6540 series instrument using ESI technique.

General procedure for the synthesis of 3a–j

To a stirred solution of appropriately substituted benzaldehydes 1a–j (1 mmol) in ethanol (3 ml) was added 0.5 ml of acetone (2) and 15% aqueous NaOH (1 ml) solution at 0 °C. The reaction was allowed to stir at room temperature till it was completed. The reaction mixture was evaporated to dryness, extracted twice with ethyl acetate, and the combined organic layers were dried over anhydrous Na2SO4 and concentrated under reduced pressure. The crude product was purified by column chromatography (Silica gel, 60–120 mesh, 9:1 hexane/ethyl acetate) to obtain the desired chalcones 3a–j in good to very good yields.

General procedure for the synthesis of 5a–j

To a stirred solution of chalcone 3a–j (0.5 mmol) in ethanol (3 ml) was added 15% aqueous NaOH (1 ml) solution and aldehyde 4 (0.5 mmol) at 0 °C. The resulting solution was stirred at room temperature till the complete consumption of starting materials was observed. The reaction mixture was evaporated to dryness, extracted twice with ethyl acetate, and the combined organic layers were dried over anhydrous Na2SO4 and concentrated under reduced pressure. The crude mass was purified by recrystallization in ethanol or ethyl acetate to give the pure product 5a–j in yields of 35–45%.

4-((1E,4E)-5-(4-isopropoxyphenyl)-3-oxopenta-1,4-dien-1-yl)benzenesulfonamide (5a)

Yellow solid, yield 42%; mp: 163–165 °C; IR (KBr, cm−1): νmax 3309.1, 3249.8, 2975.3, 2937.2, 1666.6, 1622.4, 1600.6, 1337.7, 1158.0, 1095.2; 1H NMR (500 MHz, DMSO-d6): δ 7.98 (d, J = 8.5 Hz, 2H), 7.88 (d, J = 8.4 Hz, 2H), 7.83–7.72 (m, 4H), 7.50–7.41 (m, 3H), 7.20 (d, J = 16.0 Hz, 1H), 7.01 (d, J = 8.8 Hz, 2H), 4.76–4.69 (m, 1H), 1.30 (d, J = 6.0 Hz, 6H); 13C NMR (125 MHz, DMSO-d6): δ 189.1, 156.3, 144.2, 143.8, 138.5, 131.0, 130.0, 129.2, 126.6, 125.9, 125.2, 123.7, 116.3, 69.9, 22.3, 22.2; HRMS (ESI): m/z Calcd for C20H22NO4S 372.1264, found 372.1262 [M + H]+.

4-((1E,4E)-5-(4-methoxyphenyl)-3-oxopenta-1,4-dien-1-yl)benzenesulfonamide (5b)

Yellow solid, yield 39%; mp: 192–193 °C; IR (KBr, cm−1): νmax 3356.5, 3289.2, 3071.1, 2934.3, 2838.6, 1645.3, 1597.0, 1331.3, 1161.3, 1094.7; 1H NMR (500 MHz, DMSO-d6): δ 7.98 (d, J = 8.4 Hz, 2H), 7.89 (d, J = 8.5 Hz, 2H), 7.84–7.75 (m, 4H), 7.52–7.43 (m, 3H), 7.21 (d, J = 16.0 Hz, 1H), 7.05 (d, J = 8.8 Hz, 2H), 3.83 (s, 3H); 13C NMR (125 MHz, DMSO-d6): δ 188.7, 161.8, 145.5, 143.8, 140.8, 138.5, 130.9, 129.3, 128.2, 127.6, 126.6, 123.9, 115.0, 55.8; HRMS (ESI): m/z Calcd for C18H18NO4S 344.0951, found 344.0950 [M + H]+.

4-((1E,4E)-5-(2,4-dimethoxyphenyl)-3-oxopenta-1,4-dien-1-yl)benzenesulfonamide (5c)

Yellow solid, yield 40%; mp: 185–186 °C; IR (KBr, cm−1): νmax 3321.3, 3258.6, 2939.6, 2838.9, 1644.6, 1600.6, 1328.7, 1159.9, 1098.9; 1H NMR (500 MHz, DMSO-d6): δ 7.97 (d, J = 8.6 Hz, 2H), 7.93 (s, 1H), 7.88 (d, J = 8.5 Hz, 3H), 7.79–7.74 (m, 3H), 7.39 (d, J = 16.1 Hz, 2H), 7.28 (d, J = 16.0 Hz, 2H), 3.91 (s, 3H), 3.85 (s, 3H); 13C NMR (125 MHz, DMSO-d6): δ 188.7, 163.6, 160.5, 145.5, 140.5, 138.6, 138.5, 130.8, 129.2, 129.0, 126.6, 123.5, 116.3, 106.9, 98.9, 56.3, 56.0; HRMS (ESI): m/z Calcd for C19H20NO5S 374.1057, found 374.1061 [M + H]+.

4-((1E,4E)-5-(2,5-dimethoxyphenyl)-3-oxopenta-1,4-dien-1-yl)benzenesulfonamide (5d)

Yellow solid, yield 38%; mp: 202–203 °C; IR (KBr, cm−1): νmax 3364.1, 3258.2, 2943.2, 2836.0, 1654.0, 1597.9, 1326.5, 1158.5, 1096.1; 1H NMR (500 MHz, DMSO-d6): δ 7.98 (d, J = 8.6 Hz, 2H), 7.87 (d, J = 8.5 Hz, 2H), 7.83–7.77 (m, 3H), 7.40 (d, J = 16.0 Hz, 2H), 7.21 (d, J = 16.1 Hz, 1H), 7.08–7.04 (m, 3H), 3.85 (s, 3H), 3.79 (s, 3H); 13C NMR (125 MHz, DMSO-d6): δ 188.7, 153.7, 153.2, 144.3, 143.3, 141.3, 138.0, 129.3, 126.6, 123.5, 118.5, 113.6, 113.5, 113.2, 56.5, 56.1; HRMS (ESI): m/z Calcd for C19H20NO5S 374.1057, found 374.1060 [M + H]+.

4-((1E,4E)-5-(3,4-dimethoxyphenyl)-3-oxopenta-1,4-dien-1-yl)benzenesulfonamide (5e)

Yellow solid, yield 40%; mp: 198–200 °C; IR (KBr, cm−1): νmax 3321.7, 3252.6, 3091.3, 2940.1, 2839.5, 1619.5, 1596.1, 1330.1, 1158.8, 1095.2; 1H NMR (500 MHz, DMSO-d6): δ 7.99 (d, J = 8.4 Hz, 2H), 7.89 (d, J = 8.5 Hz, 2H), 7.84–7.74 (m, 4H), 7.51 (d, J = 16.0 Hz, 2H), 7.23 (d, J = 16.0 Hz, 2H), 7.06 (d, J = 8.4 Hz, 1H), 3.85 (s, 3H), 3.83 (s, 3H); 13C NMR (125 MHz, DMSO-d6): δ 188.7, 151.7, 149.5, 145.5, 144.3, 140.7, 138.5, 130.5, 129.2, 128.1, 127.8, 126.8, 126.6, 124.2, 123.9, 112.1, 111.0, 56.5, 56.0; HRMS (ESI): m/z Calcd for C19H20NO5S 374.1057, found 374.1055 [M + H]+.

4-((1E,4E)-3-oxo-5-(2,3,4-trimethoxyphenyl)penta-1,4-dien-1-yl)benzenesulfonamide (5f)

Yellow solid, yield 35%; mp: 159–161 °C; IR (KBr, cm−1): νmax 3305.9, 3241.6, 2974.1, 2948.4, 1652.1, 1592.9, 1334.4, 1156.5, 1090.8; 1H NMR (500 MHz, DMSO-d6): δ 7.98 (d, J = 8.4 Hz, 2H), 7.89–7.84 (m, 3H), 7.80 (d, J = 16.1 Hz, 1H), 7.63 (d, J = 8.9 Hz, 1H), 7.47 (s, 2H), 7.42–7.36 (m, 1H), 7.33 (d, J = 16.0 Hz, 1H), 6.95 (d, J = 8.9 Hz, 1H), 3.88 (s, 6H), 3.78 (s, 3H); 13C NMR (125 MHz, DMSO-d6): δ 188.7, 156.2, 153.5, 145.5, 142.3, 140.9, 138.4, 138.1, 129.3, 129.0, 126.6, 124.5, 123.9, 121.3, 109.0, 61.9, 60.9, 56.5; HRMS (ESI): m/z Calcd for C20H22NO6S 404.1162, found 404.1169 [M + H]+.

4-((1E,4E)-3-oxo-5-(2,4,6-trimethoxyphenyl)penta-1,4-dien-1-yl)benzenesulfonamide (5g)

Yellow solid, yield 43%; mp: 249–251 °C; IR (KBr, cm−1): νmax 3304.9, 3253.2, 2969.1, 2951.9, 1616.8, 1591.7, 1335.2, 1157.2, 1090.1; 1H NMR (500 MHz, DMSO-d6): δ 8.07–8.01 (m, 1H), 7.97 (d, J = 8.4 Hz, 2H), 7.86 (d, J = 8.4 Hz, 2H), 7.64 (d, J = 16.0 Hz, 1H), 7.46–7.39 (m, 3H), 7.35 (d, J = 16.0 Hz, 1H), 6.32 (s, 2H), 3.91 (s, 6H), 3.86 (s, 3H); 13C NMR (125 MHz, DMSO-d6): δ 189.4, 161.8, 160.4, 145.1, 141.1, 137.3, 130.9, 129.2, 128.8, 126.6, 123.6, 115.9, 104.8, 90.1, 55.3, 54.5; HRMS (ESI): m/z Calcd for C20H22NO6S 404.1162, found 404.1166 [M + H]+.

4-((1E,4E)-3-oxo-5-(3,4,5-trimethoxyphenyl)penta-1,4-dien-1-yl)benzenesulfonamide (5h)

Yellow solid, yield 45%; mp: 242–244 °C; IR (KBr, cm−1): νmax 3301.8, 3240.4, 2940.2, 2834.7, 1639.5, 1590.5, 1319.8, 1157.0, 1093.9; 1H NMR (500 MHz, DMSO-d6): δ 8.11–7.99 (m, 2H), 7.97 (d, J = 8.5 Hz, 2H), 7.85 (d, J = 8.4 Hz, 2H), 7.63 (d, J = 15.9 Hz, 1H), 7.48–7.42 (m, 2H), 7.33 (d, J = 16.0 Hz, 1H), 6.93 (s, 2H), 3.91 (s, 6H), 3.86 (s, 3H); 13C NMR (125 MHz, DMSO-d6): δ 188.6, 153.2, 143.3, 142.7, 138.6, 130.5, 129.3, 127.8, 126.8, 123.7, 119.1, 104.1, 60.6, 56.5; HRMS (ESI): m/z Calcd for C20H22NO6S 404.1162, found 404.1164 [M + H]+.

4-((1E,4E)-5-(4-chlorophenyl)-3-oxopenta-1,4-dien-1-yl)benzenesulfonamide (5i)

Yellow solid, yield 43%; mp: 205–206 °C; IR (KBr, cm−1): νmax 3306.9, 3238.4, 2962.1, 2881.8, 1611.9, 1597.2, 1327.0, 1157.7, 659.6; 1H NMR (500 MHz, DMSO-d6): δ 7.99 (d, J = 8.2 Hz, 2H), 7.89 (d, J = 8.2 Hz, 2H), 7.86–7.81 (m, 4H), 7.55 (dd, J = 8.2, 3.3 Hz, 2H), 7.50–7.44 (m, 2H), 7.39 (dd, J = 16.0, 12.2 Hz, 2H); 13C NMR (125 MHz, DMSO-d6): δ 188.9, 145.6, 142.3, 142.0, 141.6, 138.3, 135.6, 135.5, 134.1, 130.7, 129.5, 129.3, 128.1, 126.7, 126.6; HRMS (ESI): m/z Calcd for C17H15ClNO3S 348.0456, found 348.0448 [M + H]+.

4-((1E,4E)-5-(2-fluoro-4-methoxyphenyl)-3-oxopenta-1,4-dien-1-yl)benzenesulfonamide (5j)

Yellow solid, yield 40%; mp: 167–169 °C; IR (KBr, cm−1): νmax 3289.8, 3241.1, 2944.6, 2845.9, 1628.8, 1615.3, 1320.3, 1272.0, 1157.4, 1091.3; 1H NMR (500 MHz, DMSO-d6): δ 7.99 (d, J = 8.5 Hz, 2H), 7.92–7.86 (m, 3H), 7.81 (d, J = 16.1 Hz, 2H), 7.50–7.43 (m, 3H), 7.31 (d, J = 16.1 Hz, 1H), 6.99 (dd, J = 13.0, 2.5 Hz, 1H), 6.94–6.90 (m, 1H), 3.85 (s, 3H); 13C NMR (125 MHz, DMSO-d6): δ 188.7, 161.7, 158.6, 145.6, 143.3, 141.5, 138.5, 130.8, 128.3, 126.6, 123.7, 116.8, 111.2, 106.7, 55.7; HRMS (ESI): m/z Calcd for C18H17FNO4S 362.0857, found 362.0856 [M + H]+.

CA inhibition assay

An Applied Photophysics stopped-flow instrument has been used for assaying the CA catalyzed CO2 hydration activityCitation29. Phenol red (at a concentration of 0.2 mM) has been used as indicator, working at the absorbance maximum of 557 nm, with 20 mM Hepes (pH 7.5) as buffer, and 20 mM Na2SO4 (for maintaining constant the ionic strength), following the initial rates of the CA-catalysed CO2 hydration reaction for a period of 10–100 s. The CO2 concentrations ranged from 1.7 to 17 mM for the determination of the kinetic parameters and inhibition constants. For each inhibitor at least six traces of the initial 5–10% of the reaction have been used for determining the initial velocity. The uncatalysed rates were determined in the same manner and subtracted from the total observed rates. Stock solutions of inhibitor (0.1 mM) were prepared in distilled–deionized water and dilutions up to 0.01 nM were done thereafter with the assay buffer. Inhibitor and enzyme solutions were preincubated together for 15 min at room temperature prior to assay, in order to allow for the formation of the E–I complex. The inhibition constants were obtained by non-linear least-squares methods using PRISM 3 and the Cheng–Prusoff equation, as reported earlierCitation30–32, and represent the mean from at least three different determinations. All CA isofoms were recombinant ones obtained in-house as reported earlierCitation30–32.

Results and discussion

Chemistry

The synthetic route for the preparation of the target compounds, 4-((1E,4E)-3-oxo-5-phenylpenta-1,4-dien-1-yl)benzenesulfonamides (5a–j) is illustrated in Scheme 1. In the first step, chalcones (3a–j) were prepared by NaOH-catalyzed Claisen–Schmidt condensation of a variety of benzaldehydes (1) with acetoneCitation28. The corresponding chalcones were reacted with 4-sulfamoyl benzaldehyde (4) using 15% NaOH by Claisen–Schmidt condensation to afford the target compounds 5a–j in good yields. All the synthesized compounds were purified by recrystallization in hot ethanol and well characterized by spectroscopic techniques such as 1H, 13C NMR, FT-IR and HRMS etc., which were in full accordance with the depicted structures. The 1H NMR spectrum of 4-((1E,4E)-5-(4-isopropoxyphenyl)-3-oxopenta-1,4-dien-1-yl)benzenesulfonamide (5a) showed a multiplet of isopropoxy (C–H) proton at δ 4.76–4.69, a doublet of two methyl group’s protons (total 6) of isopropoxy group at δ 1.30 and rest all protons appeared in the aromatic region in the range of δ 7.98–7.01. In the 13C NMR spectrum of 5a, the carbonyl carbon appeared at δ 189.1 and the remaining aromatic carbons appeared in the range of δ 156.3–116.3. Nearly similar pattern was observed in 1H and 13C NMR spectra of all the other compounds (5b–j) of this series. In the FT-IR spectrum, bands at 3309.1, 3249.8, 1337.7 and 1158.0 cm−1 confirmed the presence of sulfonamide whereas a band at 1666.6 cm−1 confirmed the presence of ketone functionality in compound 5a. The HRMS (ESI) of all the compounds showed an [M + H]+ peak equivalent to their molecular formulae.

Scheme 1. Synthesis of curcumin inspired sulfonamide derivatives (5a–j); Reagents and conditions: (i) 15% NaOH, ethanol, 0 °C – rt, 2–3 h, 72–85%; (ii) 15% NaOH, ethanol, 0 °C – rt, 1–2 h, 35–45%.

Scheme 1. Synthesis of curcumin inspired sulfonamide derivatives (5a–j); Reagents and conditions: (i) 15% NaOH, ethanol, 0 °C – rt, 2–3 h, 72–85%; (ii) 15% NaOH, ethanol, 0 °C – rt, 1–2 h, 35–45%.

CA inhibition

All compounds 5a–j were tested in vitro for their inhibitory activity against the physiologically relevant hCA isoforms I, II, IX, and XII by means of the stopped-flow carbon dioxide hydration assayCitation29 and their activities were compared to the standard CA inhibitor (CAI) acetazolamide (AAZ) ().

Table 1. In vitro CA I, II, IX, and XII inhibition with compounds 5a–j and acetazolamide (AAZ) as standard, by a stopped-flow, CO2 hydrase assay.Citation29

We have investigated the novel series of sulfonamide derivatives for their interaction with four hCAs of pharmacologic interest (i.e. isoforms hCA I, II, IX, and XII), using a period of 15 min of incubation of the enzyme and inhibitor solutionsCitation7,Citation8,Citation15. The following structure–activity-relationship (SAR) may be noted regarding the inhibition data of :

  1. Against the slow cytosolic isoform hCA I, almost all tested sulfonamide derivatives exhibited moderate inhibitory activity with KI in the range 191.8–904.2 nM. Methoxy substituents on the benzene ring showed an important role for the modulation of inhibition. Among them, compound 5f incorporating three MeO moieties in the 2, 3, 4 positions on the ring, prove to possess the highest inhibitory potency with a KI of 191.8 nM.

  2. hCA II, the dominant physiologic isoform, was effectively inhibited by all compounds here considered, in the low nanomolar range with KIs of 0.75–8.8 nM. Also for this isoform, the presence and position of methoxy groups in the chalcone synthon seems to play a crucial role for the inhibition potency. Among compounds with two OCH3 groups, derivative 5c showed the “worst” inhibition profile, with a 5 times decreased activity compared to 5d (4.7 nM and 0.89 nM), the two compounds only differing by the position of the two methoxy moieties in the second aryl functionality, the one coming from the chalcone. In addition, the position of OCH3 groups for compounds 5f–h, with three substituents, plays an important role in modulating the activity. However, all these compounds were highly effective hCA II inhibitors (all of them better than AAZ), with a very small variation of the KIs, and more detailed structure activity relationship is impossible to be delineated at this pot for a relatively small group of derivatives.

  3. hCA IX, the tumor-associated isoform, was inhibited by almost all compounds reported here, in the low nanomolar range (KI 2.3–8.6 nM) except for compounds 5i and 5j which were active in the medium nanomolar range (KI 61–87.3 nM). For this transmembrane isoform, the methoxy groups present on the scaffold did not influence significantly the inhibition profile. The interesting case for hCA IX inhibition was the presence of halogens in compounds 5i and 5j, that decreased the inhibition potency near 10 times compared to the other sulphonamides investigated here, which were devoid of such moieties.

  4. The last membrane isoform here considered here, hCA XII, was inhibited by compounds 5a–d and 5g in low nanomolar range (KI 6.1–7.4 nM). On the other hand, compounds 5e–f and 5h–j inhibited this isoform in medium nanomolar range (KI 31.1–71.8 nM). In analogy with hCA IX, the methoxy groups in the scaffold did not influence significantly the potency of inhibition.

Conclusions

In summary, a series of curcumin inspired sulfonamide derivatives (5a–j) were synthesized from chalcones and 4-sulfamoyl benzaldehyde by using Claisen–Schmidt condensation. The new sulfonamides were evaluated as inhibitors of four such isoforms, i.e. hCA I, II, IX, and XII. Interesting inhibitory activities were observed against almost all these isoforms as follows. These analogues inhibited hCA I (involved in some eye diseases) moderately with KIs in the range of 191.8–904.2 nM, hCA II (an antiglaucoma drug target) very potently with KIs in the range of 0.75–8.8 nM, hCA IX (an isoform involved in cancer) significantly with KIs in the range of 2.3–87.3 nM and hCA XII (antiglaucoma and anticancer drug target) with KIs in the range of 6.1–71.8 nM. Interestingly, 5d was found to be almost 9 times more selective against hCA II isoform (KI = 0.89 nM) over hCA IX and hCA XII isoforms and 5e was nearly 3 and 70 times more selective against hCA II isoform (KI = 0.75 nM) over hCA IX and hCA XII isoforms respectively.

Acknowledgements

PVSR and SA are thankful to DoP, Ministry of Chemicals & Fertilizers, Govt. of India, New Delhi, for the award of a NIPER fellowship.

Disclosure statement

No potential conflict of interest was reported by the authors.

References

  • (a) Supuran CT. Structure and function of carbonic anhydrases. Biochem J 2016;473:2023–32. (b) Supuran CT. Carbonic anhydrase inhibition and the management of neuropathic pain. Expert Rev Neurother 2016;16:961–8.
  • (a) Ekinci D, Al-Rashida M, Abbas G, et al. Chromone containing sulfonamides as potent carbonic anhydrase inhibitors. J Enzyme Inhib Med Chem 2012;27:744–7. (b) Sentürk M, Ekinci D, Goksu S, Supuran CT Effects of dopaminergic compounds on carbonic anhydrase isozymes I, II, and VI. J Enzyme Inhib Med Chem 2012;27:365–9. (c) Khalifah RG. Carbon dioxide hydration activity of carbonic anhydrase: paradoxical consequences of the unusually rapid catalysis. Proc Natl Acad Sci U S A 1973;70:1986–9.
  • (a) Supuran CT. Carbonic anhydrases: novel therapeutic applications for inhibitors and activators. Nat Rev Drug Discov 2008;7:168–81. (b) Margheri F, Ceruso M, Carta F, et al. Overexpression of the transmembrane carbonic anhydrase isoforms IX and XII in the inflamed synovium. J Enzyme Inhib Med Chem 2016;31:60–3. (c) Bua S, Di Cesare Mannelli L, Vullo D, et al. Design and synthesis of novel nonsteroidal anti-inflammatory drugs and carbonic anhydrase inhibitors hybrids (NSAIDs-CAIs) for the treatment of rheumatoid arthritis. J Med Chem 2017;60:1159–70.
  • Supuran CT. Advances in structure-based drug discovery of carbonic anhydrase inhibitors. Expert Opin Drug Discov 2017;12:61–88.
  • (a) Çavdar H, Ekinci D, Talaz O, et al. α-Carbonic anhydrases are sulfatases with cyclic diol monosulfate esters. J Enzyme Inhib Med Chem 2012;27:148–54. (b) Supuran CT, Capasso C. New light on bacterial carbonic anhydrases phylogeny based on the analysis of signal peptide sequences. J Enzyme Inhib Med Chem 2016;31:1254–60. (c) Capasso C, Supuran CT. An overview of the alpha-, beta- and gamma-carbonic anhydrases from Bacteria: can bacterial carbonic anhydrases shed new light on evolution of bacteria? J Enzyme Inhib Med Chem 2015;30:325–32.
  • (a) Ekinci D, Karagoz L, Ekinci D, et al. Carbonic anhydrase inhibitors: in vitro inhibition of α isoforms (hCA I, hCA II, bCA III, hCA IV) by flavonoids. J Enzyme Inhib Med Chem 2013;28:283–8. (b) Clare BW, Supuran CT. Carbonic anhydrase activators. 3: structure-activity correlations for a series of isozyme II activators. J Pharm Sci 1994;83:768–73. (c) Briganti F, Mangani S, Orioli P, et al. Carbonic anhydrase activators: X-ray crystallographic and spectroscopic investigations for the interaction of isozymes I and II with histamine. Biochemistry 1997;36:10384–92.
  • (a) Maresca A, Vullo D, Scozzafava A, et al. Inhibition of the β-class carbonic anhydrases from Mycobacterium tuberculosis with carboxylic acids. J Enzyme Inhib Med Chem 2013;28:392–6. (b) Güzel-Akdemir O, Akdemir A, Pan P, et al. A class of sulfonamides with strong inhibitory action against the α-carbonic anhydrase from Trypanosoma cruzi. J Med Chem 2013;56:5773–81. (c) Vullo D, Del Prete S, Osman SM, et al. Sulfonamide inhibition studies of the δ-carbonic anhydrase from the diatom Thalassiosira weissflogii. Bioorg Med Chem Lett 2014;24:275–9. (d) Capasso C, Supuran CT. Sulfa and trimethoprim-like drugs – antimetabolites acting as carbonic anhydrase, dihydropteroate synthase and dihydrofolate reductase inhibitors. J Enzyme Inhib Med Chem 2014;29:379–87. (e) Maresca A, Scozzafava A, Vullo D, Supuran CT. Dihalogenated sulfanilamides and benzolamides are effective inhibitors of the three β-class carbonic anhydrases from Mycobacterium tuberculosis. J Enzyme Inhib Med Chem 2013;28:384–7. (f) Vullo D, Del Prete S, Osman SM, et al. Sulfonamide inhibition studies of the γ-carbonic anhydrase from the oral pathogen Porphyromonas gingivalis. Bioorg Med Chem Lett 2014;24:240–4. (g) De Luca V, Vullo D, Scozzafava A, et al. An alpha-carbonic anhydrase from the thermophilic bacterium Sulphurihydrogenibium azorense is the fastest enzyme known for the CO2 hydration reaction. Bioorg Med Chem 2013;21:1465–9.
  • (a) Del Prete S, Vullo D, Fisher GM, et al. Discovery of a new family of carbonic anhydrases in the malaria pathogen Plasmodium falciparum—the η-carbonic anhydrases. Bioorg Med Chem Lett 2014;24:4389–96. (b) Maresca A, Vullo D, Scozzafava A, Supuran CT. Inhibition of the alpha- and beta-carbonic anhydrases from the gastric pathogen Helicobacter pylori with anions. J Enzyme Inhib Med Chem 2013;28:388–91. (c) Del Prete S, De Luca V, Scozzafava A, et al. Biochemical properties of a new α-carbonic anhydrase from the human pathogenic bacterium, Vibrio cholerae. J Enzyme Inhib Med Chem 2014;29:23–7.
  • Alterio V, Di Fiore A, D’Ambrosio K, et al. Multiple binding modes of inhibitors to carbonic anhydrases: how to design specific drugs targeting 15 different isoforms? Chem Rev 2012;112:4421–68.
  • Supuran CT. Structure-based drug discovery of carbonic anhydrase inhibitors. J Enzyme Inhib Med Chem 2012;27:759–72.
  • De Simone G, Alterio V, Supuran CT. Exploiting the hydrophobic and hydrophilic binding sites for designing carbonic anhydrase inhibitors. Expert Opin Drug Discov 2013;8:793–810.
  • Maren TH. Carbonic anhydrase: chemistry, physiology, and inhibition. Physiol Rev 1967;47:595–781.
  • Aggarwal M, Kondeti B, McKenna R. Insights towards sulfonamide drug specificity in α-carbonic anhydrases. Bioorg Med Chem 2013;21:1526–33.
  • (a) Supuran CT. How many carbonic anhydrase inhibition mechanisms exist? J Enzyme Inhib Med Chem 2016;31:345–60. (b) Carta F, Supuran CT, Scozzafava A. Sulfonamides and their isosters as carbonic anhydrase inhibitors. Future Med Chem 2014;6:1149–65.
  • (a) Neri D, Supuran CT. Interfering with pH regulation in tumours as a therapeutic strategy. Nat Rev Drug Discov 2011;10:767–77. (b) De Simone G, Supuran CT. Carbonic anhydrase IX: biochemical and crystallographic characterization of a novel antitumor target. Biochim Biophys Acta Proteins Proteomics 2010;1804:404–9. (c) Aggarwal M, Boone CD, Kondeti B, McKenna R. Structural annotation of human carbonic anhydrases. J Enzyme Inhib Med Chem 2013;28:267–77. (d) Supuran CT. Carbonic anhydrases: from biomedical applications of the inhibitors and activators to biotechnological use for CO2 capture. J Enzyme Inhib Med Chem 2013;28:229–30. (e) Supuran CT, Clare BW. Carbonic anhydrase inhibitors – part 57: quantum chemical QSAR of a group of 1,3,4-thiadiazole- and 1,3,4-thiadiazoline disulfonamides with carbonic anhydrase inhibitory properties. Eur J Med Chem 1999;34:41–50. (f) Winum JY, Supuran CT. Recent advances in the discovery of zinc-binding motifs for the development of carbonic anhydrase inhibitors. J Enzyme Inhib Med Chem 2015;30:321–4. (g) Capasso C, Supuran CT. An overview of the alpha-, beta- and gamma-carbonic anhydrases from bacteria: can bacterial carbonic anhydrases shed new light on evolution of bacteria? J Enzyme Inhib Med Chem 2015;30:325–32. (h) Supuran CT. Carbonic anhydrase inhibitors and activators for novel therapeutic applications. Future Med Chem 2011;3:1165–80.
  • (a) Maresca A, Temperini C, Pochet L, et al. Deciphering the mechanism of carbonic anhydrase inhibition with coumarins and thiocoumarins. J Med Chem 2010;53:335–44. (b) D’Ambrosio K, Carradori S, Monti SM, et al. Out of the active site binding pocket for carbonic anhydrase inhibitors. Chem Commun 2015;51:302–5. (c) Maresca A, Temperini C, Vu H, et al. Non-zinc mediated inhibition of carbonic anhydrases: coumarins are a new class of suicide inhibitors. J Am Chem Soc 2009;131:3057–62. (d) Angapelly S, Ramya PVS, Angeli A, et al. Discovery of 4-sulfamoyl-phenyl-β-lactams as a new class of potent carbonic anhydrase isoforms I, II, IV and VII inhibitors: the first example of subnanomolar CA IV inhibitors. Bioorg Med Chem 2017;25:539–44.
  • Winum J-Y, Scozzafava A, Montero J-L, Supuran CT. New zinc binding motifs in the design of selective carbonic anhydrase inhibitors. Mini Rev Med Chem 2006;6:921–36.
  • (a) Winum J-Y, Poulsen S-A, Supuran CT. Therapeutic applications of glycosidic carbonic anhydrase inhibitors. Med Res Rev 2009;29:419–35. (b) Bozdag M, Ferraroni M, Nuti E, et al. Combining the tail and the ring approaches for obtaining potent and isoform-selective carbonic anhydrase inhibitors: solution and X-ray crystallographic studies. Bioorg Med Chem 2014;22:334–40. (c) Scozzafava A, Menabuoni L, Mincione F, et al. Carbonic anhydrase inhibitors. synthesis of water-soluble, topically effective, intraocular pressure-lowering aromatic/heterocyclic sulfonamides containing cationic or anionic moieties: is the tail more important than the ring? J Med Chem 1999;42:2641–50. (d) Pacchiano F, Carta F, McDonald PC, et al. Ureido-substituted benzenesulfonamides potently inhibit carbonic anhydrase IX and show antimetastatic activity in a model of breast cancer metastasis. J Med Chem 2011;54:1896–902. (e) ClinicalTrials.gov. https://clinicaltrials.gov/ct2/results?term=slc-0111.
  • (a) Lopez M, Salmon AJ, Supuran CT, Poulsen S-A. Carbonic anhydrase inhibitors developed through ‘click tailing’. Curr Pharm Des 2010;16:3277–87. (b) Pala N, Micheletto L, Sechi M, et al. Carbonic anhydrase inhibition with benzenesulfonamides and tetrafluorobenzenesulfonamides obtained via click chemistry. ACS Med Chem Lett 2014;5:927–30. (c) Wilkinson BL, Bornaghi LF, Houston TA, et al. Carbonic anhydrase inhibitors: inhibition of isozymes I, II, and IX with triazole-linked O-glycosides of benzene sulfonamides. J Med Chem 2007;50:1651–7. (d) Wilkinson BL, Bornaghi LF, Houston TA, et al. A novel class of carbonic anhydrase inhibitors: glycoconjugate benzene sulfonamides prepared by “click-tailing”. J Med Chem 2006;49:6539–48. (e) Nocentini A, Carta F, Ceruso M, et al. Click-tailed coumarins with potent and selective inhibitory action against the tumor-associated carbonic anhydrases IX and XII. Bioorg Med Chem 2015;23:6955–66.
  • Supuran CT. Carbonic anhydrase inhibitors. Bioorg Med Chem Lett 2010;20:3467–74.
  • (a) Briganti F, Pierattelli R, Scozzafava A, Supuran CT, Carbonic anhydrase inhibitors. Part 37. Novel classes of carbonic anhydrase inhibitors and their interaction with the native and cobalt-substituted enzyme: kinetic and spectroscopic investigations. Eur J Med Chem 1996;3:1001–10. (b) Garaj V, Puccetti L, Fasolis G, et al. Carbonic anhydrase inhibitors: synthesis and inhibition of cytosolic/tumor-associated carbonic anhydrase isozymes I, II, and IX with sulfonamides incorporating 1,2,4-triazine moieties. Bioorg Med Chem Lett 2004;14:5427–33. (c) Garaj V, Puccetti L, Fasolis G, et al. Carbonic anhydrase inhibitors: novel sulfonamides incorporating 1,3,5-triazine moieties as inhibitors of the cytosolic and tumour-associated carbonic anhydrase isozymes I, II and IX. Bioorg Med Chem Lett 2005;15:3102–8. (d) Carta F, Garaj V, Maresca A, et al. Sulfonamides incorporating 1,3,5-triazine moieties selectively and potently inhibit carbonic anhydrase transmembrane isoforms IX, XII and XIV over cytosolic isoforms I and II: solution and X-ray crystallographic studies. Bioorg Med Chem 2011;19:3105–19.
  • (a) Sahin H, Can Z, Yildiz O, et al. Inhibition of carbonic anhydrase isozymes I and II with natural products extracted from plants, mushrooms and honey. J Enzyme Inhib Med Chem 2012;27:395–402. (b) Maresca A, Supuran CT, Coumarins incorporating hydroxy- and chloro-moieties selectively inhibit the transmembrane, tumor-associated carbonic anhydrase isoforms IX and XII over the cytosolic ones I and II. Bioorg Med Chem Lett 2010;2:4511–14. (c) Carta F, Vullo D, Maresca A, et al. New chemotypes acting as isozyme-selective carbonic anhydrase inhibitors with low affinity for the offtarget cytosolic isoform II. Bioorg Med Chem Lett 2012;22:2182–5. (d) Ferraroni M, Carta F, Scozzafava A, Supuran CT. Thioxocoumarins show an alternative carbonic anhydrase inhibition mechanism compared to coumarins. J Med Chem 2016;59:462–73.
  • (a) Şentürk M, Gülçin I, Beydemir Ş, et al. In vitro inhibition of human carbonic anhydrase I and II isozymes with natural phenolic compounds. Chem Biol Drug Des 2011;77:494–9. (b) Carta F, Aggarwal M, Maresca A, et al. Dithiocarbamates: a new class of carbonic anhydrase inhibitors. Crystallographic and kinetic investigations. Chem Commun (Camb) 2012;48:1868–70. (c) Langella E, D'Ambrosio K, D'Ascenzio M, et al. A combined crystallographic and theoretical study explains the capability of carboxylic acids to adopt multiple binding modes in the active site of carbonic anhydrases. Chemistry 2016;22:97–100. (d) D'Ambrosio K, Carradori S, Monti SM, et al. Out of the active site binding pocket for carbonic anhydrase inhibitors. Chem Commun (Camb) 2015;51:302–5.
  • Liu X-F, Hao J-L, Xie T, et al. Curcumin, a potential therapeutic candidate for anterior segment eye diseases: a review. Front Pharmacol 2017;8:66–78.
  • Aditama R, Eryanti Y, Mujahidin D, et al. Determination of activities of human carbonic anhydrase II inhibitors from curcumin analogs. Trop J Pharm Res 2017;16:849–54.
  • Gençer N, Bilen C, Demir D, et al. In vitro inhibition effect of some chalcones on erythrocyte carbonic anhydrase I and II. Artif Cells Nanomed Biotechnol 2013;41:384–8.
  • Ramya PVS, Angapelly S, Guntuku L, et al. Synthesis and biological evaluation of curcumin inspired indole analogues as tubulin polymerization inhibitors. Eur J Med Chem 2017;127:100–14.
  • Qin N, Li C-B, Jin M-N, et al. Synthesis and biological activity of novel tiliroside derivants. Eur J Med Chem 2011;46:5189–95.
  • Khalifah RG. The carbon dioxide hydration activity of carbonic anhydrase I. Stop-flow kinetic studies on the native human isoenzymes B and C. J Biol Chem 1971;246:2561–73.
  • (a) Supuran CT, Barboiu M, Luca C, et al. Carbonic anhydrase activators. Part 14. Syntheses of mono and bis pyridinium salt derivatives of 2-amino-5-(2-aminoethyl)- and 2-amino-5-(3-aminopropyl)-1,3,4-thiadiazole and their interaction with isozyme II. Eur J Med Chem 1996;31:597–606. (b) Carta F, Aggarwal M, Maresca A, et al. Dithiocarbamates strongly inhibit carbonic anhydrases and show antiglaucoma action in vivo. J Med Chem 2012;55:1721–30. (c) Supuran CT, Nicolae A, Popescu A. Carbonic anhydrase inhibitors. Part 35. Synthesis of Schiff bases derived from sulfanilamide and aromatic aldehydes: the first inhibitors with equally high affinity towards cytosolic and membrane-bound isozymes. Eur J Med Chem 1996;31:431–8. (d) Nishimori I, Onishi S, Takeuchi H, Supuran CT. The alpha and beta classes carbonic anhydrases from Helicobacter pylori as novel drug targets. Curr Pharm Des 2008;14:622–30.
  • (a) Vullo D, Voipio J, Innocenti A, et al. Carbonic anhydrase inhibitors. Inhibition of the human cytosolic isozyme VII with aromatic and heterocyclic sulfonamides. Bioorg Med Chem Lett 2005;15:971–6. (b) Nishimori I, Vullo D, Innocenti A, et al. Carbonic anhydrase inhibitors. The mitochondrial isozyme VB as a new target for sulfonamide and sulfamate inhibitors. J Med Chem 2005;48:7860–6. (c) Scozzafava A, Menabuoni L, Mincione F, Supuran CT. Carbonic anhydrase inhibitors. A general approach for the preparation of water-soluble sulfonamides incorporating polyamino − polycarboxylate tails and of their metal complexes possessing long-lasting, topical intraocular pressure-lowering properties. J Med Chem 2002;45:1466–76. (d) Scozzafava A, Menabuoni L, Mincione F, et al. Carbonic anhydrase inhibitors: synthesis of sulfonamides incorporating dtpa tails and of their zinc complexes with powerful topical antiglaucoma properties. Bioorg Med Chem Lett 2001;11:575–82.
  • (a) Mishra CB, Kumari S, Angeli A, et al. Design, synthesis and biological evaluation of N-(5-methyl-isoxazol-3-yl/1,3,4 thiadiazol-2-yl)-4-(3-substitutedphenylureido) benzenesulfonamides as human carbonic anhydrase isoenzymes I, II, VII and XII inhibitors. J Enzyme Inhib Med Chem 2016;31:174–179. (b) Diaz JRA, Baldo MF, Echeverría G, et al. A substituted sulfonamide and its Co (II), Cu (II), and Zn (II) complexes as potential antifungal agents. J Enzyme Inhib Med Chem 2016;31:51–62. (c) Nocentini A, Carta F, Ceruso M, et al. Click-tailed coumarins with potent and selective inhibitory action against the tumor-associated carbonic anhydrases IX and XII. Bioorg Med Chem 2015;23:6955–66.