1,386
Views
0
CrossRef citations to date
0
Altmetric
Brief Report

Diversely substituted sulfamides for fragment-based drug discovery of carbonic anhydrase inhibitors: synthesis and inhibitory profile

, , , ORCID Icon, ORCID Icon, , , ORCID Icon & show all
Pages 857-865 | Received 21 Feb 2022, Accepted 03 Mar 2022, Published online: 16 Mar 2022

Abstract

A series of sulfamide fragments has been synthesised and investigated for human carbonic anhydrase inhibition. One of the fragments showing greater selectivity for cancer-related isoforms hCA IX and XII was co-crystalized with hCA II showing significant potential for fragment periphery evolution via fragment growth and linking. These opportunities will be identified in the future via the screening of this fragment structure for co-operative carbonic anhydrase binding with other structurally diverse fragments.

Graphical Abstract

1. Introduction

The carbonic anhydrase (CA) family of Zn(II) metalloenzymes (EC 4.2.1.1) catalyses the reversible hydration of carbon dioxide to bicarbonate anion, a fundamental reaction that controls physiological processes requiring pH control as well as ion transport and fluid secretionCitation1. Hyperactivity of specific CA isoformsCitation2 in various disease states makes these enzymes potential (and sometimes already validated) targets for therapeutic intervention with small molecule carbonic anhydrase inhibitors (CAIs)Citation3.

Clinically validated applications of CAIs currently include, among other diseases, the treatment of glaucomaCitation4, idiopathic intracranial hypertensionCitation5, high-altitude sicknessCitation6, congestive heart failureCitation7, peptic ulcersCitation8 and epilepsyCitation9. Another important potential application of CAIs (specifically, of related human (h) CA IX and XII isoformsCitation10) is in neoplastic therapyCitation11. The current state of development in this field is underscored by the hCA IX-selective drug SLC-0111Citation12 which is undergoing phase 1 b clinical study for tumours overexpressing hCA IXCitation13 and non-selective inhibitor E7070 (indisulam) developed by Eisai Co., Ltd. which successfully completed phase II clinical studyCitation14. The highly promising application of CAIs as antibacterials is based on the premise of selective inhibition of microbial CAs (crucial for the survival of bacteria) without affecting the CAs in the same concentration rangeCitation15. Thus, CA inhibition from such microorganisms as Vibrio choleraeCitation16, Burkholderia pseudomalleiCitation17, Mycobacterium tuberculosisCitation18, Salmonella entericaCitation19, Helicobacter pyloriCitation20, Escherichia coliCitation21 and many othersCitation15.

Considering the plethora of validated and potential therapeutic applications of CAIs, the discovery of new chemotypes endowed with CA inhibitory activity will continue to be a significant aim. CA contains a Zn2+ ion in its active site which mandates the zinc-binding nature of active site targeting pharmacophoric groups which can be employed in the design of new CAIs. Indeed, most of the clinically investigated (SLC-0111 and E7070) and used (e.g. acetazolamide, methazolamide, dorzolamide, brinzolamide and zonisamide) are primary sulphonamides (CSO2NH2) in which the sulphonamide group anchors to the prosthetic zinc ion and the molecular periphery defines the potency and isoform selectivity of these CAIsCitation2 ().

Figure 1. CAIs in clinical development and clinical use.

Figure 1. CAIs in clinical development and clinical use.

Many other zinc-binding motifs have been implicated as warheads in the CAI designCitation22. Among them, sulfamides (NSO2NH2) appears as an attractive alternative to the frequently studied sulphonamides. Indeed, sulphonamides are expected to have greater polarity and solubility compared to sulphonamides, due to the presence of an additional nitrogen atom. Moreover, primary sulphonamide group is found in such drugs as anticancer epacadostatCitation23 and gastric ulcer medication famotidineCitation24, both of which (epacadostatCitation25 and famotidineCitation26) were also found to inhibit various CA isoforms ().

Figure 2. Examples of clinically used sulfamide drugs.

Figure 2. Examples of clinically used sulfamide drugs.

The discovery of novel sulfamide CAIs would traditionally entail synthesis of structurally diverse libraries of compounds and their screening against an isoform panel of CAs. We have recently validated an approachCitation27 to the discovery of new sulphonamide CAIs based on the simultaneous screening of a diverse set of chemical fragments (i.e. small, Mw <∼250 and polar, cLopP< 3.0Citation28) along with sulphonamide zink-binding wasread (in particular, benzenesulfonamide or BSA). This led not only to the discovery of over 100 fragment hits which potentiated the binding of BSA but also to rediscovery of BSA-based CAIs with the molecular periphery replicating the fragment co-binders. In our intent aimed at the discovery of novel sulfamide CAIs, we decided to take a similar approach. Realisation of such an approach would require synthesis of a library of fragment-like sulfamides and profiling them against a panel of human CAs (in this case, anti-glaucoma target hCA II, two membrane-associated cancer-related targets hCA IX and XII and the usual cytosolic off-target hCA I). In this work, we aimed at the realisation of this approach and selection of a suitable sulfamide zinc-binding warhead for fragment-based discovery of novel sulfamide-type CAIs, a chemotype much less studied in the context of CA inhibition compared to sulphonamidesCitation29. Moreover, in this study, we were looking to identify: fragments that do not display apparently high intrinsic selectivity towards specific CA isoforms (mindful that such a selectivity will be gained in the future from co-operative screening of specific “tail” fragmentsCitation30) and yet would show a tendency to inhibit cancer-related isoforms hCA IX and XII over cytosolic hCA I and II. Of particular interest would be fragments that do not display a pronounce potency against CA isoforms of interest, ideally in the 10−7 M range of Ki values (so that the future contribution from co-operative fragment binding would be more pronounced). Specific emphasis was put on conformationally constrained fragments which would be structurally close to the classical BSA zinc-binding motif and would co-crystallize with any of the isoforms (e.g. the most readily available hCA II) to further guide further fragment evolution via growing, linking and mergingCitation31. Herein, we report on the successful realisation of this strategy.

2. Results and discussion

Seventeen non-symmetrically substituted primary sulfamides were synthesised from inorganic sulfamide 1 via direct nucleophilic substitution at the sulphur atom, via the thermally promoted reaction in dioxane with a four-fold excess of 1, conducted at 110 °C over 48 hCitation32. The yields of the resulting compounds 2a–q were generally modest to good (Scheme 1).

Scheme 1. Synthesis of unsymmetrically substituted primary sulfamides 2a–q.

Scheme 1. Synthesis of unsymmetrically substituted primary sulfamides 2a–q.

The same thermally promoted protocol, not unexpectedly, did not work for less reactive (hetero)aromatic amines. Thus, an alternative approach was takenCitation33. Instead of sulfamide, commercially available chlorosulfonyl isocyanate 3 dissolved in dichloromethane at 0 °C was reacted with 1 equiv of tert-butanol to give the Boc-protected amino-sulfonyl-chloride (4), which was subsequently added slowly to a solution of 1 equiv of the respective hetero(aromatic) amine in the presence of 2 equiv of triethylamine in dichloromethane at 0 °C. In this case, again, the yields of unsymmetrically substituted primary sulfamides 2r–w were modest to good over two steps (Scheme 2).

Scheme 2. Synthesis of unsymmetrically (hetero)aromatic amine-substituted sulfamides 2r–w.

Scheme 2. Synthesis of unsymmetrically (hetero)aromatic amine-substituted sulfamides 2r–w.

From the physicochemical data summarised for fragments 2a–w in , one can appreciate their being distinctly fragment-like (Mw = 165.2 … 257.4, cLogP = −1.87 … 1.71). Furthermore, the inhibitory data reveal the absence of apparent isoform selectivity displayed by these fragments which is perfectly in line with the limited size of the molecular periphery (typically responsible for making additional contacts with the protein and ensuring higher potency and isoform selectivity). With our initial focus on the cancer-related, membrane-bound CA isoforms hCA IX and XII compounds displaying greater selectivity towards these isoforms against cytosolic hCA I and II (structural homologs of each other 2r and 2v) received our priority attention.

Table 1. Calculated physicochemical properties and hCA I, II, IX and XII inhibitory profile of compounds 2a–w.

After much experimentation, compound 2v was co-crystallized with recombinant hCA II and its structure was resolved ().

Figure 3. Co-crystal structure of fragment sulfamide 2v with hCA II (PDB code 7QSI).

Figure 3. Co-crystal structure of fragment sulfamide 2v with hCA II (PDB code 7QSI).

As one can see from the crystal structure of 2v with hCA II isoform, the small (Mw = 198.2) N-(aminosulfonyl)indoline fragment displayed two finding poses within the hCA II active site. In both poses, the sulfonylamino groups are anchored to the prosthetic zinc ion (displayed as a grey sphere). The validity of the two binding poses signifies the fact that the binding of 2v leaves a significant room for the periphery growth around this fragment and makes it a highly suitable candidate for co-operative screening with other structurally diverse fragments in order to identify the starting points for the structural evolution of this fragment.

3. Materials and methods

3.1. Chemical synthesis – general

NMR spectra were recorded on a Bruker Avance III 400 spectrometer (1H: 400.13 MHz; 13С: 100.61 MHz; chemical shifts are reported as parts per million (δ, ppm); the residual solvent peaks were used as internal standards: δ 7.28 1H in CDCl3, δ 77.02 ppm for 13 C in CDCl3; multiplicities are abbreviated as follows: s = singlet, d = doublet, t = triplet, q = quartette, m = multiplet, br = broad; coupling constants, J, are reported in Hz. Mass spectra were recorded on a Bruker microTOF spectrometer (ESI ionization).

3.1.1. General procedure for the synthesis of compounds 2a–q

A mixture of corresponding amine (2 mmol) and sulphuric diamide (8 mmol, 768 mg) in dry 1.4-dioxane (4 mL) was stirred at 110 oC for 48 h. If amine was in the form of hydrochloride salt, an additional equivalent of triethylamine was added. CH2Cl2 (5 mL) was added and the resulting precipitate was filtered off, washed with ethyl acetate (5 mL). The filtrate and the washings were combined and solvent was evaporated under reduced pressure. The crude product was purified by column chromatography on silica gel using CH2Cl2:2-propanol 20:1–10:1 gradient as eluent (Rf values are given for solvent system indicated).

3.1.2. 4-Hydroxypiperidine-1-sulphonamide (2a)

Rf = 0.30 (EtOAc); 1H NMR (400 MHz, DMSO-d6) δ = 6.66 (s, 2H), 4.69 (d, J = 3.9 Hz, 1H), 3.58 (dq, J = 7.8, 4.0 Hz, 1H), 3.22 (ddd, J = 11.0, 6.5, 3.8 Hz, 2H), 2.74 (ddd, J = 12.0, 9.0, 3.3 Hz, 2H), 1.77 (ddt, J = 13.7, 7.2, 3.6 Hz, 2H), 1.46 (dtd, J = 12.5, 8.6, 3.7 Hz, 2H) ppm; 13 C NMR (101 MHz, DMSO-d6, Dept) δ = 65.049 (C-OH), 43.80 and 33.47 (2 CH2) ppm; HRMS (ESI) C5H11N2O3S+ m/z: [M – H] 179.0492 (calc 179.0485).

3.1.3. 4-Methylpiperazine-1-sulphonamide (2 b)

Rf = 0.61 (MeOH); 1H NMR (400 MHz, DMSO-d6) δ = 6.76 (s, 2H), 2.95 (t, J = 4.9 Hz, 4H), 2.38 (t, J = 5.0 Hz, 4H), 2.19 (s, 3H) ppm; 13 C NMR (101 MHz, DMSO-d6, DEPT) δ = 54.08 (CH2), 46.14 (CH2), 45.89 (CH3) ppm; HRMS (ESI) C5H12N3O2S+ m/z: [M − H] 178.0644 (calc 178.0645).

3.1.4. N-(4-Fluorobenzyl)sulfamide (2c)

1H NMR (400 MHz, DMSO-d6) δ = 7.39 (dd, J = 8.5, 5.7 Hz, 2H), 7.15 (t, J = 8.9 Hz, 2H), 7.07 (t, J = 6.5 Hz, 1H), 6.63 (s, 2H), 4.07 (d, J = 6.5 Hz, 2H) ppm; 13 C NMR (101 MHz, DMSO-d6, DEPT) δ = 161.70 (d, J = 242.2 Hz), 135.39 (d, J = 2.9 Hz), 130.05 (d, J = 8.1 Hz, CH), 115.29 (d, J = 21.3 Hz, CH), 45.77 (s, CH2) ppm; 19 F NMR (470 MHz, DMSO-d6) δ = −116.15 ppm; HRMS (ESI) C7H9FN2NaO2S+ m/z: [M + Na]+ 227.0272 (calc 227.0261).

3.1.5. N-(1,2,3,4-Tetrahydronaphthalen-1-yl)sulfamide (2d)

Rf = 0.80 (CH2Cl2-CH3OH 4:1); 1H NMR (400 MHz, DMSO-d6) δ = 7.57 − 7.50 (m, 1H), 7.19 − 7.11 (m, 2H), 7.08 − 7.02 (m, 1H), 6.96 (d, J = 8.9 Hz, 1H), 6.65 (s, 2H), 4.39 (td, J = 8.0, 5.2 Hz, 1H), 2.79 − 2.57 (m, 2H), 2.11 − 2.00 (m, 1H), 1.96 − 1.80 (m, 2H), 1.74 − 1.60 (m, 1H) ppm;

13 C NMR (101 MHz, DMSO-d6, DEPT) δ = 138.15, 137.41, 129.43, 128.94, 127.07, 126.02, 51.32 (CH), 30.89, 29.23 and 20.39 (3CH2) ppm; HRMS (ESI) C10H14N2NaOS+ m/z: [M + Na]+ 249.0671 (calc 249.0668).

3.1.6. N-((tetrahydrofuran-2-yl)methyl)sulfamide (2e)

Rf = 0.59 (DCM:i-PrOH = 9:1); 1H NMR (400 MHz, CDCl3) δ = 4.94 (br. t, J = 6.0 Hz, 1H), 4.88 (br. s, 2H), 4.08 (ddt, J = 10.5, 7.3, 3.4 Hz, 1H), 3.88 (dt, J = 8.3, 6.7 Hz, 1H), 3.82 − 3.73 (m, 1H), 3.28 (ddd, J = 13.3, 6.4, 3.2 Hz, 1H), 3.14 (ddd, J = 13.4, 7.8, 5.8 Hz, 1H), 2.05 − 1.85 (m, 3H), 1.68 − 1.58 (m, 1H) ppm; 13 C NMR (101 MHz, CDCl3, DEPT) δ = 77.65 (CH), 68.21, 47.52, 28.61 and 25.76 (4 CH2) ppm; HRMS (ESI) C5H12N2NaO3S+ m/z: [M + Na]+ 203.0464 (calc 203.0461).

3.1.7. Methyl sulfamoylprolinate (2f)

Rf = 0.75 (EtOAc); 1H NMR (400 MHz, DMSO-d6) = 6.82 (s, 2H), 4.18 (dd, J = 8.9, 4.1 Hz, 1H), 3.27 (dq, J = 6.2, 3.5, 2.9 Hz, 2H), 2.18 − 2.07 (m, 1H), 1.92 − 1.80 (m, 3H) ppm; 13 C NMR (101 MHz, DMSO-d6, DEPT) δ = 173.24, 60.47 and 52.28 (CH and CH3), 49.11, 31.06 and 24.92 (3 CH2) ppm; HRMS (ESI) C6H13N2O4S+m/z: [M + H]+ 209.0599 (calc 209.0591).

3.1.8. (1-Methylpiperidin-4-yl)sulfamide (2g)

Rf = 0,13 (MeOH); 1H NMR (400 MHz, DMSO-d6) δ = 6.47 (s, 3H), 3.01 (tt, J = 10.6, 4.2 Hz, 1H), 2.68 (d, J = 11.9 Hz, 2H), 2.12 (s, 3H), 1.85 (qd, 11.9, 2.0 Hz, 4H), 1.42 (qd, J = 12.5, 4.2 Hz, 2H) ppm; 13 C NMR (101 MHz, DMSO-d6) δ = 54.78, 50.36, 46.41, 32.86 ppm; HRMS (ESI) C6H16N3O2S+ m/z: [M + H]+ 194.0955 (calc 194.0958).

3.1.9. 4-Phenylpiperidine-1-sulphonamide (2 h)

Rf = 0.86 (EtOAc); 1H NMR (400 MHz, DMSO-d6) δ = 7.34 − 7.25 (m, 4H), 7.21 (td, J = 6.8, 1.7 Hz, 1H), 6.75 (br.s, 2H), 3.59 (br.d, J = 12.1 Hz, 2H), 2.63 (td, J = 12.2, 2.7 Hz, 2H), 2.59 − 2.53 (m, 1H), 1.86 (br.d, J = 12.2 Hz, 2H), 1.69 (qd, J = 12.5, 4.0 Hz, 2H) ppm;13C NMR (101 MHz, DMSO-d6, DEPT) δ = 145.97, 128.84 (CArH), 127.20 (CArH), 126.67 (CArH), 46.96 (CH2), 41.45 (CH), 32.45 (CH2) ppm; HRMS (ESI) C11H16N2NaO2S+ m/z: [M + Na]+ 263.0830 (calc 263.0825).

3.1.10. Benzyl (2-(dimethylamino)ethyl)sulfamide (2i)

Rf = 0.58 (MeOH); 1H NMR (400 MHz, DMSO-d6) δ = 7.40 − 7.33 (m, 4H), 7.31 − 7.26 (m, 1H), 6.91 (s, 2H), 4.26 (s, 2H), 3.09 (t, J = 6.8 Hz, 2H), 2.30 (t, J = 6.8 Hz, 2H), 2.07 (s, 6H) ppm; 13 C NMR (101 MHz, DMSO-d6, DEPT) δ = 138.06, 128.76, 128.56 and 127.74 (3 CArH), 57.05, 51.41 and 45.72 (3 CH2), 45.52 (CH3) ppm; HRMS (ESI) C11H20N3O2S+ m/z: [M + H]+ 258.1276 (calc 258.1271).

3.1.11. (Pyridin-3-ylmethyl)sulfamide (2j)

Rf = 0.90 (MeOH); 1H NMR (400 MHz, CD3CN) δ = 8.55 (d, J = 2.2 Hz, 1H), 8.48 (dd, J = 4.8, 1.6 Hz, 1H), 7.76 (dt, J = 7.9, 1.9 Hz, 1H), 7.33 (dd, J = 7.8, 4.8 Hz, 1H), 5.62 (br. s, 1H), 5.31 (br. s, 2H), 4.20 (d, J = 6.3 Hz, 2H) ppm; 13 C NMR (101 MHz, CD3CN) δ = 149.23, 148.59, 135.72, 133.71, 123.46, 44.32 ppm; HRMS (ESI) C6H9N3NaO2S+ m/z: [M + Na]+ 210.0308 (calc 210.0308).

3.1.12. (Thiophen-2-ylmethyl) sulfamide (2k)

Rf = 0.49 (DCM:i-PrOH = 20:1); 1H NMR (400 MHz, Acetone-d6) = 7.37 (dd, J = 5.2, 1.2 Hz, 1H), 7.07 (dq, J = 3.3, 1.1 Hz, 1H), 6.97 (dd, J = 5.1, 3.5 Hz, 1H), 6.20 (s, 1H), 5.99 (s, 2H), 4.45 (dd, J = 6.4, 1.0 Hz, 2H) ppm; 13 C NMR (101 MHz, Acetone-d6, DEPT) δ = 141.21 (>C=), 126.62, 125.80, and 125.08 (3 CArH), 41.97 (CH2) ppm; HRMS (ESI) C5H8N2NaO2S2+ m/z: [M + Na]+ 214.9919 and 216.9877 (calc 214.9919 and 216.9878).

3.1.13. (2-(Diethylamino)ethyl)(methyl)sulfamide (2 l)

Rf = 0.58 (MeOH); 1H NMR (400 MHz, DMSO-d6) = 6.73 (s, 2H), 3.05 (t, J = 6.8 Hz, 2H), 2.68 (s, 2H), 2.55 (t, J = 6.8 Hz, 2H), 2.49 (q, J = 7.1 Hz, 4H), 0.96 (t, J = 7.1 Hz, 6H) ppm; 13 C NMR (101 MHz, DMSO-d6, DEPT) δ = 50.53, 48.57 and 46.95(3 CH2), 35.57 and 12.01 (2 CH3) ppm; HRMS (ESI) C7H20N3O2S+ m/z: [M + H]+ 210.1274 (calc 210.1271).

3.1.14. Ethyl 1-sulfamoylpiperidine-4-carboxylate (2 m)

Rf = 0.81 (EtOAc); 1H NMR (400 MHz, DMSO-d6) δ = 6.71 (s, 2H), 4.07 (q, J = 7.1 Hz, 2H), 3.36 (dt, J = 12.3, 3.8 Hz, 2H), 2.62 (td, J = 11.6, 2.8 Hz, 2H), 2.42 (tt, J = 10.7, 4.0 Hz, 1H), 1.90 (br. dd, J = 13.5, 3.8 Hz, 2H), 1.60 (qd, J = 11.1, 3.8 Hz, 2H), 1.18 (t, J = 7.1 Hz, 3H) ppm; 13 C NMR (101 MHz, DMSO-d6) δ = 174.25, 60.44, 45.55, 27.47, 14.54 ppm; HRMS (ESI) C8H16N2NaO4S+ m/z: [M + Na]+ 259.0722 (calc 259.0723).

3.1.15. [1,4'-Bipiperidine]-1'-sulphonamide (2n)

Rf = 0.24 (MeOH); 1H NMR (400 MHz, DMSO-d6) δ = 6.68 (br.s, 2H), 3.48 (br.d, J = 12.4 Hz, 2H), 2.53 − 2.41 (m, 3H), 2.44 (t, J = 5.1 Hz, 3H), 2.24 (tt, J = 11.3, 3.6 Hz, 1H), 1.76 (br.d, J = 11.7 Hz, 2H), 1.52 − 1.43 (m, 6H), 1.38 (q, J = 5.9 Hz, 2H) ppm; 13 C NMR (101 MHz, DMSO-d6, DEPT) δ = 61.56, 50.16, 46.22, 27.21, 26.52, 25.01; HRMS (ESI) C10H22N3O2S+ m/z: [M + H]+ 248.1431 (calc 248.1428).

3.1.16. Morpholine-4-sulphonamide (2o)

Rf = 0.56 (EtOAc); 1H NMR (400 MHz, DMSO-d6) δ = 6.82 (s, 2H), 3.86 − 3.50 (m, 4H), 3.02 − 2.77 (m, 4H) ppm; 13 C NMR (101 MHz, DMSO-d6) δ = 65.73, 46.43 ppm; HRMS (ESI) C4H9N2O3S+ m/z: [M − H] 165.0326 (calc 165.0339)

3.1.17. 1-Sulfamoylpiperidine-4-carboxamide (2p)

Rf = 0.61 (EtOAc:MeOH 3:1); 1H NMR (400 MHz, DMSO-d6) δ = 7.27 (s, 1H), 6.82 (s, 1H), 6.79 − 6.44 (m, 2H), 3.52 − 3.39 (m, 2H), 2.21 − 2.05 (m, 1H), 1.78 (d, J = 13.1 Hz, 2H), 1.57 (t, J = 12.7 Hz, 2H) ppm; 13 C NMR (101 MHz, DMSO-d6) δ = 176.30, 45.96, 41.12, 28.07 ppm; HRMS (ESI) C6H13N3NaO3S+ m/z: [M + Na]+ 230.0565 (calc 230.0570)

3.1.18. Piperazine-1-sulphonamide hydrochloride (2q)

1H NMR (400 MHz, DMSO-d6) δ = 9.60 (s, 2H), 7.08 (s, 2H), 3.21 (dd, J = 7.1, 3.6 Hz, 4H), 3.14 (dd, J = 7.0, 3.7 Hz, 4H) ppm; 13 C NMR (101 MHz, DMSO-d6) δ = 43.27, 42.37 ppm; HRMS (ESI) C4H12N3O2S+ m/z: [M + H] 166.0649 (calc 166.0645).

3.1.19. 2-Methylindoline-1-sulphonamide (2r)

Rf = 0.81 (n-Hexane:EtOAc 1:1); 1H NMR (400 MHz, CDCl3) = 7.40 (d, J = 8.2 Hz, 1H), 7.19 (dt, J = 7.7, 3.7 Hz, 2H), 7.04 (t, J = 7.4 Hz, 1H), 4.46 (s br., 1H), 4.07 (s br., 2H), 3.52 − 3.43 (m, 2H), 2.67 (dd, J = 15.9, 3.5 Hz, 1H), 1.43 (d, J = 6.1 Hz, 3H) ppm; 13 C NMR (101 MHz, CDCl3) δ = 141.32, 131.15, 127.79, 125.40, 124.10, 115.77, 59.19, 36.64, 22.63 ppm; HRMS (ESI) C9H12N2NaO2S+ m/z: [M + Na]+ 235.0515 (calc 235.0512).

3.1.20. N-(Pyrimidin-2-yl)sulfamide (2s)

1H NMR (400 MHz, DMSO-d6) = 8.56 (d, J = 4.8 Hz, 4H), 7.06 (t, J = 4.9 Hz, 2H) ppm; 13 C NMR (101 MHz, DMSO-d6) δ = 158.73, 158.26, 115.28 ppm; HRMS (ESI) C4H6N4NaO2S+ m/z: [M + Na]+ 197.1668 (calc 197.1672)

3.1.21. 3,4-Dihydroquinoline-1(2H)-sulphonamide (2t)

Rf = 0.55 (n-Hexane:EtOAc 3:1); 1H NMR (400 MHz, CDCl3) δ = 7.67 (d, J = 8.2 Hz, 1H), 7.17 (t, J = 7.6 Hz, 1H), 7.14 − 7.02 (m, 2H), 4.56 (s, 2H), 3.87 − 3.53 (m, 2H), 2.86 (t, J = 6.6 Hz, 2H), 2.06 (p, J = 6.5 Hz, 2H) ppm; 13 C NMR (400 MHz, CDCl3, DEPT) δ = 137.20, 129.92, 129.53, 126.60, 124.65 and 123.61 (4 CArH), 47.28, 26.94 and 21.87 (3 CH2) ppm; HRMS (ESI) C9H12N2NaO2S+ m/z: [M + Na]+ 235.05110 (calc 235.0511)

3.1.22. N-(Naphthalen-1-yl)sulfamide (2 u)

Rf = 0.44 (n-Hexane:EtOAc 1:1); 1H NMR (400 MHz, DMSO-d6) δ = 9.27 (s, 1H), 8.35 − 8.25 (m, 1H), 7.96 − 7.87 (m, 1H), 7.75 (d, J = 8.1 Hz, 1H), 7.59 (dd, J = 7.6, 1.2 Hz, 1H), 7.55 − 7.47 (m, 3H), 7.00 (s, 2H) ppm; 13 C NMR (101 MHz, DMSO-d6, Dept) δ = 134.74, 134.36, 129.17, 128.26, 126.43, 126.10, 125.49, 124.10, 121.57 ppm; HRMS (ESI) C10H10N2NaO2S+ m/z: [M + Na]+ 245.0359 (calc 245.0356)

3.1.23. Indoline-1-sulphonamide (2v)

Rf = 0.77 (EtOAc); 1H NMR (400 MHz, DMSO-d6) = 7.28 (d, J = 8.0 Hz, 1H), 7.25 − 7.20 (m, 1H), 7.23 (s, 2H), 7.16 (t, J = 7.7 Hz, 1H), 6.96 (t, J = 7.4 Hz, 1H), 3.80 (t, J = 8.4 Hz, 2H), 3.05 (t, J = 8.4 Hz, 2H) ppm; 13 C NMR (101 MHz, DMSO-d6, DEPT) δ = 143.51 and 131.99 (2 >CAr=), 127.63, 125.40, 122.83 and 114.21 (4 CArH), 50.39 (CH2), 27.76 (CH2) ppm; HRMS (ESI) C8H10N2NaO2S+ m/z: [M + Na]+ 221.0357 (calc 221.0355)

3.1.24. N-(2,4-Difluorophenyl)sulfamide (2w)

1H NMR (400 MHz, CDCl3) δ = 7.40 (d, J = 8.2 Hz, 1H), 7.19 (dt, J = 7.7, 3.7 Hz, 2H), 7.04 (t, J = 7.4 Hz, 1H), 4.46 (s br., 1H), 4.07 (s br., 2H), 3.52 − 3.43 (m, 2H), 2.67 (dd, J = 15.9, 3.5 Hz, 1H), 1.43 (d, J = 6.1 Hz, 3H) ppm; 13 C NMR (101 MHz, CDCl3) δ = 160.29 (dd, J = 248.7, 11.3 Hz), 154.88 (dd, J = 247.7, 12.1 Hz), 125.60 (dd, J = 9.6, 1.4 Hz, CArH), 121.02 (dd, J = 12.5, 3.7 Hz), 111.97 (dd, J = 22.3, 3.8 Hz), 104.40 (dd, J = 26.7, 23.7 Hz) ppm; 19 F NMR (376 MHz, CDCl3) δ = −111.98, −123.53 ppm; HRMS (ESI) C6H6F2N2NaO2S+ m/z: [M + Na]+ 231.0010 (calc 231.0011)

3.2. Carbonic anhydrase inhibition testing

An Applied Photophysics stopped-flow instrument has been used for assaying the CA catalysed CO2 hydration activityCitation35. 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 uncatalyzed 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 deionised 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 EI complex. The inhibition constants were subsequently obtained by nonlinear least-squares methods using PRISM 3 and the ChengPrusoff equation, as reported earlier, and represent the mean from at least three different determinations. All CA isoforms were recombinant ones obtained in-house as reported earlierCitation36–39.

3.3. Co-crystallization and X-ray data collection

Crystals of hCA II complexed with compound 2v were obtained using the sitting drop vapour diffusion method. An equal volume of 0.8 mM solution of hCA II in Tris pH = 8.0 and 1.6 mM of the inhibitors in Hepes 20 mM pH = 7.4 was mixed and incubated for 15 min. 2 mL of the complex solution were mixed with 2 mL of a solution of 1.6 M sodium citrate, 50 mM Tris pH 8.0 and were equilibrated against the same solution at 296 K. Crystals of the complex grew in a few days. The crystals were flash-frozen at 100 K using a solution obtained by adding 25% (v/v) glycerol to the mother liquor solution as cryoprotectant. A data set on a crystal of the complex with the inhibitor 2v was collected at the Centro di Cristallografia Strutturale (CRIST) in Florence using an Oxford Diffraction instrument equipped with a sealed tube Enhance Ultra (Cu) and a Onyx CCD detector. Data were integrated and scaled using the program XDS.24 Data processing statistics are showed in .

Table 2. Summary of data collection and atomic model refinement statistics for hCAII.a

3.4. Structure determination

The crystal structure of hCA II (PDB accession code: 7QSI) without solvent molecules and other heteroatoms was used to obtain initial phases of the structures using Refmac5Citation40. 5% of the unique reflections were selected randomly and excluded from the refinement data set for the purpose of Rfree calculations. The initial |Fo ̶ Fc| difference electron density maps unambiguously showed the inhibitor molecules. An electron density, which could be interpreted as a second molecule of inhibitor 2v, was present near the N-terminus of the protein. Thus, a second 2v molecule was introduced in the model with 0.75 occupancy. Atomic model for the inhibitor was calculated and energy minimised using the program JLigand 1.0.39. Refinements proceeded using normal protocols of positional, isotropic atomic displacement parameters alternating with manual building of the models using COOTCitation41. Solvent molecules were introduced automatically using the program ARPCitation42. The quality of the final model was assessed with COOT and RampageCitation43. Crystal parameters and refinement data are summarised in . Atomic coordinates were deposited in the Protein Data Bank (PDB accession code: 7QSI). Graphical representations were generated with ChimaeraCitation44.

4. Conclusions

We have described the synthesis and testing against a panel of human carbonic anhydrases (hCA I, II, IX and XII) of a series of hydrophilic, fragment sulfamides intended for fragment-based drug discovery of isoform-selective carbonic anhydrase inhibitors via cooperative screening with other, non-zinc-binding fragments. As expected from the minimal-periphery zinc-binding moieties, these fragment sulfamides demonstrated little selectivity across the panel of hCAs. However, for one of the fragment inhibitors (2v) which showed higher selectivity towards the cancer-related hCA isoforms (IX and CII), we obtained a crystal structure with the most abundant cytosolic isoform hCA II which showed two possible binding modes and thus significant room for cooperative fragment biding and subsequent periphery evolution.

Supplementary materials

Copies of 1H and 13 C NMR spectra of compounds 2a–w.

Supplemental material

Supplemental Material

Download PDF (2.7 MB)

Acknowledgements

We are grateful to the Research Centre for Magnetic Resonance and the Centre for Chemical Analysis and Materials Research of Saint Petersburg State University Research Park for the analytical data.

Disclosure statement

No potential conflict of interest was reported by all author(s) except CTS. CT Supuran is Editor-in-Chief of the Journal of Enzyme Inhibition and Medicinal Chemistry. He was not involved in the assessment, peer review, or decision-making process of this paper. The authors have no relevant affiliations of financial involvement with any organisation or entity with a financial interest in or financial conflict with the subject matter or materials discussed in the manuscript. This includes employment, consultancies, honoraria, stock ownership or options, expert testimony, grants or patents received or pending, or royalties. This article was submitted before the start of the Russia-Ukraine military conflict on February 24, 2022.

Correction Statement

This article has been republished with minor changes. These changes do not impact the academic content of the article.

Additional information

Funding

This research was supported by the Ministry of Education of the Russian Federation [government contract 073–00077-21–02 “Development of an innovative glaucoma drug based on selective inhibition of carbonic anhydrase II,” registry number 730000Ф.99.1.БВ10АА00006].

References

  • Pastorekova S, Parkkila S, Pastorek J, Supuran CT. Carbonic anhydrases: current state of the art, therapeutic applications and future prospects. J Enzyme Inhib Med Chem 2004;19:199–229.
  • 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.
  • Scozzafava A, Supuran CT. Glaucoma and the applications of carbonic anhydrase inhibitors. Subcell Biochem 2014;75:349–59.
  • Supuran CT. Acetazolamide for the treatment of idiopathic intracranial hypertension. Expert Rev Neurother 2015;15:851–6.
  • Swenson ER. Carbonic anhydrase inhibitros and high altitude ilnesses. Subcell Biochem 2014;75:361–86.
  • Wongboonsin J, Thongprayoon C, Bathini T, et al. Acetazolamide therapy in patients with heart failure: a meta-analysis. J Clin Med 2019;8:349.
  • Buzás GM, Supuran CT. The history and rationale of using carbonic anhydrase inhibitors in the treatment of peptic ulcers. In memoriam Ioan Puşcaş (1932–2015). J Enzyme Inhib Med Chem 2016;31:527–33.
  • Ciccone L, Cerri C, Nencetti S, Orlandini E. Carbonic anhydrase and epilepsy: state of the art and future perspectives. Molecules 2021;26:6380.
  • Mboge MY, McKenna R, Frost SC. Advances in anti-cancer drug development targeting carbonic anhydrase IX and XII. Top Anticancer Res 2015;5:3–42.
  • Ward C, Meehan J, Gray ME, et al. The impact of tumour pH on cancer progression: strategies for clinical intervention. Explor Target Antitumor Ther 2020;1:71–100.
  • Carta F, Vullo D, Osman SM, et al. Synthesis and carbonic anhydrase inhibition of a series of SLC-0111 analogs. Bioorg Med Chem 2017;25:2569–76.
  • https://clinicaltrials.gov/ct2/show/NCT03450018 [last accessed 14 Jan 2022].
  • Assi R, Kantarjian H, Kadia TM, et al. Final results of a phase 2, open-label study of Indisulam, idarubicin, and cytarabine in patients with relapsed or refractory acute myeloid leukemia and high-risk myelodysplastic syndrome. Cancer 2018;124:2758–65.
  • Supuran CT, Capasso C. Antibacterial carbonic anhydrase inhibitors: an update on the recent literature. Exp Opin Ther Pat 2020;30:963–82.
  • De Vita D, Angeli A, Pandofi F, et al. Inhibition of the α-carbonic anhydrase from Vibrio cholerae with amides and sulfonamides incorporating imidazole moieties. J Enzyme Inhib Med Chem 2017;32:798–804.
  • Rasti B, Mazraedoost S, Panahi H, et al. New insights into the selective inhibition of the β-carbonic anhydrases of pathogenic bacteria Burkholderia pseudomallei and Francisella tularensis: a proteochemometrics study. Mol Divers 2019;23:263–73.
  • Aspatwar A, Winum J-Y, Carta F, et al. Carbonic anhydrase inhibitors as novel drugs against mycobacterial β-carbonic anhydrases: an update on in vitro and in vivo studies. Molecules 2018;23:2911.
  • Nishimori I, Minakuchi T, Vullo D, et al. Inhibition studies of the β-carbonic anhydrases from the bacterial pathogen Salmonella enterica serovar Typhimurium with sulfonamides and sulfamates. Bioorg Med Chem 2011;19:5023–30.
  • Grande R, Carradori S, Puca V, et al. Selective inhibition of Helicobacter pylori carbonic anhydrases by carvacrol and thymol could impair biofilm production and the release of outer membrane vesicles. Int J Mol Sci 2021;22:11583.
  • Del Prete S, De Luca V, Bua S, et al. The effect of substituted benzene-sulfonamides and clinically licensed drugs on the catalytic activity of CynT2, a carbonic anhydrase crucial for Escherichia coli life cycle. Int J Mol Sci 2020;21:4175.
  • 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 Chen 2006;6:921–36.
  • Yue EW, Sparks R, Polam P, et al. INCB24360 (Epacadostat), a highly potent and selective indoleamine-2,3-dioxygenase 1 (IDO1) inhibitor for immuno-oncology. ACS Med Chem Lett 2017;8:486–91.
  • Langtry HD, Grant SM, Goa KL. Famotidine. An updated review of its pharmacodynamic and pharmacokinetic properties, and therapeutic use in peptic ulcer disease and other allied diseases. Drugs 1989;38:551–90.
  • Angeli A, Ferraroni M, Nocentini A, et al. Polypharmacology of epacadostat: a potent and selective inhibitor of the tumor associated carbonic anhydrases IX and XII. Chem Commun (Camb) 2019;55:5720–3.
  • Angeli A, Ferraroni M, Supuran CT. Famotidine, an antiulcer agent, strongly inhibits Helicobacter pylori and human carbonic anhydrases. ACS Med Chem Lett 2018;9:1035–8.
  • Krasavin M, Kalinin S, Zozulya S, et al. Screening of benzenesulfonamide in combination with chemically diverse fragments against carbonic anhydrase by differential scanning fluorimetry. J Enzyme Inhib Med Chem 2020;35:306–10.
  • Li Q. Application of fragment-based drug discovery to versatile targets. Front Mol Biosci 2020;7:180.
  • According to the Reaxys database, only 9,106 sulfamides (NSO2NH2) were associated with the keywords “carbonic anhydrase inhibitor” compared to 108,988 sulfonamides searchable in the same context (information retrieved on January 18, 2022).
  • Bonardi A, Nocentini A, Bua S, et al. Sulfonamide inhibitors of human carbonic anhydrases designed through a three-tails approach: improving ligand/isoform matching and selectivity of action. J Med Chem 2020;63:7422–44.
  • Ferreira LG, Andricopulo AD. From protein structure to small-molecules: recent advances and applications to fragment-based drug discovery. Curr Top Med Chem 2017;17:2260–70.
  • Xu Z, Tice CM, Zhao W, et al. Structure-based design and synthesis of 1,3-oxazinan-2-one inhibitors of 11β-hydroxysteroid dehydrogenase type 1. J Med Chem 2011;54:6050–62.
  • Casini A, Winum J-Y, Montero J-L, et al. Carbonic anhydrase inhibitors: inhibition of cytosolic isozymes I and II with sulfamide derivatives. Bioorg. Med. Chem. Lett 2003;13:837–40.
  • https://www.molinspiration.com/.
  • 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.
  • 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.
  • Lopez M, Salmon AJ, Supuran CT, Poulsen S-A. Carbonic anhydrase inhibitors developed through ‘click tailing’. Curr Pharm Des 2010;16:3277–87.
  • 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.
  • 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.
  • Murshudov GN, Vagin AA, Dodson EJ. Refinement of macromolecular structures by the maximum-likelihood method. Acta Crystallogr D Biol Crystallogr 1997;53:240–55.
  • Emsley P, Lohkamp B, Scott W, Cowtan K. Features and development of Coot. Acta Crystallogr D Biol Crystallogr 2010;66:486–501.
  • Lamzin VS, Perrakis A, Wilson KS. The ARP/WARP suite for automated construction and refinement of protein models. In: Rossmann MG, Arnold E, eds. Int. Tables for crystallography. Vol. F: Crystallography of biological macromolecules. Dordrecht, The Netherlands: Kluwer Academic Publishers; 2001:720–722.
  • Lovell SC, Davis IW, Arendall WB, III, et al. Structure validation by Cα geometry: ϕ,ψ and Cβ deviation. Proteins Struct Funct Genet 2003;50:437–50.
  • Pettersen EF, Goddard TD, Huang CC, et al. UCSF Chimera-a visualization system for exploratory research and analysis. J Comput Chem 2004;25:1605–12.