Next Article in Journal
Hippeastrum reticulatum (Amaryllidaceae): Alkaloid Profiling, Biological Activities and Molecular Docking
Next Article in Special Issue
Activation Profile Analysis of CruCA4, an α-Carbonic Anhydrase Involved in Skeleton Formation of the Mediterranean Red Coral, Corallium rubrum
Previous Article in Journal
Recent Development of Non-Peptide GnRH Antagonists
Previous Article in Special Issue
Development of a Non-Hydroxamate Dual Matrix Metalloproteinase (MMP)-7/-13 Inhibitor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Five- and Six-Membered Nitrogen-Containing Compounds as Selective Carbonic Anhydrase Activators

1
Department of Pharmacy, “G. d’Annunzio” University of Chieti-Pescara, Via dei Vestini 31, 66100 Chieti, Italy
2
Department of Pharmacology, Bezmialem Vakif University, Vatan Caddesi, 34093 Fatih/Istanbul, Turkey
3
Neurofarba Department, Università degli Studi di Firenze, Via U. Schiff 6, 50019 Sesto Fiorentino (Florence), Italy
*
Author to whom correspondence should be addressed.
Molecules 2017, 22(12), 2178; https://doi.org/10.3390/molecules22122178
Submission received: 27 October 2017 / Revised: 30 November 2017 / Accepted: 5 December 2017 / Published: 9 December 2017
(This article belongs to the Special Issue Metalloenzyme Inhibitors and Activators)

Abstract

:
It has been proven that specific isoforms of human carbonic anhydrase (hCA) are able to fine-tune physiological pathways connected to signal processing, and that decreased CAs expression negatively influences cognition, leading to mental retardation, Alzheimer’s disease, and aging-related cognitive dysfunctions. For this reason, a small library of natural and synthetic nitrogen containing cyclic derivatives was assayed as activators of four human isoforms of carbonic anhydrase (hCA I, II, IV and VII). Most of the compounds activated hCA I, IV and VII in the micromolar range, with KAs ranging between 3.46 and 80.5 μM, whereas they were not active towards hCA II (KAs > 100 μM). Two natural compounds, namely l-(+)-ergothioneine (1) and melatonin (2), displayed KAs towards hCA VII in the nanomolar range after evaluation by a CO2 hydration method in vitro, showing a rather efficient and selective activation profile with respect to histamine, used as a reference compound. Corroborated with the above in vitro findings, a molecular modelling in silico approach has been performed to correlate these biological data, and to elucidate the binding interaction of these activators within the enzyme active site.

Graphical Abstract

1. Introduction

Carbonic anhydrases (CAs, EC 4.2.1.1) are a large family of zinc-containing metalloenzymes found in prokaryotes and eukaryotes, involved in the reversible conversion of carbon dioxide into bicarbonate and a proton. The specific activation of this enzyme has been explored in animal models of neurodegenerative or age-related diseases with the aim of proposing new innovative pathways able to contrast cognitive impairment and brain functional loss [1,2,3,4]. As evidence of this concept, the management of such pathological conditions in which learning and memory are altered relies on a large plethora of studies demonstrating how neuronal responses, especially in the hippocampus, were strictly modulated by an increased bicarbonate conductance through synaptic GABAA receptor channels and by extracellular signal-regulated kinase (ERK) phosphorylation in the amygdala [5,6]. Carbonic anhydrase could be genetically dysregulated in several deficiency syndromes, where mutations of CA-related genes led to clinical manifestations of autosomal recessive osteopetrosis, renal tubular acidosis, increased frequency of fractures and dental abnormalities, and delay of development [7]. Impaired neuronal functioning was often coupled to brain calcification affecting oligodendrocytes, astrocytes, and some pyramidal neurons [8]. Many isoforms of CAs are expressed in the central nervous system (CNS) contributing to signal processing, better performance in the Morris maze test, fear memory, attention gating of memory storage, spatial learning, cerebrospinal fluid composition and turnover, and long term synaptic transmission especially in the cortex and hippocampus [9,10]. Moreover, local pH fluctuations were shown to modulate physiological responses sustained by the NMDA receptors [11].
Up to date, CA activators are chemically represented by natural endogenous compounds (histamine, catecholamines, amino acids/peptides, carnosine) [12,13,14,15,16,17,18,19,20] and some synthetic derivatives (oligopeptides, pyridinium-azoles, selective serotonin reuptake inhibitors, histidine and histamine derivatives, imidazole/triazole/indazole/pyrazole/oxazole/thiadiazole derivatives) [21,22,23,24,25,26,27,28,29,30,31,32] presenting common features such as:
  • Steric factors leading to a proper orientation for facilitating the proton shuttle from the entrance to the active site;
  • Electronic factors like pKa values in the range 6.5–8.0 for protonatable moieties suitable to establish a network of H-bonds with water molecules or amino acid residues;
  • Structural characteristics like nitrogen-containing rings or chains to provide the protonatable moiety.
Collectively, this proposed pharmacophore would allow the binding of the activator to the entrance of the active site cavity differently from the classical inhibitors binding site, the establishment of a re-organization of H-bonds within this area, and the corresponding enzymatic activity facilitating the His64-mediated (His66 in CA VII according to the numbering of amino acid residues) proton shuttle transfer reaction.
The complete elucidation of this mechanism was carefully demonstrated by Temperini et al. with the support of in-depth kinetic, spectroscopic, and crystallographic studies [33]. For this reason, CA activators (CAAs) were widely designed as important pharmacological and diagnostic tools [34] showing CA activation profiles up to the nanomolar range, and with interesting isoform specificity [35,36].
The most promising results in terms of activation potency and isoform selectivity were disclosed in order to extrapolate the structure-activity relationship (SAR) useful for the design of novel CAAs. On the basis of the enzyme-activator adducts and keeping in mind the CNS distribution of hCA I, II, IV, and VII isozymes, we propose novel chemotypes incorporating all the above mentioned characteristics in five- or six-membered nitrogen containing structures to elicit this peculiar biological activity and to explore the chemical space around these positions.
We selected important derivatives such as l-(+)-ergothioneine (1), an anti-oxidant and chelating naturally-occurring compound containing a sulfur atom in the imidazole ring of histidine, melatonin (2), which is a first-line defense hormone against oxidative stress in the CNS, 4-imidazole-acrylic acid (3) and indazole-3-carboxylic acid (4), both incorporating acid and basic functions, TIC.HCl ((S)-1,2,3,4-tetrahydro-3-isoquinolinecarboxylic acid) (5) and spinacine ((S)-4,5,6,7-tetrahydro-1H-imidazo[4,5-c]pyridine-6-carboxylic acid) (6) in order to evaluate the impact of the presence/absence of an imidazole ring. Lastly, we synthesized and purified diketopiperazine derivatives (79) characterized by an imidazole ring linked to a nitrogen-containing six-membered nucleus (Figure 1).
All the compounds (19) were subjected to in vitro and in silico exhaustive evaluations as CA activators. This experimental design let us to discover innovative lead compounds not only selective and potent activators, but also endowed with appropriate physical-chemical characteristics such as balanced hydro/liposolubility and appropriate water solubility for the steps towards commercial drugs.

2. Chemistry

Derivatives 16 were commercially available. Conversely, a synthetic strategy to prepare derivatives (79) has been developed in our laboratory [37] involving Boc-solution chemistry (Scheme 1). The starting common dipeptide precursor was obtained by coupling Boc-(L)HisOH with L-PheOCH3, l-MetOCH3 and l-ProOCH3 in N,N-dimethylformamide (DMF). The obtained product was then deprotected and cyclized into the diketopiperazine derivatives (79) in high yield as reported in Scheme 1.

3. Biological Evaluation

All the synthesized and purchased (high purity grade) compounds were subjected to tests to measure their activation of the most important cytosolic isoforms, hCA I, II and VII, and the membrane-associated hCA IV by a stopped-flow, CO2 hydrase assay method. For comparison, we also discussed the activation profile of the standard histamine (Hst) against the same isoforms (Table 1).

4. Results and Discussion

Collectively, all compounds, except compound 2 (melatonin) have the possibility for accepting a proton, and can function as proton shuttles. From the data in Table 1, the following relationships between structure and activation profile can be concluded for this small library of nitrogen containing cyclic compounds and the first demonstrated standard activator (histamine):
(i)
The abundant and cytosolic hCA I isozyme was moderately activated by compounds 39 similarly to histamine (micromolar range), whereas compounds 1 and 2 did not affect its enzymatic activity up to 100 μM;
(ii)
As histamine, all compounds reported in this work were totally inactive against hCA II isozyme, usually abundant in the choroid plexus, oligodendrocytes, astrocytes, myelinated tracts, and myelin sheets;
(iii)
With regards to the membrane-bound hCA IV isoform located on the luminal surface of cerebral capillaries and within the cortex, the hippocampus and thalamus, compounds 29 were poor activators (KAs = 53.4–80.5 μM) and with the same activation profile of histamine in the micromolar range. Conversely, compound 1 was totally inactive (KA > 100 μM);
(iv)
The last cytosolic hCA VII isoform displayed the most promising results in terms of activation and selectivity. Derivatives 39 activated this isozyme in the micromolar range, but compounds 1 and 2 were demonstrated to be rather efficient activators with KA values of 820 and 120 nM, respectively (best-in-class activators, more selective up to two orders of magnitude). Compound 1 showed a very selective activation of only this isoform;
(v)
Collectively, the synthetic diketopiperazines (79) were as potent micromolar activators as histamine.
The activation of the brain-associated hCA VII isoform is of particular interest not only because it is present in high levels in the cortex, thalamus and hippocampus, but also owing to the discovery of its pharmacological interaction with anti-neuropathic and anti-epilepsy agents.

5. Docking Studies

Compounds 19 and histamine were docked into the active site of hCA VII, the enzyme against which the ligands show the lowest measured KA values, using the GOLD docking software and the ChemScore scoring function. In our previous studies, the ChemScore scoring function showed better results with respect to the suggested docked poses and hydrogen bond interactions to the hCA active sites.
l-(+)-Ergothioneine (1) showed a potent activation constant for hCA VII, whereas it displayed KA > 100 μM for all other tested isozymes. It can adopt different docked poses for both its stereoisomers in which the acidic group of the ligand may be involved in the proton transfer from the zinc-bound water molecule to the solvent, as graphically described in Figure 2. In the first pose (Figure 2A), the carboxylic acid points towards to the zinc-bound water molecule (the shortest distance between carboxylic acid and oxygen of water molecule: 4.293 Å). Hydrogen bonds are formed with the side chains of Gln69 and Gln94, and the backbone carbonyl group of Pro203. In addition, the ligand’s sulfur atom may form an interaction with the side chain of Trp7. In the second pose (Figure 2B), the carboxyl group of the ligand forms a direct hydrogen bond to the zinc-bound water molecule. Additional hydrogen bonds are formed with the Thr202 (side chain and backbone) and Pro203. The cationic amine group of the ligand is water accessible.
Melatonin (2) formed hydrogen bonds with Gln69 and Gln94 through its carbonyl group, and with the Pro203 through its indole amine group (Figure 3). The indole ring established hydrophobic interactions with the side chain of Leu200. This docked pose placed the methoxy group of melatonin close to the zinc-bound water molecule. The hydrogen bond, that may be formed, could further assist in the transportation of the proton to the solvent.
Compounds 36 showed similar docked poses in which the carboxylic acid moiety approached the zinc-bound water molecule in a comparable way, as observed for compound 1. Conversely, various docked poses have been obtained for diketopiperazines 79, in which the imidazole moiety approached the zinc-bound water molecule, or was located close to the proton shuttle His66 (Figure 4). In one docked pose for compound 7 (Figure 4A), the imidazole formed a hydrogen bond with the zinc-bound water molecule, whereas the unsubstituted phenyl group formed hydrophobic interactions with Phe133. Additional hydrogen bonds were formed with the side chain of Gln69. In another docked pose for compound 8 (Figure 4B), the imidazole ring of the ligand was pointing towards the solvent and established a hydrogen bond with the backbone of Pro203. The aliphatic tail of the ligand entered a hydrophobic pocket near Val125, Leu143 and Leu200. A hydrogen bond was also formed between the carbonyl group of the ligand and the side chain of Thr202.
Interestingly, similar docked poses to the ones obtained for hCA VII (Figure 2, Figure 3 and Figure 4) have also been observed for hCA I, II and IV despite some slight differences in their active sites as reported in Table 2. Many of the residues of hCA VII that are involved in hydrogen bonding interaction with the activators are conserved amongst hCA I, II and IV (Table 2; Trp7, Gln94 and Pro203). Only Gln69 is not conserved amongst the residues that form hydrogen bonding interactions to the ligands. As such, even though similar binding interactions may be possible, the binding strength for hCA VII is expected to be higher compared to hCA I, II and IV.

6. Experimental Protocols

All reactions involving air- or moisture-sensitive compounds were performed under a nitrogen atmosphere using dried glassware and syringe techniques to transfer solutions. Compounds 16 were provided by Sigma-Aldrich (Milan, Italy) and used in the biological assays without further purification (purity ≥ 98%). All other synthesized compounds (79) were correctly characterized by analytical and spectral data to ensure structure elucidation and purity. Column chromatography was carried out using Sigma-Aldrich® silica gel (high purity grade, pore size 60 Å, 200–425 mesh particle size). Analytical thin-layer chromatography was carried out on Sigma-Aldrich® silica gel on aluminum foils with fluorescent indicator. Visualization of the spots was carried out under ultra-violet (UV) irradiation (254 and 365 nm) and under development in an iodine chamber. Ashless Whatman Grade 42 (Sigma-Aldrich, Milan, Italy) filter paper was used for vacuum filtration. 1H-NMR (nuclear magnetic resonance) spectra were recorded on a Bruker AV300 (Bruker Biospin Corp., Billerica, MA, USA). The assignment of exchangeable protons was confirmed by the addition of D2O. Chemical shifts are reported in parts per million (ppm) and referenced to the characteristic non-deuterated solvent peak (2.50 ppm for hexadeuterated dimethylsulfoxide, DMSO-d6). Temperatures are reported in °C. Systematic compound names are extrapolated by ChemBioDraw Ultra® 12.0 (https://chemistry.com.pk/software/free-download-chemdraw-ultra-12/) following IUPAC conventions. List of abbreviations: DMF, N,N-dimethylformamide; DCC, N,N'-dicyclohexylcarbodiimide; s.s., saturate solution; DIPEA, N,N-diisopropylethylamine; TFA, trifluoroacetic acid; DCM, dichloromethane; NMM, N-methylmorpholin.
Boc(L)His-PheOMe: To an ice cooled mixture of Boc(L)His-OH (1.0 equiv.) in DMF, DCC (1.1 equiv.), HOBt·H2O (1.1 equiv.), DIPEA (2.2 equiv.) and HCl·NH2Phe-OMe (1.0 equiv.) were added under anhydrous conditions. The mixture was stirred for 20 min and then was left to warm at room temperature (r.t.) overnight. The solvent was evaporated under reduced pressure and the solid residue was crystallized from EtOAc. All the collected organic phases were evaporated under reduced pressure. The residue was dissolved in EtOAc/1-butanol (2:1, v:v) and washed with 5% citric acid (three times), NaHCO3 s.s. (three times) and brine (three times). The organic phase was dried on anhydrous Na2SO4, filtered and the solvent removed under reduced pressure to give the crude product which was purified by silica gel column chromatography (gradient from CHCl3/MeOH = 97:3 to CHCl3/MeOH = 93:7, v:v) to obtain the final product in 70% yield. 1H-NMR (300 MHz, DMSO-d6) δ: 1.30 (s, 9H, Boc-His), 2.61–2.74 (m, 2H, Phe-CH2), 2.86–2.99 (m, 2H, His-CH2), 3.54 (s, 3H, Phe-OCH3), 4.10–4.14 (m, 1H, His-αCH), 4.43–4.47 (m, 1H, Phe-αCH), 6.69 (t, 1H, Boc-NH), 6.84 (d, 1H, CCHNH), 7.17–7.27 (m, 5H, Phe aromatic ring), 7.51 (d, 1H, amide), 8.18 (d, 1H, NCHNH).
(3S,6S)-3-((1H-Imidazol-5-yl)methyl)-6-benzylpiperazine-2,5-dione (7): Boc(L)His-PheOMe was deprotected with a mixture of TFA/CH2Cl2 = 1:1 (v:v) at r. t. for 1 h. The intermediate TFA salt was used for the following reaction without further purification. The so obtained product was dissolved in 0.1 M AcOH/2-butanol solution (1.5 equiv.), then NMM (1 equiv.) was added and stirred under reflux at 120 °C overnight. The solvent was removed in rotavapor, the solid residue was collected by filtration and washed with a small amount of 0.1 M AcOH/2-butanol solution, to afford the desired product in 68% yield. 1H-NMR (300 MHz, DMSO-d6) δ: 2.72–2.84 (m, 2H, Phe-CH2), 2.94–3.04 (m, 2H, His-CH2), 3.47–3.89 (m, 1H, His-αCH), 4.04–4.19 (m, 1H, Phe-αCH), 6.84 (d, 1H, CCHNH), 7.17–7.27 (m, 5H, Phe aromatic ring), 7.51 (d, 1H, His amide), 8.18 (d, 1H, NCHNH), 8.79 (d, 1H, Phe amide).
Boc(L)His-MetOMe: To an ice cooled mixture of Boc(L)HisOH (1.0 equiv.) in DMF, DCC (1.1 equiv.), HOBt·H2O (1.1 equiv.), DIPEA (2.2 equiv.) and HCl·NH2(L)MetOMe (1.0 equiv.) were added. The reaction mixture was stirred at 0 °C for 20 min and then it was allowed to warm at r.t. overnight. Solvent was removed under reduced pressure, the solid residue was then dissolved in EtOAc/1-butanol (2:1, v:v) and washed with 5% citric acid solution (three times), saturated solution of NaHCO3 (three times) and brine (three times). The combined organic phases were dried on anhydrous Na2SO4, filtered and evaporated under reduced pressure to give the desired crude product. The crude product was purified by silica gel column chromatography (eluent: CHCl3/MeOH = 97:3, v:v) to obtain the final product in 58% yield. 1H-NMR (300 MHz, DMSO-d6) δ: 1.31 (s, 9H, Boc-His), 1.85–1.92 (m, 2H, Met-CH2), 1.99 (s, 3H, S-CH3), 2.44–2.58 (m, 2H, Met-CH2), 2.71–2.79 (m, 2H, His-CH2), 3.58 (s, 3H, Met-OCH3), 4.08–4.14 (m, 1H, His-αCH), 4.32–4.37 (m, 1H, Met-αCH), 6.73 (d, 1H, CCHNH), 7.50 (d, 1H, NCHNH), 8.21 (d, 1H, Boc-NH), 8.28 (d, 1H, amide).
(3S,6S)-3-((1H-Imidazol-5-yl)methyl)-6-(2-(methylthio)ethyl)piperazine-2,5-dione (8): Boc(L)His-MetOMe was deprotected by a mixture of TFA/CH2Cl2 = 1:1 (v:v) at r.t. for 1 h. The intermediate TFA salt was used for the following cyclization without further purification. His-MetOMe as TFA salt was dissolved in 0.1 M AcOH/2-butanol solution (1.5 equiv.) and then NMM (1 equiv.) was added. The reaction mixture was stirred under reflux at 120 °C overnight, then it was concentrated in rotavapor and filtered; the solid was washed with a solution of 0.1 M AcOH/2-butanol and dried in rotavapor under reduced pressure. The crude product was purified by silica gel column chromatography (eluent: CHCl3/MeOH/H2O/AcOH = 70:10:10:10, v:v:v:v) to obtain the final product in 72% yield. 1H-NMR (300 MHz, DMSO-d6) δ: 1.99 (s, 3H, S-CH3), 2.25–2.32 (m, 2H, Met-CH2), 2.86–2.93 (m, 2H, Met-CH2), 3.79–3.86 (m, 1H, His-αCH), 4.01–4.05 (m, 1H, Met-αCH), 6.76 (d, 1H, CCHNH), 7.51 (d, 1H, NCHNH), 8.06 (d, 1H, His-amide), 8.14 (d, 1H, Met-amide).
Boc(L)His-ProOMe: To an ice cooled mixture of Boc(L)HisOH (1.0 equiv.) in DMF, DCC (1.1 equiv.), HOBt·H2O (1.1 equiv.), DIPEA (2.2 equiv.) and HCl·NH-ProOMe (1.0 equiv.) were added. The reaction mixture was stirred at 0 °C for 20 min and then it was allowed to warm at r.t. overnight. The reaction mixture was evaporated in rotavapor. The residue was dissolved in EtOAc/1-butanol (2:1, v:v) and washed with three portions of 5% citric acid solution (three times), NaHCO3 s.s. (three times) and brine (three times). The combined organic phases were dried on anhydrous Na2SO4, filtered and evaporated under reduced pressure to give the crude product which was purified by silica gel column chromatography (eluent: CHCl3/MeOH = 97:3, v:v) to obtain the final product in 77% yield. 1H-NMR (300 MHz, DMSO-d6) δ: 1.30 (s, 9H, Boc-His), 1.82–1.91 (m, 3H, Pro-CH2), 2.14–2.23 (m, 1H, Pro-CH2), 2.67–2.74 (m, 3H, His-CH2 + Pro-CH), 3.56–3.62 (m, 4H, 1H-Pro-CH + 3H Pro-OCH3), 4.29 (m, 1H, Pro-αCH), 4.32 (m, 1H, His-αCH), 6.80 (d, 1H, CCHNH), 7.13 (d, 1H, Boc-NH), 7.52 (d, 1H, NCHNH).
(3S,8aS)-3-((1H-Imidazol-5-yl)methyl)hexahydropyrrolo[1,2-a]pyrazine-1,4-dione (9): Boc(L)His-ProOMe was deprotected with a mixture of TFA/CH2Cl2 = 1:1 (v:v) at r.t. for 1 h. The intermediate TFA salt was used for cyclization reaction without further purification. (L)His-ProOMe as TFA salt was dissolved in 0.1 M AcOH/2-butanol solution (1.5 equiv.) and NMM (1 equiv.) was added. The resulting solution was stirred under reflux at 120 °C overnight. The solvent was concentrated in rotavapor, followed by trituration with diethyl ether. The crude product was purified by silica gel column chromatography (eluent: CHCl3/MeOH/H2O/AcOH = 75:27:5:0.5, v:v:v:v) to obtain the final product in 71% yield. 1H-NMR (300 MHz, DMSO-d6) δ: 1.82–1.91 (m, 2H, Pro-CH2), 2.10 (m, 1H, Pro-CH2), 2.71–2.70 (m, 3H, His-CH2 + Pro-CH), 3.60 (m, 1H, 1H-Pro-CH), 4.17–4.10 (m, 2H, Pro-αCH + His-αCH), 4.32 (m, 1H, His-αCH), 6.87 (d, 1H, CCHNH), 7.56 (d, 1H, NCHNH), 8.06 (d, 1H, amide).

6.1. Carbonic Anhydrase Assays

A stopped-flow method has been used for assaying the CA-catalyzed CO2 hydration activity with Phenol red as an indicator, working at the absorbance maximum of 557 nm, following the initial rates of the CA-catalyzed CO2 hydration reaction for 10–100 s [38]. For each activator, at least six traces of the initial 5–10% of the reaction have been used for assessing the initial velocity. The uncatalyzed rates were measured in the same manner and subtracted from the total observed rates. Stock solutions of activator (0.1 mM) were done in distilled-deionized water and successive dilutions up to 0.1 nM were then performed with the assay buffer. The activation constant (KA), defined similarly to the inhibition constant KI, was calculated by the classical Michaelis-Menten equation (Equation (1)) fitted by non-linear least squares by using PRISM 3:
v = vmax/{1 + KM/[S] (1 + [A]f/KA)}
where [A]f is the free concentration of activator.
Working at substrate concentrations markedly lower than KM ([S] << KM), and keeping in mind that [A]f can be expressed as the total concentration of the activator ([A]t) and enzyme ([E]t), the obtained competitive steady-state equation for measuring the activation constant is provided by Equation (2):
v = v0·KA/{KA + ([A]t − 0.5 {([A]t + [E]t + KA) − ([A]t + [E]t + KA)2 − 4[A]t·[E]t)1/2}}
where v0 is the initial velocity of the enzymatic reaction in the absence of activator. All carbonic anhydrase isoforms used in these experiments were obtained in-house as purified recombinant proteins [39,40,41,42,43,44].

6.2. Molecular Modelling Studies

6.2.1. Preparation of Files for Docking Studies

Crystal structures of hCA I (pdb: 3lxe, 1.9 Å, in complex with topiramate), II (pdb: 4e3d, 1.6 Å, in complex with 2,5-dihydroxybenzoic acid), IV (pdb: 3fw3, 1.72 Å, in complex with dorzolamide) and VII (pdb: 3mdz, 2.32 Å, in complex with ethoxzolamide) were obtained from the Brookhaven Protein Data Bank. All ligands (topiramate, 2,5-dihydroxybenzoic acid, ethoxzolamide and dorzolamide) and the zinc-bound water molecule of hCA II were retained and all other non-protein atoms were deleted. Hydrogen atoms were added with the “protonate 3D” tool, and subsequently a steepest-descent energy minimization was performed using the AMBER12:EHT force field (MOE software package, version 2015.10, chemical computing group, Inc., Montreal, QC, Canada). The four protein structures were superposed on the backbone atoms of hCA I (Cα atoms, RMSD: 1.395 Å, for 236 residues). The coordinates of the hCA II zinc-bound water molecule were copied into the other hCA structures.
The molecular structures of the activators were prepared with the MOE software package. All stereoisomers were generated. All strong bases were protonated and all strong acids were deprotonated. Subsequently, the ligand structures were energy minimized (MMFF94x force field) and the ligands were saved as multi-mol2 files.

6.2.2. Docking Studies

Docking calculations were performed with the GOLD software package (v5.4, CCDC, Cambridge, UK) using the ChemScore scoring function (50 dockings) and default settings [45,46]. The binding pocket was defined as within 16 Å around a centroid (x: −18.459, y: 35.3775, z: 43.564; OAR atom of topiramate, 3lxe structure) as previously reported [45,46].

7. Conclusions

Due to the significantly diminished activity or expression of specific isoforms of carbonic anhydrases in the brain of patients suffering from ageing or Alzheimer’s diseases, we proposed novel nitrogen containing chemotypes incorporating five- or six-membered cycles able to elicit the proton shuttle, and thus to activate the catalytic activity of CA. Natural and synthetic compounds were investigated against four isoforms of CA usually involved in CNS development and functioning. We obtained promising results in terms of activation activity (up to nanomolar range) and isoform selectivity (especially towards hCA VII).

Acknowledgments

This work was financed by two FP7 EU projects (Metoxia and Dynano) to CTS.

Author Contributions

A.M. conceived the study, G.M. and A.S. synthesized, purified and characterized the compounds, S.C. wrote the study, A.K. performed the molecular modelling studies, A.A. and C.T.S. carried out the biological evaluation.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Meier-Ruge, W.; Iwangoff, P.; Reichlmeier, K. Neurochemical enzyme changes in Alzheimer’s and Pick’s disease. Arch. Gerontol. Geriatr. 1984, 3, 161–165. [Google Scholar] [PubMed]
  2. Meier-Ruge, W.; Iwangoff, P.; Reichlmeier, K.; Sandoz, P. Neurochemical findings in the aging brain. Adv. Biochem. Psychopharmacol. 1980, 23, 323–338. [Google Scholar] [PubMed]
  3. Sun, M.K.; Alkon, D.L. Pharmacological enhancement of synaptic efficacy, spatial learning, and memory through carbonic anhydrase activation in rats. J. Pharmacol. Exp. Ther. 2001, 297, 961–967. [Google Scholar] [PubMed]
  4. Sun, M.K.; Alkon, D.L. Carbonic anhydrase gating of attention: Memory therapy and enhancement. Trends Pharmacol. Sci. 2002, 23, 83–89. [Google Scholar] [PubMed]
  5. Yang, M.T.; Chien, W.L.; Lu, D.H.; Liou, H.C.; Fu, W.M. Acetazolamide impairs fear memory consolidation in rodents. Neuropharmacology 2013, 67, 412–418. [Google Scholar] [PubMed]
  6. Pasternack, M.; Voipio, J.; Kaila, K. Intracellular carbonic anhydrase activity and its role in GABA-induced acidosis in isolated rat hippocampal pyramidal neurons. Acta Physiol. Scand. 1993, 148, 229–231. [Google Scholar] [PubMed]
  7. Sly, W.S.; Hu, P.Y. Human carbonic anhydrases and carbonic anhydrase deficiencies. Annu. Rev. Biochem. 1995, 64, 375–401. [Google Scholar] [CrossRef] [PubMed]
  8. Bosley, T.M.; Salih, M.A.; Alorainy, I.A.; Islam, M.Z.; Oystreck, D.T.; Suliman, O.S.M.; al Malki, S.; Suhaibani, A.H.; Khiari, H.; Beckers, S.; et al. The neurology of carbonic anhydrase type II deficiency syndrome. Brain 2011, 134, 3502–3515. [Google Scholar] [CrossRef] [PubMed]
  9. Kida, E.; Palminiello, S.; Golabek, A.A.; Walus, M.; Wierzba-Bobrowicz, T.; Rabe, A.; Albertini, G.; Wisniewski, K.E. Carbonic anhydrase II in the developing and adult human brain. J. Neuropathol. Exp. Neurol. 2006, 7, 664–674. [Google Scholar]
  10. Canto de Souza, L.; Provensi, G.; Vullo, D.; Carta, F.; Scozzafava, A.; Costa, A.; Schmidt, S.D.; Passani, M.B.; Supuran, C.T.; Blandina, P. Carbonic anhydrase activation enhances object recognition memory in mice through phosphorylation of the extracellular signal-regulated kinase in the cortex and the hippocampus. Neuropharmacology 2017, 118, 148–156. [Google Scholar] [PubMed]
  11. Chesler, M. Regulation and modulation of pH in the brain. Physiol. Rev. 2003, 83, 1183–1221. [Google Scholar] [CrossRef] [PubMed]
  12. Saada, M.C.; Vullo, D.; Montero, J.-L.; Scozzafava, A.; Winum, J.-Y.; Supuran, C.T. Carbonic anhydrase I and II activation with mono- and dihalogenated histamine derivatives. Bioorg. Med. Chem. Lett. 2011, 21, 4884–4887. [Google Scholar] [CrossRef] [PubMed]
  13. Saada, M.C.; Vullo, D.; Montero, J.-L.; Scozzafava, A.; Supuran, C.T.; Winum, J.-Y. Mono- and di-halogenated histamine, histidine and carnosine derivatives are potent carbonic anhydrase I, II, VII, XII and XIV activators. Bioorg. Med. Chem. 2014, 22, 4752–4758. [Google Scholar] [CrossRef] [PubMed]
  14. Temperini, C.; Innocenti, A.; Scozzafava, A.; Supuran, C.T. Carbonic anhydrase activators: Kinetic and X-ray crystallographic study for the interaction of D- and L-tryptophan with the mammalian isoforms I-XIV. Bioorg. Med. Chem. 2008, 16, 8373–8378. [Google Scholar] [CrossRef] [PubMed]
  15. Temperini, C.; Scozzafava, A.; Supuran, C.T. Carbonic anhydrase activators: The first X-ray crystallographic study of an adduct of isoform I. Bioorg. Med. Chem. Lett. 2006, 16, 5152–5156. [Google Scholar] [CrossRef] [PubMed]
  16. Temperini, C.; Innocenti, A.; Scozzafava, A.; Mastrolorenzo, A.; Supuran, C.T. Carbonic anhydrase activators: L-adrenaline plugs the active site entrance of isozyme II, activating better isoforms I, IV, VA, VII, and XIV. Bioorg. Med. Chem. Lett. 2007, 17, 628–635. [Google Scholar] [CrossRef] [PubMed]
  17. Vullo, D.; Nishimori, I.; Innocenti, A.; Scozzafava, A.; Supuran, C.T. Carbonic anhydrase activators: An activation study of the human mitochondrial isoforms VA and VB with amino acids and amines. Bioorg. Med. Chem. Lett. 2007, 17, 1336–1340. [Google Scholar] [PubMed]
  18. Pastorekova, S.; Vullo, D.; Nishimori, I.; Scozzafava, A.; Pastork, J.; Supuran, C.T. Carbonic anhydrase activators: Activation of the human tumor-associated isozymes IX and XII with amino acids and amines. Bioorg. Med. Chem. 2008, 16, 3530–3536. [Google Scholar] [CrossRef] [PubMed]
  19. Vullo, D.; Nishimori, I.; Scozzafava, A.; Supuran, C.T. Carbonic anhydrase activators: Activation of the human cytosolic isozyme III and membrane-associated isoform IV with amino acids and amines. Bioorg. Med. Chem. Lett. 2008, 18, 4303–4307. [Google Scholar] [CrossRef] [PubMed]
  20. Dave, K.; Ilies, M.A.; Scozzafava, A.; Temperini, C.; Vullo, D.; Supuran, C.T. An inhibitor-like binding mode of a carbonic anhydrase activator within the active site of isoform II. Bioorg. Med. Chem. Lett. 2011, 21, 2764–2768. [Google Scholar] [CrossRef] [PubMed]
  21. Le Duc, Y.; Licsandru, E.; Vullo, D.; Barboiu, M.; Supuran, C.T. Carbonic anhydrases activation with 3-amino-1H-1,2,4-triazole-1-carboxamides: Discovery of subnanomolar isoform II activators. Bioorg. Med. Chem. 2017, 25, 1681–1686. [Google Scholar] [PubMed]
  22. Supuran, C.T. Carbonic anhydrases: Novel therapeutic applications for inhibitors and activators. Nat. Rev. Drug Discov. 2008, 7, 168–181. [Google Scholar] [CrossRef] [PubMed]
  23. Supuran, C.T. How many carbonic anhydrase inhibition mechanisms exist? J. Enzym. Inhib. Med. Chem. 2016, 31, 345–360. [Google Scholar] [CrossRef] [PubMed]
  24. Scozzafava, A.; Supuran, C.T. Carbonic anhydrase activators: Human isozyme II is strongly activated by oligopeptides incorporating the carboxyterminal sequence of the bicarbonate anion exchanger AE1. Bioorg. Med. Chem. Lett. 2002, 12, 1177–1180. [Google Scholar] [CrossRef]
  25. Temperini, C.; Scozzafava, A.; Puccetti, L.; Supuran, C.T. Carbonic anhydrase activators: X-ray crystal structure of the adduct of human isozyme II with L-histidine as a platform for the design of stronger activators. Bioorg. Med. Chem. Lett. 2005, 15, 5136–5141. [Google Scholar] [CrossRef] [PubMed]
  26. Abdo, M.R.; Vullo, D.; Saada, M.C.; Montero, J.L.; Scozzafava, A.; Winum, J.Y.; Supuran, C.T. Carbonic anhydrase activators: Activation of human isozyme I, II and IX with phenylsulfonylhydrazido L-histidine derivatives. Bioorg. Med. Chem. Lett. 2009, 19, 2440–2443. [Google Scholar] [CrossRef] [PubMed]
  27. Ilies, M.; Banciu, M.D.; Ilies, M.A.; Scozzafava, A.; Caproiu, M.T.; Supuran, C.T. Carbonic anhydrase activators: Design of high affinity isozymes I, II, and IV activators, incorporating tri/tetrasubstituted-pyridinium-azole moieties. J. Med. Chem. 2002, 45, 504–510. [Google Scholar] [PubMed]
  28. Casini, A.; Caccia, S.; Scozzafava, A.; Supuran, C.T. Carbonic anhydrase activators. The selective serotonin reuptake inhibitors fluoxetine, sertraline and citalopram are strong activators of isozymes I and II. Bioorg. Med. Chem. Lett. 2003, 13, 2765–2768. [Google Scholar] [CrossRef]
  29. Licsandru, E.; Tanc, M.; Kocsis, I.; Barboiu, M.; Supuran, C.T. A class of carbonic anhydrase I—Selective activators. J. Enzym. Inhib. Med. Chem. 2017, 32, 37–46. [Google Scholar] [CrossRef] [PubMed]
  30. Angeli, A.; Vaiano, F.; Mari, F.; Bertol, E.; Supuran, C.T. Psychoactive substances belonging to the amphetamine class potently activate brain carbonic anhydrase isoforms VA, VB, VII, and XII. J. Enzym. Inhib. Med. Chem. 2017, 32, 1253–1259. [Google Scholar] [CrossRef] [PubMed]
  31. Draghici, B.; Vullo, D.; Akocak, S.; Walker, E.A.; Supuran, C.T.; Ilies, M.A. Ethylene bis-imidazoles are highly potent and selective activators for isozymes VA and VII of carbonic anhydrase, with a potential nootropic effect. Chem. Commun. 2014, 50, 5980–5983. [Google Scholar]
  32. Maccallini, C.; Di Matteo, M.; Vullo, D.; Ammazzalorso, A.; Carradori, S.; De Filippis, B.; Fantacuzzi, M.; Giampietro, L.; Pandolfi, A.; Supuran, C.T.; et al. Indazole, pyrazole, and oxazole derivatives targeting Nitric Oxide Synthases and Carbonic Anhydrases. ChemMedChem 2016, 11, 1695–1699. [Google Scholar] [CrossRef] [PubMed]
  33. Temperini, C.; Scozzafava, A.; Supuran, C.T. Carbonic anhydrase activation and the drug design. Curr. Pharm. Des. 2008, 14, 708–715. [Google Scholar] [CrossRef] [PubMed]
  34. Supuran, C.T. Carbonic anhydrases: From biomedical applications of the inhibitors and activators to biotechnological use for CO2 capture. J. Enzym. Inhib. Med. Chem. 2013, 28, 229–230. [Google Scholar] [CrossRef] [PubMed]
  35. Temperini, C.; Scozzafava, A.; Vullo, D.; Supuran, C.T. Carbonic anhydrase activators. activation of isoforms I, II, IV, VA, VII, and XIV with L- and D-phenylalanine and crystallographic analysis of their adducts with isozyme II:  Stereospecific recognition within the active site of an enzyme and its consequences for the drug design. J. Med. Chem. 2006, 49, 3019–3027. [Google Scholar] [PubMed]
  36. Akokak, S.; Lolak, N.; Vullo, D.; Durgun, M.; Supuran, C.T. Synthesis and biological evaluation of histamine Schiff bases as carbonic anhydrase I, II, IV, VII, and IX activators. J. Enzym. Inhib. Med. Chem. 2017, 32, 1305–1312. [Google Scholar] [CrossRef] [PubMed]
  37. Mollica, A.; Costante, R.; Mirzaie, S.; Carradori, S.; Macedonio, G.; Stefanucci, A.; Novellino, E. Preparation of constrained unnatural aromatic amino acids via unsaturated diketopiperazine intermediate. J. Heterocycl. Chem. 2016, 53, 2106–2110. [Google Scholar] [CrossRef]
  38. Khalifah, R.G. 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, 2561–2573. [Google Scholar] [PubMed]
  39. Vullo, D.; Innocenti, A.; Nishimori, I.; Scozzafava, A.; Kaila, K.; Supuran, C.T. Carbonic anhydrase activators: Activation of the human isoforms VII (cytosolic) and XIV (transmembrane) with amino acids and amines. Bioorg. Med. Chem. Lett. 2007, 17, 4107–4112. [Google Scholar] [PubMed]
  40. Parkkila, S.; Vullo, D.; Puccetti, L.; Parkkila, A.K.; Scozzafava, A.; Supuran, C.T. Carbonic anhydrase activators: Activation of isozyme XIII with amino acids and amines. Bioorg. Med. Chem. Lett. 2006, 16, 3955–3959. [Google Scholar] [CrossRef] [PubMed]
  41. Nishimori, I.; Onishi, S.; Vullo, D.; Innocenti, A.; Scozzafava, A.; Supuran, C.T. Carbonic anhydrase activators. The first activation study of the human secretory isoform VI. Bioorg. Med. Chem. 2007, 15, 5351–5357. [Google Scholar] [CrossRef] [PubMed]
  42. Mollica, A.; Locatelli, M.; Macedonio, G.; Carradori, S.; Sobolev, A.P.; De Salvador, R.F.; Monti, S.M.; Buonanno, M.; Zengin, G.; Angeli, A.; et al. Microwave-assisted extraction, HPLC analysis, and inhibitory effects on carbonic anhydrase I, II, VA, and VII isoforms of 14 blueberry Italian cultivars. J. Enzym. Inhib. Med. Chem. 2016, 31, 1–6. [Google Scholar] [CrossRef] [PubMed]
  43. Mocan, A.; Carradori, S.; Locatelli, M.; Secci, D.; Cesa, S.; Mollica, A.; Riga, S.; Angeli, A.; Supuran, C.T.; Celia, C.; et al. Bioactive isoflavones from Pueraria lobata root and starch: Different extraction techniques and carbonic anhydrase inhibition. Food Chem. Toxicol. 2017. [Google Scholar] [CrossRef] [PubMed]
  44. Nescatelli, R.; Carradori, S.; Marini, F.; Caponigro, V.; Bucci, R.; De Monte, C.; Mollica, A.; Mannina, L.; Ceruso, M.; Supuran, C.T.; et al. Geographical characterization by MAE-HPLC and NIR methodologies and carbonic anhydrase inhibition of Saffron components. Food Chem. 2017, 221, 855–863. [Google Scholar] [CrossRef] [PubMed]
  45. Akdemir, A.; De Monte, C.; Carradori, S.; Supuran, C.T. Computational investigation of the selectivity of salen and tetrahydrosalen compounds towards the tumor-associated hCA XII isozyme. J. Enzyme Inhib. Med. Chem. 2015, 30, 114–118. [Google Scholar] [CrossRef] [PubMed]
  46. D’Ascenzio, M.; Carradori, S.; Secci, D.; Vullo, D.; Ceruso, M.; Akdemir, A.; Supuran, C.T. Selective inhibition of human carbonic anhydrases by novel amide derivatives of probenecid: Synthesis, biological evaluation and molecular modelling studies. Bioorg. Med. Chem. 2014, 22, 3982–3988. [Google Scholar] [CrossRef] [PubMed]
Sample Availability: Samples of the compounds 16 and the exopolysaccharide are available from the authors.
Figure 1. Chemical structures of histamine (Hst, reference drug), commercial products 16, and synthetic compounds 79.
Figure 1. Chemical structures of histamine (Hst, reference drug), commercial products 16, and synthetic compounds 79.
Molecules 22 02178 g001
Scheme 1. Synthesis of compounds 79. Reagents and Conditions: (a) 1.1 equiv. DCC, 1.1 equiv. HOBt·H2O, 2.2 equiv. DIPEA, 1 equiv. L-PheOCH3 or L-MetOCH3 or L-ProOCH3, DMF, room temperature (r.t.) overnight; (b) TFA:DCM = 1:1 (v:v), r.t. 1 h; (c) 1.5 equiv. 0.1 M AcOH/2-butanol solution, 1 equiv. NMM, reflux overnight, 68% yield for 7, 72% for 8, 71% for 9. For abbreviations, please refer to the Experimental Protocols section.
Scheme 1. Synthesis of compounds 79. Reagents and Conditions: (a) 1.1 equiv. DCC, 1.1 equiv. HOBt·H2O, 2.2 equiv. DIPEA, 1 equiv. L-PheOCH3 or L-MetOCH3 or L-ProOCH3, DMF, room temperature (r.t.) overnight; (b) TFA:DCM = 1:1 (v:v), r.t. 1 h; (c) 1.5 equiv. 0.1 M AcOH/2-butanol solution, 1 equiv. NMM, reflux overnight, 68% yield for 7, 72% for 8, 71% for 9. For abbreviations, please refer to the Experimental Protocols section.
Molecules 22 02178 sch001
Figure 2. Two different docked poses (A,B) of l-(+)-ergothioneine (1) in the active site of hCA VII. The proton shuttle His66 is shown in magenta.
Figure 2. Two different docked poses (A,B) of l-(+)-ergothioneine (1) in the active site of hCA VII. The proton shuttle His66 is shown in magenta.
Molecules 22 02178 g002
Figure 3. The docked pose of melatonin (2) in the active site of hCA VII. The proton shuttle His66 is shown in magenta.
Figure 3. The docked pose of melatonin (2) in the active site of hCA VII. The proton shuttle His66 is shown in magenta.
Molecules 22 02178 g003
Figure 4. Representative docked poses of compounds 7 (A) and 8 (B) in the active site of hCA VII. The proton shuttle His66 is shown in magenta.
Figure 4. Representative docked poses of compounds 7 (A) and 8 (B) in the active site of hCA VII. The proton shuttle His66 is shown in magenta.
Molecules 22 02178 g004
Table 1. Experimental activation constants (KAs) of hCA I, hCA II, hCA VII (cytosolic isozymes) and hCA IV (membrane-associated isoform) for compounds 19. Histamine (Hst) was used as reference compound.
Table 1. Experimental activation constants (KAs) of hCA I, hCA II, hCA VII (cytosolic isozymes) and hCA IV (membrane-associated isoform) for compounds 19. Histamine (Hst) was used as reference compound.
CompoundStructureKA (μM) *
hCA IhCA IIhCA IVhCA VII
1l-(+)-Ergothioneine>100>100>1000.82
2Melatonin>100>10053.40.12
34-Imidazole-acrylic acid21.3>10064.73.46
41H-Indazole-3-COOH32.4>10078.45.23
5TIC.HCl23.1>10073.53.61
6Spinacine7.21>10069.715.7
7c(HisPhe)16.3>10080.515.1
8c(HisMet)8.39>10077.021.6
9c(HisPro)7.92>10070.134.7
HstHistamine2.1012536.925.3
* Mean from three different determinations (errors in the range of 5–10% of the reported values, data not shown).
Table 2. Comparison of the active sites among the four selected hCA isoforms.
Table 2. Comparison of the active sites among the four selected hCA isoforms.
hCA I (3lxe)hCA II (4e3d)hCA IV (3fw3)hCA VII (3mdz)
Trp5 *Trp5Trp5Trp7
Val62Asn62Asn62Asn64
His64His64His64His66
His67Asn67Met67Gln69
Phe91Ile91Lys91Lys93
Gln92Gln92Gln92Gln94
His94His94His94His96
His96His96His96His98
His119His119His119His121
Ala121Val121Val121Val123
Leu131Phe131 ***--Val131--Phe133
Ala132Gly132--Lys132-- **Gly134
Ala135Val135--Gln135--Ala137
Leu141Leu141Ile141Leu143
Val143Val143Val143Val145
Leu198Leu198Leu198Leu200
Thr199Thr199Thr199Thr201
His200Thr200Thr200Thr202
Pro201Pro201Pro201Pro203
Pro202Pro202Thr202 **Pro205
Tyr204Leu204Asp204Ser206
Val207Val207Val207Val209
Trp209Trp209Trp209Trp211
* conserved but different rotamer compared to the rest; ** points away from ligand (or will point away from hCA IV ligand); *** the region near Phe131 (hCA II) is not resolved in the hCA IV (3fw3) structure.

Share and Cite

MDPI and ACS Style

Mollica, A.; Macedonio, G.; Stefanucci, A.; Carradori, S.; Akdemir, A.; Angeli, A.; Supuran, C.T. Five- and Six-Membered Nitrogen-Containing Compounds as Selective Carbonic Anhydrase Activators. Molecules 2017, 22, 2178. https://doi.org/10.3390/molecules22122178

AMA Style

Mollica A, Macedonio G, Stefanucci A, Carradori S, Akdemir A, Angeli A, Supuran CT. Five- and Six-Membered Nitrogen-Containing Compounds as Selective Carbonic Anhydrase Activators. Molecules. 2017; 22(12):2178. https://doi.org/10.3390/molecules22122178

Chicago/Turabian Style

Mollica, Adriano, Giorgia Macedonio, Azzurra Stefanucci, Simone Carradori, Atilla Akdemir, Andrea Angeli, and Claudiu T. Supuran. 2017. "Five- and Six-Membered Nitrogen-Containing Compounds as Selective Carbonic Anhydrase Activators" Molecules 22, no. 12: 2178. https://doi.org/10.3390/molecules22122178

Article Metrics

Back to TopTop