Next Article in Journal
Mesenchymal Stem Cell Treatment Perspectives in Peripheral Nerve Regeneration: Systematic Review
Next Article in Special Issue
In Silico Investigation of Potential Applications of Gamma Carbonic Anhydrases as Catalysts of CO2 Biomineralization Processes: A Visit to the Thermophilic Bacteria Persephonella hydrogeniphila, Persephonella marina, Thermosulfidibacter takaii, and Thermus thermophilus
Previous Article in Journal
Cartilage Targets of Knee Osteoarthritis Shared by Both Genders
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Sulfonamides and Their Structurally Related Derivatives on the Activity of ι-Carbonic Anhydrase from Burkholderia territorii

1
Institute of Biosciences and Bioresources, CNR, via Pietro Castellino 111, 80131 Napoli, Italy
2
Proteomics & Mass Spectrometry Laboratory, ISPAAM, CNR, via Argine 1085, 80147 Napoli, Italy
3
Section of Pharmaceutical and Nutraceutical Sciences, Department of Neurofarba, University of Florence, via U. Schiff 6, 50019 Florence, Italy
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(2), 571; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22020571
Submission received: 23 November 2020 / Revised: 30 December 2020 / Accepted: 5 January 2021 / Published: 8 January 2021
(This article belongs to the Special Issue Carbonic Anhydrases: A Superfamily of Ubiquitous Enzymes 2.0)

Abstract

:
Carbonic anhydrases (CAs) are essential metalloenzymes in nature, catalyzing the carbon dioxide reversible hydration into bicarbonate and proton. In humans, breathing and many other critical physiological processes depend on this enzymatic activity. The CA superfamily function and inhibition in pathogenic bacteria has recently been the object of significant advances, being demonstrated to affect microbial survival/virulence. Targeting bacterial CAs may thus be a valid alternative to expand the pharmacological arsenal against the emergence of widespread antibiotic resistance. Here, we report an extensive study on the inhibition profile of the recently discovered ι-CA class present in some bacteria, including Burkholderia territorii, namely BteCAι, using substituted benzene-sulfonamides and clinically licensed sulfonamide-, sulfamate- and sulfamide-type drugs. The BteCAι inhibition profile showed: (i) several benzene-sulfonamides with an inhibition constant lower than 100 nM; (ii) a different behavior with respect to other α, β and γ-CAs; (iii) clinically used drugs having a micromolar affinity. This prototype study contributes to the initial recognition of compounds which efficiently and selectively inhibit a bacterial member of the ι-CA class, for which such a selective inhibition with respect to other protein isoforms present in the host is highly desired and may contribute to the development of novel antimicrobials.

1. Introduction

Carbonic anhydrases (CAs, EC 4.2.1.1) are essential metalloenzymes in nature, which speed up a fundamental reaction for all living organisms, the hydration of a molecule of carbon dioxide (CO2) into bicarbonate (HCO3) and proton (H+), according to the following chemical reaction [1,2,3,4,5,6,7]: CO2 + H2O ⇋ HCO3 + H+. The CA superfamily is grouped into eight CA classes indicated with the Greek letters (α, β, γ, δ, ζ, η, θ and ι), whose distribution is very variegated from the most complex organisms (plants and animals) to the simplest ones (bacteria and archaea) [1,2,3,4,5]. The genome of mammals, for example, encodes only for the α-CA class, of which 15 isoforms have been identified, which accomplish specialized functions in various tissues and organs [8,9,10,11,12]. Mammal breathing is a compelling example of the importance of these enzymes. In mammals, at the peripheral tissue level, the CO2 produced by aerobic metabolism leaves the cells and enters the bloodstream due to a pressure gradient effect [13]. About 90% of CO2 flows into red blood cells and is converted into bicarbonate by CAs. Then, the produced HCO3 comes out from the red cell through an anion exchanger (AE) protein and is transported by the bloodstream to the lungs [13]. At the alveolar level, the concentration of CO2 is lower than in peripheral tissues, while there is a higher concentration of bicarbonate that is pumped into the red blood cell. Here, through the action of the inverse reaction catalyzed by CAs, the bicarbonate is transformed into water and CO2 [13]. The CO2 produced in this way is released into the bloodstream and, passing through the alveolus walls, is exhaled [13]. These reactions can also occur without the enzyme but carbonic anhydrase increases the reaction speed up to a million times (kcat = 106 s−1) [7]. At a physiological pH value, the spontaneous reversible CO2 hydration reaction in the absence of the catalyst has an effective first-order rate constant of 0.15 s−1, while the reverse reaction shows a rate constant of 50 s−1 [7]. Mammalian CAs are also involved in other important physiological processes, such as gluconeogenesis, lipogenesis and ureagenesis; transport of CO2/bicarbonate; electrolyte secretion in a variety of tissues/organs; bone resorption; calcification; renal and male reproductive tract acidification, signal transduction and formation of gastric acid [8,9,10,11,12].
In plants, α and β-CAs have an essential role in photosynthesis and biosynthetic reactions linked to it, in addition to few selected processes already mentioned above [14]. In simpler organisms, such as bacteria, Archaea and cyanobacteria α, β, γ and ι-CAs are present, with the function to balance the CO2/HCO3 concentration ratio and a role in the carbon dioxide fixation [5,6,7,14,15,16]. Marine diatoms encode for α- 𝛿-, 𝜁-, θ- and 𝜄-CAs, which are involved in carbon dioxide fixation and metabolism [17,18,19]. In protozoa are present α-CAs and the recently discovered class, the η-CA, which is involved in de novo purine/pyrimidine biosynthetic pathways [20].
The significant progress made in the DNA sequencing approach allowed the identification of genes encoding for CAs in pathogenic and non-pathogenic microorganisms [21]. Intriguing, the understanding of the function of the bacterial CAs has increased significantly [6,22], confirming that the activity of CAs is connected to the survival as well as the virulence of pathogens because the metabolic reaction catalyzed by these enzymes is essential for supporting numerous physiological functions involving dissolved inorganic carbon [6,22]. Based on these considerations and in the time of emerging antibiotic resistance, targeting bacterial CAs towards a new generation of antibacterial drugs might represent a valid alternative to reinforce the pharmacological arsenal against these pathogens. Thus, scientists heterologously produced in vitro the CA-classes encoded by bacteria as well as other pathogens. Many data are available on kinetic parameters and inhibition profiles of the α-, β- and 𝛾-CAs encoded by pathogenic and non-pathogenic bacteria, while very few data are available on the latest discovered bacterial CA-class, the 𝜄-CA. This CA class was identified for the first time in the marine diatom Thalassiosira pseudonana [23] and, surprisingly, showed to prefer Mn2+ to Zn2+ as metal ion cofactor [23]. Generally, bacterial α- and β-CAs use the Zn2+ ion as catalytic metal, while γ-CAs are Fe2+-dependent enzymes but they are functional enzymes also with bound Zn2+ or Co2+ ions [24,25,26,27].
In a previous work, we have cloned, expressed and purified the first recombinant bacterial ι-CA (acronym BteCAι) identified in the genome of Burkholderia territorii, a Gram-negative bacterium found in soil and water, which is often resistant to common antibiotics [28,29]. The recombinant BteCAι was shown to be a suitable catalyst for the hydration of CO2 to bicarbonate and proton, with a kcat of 3.0 × 105 s−1 and kcat/KM of 3.9 × 107 M−1 s−1 and was also sensitive to inhibition by the sulfonamide acetazolamide [29]. Here, we carried out an extensive study on the inhibition profiles of BteCAι using the substituted benzene-sulfonamides and clinically licensed drugs, which, among the groups of the classical CAIs, generally inhibit other CAs in the nanomolar range and have been clinically used for decades as antiglaucoma [30], diuretic [15], antiepileptic [11], antiobesity [8,31] and anticancer agents [10]. Besides, a comparative analysis of the sulfonamide inhibition profiles was performed comparing BteCAι results with those obtained from CAs belonging to different classes, such as the two human α-CA isoforms (hCA I and hCA II) and the β-and γ-CAs from Escherichia coli. This study gives useful information for designing new antibacterials to disarm the pathogen or bypass their resistance to conventional antimicrobials by inhibiting enzymes involved in the CO2/HCO3 balancing pathway.

2. Results and Discussion

2.1. Production, Validation and Qualitative Activity Assessment of BteCA𝜄

The ι-CA encoded by the genome of B. territorii was heterologously expressed in bacteria with the aim to produce a sufficient amount of protein for the determination of the corresponding inhibition profile with substituted benzene-sulfonamides and clinically licensed drugs. Sodium dodecyl-sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) and protonography were used for the evaluation of the enzyme homogeneity and purity, as well as to verify the enzyme activity of BteCAι. As shown in Figure 1, SDS-PAGE analysis showed a protein migration corresponding to a molecular mass of 19.0 kDa. The BteCAι band was also active when subjected to protonography analysis, as demonstrated by the corresponding yellow color developed by the production of ions (H+) during the CO2 hydration reaction (Figure 1). Moreover, the protein concentration of the recombinant BteCAι was obtained from the SDS-PAGE, analyzing the gel by Image J and using a known concentration of the commercial bovine CA (bCA). Protein quantification by densitometry revealed that the E. coli cells overexpressed BteCAι with a yield of about 0.8 mg/mL from a bacterial culture of 1 L. The protein concentration calculated by densitometry resulted about 10% less with respect to that ascertained by the Bradford method. The BteCAι activity of the purified protein was also measured in solution and expressed in Wilbur-Anderson Units (WAU). BteCAι solution showed a specific activity of 80 ± 8.0 WAU mg−1.

2.2. Comparison of the BteCAι Kinetic Parameters with Those of Other Bacterial CA Classes

The CO2 hydratase activity and the kinetic constants of the purified BteCAι were determined using the stopped-flow technique. The enzyme had a high catalytic activity (kcat 3.0 × 105 s−1) for the physiological reaction of CO2 hydration to bicarbonate and protons. The enzyme was also well inhibited by the sulfonamide acetazolamide (KI = 64.9 nM), a classical CA inhibitor. Table 1 shows a comparison of the BteCAι kinetic behavior (kcat and kcat/KM) with those of other CAs belonging to different classes, such as two α-CA isoforms from Homo sapiens (hCAI and hCAII) and the β- and γ-CAs from E. coli.
In Table 1, it is readily apparent that considered enzymes showed a catalytic constant (kcat, which indicates the maximum rate of the reaction at saturating substrate concentration) in the same order of magnitude, except for the human isoform CA II. Intriguing, the KM value of BteCAι, the substrate concentration at which the reaction rate is half of Vmax, is one order of magnitude lower than that shown by the other two bacterial enzymes belonging to a different class (EcoCAβ and EcoCAγ). KM is a measure of the enzyme’s affinity for its substrate and a lower KM indicates that the enzyme accomplishes its function at a lower substrate concentration. Considering this aspect, the affinity for the substrate (CO2) of BteCAι is higher than that of the two human isoforms (hCA I and hCA II) as well as of the bacterial enzymes (β and γ). In particular, the affinity of BteCAι is 1.3- and 3.0-fold higher than that of hCA I and hCA II, respectively, whereas it is 4.2 to 26-fold higher compared with that of the other two bacterial enzymes (Figure 2).
In a previous work, we carried out a phylogenetic analysis to better evidence the amino acid sequence relationship of bacterial ι-CAs with other microorganism CA-classes (α, β and γ- CAs), demonstrating that the 𝜄-class is closer to the γ-CAs [29]. In the literature, it has been reported that ι-CA from the marine diatom Thalassiosira pseudonana shows a promiscuous esterase activity, which is a feature of the CAs belonging to the α-class [23]. Generally, the enzyme catalytic pocket is small for the γ-CAs and gets more ample for the β-CAs, being quite large in the α-CAs [6]. Based on these observations, it is intriguing to hypothesize that the ι-CAs have an active site spatial organization similar to that of the α-CA, for example, hCA I. However, the BteCAι spatial organization of the catalytic pocket probably allows a better entry of CO2 in the catalytic pocket as well as its tightly binding and thus a higher affinity for this molecule, as demonstrated by the low KM value of BteCAι with respect to that of the human and the other two bacterial enzymes (β and γ-CAs) (Table 1 and Figure 2). Interestingly, although the ι-CA might have a catalytic pocket similar to the human CAs, the classical CA inhibitor, acetazolamide, inhibited BteCAι with a KI of 519 nM, which is a value higher than those determined for other CAs considered in this study. All these findings suggest that even though these enzymes catalyze the same reaction, there are differences in the catalytic pocket, explaining the distinctive KM and the diverse degree of inhibition shown by the various CA classes. Therefore, we are trying to crystallize the BteCAι with the aim to compare its three-dimensional structure with that reported in the literature for other known CAs. This analysis will allow the identification of the ι-CA structural elements responsible for such differences with respect to other CA classes.

2.3. Effects of Simple Aromatic/Heterocyclic Sulfonamide Inhibitors on BteCAι Activity

The first sulfonamide showing a significant antibacterial activity was Prontosil, a sulfanilamide prodrug, which is isosteric/isostructural with p-aminobenzoic acid (PABA), the substrate of dihydropteroate synthase (DHPS) [32,33]. After sulfanilamide was demonstrated to be an effective antibacterial agent, a range of molecular analogs constituting the so-called sulfa drugs entered in clinical use [30]. The presence of primary sulfonamide moieties in sulfanilamide characterizes most investigated CA inhibitors (CAIs) [8,34,35,36,37]. Figure 3 shows some of these sulfonamide inhibitors (simple derivatives 1–24 and clinically used drugs) [3,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52]. Acetazolamide (AAZ), methazolamide (MZA), ethoxzolamide (EZA) and dichlorphenamide (DCP) are systemically acting antiglaucoma CAIs. Dorzolamide (DZA) and brinzolamide (BRZ) are antiglaucoma agents that function topically; benzolamide (BZA) is an orphan drug of this pharmacological class. Zonisamide (ZNS), sulthiame (SLT) and topiramate (TPM) are widely used antiepileptic drugs. Sulpiride (SLP) and indisulam (IND) also belong to this class of pharmacological agents, together with the COX-2 selective inhibitors celecoxib (CLX) and valdecoxib (VLX). Saccharin (SAC) and the diuretic hydrochlorothiazide (HCT) are also known to act as CAIs [11,53,54]. Famotidine (FAM) is a competitive histamine H2-receptor antagonist [53] and epacadostat (EPA) is an inhibitor of the heme-containing enzyme, indoleamine 2,3-dioxygenase-1 (IDO1) but they also act as CAIs [54]. The aromatic/heterocyclic part of the inhibitor interacts with the hydrophilic and hydrophobic residues of the catalytic cavity. Its −SO2NH2 group binds in a tetrahedral geometry to the Zn2+ ion in the deprotonated state, with the nitrogen atom of the sulfonamide moiety coordinated to Zn2+ and an extended network of hydrogen bonds, involving amino acids of the enzyme, also participating in the anchoring of the inhibitor molecule to the metal ion, as shown by X-ray crystallographic data of enzyme-inhibitor adducts [11,53,54].
Numerous pieces of evidence support the involvement of CA activity in the survival, pathogenicity and virulence of several species of human pathogens, such as Helicobacter pylori [55,56,57], Vibrio cholerae [58], Brucella suis [52,59,60,61], Salmonella enterica [62] and Pseudomonas aeruginosa [63]. For example, ethoxzolamide (EZA) was demonstrated to inhibit the V. cholerae virulence in vivo by blocking the cholera toxin gene expression, induced by the bicarbonate produced by the Vibrio CA activity [58]; this inhibitor can also prejudice the virulence of M. tuberculosis [64]. Recently, it has been demonstrated that AAZ and some AAZ-based sulfonamides act as potent inhibitors of vancomycin-resistant Enterococcus [65], which are the leading causes of drug-resistant healthcare-associated infections. All these findings explain the proof of concept that bacterial CAs are promising targets for developing new drugs. Since CAs are valuable targets for compromising the microbial vitality or virulence, the BteCAι inhibition profile with these compounds can be considered an initial information for recognizing efficient and selective inhibitors of bacterial members of the ι-CA class in pathogens, with respect to other protein isoforms present in the host. This information is highly desired for obtaining new pharmacological agents, which may impair the bacterial growth/virulence. Table 2 compares the inhibition profiles of three classes of bacterial CAs and those obtained for the two human isoforms (hCA I and hCA II).
From the data of Table 2, the following results can be observed:
  • Among the sulfonamides and sulfamate used to determine the BteCAι inhibition profile, only three inhibitors resulted in a KI lower than 100 nM. This is the case of the compounds 5, 9 and 19. These results confirm how different is the spatial organization of the catalytic pocket of the different CA classes. For the same compounds, the other two bacterial enzymes showed KI values in the range 160–2840 nM, while the corresponding KI values of two human isoforms were between 33–25,000 nM. The two human isoforms, hCA I and hCA II, resulted in a variegate behavior since hCA I was very susceptible to the inhibitors 9 and 19 with a KI values of 40 and 33 nM, respectively. Again, both human isoforms showed many nanomolar inhibitors with a KI below 100 nM, such as compounds 20, 21, 24, MZA, EZA, BZA, ZNS, IND. Once more, there were compounds such as 3, 7, 10, 11, 12, 13, 14, 15, 16, DCP, DZA, BRZ, SLP, VLX and CLX, which were potent inhibitors of hCA II with KI values in the range 3–94 nM but were very week inhibitors of hCA I (KIS > 5500 nM). Only the study of the three-dimensional structures of BteCAι (not available at this moment) will explain the structural factors responsible for the KI variations.
  • Most of the inhibitors considered in Table 2 were moderate inhibitors of BteCAι with KIs in the range 239–955 nM, such as compounds 1, 2, 3, 4, 6, 7, 8, 10, 11, 12, 13, 14, 15, 16, 17, 20, 21, 22, 23, 24, AAZ, DCP, DZA, BZA, TMP, ZNS, SLP, SLT, IND, VLX, CLX, SLT, HTC, FAM and EPA. A good number of these compounds, such as 1, 2, 3, 14, 20, 21, AAZ, BZA and SLT, resulted in moderate inhibitors for the other two bacterial enzymes (EcoCAβ and EcoCAγ), too. It is important to note that some of these inhibitors were very sensitive versus the human isoform hCA II but not versus the human isoform hCA I (KIS > 10,000 nM). The zonisamide (ZNS), an aliphatic primary sulfonamide, was also a very weak inhibitor for the bacterial enzymes (KIs = 755–3189 nM) but effective towards the human isoenzymes (KIs = 35–56 nM).
  • Some substituted benzene-sulfonamides, such as MZA, EZA, BRZ and SAC, were rather ineffective, weak inhibitors of BteCAι, showing KI values in the range of 5024–8466 nM. Moreover, MZA inhibited the E. coli EcoCAβ and EcoCAγ enzymes with KI of 480 and 921 nM, respectively.
The results reported in Table 2 showed substantial differences in the inhibition constants between the various CA-classes. The inhibition pattern differences must be considered useful for the development of new specific drugs since it means that the spatiality and the amino acids surrounding the catalytic pocket of the CA enzymes are different. This diversity allows the possibility to design efficient and selective inhibitors of the bacterial enzymes without interfering with the activity of the human CAs, even if they show a high percentage of amino acid sequence identity.

3. Materials and Methods

3.1. Chemicals and Instruments

All the chemicals used in this study were of reagent grade and purchased from Sigma. The Affinity column (His-Trap FF) and the AKTA-Prime purification system were bought from GE Healthcare. SDS–PAGE apparatus was procured from Bio-Rad (Hercules, CA, USA).

3.2. Heterologous Expression and Purification of the Recombinant Enzyme

The synthetic B. territorii gene encoding for the BteCAι was cloned, expressed and purified, as described by Del Prete et al. [29]. Briefly, the synthetic gene contained in the expression vector pET100D-Topo/BteCAι was heterologously overexpressed, transforming the competent E. coli BL21 (DE3) codon plus cells (Agilent), using as Isopropyl β-D-1-thiogalactopyranoside (IPTG) as inducer. The produced cytoplasmic protein was purified by using a resin functionalized with Ni2+, which has a high affinity for the polyhistidine-tag (His-Tag) added by genetic engineering to the amino terminus of the neo-synthetized recombinant protein. The protein concentration was determined using the Bradford method (Bio-Rad) [66] and by densitometry using the Gel Analyzer tool of ImageJ [67].

3.3. Enzyme Activity, SDS-PAGE and Protonography

Throughout the purification, the proteolytic activity of BteCAι was achieved as described by Capasso et al. [68]. Twelve % SDS-PAGE performed as described by Laemmli [69] and protonography [70,71,72,73] carried out as reported by Del Prete et al. [29] were used to monitor the apparent molecular mass of the purified recombinant protein on the polyacrylamide gel and the corresponding hydratase activity, respectively.

3.4. Kinetic Parameters and Inhibition Constants

The CO2 hydration activity exerted by BteCAι was monitored using an Applied Photophysics stopped-flow instrument [74]. Phenol red (at a concentration of 0.2 mM) was used as an indicator, working at the absorbance maximum of 557 nm, with 20 mM TRIS (pH 8.3) as a buffer and 20 mM NaClO4 (for maintaining constant the ionic strength), following the initial rates of the CA-catalyzed CO2 hydration reaction for a period of 10–100 s. To determine the kinetic parameters by Lineweaver-Burk plots and the inhibition constants, a concentration of CO2 between 1.7 to 17 mM was used. For each inhibitor, at least three measurements were used to assess the initial velocity at all inhibitor concentrations tested. The uncatalyzed rates were identically determined and detracted from the total observed rates. Stock inhibitor solutions (10 mM) were prepared in distilled-deionized water and dilutions up to 0.1 nM were done with the buffer test. Inhibitor and enzyme solutions were preincubated together for 15 min, at room temperature, before the assay to allow for the formation of the E-I complex or the eventual active site mediated hydrolysis of the inhibitor. The inhibition constants were obtained by non-linear least-squares method using PRISM 6 and the Cheng-Prusoff equation, as reported earlier [75,76,77] and represent the mean from at least three different determinations. hCA I, hCA II and the bacterial enzymes were recombinant proteins obtained in-house. Their concentrations in the assay system were of 5–14 nM.

4. Conclusions

In this context, a broad range of substituted benzene-sulfonamides and clinically licensed drugs were used to determine the inhibition profile of BteCAι and the possible off-targets hCA I and hCA II. Among the sulfonamides and the one sulfamate used as inhibitors, only three of them resulted in having a KI value lower than 100 nM (compounds 5, 9 and 19). All the other inhibitors had KIs > 100 nM. Surprisingly, the results reported showed substantial differences in the inhibition constants between the various CA-classes considered in this study (α, β, γ and ι). For example, for some compounds EcoCAγ showed KIs > 2000 nM, evidencing that this enzyme form or others E. coli froms were less or not inhibited by some of the substituted benzene-sulfonamides and clinically licensed drugs. These differences in the sulfonamide inhibition pattern represent an aspect of the CA inhibition useful for the development of specific and selective drugs versus the bacterial enzymes. It means that the spatiality and the amino acids surrounding the catalytic pocket of the CA enzymes are different. This diversity will allow the possibility to design inhibitors of the bacterial enzymes, which are efficient and selective versus the bacterial enzymes without interfering with the activity of human CAs, even if they show a significant percentage of amino acid sequence identity.

Author Contributions

Investigation, V.D.L., A.P., A.N.; Data curation, A.S., C.C.; Supervision, C.T.S. and C.C.; Writing—original draft, C.C.; Writing—review & editing, A.S., C.T.S. and C.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The data presented in this study are openly available in https://pubmed.ncbi.nlm.nih.gov.

Acknowledgments

We are grateful to Giovanni Del Monaco for technical assistance.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Annunziato, G.; Angeli, A.; D’Alba, F.; Bruno, A.; Pieroni, M.; Vullo, D.; De Luca, V.; Capasso, C.; Supuran, C.T.; Costantino, G. Discovery of New Potential Anti-Infective Compounds Based on Carbonic Anhydrase Inhibitors by Rational Target-Focused Repurposing Approaches. ChemMedChem 2016, 11, 1904–1914. [Google Scholar] [CrossRef] [PubMed]
  2. Guler, O.O.; Capasso, C.; Supuran, C.T. A magnificent enzyme superfamily: Carbonic anhydrases, their purification and characterization. J. Enzym. Inhib. Med. Chem. 2016, 31, 689–694. [Google Scholar] [CrossRef] [PubMed]
  3. Del Prete, S.; Vullo, D.; De Luca, V.; Carginale, V.; Ferraroni, M.; Osman, S.M.; AlOthman, Z.; Supuran, C.T.; Capasso, C. Sulfonamide inhibition studies of the beta-carbonic anhydrase from the pathogenic bacterium Vibrio cholerae. Bioorg. Med. Chem. 2016, 24, 1115–1120. [Google Scholar] [CrossRef] [PubMed]
  4. Del Prete, S.; De Luca, V.; De Simone, G.; Supuran, C.T.; Capasso, C. Cloning, expression and purification of the complete domain of the eta-carbonic anhydrase from Plasmodium falciparum. J. Enzym. Inhib. Med. Chem. 2016, 31, 54–59. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Capasso, C.; Supuran, C.T. An Overview of the Carbonic Anhydrases from Two Pathogens of the Oral Cavity: Streptococcus mutans and Porphyromonas gingivalis. Curr. Top Med. Chem. 2016, 16, 2359–2368. [Google Scholar] [CrossRef]
  6. Capasso, C.; Supuran, C.T. An overview of the alpha-, beta- and gamma-carbonic anhydrases from Bacteria: Can bacterial carbonic anhydrases shed new light on evolution of bacteria? J. Enzym. Inhib. Med. Chem. 2015, 30, 325–332. [Google Scholar] [CrossRef] [Green Version]
  7. Supuran, C.T.; Capasso, C. An Overview of the Bacterial Carbonic Anhydrases. Metabolites 2017, 7, 56. [Google Scholar] [CrossRef] [Green Version]
  8. Supuran, C.T. Structure and function of carbonic anhydrases. BioChem. J. 2016, 473, 2023–2032. [Google Scholar] [CrossRef]
  9. McKenna, R.; Supuran, C.T. Carbonic anhydrase inhibitors drug design. Subcell BioChem. 2014, 75, 291–323. [Google Scholar]
  10. Neri, D.; Supuran, C.T. Interfering with pH regulation in tumours as a therapeutic strategy. Nat. Rev. Drug Discov. 2011, 10, 767–777. [Google Scholar] [CrossRef] [Green Version]
  11. Supuran, C.T. Carbonic anhydrases: Novel therapeutic applications for inhibitors and activators. Nat. Rev. Drug Discov. 2008, 7, 168–181. [Google Scholar] [CrossRef]
  12. Supuran, C.T. Carbonic anhydrases—An overview. Curr. Pharm. Des. 2008, 14, 603–614. [Google Scholar] [CrossRef] [PubMed]
  13. Boron, W.F. Evaluating the role of carbonic anhydrases in the transport of HCO3-related species. Biochim Biophys Acta 2010, 1804, 410–421. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Supuran, C.T.; Capasso, C. Biomedical applications of prokaryotic carbonic anhydrases. Expert Opin Ther. Patents 2018, 28, 745–754. [Google Scholar] [CrossRef] [PubMed]
  15. Supuran, C.T.; Capasso, C. Carbonic Anhydrase from Porphyromonas Gingivalis as a Drug Target. Pathogens 2017, 6, 30. [Google Scholar] [CrossRef] [Green Version]
  16. Capasso, C.; Supuran, C.T. An Overview of the Selectivity and Efficiency of the Bacterial Carbonic Anhydrase Inhibitors. Curr. Med. Chem. 2015, 22, 2130–2139. [Google Scholar] [CrossRef] [PubMed]
  17. Rogato, A.; Del Prete, S.; Nocentini, A.; Carginale, V.; Supuran, C.T.; Capasso, C. Phaeodactylum tricornutum as a model organism for testing the membrane penetrability of sulphonamide carbonic anhydrase inhibitors. J. Enzym. Inhib. Med. Chem. 2019, 34, 510–518. [Google Scholar] [CrossRef] [Green Version]
  18. Angeli, A.; Pinteala, M.; Maier, S.S.; Del Prete, S.; Capasso, C.; Simionescu, B.C.; Supuran, C.T. Inhibition of alpha-, beta-, gamma-, delta-, zeta- and eta-class carbonic anhydrases from bacteria, fungi, algae, diatoms and protozoans with famotidine. J. Enzym. Inhib. Med. Chem. 2019, 34, 644–650. [Google Scholar] [CrossRef] [Green Version]
  19. Berrino, E.; Bozdag, M.; Del Prete, S.; Alasmary, F.A.S.; Alqahtani, L.S.; AlOthman, Z.; Capasso, C.; Supuran, C.T. Inhibition of alpha-, beta-, gamma-, and delta-carbonic anhydrases from bacteria and diatoms with N′-aryl-N-hydroxy-ureas. J. Enzym. Inhib. Med. Chem. 2018, 33, 1194–1198. [Google Scholar] [CrossRef] [Green Version]
  20. Angeli, A.; Del Prete, S.; Alasmary, F.A.S.; Alqahtani, L.S.; AlOthman, Z.; Donald, W.A.; Capasso, C.; Supuran, C.T. The first activation studies of the eta-carbonic anhydrase from the malaria parasite Plasmodium falciparum with amines and amino acids. Bioorg. Chem. 2018, 80, 94–98. [Google Scholar] [CrossRef]
  21. Payne, D.J.; Gwynn, M.N.; Holmes, D.J.; Pompliano, D.L. Drugs for bad bugs: Confronting the challenges of antibacterial discovery. Nat. Rev. Drug Discov. 2007, 6, 29–40. [Google Scholar] [CrossRef] [PubMed]
  22. Supuran, C.T.; Capasso, C. New light on bacterial carbonic anhydrases phylogeny based on the analysis of signal peptide sequences. J. Enzym. Inhib. Med. Chem. 2016, 31, 1254–1260. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Jensen, E.L.; Clement, R.; Kosta, A.; Maberly, S.C.; Gontero, B. A new widespread subclass of carbonic anhydrase in marine phytoplankton. ISME J. 2019, 13, 2094–2106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. De Simone, G.; Monti, S.M.; Alterio, V.; Buonanno, M.; De Luca, V.; Rossi, M.; Carginale, V.; Supuran, C.T.; Capasso, C.; Di Fiore, A. Crystal structure of the most catalytically effective carbonic anhydrase enzyme known, SazCA from the thermophilic bacterium Sulfurihydrogenibium azorense. Bioorg. Med. Chem. Lett. 2015, 25, 2002–2006. [Google Scholar] [CrossRef] [PubMed]
  25. Di Fiore, A.; Capasso, C.; De Luca, V.; Monti, S.M.; Carginale, V.; Supuran, C.T.; Scozzafava, A.; Pedone, C.; Rossi, M.; De Simone, G. X-ray structure of the first ‘extremo-alpha-carbonic anhydrase’, a dimeric enzyme from the thermophilic bacterium Sulfurihydrogenibium yellowstonense YO3AOP1. Acta Cryst. D Biol. Cryst. 2013, 69, 1150–1159. [Google Scholar] [CrossRef] [PubMed]
  26. Lomelino, C.L.; Mahon, B.P.; McKenna, R.; Carta, F.; Supuran, C.T. Kinetic and X-ray crystallographic investigations on carbonic anhydrase isoforms I, II, IX and XII of a thioureido analog of SLC-0111. Bioorg. Med. Chem. 2016, 24, 976–981. [Google Scholar] [CrossRef]
  27. Ferraroni, M.; Del Prete, S.; Vullo, D.; Capasso, C.; Supuran, C.T. Crystal structure and kinetic studies of a tetrameric type II beta-carbonic anhydrase from the pathogenic bacterium Vibrio cholerae. Acta Cryst. D Biol. Cryst. 2015, 71, 2449–2456. [Google Scholar] [CrossRef] [PubMed]
  28. De Smet, B.; Mayo, M.; Peeters, C.; Zlosnik, J.E.; Spilker, T.; Hird, T.J.; LiPuma, J.J.; Kidd, T.J.; Kaestli, M.; Ginther, J.L.; et al. Burkholderia stagnalis sp. nov. and Burkholderia territorii sp. nov., two novel Burkholderia cepacia complex species from environmental and human sources. Int. J. Syst. Evol. Microbiol. 2015, 65, 2265–2271. [Google Scholar] [CrossRef]
  29. Del Prete, S.; Nocentini, A.; Supuran, C.T.; Capasso, C. Bacterial iota-carbonic anhydrase: A new active class of carbonic anhydrase identified in the genome of the Gram-negative bacterium Burkholderia territorii. J. Enzym. Inhib. Med. Chem. 2020, 35, 1060–1068. [Google Scholar] [CrossRef] [Green Version]
  30. Capasso, C.; Supuran, C.T. Sulfa and trimethoprim-like drugs—Antimetabolites acting as carbonic anhydrase, dihydropteroate synthase and dihydrofolate reductase inhibitors. J. Enzym. Inhib. Med. Chem. 2014, 29, 379–387. [Google Scholar] [CrossRef]
  31. Supuran, C.T. Structure-based drug discovery of carbonic anhydrase inhibitors. J. Enzym. Inhib. Med. Chem. 2012, 27, 759–772. [Google Scholar] [CrossRef] [PubMed]
  32. Otten, H. Domagk and the development of the sulphonamides. J. Antimicrob. Chemother. 1986, 17, 689–696. [Google Scholar] [CrossRef] [PubMed]
  33. Achari, A.; Somers, D.O.; Champness, J.N.; Bryant, P.K.; Rosemond, J.; Stammers, D.K. Crystal structure of the anti-bacterial sulfonamide drug target dihydropteroate synthase. Nat. Struct. Biol. 1997, 4, 490–497. [Google Scholar] [CrossRef] [PubMed]
  34. Supuran, C.T. Special Issue: Sulfonamides. Molecules 2017, 22, 1642. [Google Scholar]
  35. Supuran, C.T. Carbonic anhydrase inhibition and the management of neuropathic pain. Expert Rev. Neurother. 2016, 16, 961–968. [Google Scholar] [CrossRef]
  36. Supuran, C.T. Drug interaction considerations in the therapeutic use of carbonic anhydrase inhibitors. Expert Opin. Drug Metab. Toxicol. 2016, 12, 423–431. [Google Scholar] [CrossRef]
  37. Supuran, C.T. Advances in structure-based drug discovery of carbonic anhydrase inhibitors. Expert Opin. Drug Discov. 2017, 12, 61–88. [Google Scholar] [CrossRef]
  38. Vullo, D.; Del Prete, S.; Fisher, G.M.; Andrews, K.T.; Poulsen, S.A.; Capasso, C.; Supuran, C.T. Sulfonamide inhibition studies of the eta-class carbonic anhydrase from the malaria pathogen Plasmodium falciparum. Bioorg. Med. Chem. 2015, 23, 526–531. [Google Scholar]
  39. Vullo, D.; De Luca, V.; Del Prete, S.; Carginale, V.; Scozzafava, A.; Capasso, C.; Supuran, C.T. Sulfonamide inhibition studies of the gamma-carbonic anhydrase from the Antarctic bacterium Pseudoalteromonas haloplanktis. Bioorg. Med. Chem. Lett. 2015, 25, 3550–3555. [Google Scholar] [CrossRef] [Green Version]
  40. Vullo, D.; De Luca, V.; Del Prete, S.; Carginale, V.; Scozzafava, A.; Capasso, C.; Supuran, C.T. Sulfonamide inhibition studies of the gamma-carbonic anhydrase from the Antarctic cyanobacterium Nostoc commune. Bioorg. Med. Chem. 2015, 23, 1728–1734. [Google Scholar] [CrossRef]
  41. Dedeoglu, N.; DeLuca, V.; Isik, S.; Yildirim, H.; Kockar, F.; Capasso, C.; Supuran, C.T. Sulfonamide inhibition study of the beta-class carbonic anhydrase from the caries producing pathogen Streptococcus mutans. Bioorg. Med. Chem. Lett. 2015, 25, 2291–2297. [Google Scholar] [CrossRef]
  42. Alafeefy, A.M.; Ceruso, M.; Al-Tamimi, A.M.; Del Prete, S.; Supuran, C.T.; Capasso, C. Inhibition studies of quinazoline-sulfonamide derivatives against the gamma-CA (PgiCA) from the pathogenic bacterium, Porphyromonas gingivalis. J. Enzym. Inhib. Med. Chem. 2015, 30, 592–596. [Google Scholar] [CrossRef]
  43. Alafeefy, A.M.; Abdel-Aziz, H.A.; Vullo, D.; Al-Tamimi, A.M.; Awaad, A.S.; Mohamed, M.A.; Capasso, C.; Supuran, C.T. Inhibition of human carbonic anhydrase isozymes I, II, IX and XII with a new series of sulfonamides incorporating aroylhydrazone-, [1,2,4]triazolo[3,4-b][1,3,4]thiadiazinyl- or 2-(cyanophenylmethylene)-1,3,4-thiadiazol-3(2H)-yl moieties. J. Enzym. Inhib. Med. Chem. 2015, 30, 52–56. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Diaz, J.R.; Fernandez Baldo, M.; Echeverria, G.; Baldoni, H.; Vullo, D.; Soria, D.B.; Supuran, C.T.; Cami, G.E. A substituted sulfonamide and its Co (II), Cu (II), and Zn (II) complexes as potential antifungal agents. J. Enzym. Inhib. Med. Chem. 2016, 31, 51–62. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Del Prete, S.; Vullo, D.; De Luca, V.; Carginale, V.; Osman, S.M.; AlOthman, Z.; Supuran, C.T.; Capasso, C. Comparison of the sulfonamide inhibition profiles of the alpha-, beta- and gamma-carbonic anhydrases from the pathogenic bacterium Vibrio cholerae. Bioorg. Med. Chem. Lett. 2016, 26, 1941–1946. [Google Scholar] [CrossRef] [PubMed]
  46. Del Prete, S.; Vullo, D.; De Luca, V.; Carginale, V.; Osman, S.M.; AlOthman, Z.; Supuran, C.T.; Capasso, C. Cloning, expression, purification and sulfonamide inhibition profile of the complete domain of the eta-carbonic anhydrase from Plasmodium falciparum. Bioorg. Med. Chem. Lett. 2016, 26, 4184–4190. [Google Scholar] [CrossRef]
  47. Gawad, N.M.A.; Amin, N.H.; Elsaadi, M.T.; Mohamed, F.M.; Angeli, A.; De Luca, V.; Capasso, C.; Supuran, C.T. Synthesis of 4-(thiazol-2-ylamino)-benzenesulfonamides with carbonic anhydrase I, II and IX inhibitory activity and cytotoxic effects against breast cancer cell lines. Bioorg. Med. Chem. 2016, 24, 3043–3051. [Google Scholar] [CrossRef]
  48. Supuran, C.T. Legionella pneumophila Carbonic Anhydrases: Underexplored Antibacterial Drug Targets. Pathogens 2016, 5, 44. [Google Scholar] [CrossRef] [Green Version]
  49. Nishimori, I.; Vullo, D.; Minakuchi, T.; Scozzafava, A.; Capasso, C.; Supuran, C.T. Sulfonamide inhibition studies of two beta-carbonic anhydrases from the bacterial pathogen Legionella pneumophila. Bioorg. Med. Chem. 2014, 22, 2939–2946. [Google Scholar] [CrossRef]
  50. Vullo, D.; Sai Kumar, R.S.; Scozzafava, A.; Capasso, C.; Ferry, J.G.; Supuran, C.T. Anion inhibition studies of a beta-carbonic anhydrase from Clostridium perfringens. Bioorg. Med. Chem. Lett. 2013, 23, 6706–6710. [Google Scholar] [CrossRef]
  51. Nishimori, I.; Minakuchi, T.; Maresca, A.; Carta, F.; Scozzafava, A.; Supuran, C.T. The beta-carbonic anhydrases from Mycobacterium tuberculosis as drug targets. Curr. Pharm. Des. 2010, 16, 3300–3309. [Google Scholar] [CrossRef] [PubMed]
  52. Carta, F.; Maresca, A.; Covarrubias, A.S.; Mowbray, S.L.; Jones, T.A.; Supuran, C.T. Carbonic anhydrase inhibitors. Characterization and inhibition studies of the most active beta-carbonic anhydrase from Mycobacterium tuberculosis, Rv3588c. Bioorg. Med. Chem. Lett. 2009, 19, 6649–6654. [Google Scholar] [CrossRef] [PubMed]
  53. Nguyen, K.; Ahlawat, R. Famotidine; StatPearls: Treasure Island, FL, USA, 2020. [Google Scholar]
  54. Komiya, T.; Huang, C.H. Updates in the Clinical Development of Epacadostat and Other Indoleamine 2,3-Dioxygenase 1 Inhibitors (IDO1) for Human Cancers. Front. Oncol. 2018, 8, 423. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Modak, J.K.; Tikhomirova, A.; Gorrell, R.J.; Rahman, M.M.; Kotsanas, D.; Korman, T.M.; Garcia-Bustos, J.; Kwok, T.; Ferrero, R.L.; Supuran, C.T.; et al. Anti-Helicobacter pylori activity of ethoxzolamide. J. Enzym. Inhib. Med. Chem. 2019, 34, 1660–1667. [Google Scholar] [CrossRef] [Green Version]
  56. Ronci, M.; Del Prete, S.; Puca, V.; Carradori, S.; Carginale, V.; Muraro, R.; Mincione, G.; Aceto, A.; Sisto, F.; Supuran, C.T.; et al. Identification and characterization of the alpha-CA in the outer membrane vesicles produced by Helicobacter pylori. J. Enzym. Inhib. Med. Chem. 2019, 34, 189–195. [Google Scholar] [CrossRef] [Green Version]
  57. Buzas, G.M. Helicobacter pylori—2010. Orv. Hetil. 2010, 151, 2003–2010. [Google Scholar] [CrossRef]
  58. Abuaita, B.H.; Withey, J.H. Bicarbonate Induces Vibrio cholerae virulence gene expression by enhancing ToxT activity. Infect. Immun. 2009, 77, 4111–4120. [Google Scholar] [CrossRef] [Green Version]
  59. Kohler, S.; Ouahrani-Bettache, S.; Winum, J.Y. Brucella suis carbonic anhydrases and their inhibitors: Towards alternative antibiotics? J. Enzym. Inhib. Med. Chem. 2017, 32, 683–687. [Google Scholar] [CrossRef] [Green Version]
  60. Singh, S.; Supuran, C.T. 3D-QSAR CoMFA studies on sulfonamide inhibitors of the Rv3588c beta-carbonic anhydrase from Mycobacterium tuberculosis and design of not yet synthesized new molecules. J. Enzym. Inhib. Med. Chem. 2014, 29, 449–455. [Google Scholar] [CrossRef] [Green Version]
  61. Ceruso, M.; Vullo, D.; Scozzafava, A.; Supuran, C.T. Sulfonamides incorporating fluorine and 1,3,5-triazine moieties are effective inhibitors of three beta-class carbonic anhydrases from Mycobacterium tuberculosis. J. Enzym. Inhib. Med. Chem. 2014, 29, 686–689. [Google Scholar] [CrossRef]
  62. Rollenhagen, C.; Bumann, D. Salmonella enterica highly expressed genes are disease specific. Infect. Immun. 2006, 74, 1649–1660. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Lotlikar, S.R.; Kayastha, B.B.; Vullo, D.; Khanam, S.S.; Reygan, E.B.; Murray, A.B.; McKenna, R.; Supuran, C.T.; Patrauchan, M.A. Pseudomonas aeruginosa β-carbonic anhydrase, psCA1, is required for calcium deposition and contributes to virulence. Cell Calcium. 2019, 84, 102080. [Google Scholar] [CrossRef] [PubMed]
  64. Aspatwar, A.; Kairys, V.; Rala, S.; Parikka, M.; Bozdag, M.; Carta, F.; Supuran, C.T.; Parkkila, S. Mycobacterium tuberculosis beta-Carbonic Anhydrases: Novel Targets for Developing Antituberculosis Drugs. Int. J. Mol. Sci. 2019, 20, 5153. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Kaur, J.; Cao, X.; Abutaleb, N.S.; Elkashif, A.; Graboski, A.L.; Krabill, A.D.; AbdelKhalek, A.H.; An, W.; Bhardwaj, A.; Seleem, M.N.; et al. Optimization of Acetazolamide-Based Scaffold as Potent Inhibitors of Vancomycin-Resistant Enterococcus. J. Med. Chem. 2020, 63, 9540–9562. [Google Scholar] [CrossRef] [PubMed]
  66. Bradford, M.M. A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye binding. Anal. BioChem. 1976, 72, 248–254. [Google Scholar] [CrossRef]
  67. Villela, S.M.A.; Kraiem, H.; Bouhaouala-Zahar, B.; Bideaux, C.; Aceves Lara, C.A.; Fillaudeau, L. A protocol for recombinant protein quantification by densitometry. Microbiologyopen 2020, 9, 1175–1182. [Google Scholar] [CrossRef] [Green Version]
  68. Capasso, C.; De Luca, V.; Carginale, V.; Cannio, R.; Rossi, M. Biochemical properties of a novel and highly thermostable bacterial alpha-carbonic anhydrase from Sulfurihydrogenibium yellowstonense YO3AOP1. J. Enzym. Inhib. Med. Chem. 2012, 27, 892–897. [Google Scholar] [CrossRef] [Green Version]
  69. Laemmli, U.K. Cleavage of structural proteins during the assembly of the head of bacteriophage T4. Nature 1970, 227, 680–685. [Google Scholar] [CrossRef]
  70. De Luca, V.; Del Prete, S.; Supuran, C.T.; Capasso, C. Protonography, a new technique for the analysis of carbonic anhydrase activity. J. Enzym. Inhib. Med. Chem. 2015, 30, 277–282. [Google Scholar] [CrossRef]
  71. Del Prete, S.; De Luca, V.; Iandolo, E.; Supuran, C.T.; Capasso, C. Protonography, a powerful tool for analyzing the activity and the oligomeric state of the gamma-carbonic anhydrase identified in the genome of Porphyromonas gingivalis. Bioorg. Med. Chem. 2015, 23, 3747–3750. [Google Scholar] [CrossRef]
  72. Del Prete, S.; De Luca, V.; Supuran, C.T.; Capasso, C. Protonography, a technique applicable for the analysis of eta-carbonic anhydrase activity. J. Enzym. Inhib. Med. Chem. 2015, 30, 920–924. [Google Scholar] [CrossRef] [PubMed]
  73. Del Prete, S.; Vullo, D.; Caminiti-Segonds, N.; Zoccola, D.; Tambutte, S.; Supuran, C.T.; Capasso, C. Protonography and anion inhibition profile of the alpha-carbonic anhydrase (CruCA4) identified in the Mediterranean red coral Corallium rubrum. Bioorg. Chem. 2018, 76, 281–287. [Google Scholar] [CrossRef] [PubMed]
  74. 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] [CrossRef]
  75. Del Prete, S.; Vullo, D.; De Luca, V.; Carginale, V.; di Fonzo, P.; Osman, S.M.; AlOthman, Z.; Supuran, C.T.; Capasso, C. Anion inhibition profiles of alpha-, beta- and gamma-carbonic anhydrases from the pathogenic bacterium Vibrio cholerae. Bioorg. Med. Chem. 2016, 24, 3413–3417. [Google Scholar] [CrossRef] [PubMed]
  76. Del Prete, S.; Vullo, D.; De Luca, V.; Carginale, V.; di Fonzo, P.; Osman, S.M.; AlOthman, Z.; Supuran, C.T.; Capasso, C. Anion inhibition profiles of the complete domain of the eta-carbonic anhydrase from Plasmodium falciparum. Bioorg. Med. Chem. 2016, 24, 4410–4414. [Google Scholar] [CrossRef]
  77. De Luca, V.; Vullo, D.; Del Prete, S.; Carginale, V.; Osman, S.M.; AlOthman, Z.; Supuran, C.T.; Capasso, C. Cloning, characterization and anion inhibition studies of a gamma-carbonic anhydrase from the Antarctic bacterium Colwellia psychrerythraea. Bioorg. Med. Chem. 2016, 24, 835–840. [Google Scholar] [CrossRef]
Figure 1. Combined lanes of sodium dodecyl-sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) and protonography of BteCAι. Purified recombinant BteCAι (lane 4) was subjected to protonographic analysis (lane 2) to determine the enzyme activity on the polyacrylamide gel. Lane 3, molecular markers, from the top: 50.0 kDa, 37.0 kDa, 25 kDa and 20 kDa. Lane 1 and 5, commercial bovine carbonic anhydrase (bCA) used as control in SDS-PAGE (lane 5) and protonography (lane 1). Boxes with dashed and continuous lines indicate the bands identifying the bCA and BteCAι, respectively.
Figure 1. Combined lanes of sodium dodecyl-sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) and protonography of BteCAι. Purified recombinant BteCAι (lane 4) was subjected to protonographic analysis (lane 2) to determine the enzyme activity on the polyacrylamide gel. Lane 3, molecular markers, from the top: 50.0 kDa, 37.0 kDa, 25 kDa and 20 kDa. Lane 1 and 5, commercial bovine carbonic anhydrase (bCA) used as control in SDS-PAGE (lane 5) and protonography (lane 1). Boxes with dashed and continuous lines indicate the bands identifying the bCA and BteCAι, respectively.
Ijms 22 00571 g001
Figure 2. Graphical representation of the enzyme activity. The enzyme affinity for the CO2 was reported as 1/KM to evidence that more remarkable is the bar height, more potent is the affinity of the enzyme for the substrate.
Figure 2. Graphical representation of the enzyme activity. The enzyme affinity for the CO2 was reported as 1/KM to evidence that more remarkable is the bar height, more potent is the affinity of the enzyme for the substrate.
Ijms 22 00571 g002
Figure 3. Sulfonamides and their structurally related derivatives, such as sulfamates and sulfamides, which have the general formula A-SO2NH2 (where A can be an aromatic, heterocyclic, aliphatic or sugar scaffold) and act as CAIs: simple aromatic/heterocyclic derivatives 1–24 (left); clinically used drugs or agents in clinical development (right).
Figure 3. Sulfonamides and their structurally related derivatives, such as sulfamates and sulfamides, which have the general formula A-SO2NH2 (where A can be an aromatic, heterocyclic, aliphatic or sugar scaffold) and act as CAIs: simple aromatic/heterocyclic derivatives 1–24 (left); clinically used drugs or agents in clinical development (right).
Ijms 22 00571 g003
Table 1. BteCAι kinetic parameters for the catalyzed CO2 hydration reaction and their comparison with those determined for different CA classes (α, β and γ). The kinetic measurements were carried out at 20 °C and pH 7.5 in 10 mM HEPES buffer for the hCA I, hCA II, EcoCA𝛾 and BteCAι, while a different buffer was used for the EcoCAβ enzyme (10 mM TRIS, pH 8.3, containing 20 mM NaClO4).
Table 1. BteCAι kinetic parameters for the catalyzed CO2 hydration reaction and their comparison with those determined for different CA classes (α, β and γ). The kinetic measurements were carried out at 20 °C and pH 7.5 in 10 mM HEPES buffer for the hCA I, hCA II, EcoCA𝛾 and BteCAι, while a different buffer was used for the EcoCAβ enzyme (10 mM TRIS, pH 8.3, containing 20 mM NaClO4).
OrganismEnzyme AcronymClasskcat
(s−1)
KM
(M)
kcat/KM
(M−1·s−1)
KI (Acetazolamide) (nM)
Homo sapienshCA Iα2.0 × 1054.0 × 10−35.0 × 107250
hCA IIα1.4 × 1069.3 × 10−31.5 × 10812
Escherichia coliEcoCAβ (CynT2)β5.3 × 1051.3 × 10−2 4.1 × 107227
EcoCA𝛾𝛾5.7 × 1058.2 × 10−26.9 × 106248
Burkholderia territoriiBteCAι𝜄3.0 × 1053.1 × 10−39.7 × 107519
Mean from 3 different assays performed by a stopped flow technique (errors were in the range of ±5–10% of the reported values).
Table 2. BteCAι inhibition parameters for the catalyzed CO2 hydration reaction and their comparison with those determined for different CA classes (α, β and γ). The inhibition measurements were carried out at 20 °C and pH 7.5 in 10 mM HEPES buffer for the hCA I, hCA II, EcoCAγ and BteCAι, while a different buffer was used for the EcoCAβ enzyme (20 mM TRIS, pH 8.3, containing 20 mM NaClO4).
Table 2. BteCAι inhibition parameters for the catalyzed CO2 hydration reaction and their comparison with those determined for different CA classes (α, β and γ). The inhibition measurements were carried out at 20 °C and pH 7.5 in 10 mM HEPES buffer for the hCA I, hCA II, EcoCAγ and BteCAι, while a different buffer was used for the EcoCAβ enzyme (20 mM TRIS, pH 8.3, containing 20 mM NaClO4).
KI (nM) *
InhibitorhCA IhCA IIEcoCAβEcoCAγBteCAι
128,000300705314325
225,000240790193477
379,0008457246568
478,5003203015221446
525,000170284016097
621,0001603321622786
7830060>10,000605481
89800110>10,000671346
9650040271271896
1073005485612577357
1158006362461779239
1284007543851953329
138600604122197303
14930019440712434
1555008064451013540
1695009423404238594
1721,0001255021975404
18164462052064467
1910933416189493
2062726883268
216911473819307
2216446933501365
23109333224045408
249530824262698
AAZ25012227248519
MZA50144809218466
EZA25855755385024
DCP120038>10,000889825
DZA50,00096292007794
BRZ45,0003204848423703
BZA15927694724
TPM250103359648787
ZNS56353189755806
SLP12004097914958
IND31152392387638
VLX54,000432752891934
CLX50,000211894944960
SLT3749285446954
SAC18,5405959669349037081
HCT32829050103643780
FAM922582769274943
EPA82629172560744955
* Mean from 3 different assays, by a stopped flow technique (errors were in the range of ± 5–10% of the reported values).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

De Luca, V.; Petreni, A.; Nocentini, A.; Scaloni, A.; Supuran, C.T.; Capasso, C. Effect of Sulfonamides and Their Structurally Related Derivatives on the Activity of ι-Carbonic Anhydrase from Burkholderia territorii. Int. J. Mol. Sci. 2021, 22, 571. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22020571

AMA Style

De Luca V, Petreni A, Nocentini A, Scaloni A, Supuran CT, Capasso C. Effect of Sulfonamides and Their Structurally Related Derivatives on the Activity of ι-Carbonic Anhydrase from Burkholderia territorii. International Journal of Molecular Sciences. 2021; 22(2):571. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22020571

Chicago/Turabian Style

De Luca, Viviana, Andrea Petreni, Alessio Nocentini, Andrea Scaloni, Claudiu T. Supuran, and Clemente Capasso. 2021. "Effect of Sulfonamides and Their Structurally Related Derivatives on the Activity of ι-Carbonic Anhydrase from Burkholderia territorii" International Journal of Molecular Sciences 22, no. 2: 571. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22020571

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop