Next Article in Journal
Lipid Droplet Formation Is Regulated by Ser/Thr Phosphatase PPM1D via Dephosphorylation of Perilipin 1
Next Article in Special Issue
Synthesis, Biological and In Silico Studies of Griseofulvin and Usnic Acid Sulfonamide Derivatives as Fungal, Bacterial and Human Carbonic Anhydrase Inhibitors
Previous Article in Journal
RAI14 Promotes Melanoma Progression by Regulating the FBXO32/c-MYC Pathway
Previous Article in Special Issue
Novel Insights on Human Carbonic Anhydrase Inhibitors Based on Coumalic Acid: Design, Synthesis, Molecular Modeling Investigation, and Biological Studies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

α-CAs from Photosynthetic Organisms

Istituto di Biostrutture e Bioimmagini, CNR, Via Pietro Castellino 111, 80131 Napoli, Italy
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2022, 23(19), 12045; https://doi.org/10.3390/ijms231912045
Submission received: 26 July 2022 / Revised: 19 September 2022 / Accepted: 26 September 2022 / Published: 10 October 2022
(This article belongs to the Special Issue Carbonic Anhydrase and Carbonic Anhydrase Inhibitors)

Abstract

:
Carbonic anhydrases (CAs) are ubiquitous enzymes that catalyze the reversible carbon dioxide hydration reaction. Among the eight different CA classes existing in nature, the α-class is the largest one being present in animals, bacteria, protozoa, fungi, and photosynthetic organisms. Although many studies have been reported on these enzymes, few functional, biochemical, and structural data are currently available on α-CAs isolated from photosynthetic organisms. Here, we give an overview of the most recent literature on the topic. In higher plants, these enzymes are engaged in both supplying CO2 at the Rubisco and determining proton concentration in PSII membranes, while in algae and cyanobacteria they are involved in carbon-concentrating mechanism (CCM), photosynthetic reactions and in detecting or signaling changes in the CO2 level in the environment. Crystal structures are only available for three algal α-CAs, thus not allowing to associate specific structural features to cellular localizations or physiological roles. Therefore, further studies on α-CAs from photosynthetic organisms are strongly needed to provide insights into their structure–function relationship.

Graphical Abstract

1. Introduction

Carbonic anhydrases (CAs) are ubiquitous enzymes that catalyze both the hydration of carbon dioxide and bicarbonate dehydration, as schematically reported in the following reaction:
CO 2 + H 2 O   HCO 3 + H +
Their catalytic action allows the CO2 hydration reaction to speed up to 1,000,000 times per second (Kcat is within the range 104–106 s−1) compared to the uncatalyzed reaction [1,2,3,4]. Since the CA active site contains a metal ion, these enzymes are referred to as metalloenzymes. Most of the time the metal ion is a Zn2+ ion, but some CA classes also utilize Fe2+, Mn2+, Cd2+ and Co2+ ions [5].
The known CAs are grouped into eight distinct families, namely α, β, γ, δ, ζ, η, θ, and ι, which are phylogenetically unrelated and possess little to no sequence or structural similarity [6,7]. In particular, α-class is mainly present in animals and higher plants, algae, protozoa, fungi and bacteria [8,9,10]; members of β-class are found in plants, algae, cyanobacteria and also in non-photosynthetic organisms [6,9,11,12,13,14], but are absent in animals; γ-CAs have been found in archaea, some bacteria, and plants [15], whereas δ- and ζ-CAs only in diatoms and coccoliths [7,16,17,18,19,20,21]. Finally, η-, θ- and ι-CAs are the most recently discovered classes, η-CAs being found in Plasmodium species [22], θ-CAs in chlorophytes and diatoms [23] and ι-CAs in diatoms, algae, bacteria, and archaea [24,25].
The α-class is the most populous among the different CA families. These enzymes play a key physiologic role in all living organisms where they have been found. For example, in humans, these enzymes are involved in pH and CO2 homeostasis, respiration and transport of CO2/bicarbonate, electrolyte secretion in many tissues/organs, biosynthetic reactions, bone resorption, calcification, and tumorigenicity [26,27,28,29,30]. In protozoa, fungi, and bacteria they ensure the bicarbonate necessary for the metabolism or are involved in pH regulation [26], whereas in photosynthetic organisms such as higher plants, algae and cyanobacteria, these enzymes are mainly involved in photosynthesis by participating in photosynthetic reactions [6,8,31] and/or in the carbon-concentrating mechanism (CCM) [32,33,34,35], which is an adaptive strategy for carbon acquisition, to survive when the CO2 concentration limits photosynthesis [32]. Indeed, the reaction catalyzed by CAs allows to achieve a high concentration of CO2 near Rubisco (Ribulose Bisphosphate Carboxylase/Oxygenase), the main carboxylation enzyme in photosynthetic carbon fixation, even under limiting external inorganic carbon (Ci) [36]. Consequently, the activity of Rubisco increases, enhancing the rate of carboxylation and suppressing photorespiration events [37,38,39]. The role of cyanobacterial α-CAs as sensors to detect changes in the environment’s CO2 level has also been proposed [40,41,42].
α-CAs have been widely studied and a huge number of original papers and review articles have been published on these enzymes; however, most of these studies have been dedicated to human and bacterial enzymes and focused on drug design approaches [43,44]. In this review, we fill the gap existing in the α-CAs present in photosynthetic organisms, summarizing their physiological roles, cellular locations, and biochemical and structural features.

2. Higher Plants

Higher plants generally contain multiple gene copies of α-, β- and γ-CAs within all tissues [45]. Among these, the α-CA gene family is widely present, being detected in both cytosol and chloroplast, even though so far only a few corresponding proteins have been comprehensively characterized [46,47,48]. The largest number of studies have been carried out on Arabidopsis thaliana, as this plant is widely used as a model organism. Within its genome eight genes encoding α-CAs have been revealed; among these, four have been functionally investigated and one underwent a partial biochemical characterization [49]. Clear information on the intracellular location of these α-CA isoforms is available only for α-CA1, which is located in chloroplast stroma [46], and α-CA4 and α-CA5 found in thylakoid membranes [50,51,52], whereas recent studies provided evidence of the presence of α-CA2 in thylakoid membranes too [53,54,55].
Regarding their physiological function, it has been suggested that α-CA1 is involved in the transformation in chloroplast stroma of HCO3 to CO2 to supply it at the active site of Rubisco [56,57]. Interestingly, this protein is one of the few plant proteins known to be targeted to the chloroplast through the secretory pathway. It contains several glycosylation sites that must be occupied by N-glycans for correct folding, trafficking, and functionality of the protein. In addition, the protein must be stabilized by a disulfide bridge between the conserved Cys27 and Cys191 residues for folding and endoplasmic reticulum (ER)-export [46,49].
As for α-CA2 and α-CA4, all the literature data obtained so far indicate that they are strongly associated with photosynthetic reactions. In particular, the comparison of different properties of A. thaliana wild type plants with the knockout mutants lacking α-CA2 and α-CA4 suggested that both these proteins participate, competitively, in proton exchange close to photosystem II (PSII), controlling the protonation level of the PsbS protein [53]. This protein is a key player in dissipating excess light energy via the regulation of non-photochemical quenching (NPQ), one of the main mechanisms ensuring that the photosynthetic apparatus is protected from photoinhibition [58]. In particular, α-CA4, catalyzing the CO2 hydration, under high light supplies protons to PsbS protein, determining a conformational change in the light-harvesting antenna and consequently an increase in NPQ. α-CA2 has an opposite role; indeed, this protein, catalyzing the HCO3 dehydration, ensures fast deprotonation of PsbS under low light to avoid energy dissipation when the amount of light energy is low [53]. The presence of α-CA4 in PSII membranes and its involvement in determining proton concentration in this district have also been confirmed by more recent studies [54,59]. Moreover, it has been shown that the expression levels of the gene encoding α-CA4 increased significantly in response to the increase in light intensity [57], in agreement with the hypothesis of its role in the regulation of NPQ and in the protection of photosynthetic apparatus from photoinhibition [58].
Finally, it has been recently demonstrated that α-CA5 takes part in the photophosphorylation stimulation in the presence of bicarbonate excess in thylakoids [51,52]. Moreover, its participation in the conversion of bicarbonate to CO2 with the aim of supplying the latter to the membrane-bound Rubisco has been also hypothesized, thanks to its position on the stromal surface of stromal thylakoid membranes [55].
α-CAs have been found also in other plants, such as the α-CA Sb5G039000 expressed in Sorghum bicolor anthers [8,60], and the α-CAs Mt1g059900 and Mt1g059940 expressed in Medicago trunculata root nodules [8,61], but to date, an indication of their physiological function is not available.

3. Algae

Many α-CAs have been so far identified in photosynthetic algae and some of them have been widely characterized both from a functional and structural point of view [6,62,63]. Most of the available data concern enzymes found in green and red algae and will be described in detail in the following paragraphs. Putative α-CA encoding genes have also been identified in the genome sequences of the marine diatoms Phaeodactylum tricornutum and Thalassiosira pseudonana. However, the sequence alignment of the corresponding proteins with α-CAs from other organisms shows that none of the predicted active sites possess the three zinc-coordinating histidine residues [64,65], leaving doubts about their actual classification as α-CAs. For this reason, these enzymes will not be discussed in this review.
Clamidomonas reinhardtii is a unicellular green alga that has been extensively studied in recent years, enhancing the understanding of CCM in green algae (Figure 1) [63]. This alga encodes for several CAs, belonging to α-, β- and γ-classes, which are involved in the CCM and photosynthesis. In particular, three α-CAs, namely CAH1, CAH2 and CAH3, have been identified [66,67,68,69,70,71,72]. CAH1 and CAH2 are localized in the periplasm [66,73,74], while CAH3 was found in the thylakoid lumen [75,76]. CAH1 expression is induced under low CO2 conditions in the presence of light, while CAH2 is poorly expressed under low CO2 and slightly up-regulated under high CO2 [31,69]. Finally, CAH3 is constitutively expressed, not showing a strong response to modifications in the CO2 level [62].
Studies on the inhibition of periplasmic CAs of Chlamydomonas, using acetazolamide (AZA) and the membrane impermeant CA inhibitor dextran bound sulphonamide (DBS), showed a decrease in cell affinity for Ci and the inhibition of the Ci-dependent O2 evolution [77,78]. Therefore, it was suggested that CAH1 and CAH2 are involved in CCM, facilitating the diffusion of the Ci to the plasma membrane [62,79].
CAH3 is proposed to play a critical role in CCM of Chlamydomonas, being responsible for the rapid conversion of HCO3 to CO2 in the acidic lumen of intra-pyrenoid thylakoids, thus increasing the concentration of CO2 around Rubisco and consequently enhancing photosynthetic efficiency [79,80]. However, some evidence has been provided for the involvement of CAH3 in PSII activity as well [71], suggesting that under the light this enzyme, associated with the donor side of PSII, promotes the removal of protons transferred from the active site of the PSII water-oxidizing complex (WOC) to the lumen, catalyzing the bicarbonate dehydration reaction [71]. This avoids local acidification close to the WOC active site, thus improving its functioning. The recent study by Terentyev and colleagues carried out using specific CA inhibitors, namely trifluoromethanesulfonamide (TFMSA), ethoxyzolamide (EZA), and AZA, and measuring the pH-dependent change in PSII activity, provided further support to this hypothesis [81]. On the other side, Blanco-Rivero and co-workers reported that CAH3 is post-transcriptionally regulated via phosphorylation/dephosphorylation [47]. They also hypothesized that CAH3 remains associated with PSII in stromal thylakoids in high CO2 conditions, but under low CO2 it becomes phosphorylated and is concentrated in the intra-pyrenoid thylakoid. Thus, it seems that the function of CAH3 is related to its location within thylakoids, which in turn is regulated by post-translational modifications [32].
CAH1 and CAH3 were investigated also from a structural point of view by means of X-ray crystallography [82,83]. CAH1 is a glycosylated protein composed of a small and a large subunit which are linked by a disulfide bond [82]. The protein crystallizes as a dimer, with the two monomers linked together through an intermolecular disulfide bridge and several interactions between the loop regions (Figure 2) [82]. The formation of the dimer has also been confirmed in solution, although several data suggested that the protein can be a tetramer in some physiological conditions [82]. Each monomer displays the classical α-CA fold, with the active site region and all the key residues for the catalytic activity being very conserved (Figure 3). Some differences with respect to human CA II (hCA II) structure reside in the central β-sheet core which consists of 9 strands in CAH1 instead of 10 in hCA II. Moreover, the CAH1 structure shows some additional secondary structure elements with respect to hCA II in the surroundings of the β-sheet core. The crystallographic structure also reveals the existence of a network of hydrogen bonds between residues located in the small subunit and active site residues of the large subunit, in agreement with the experimental finding that the association of the small and big subunit is required for enzyme activity [70].
CAH3 crystallizes as a dimer as well; however, in this case, the dimer is stabilized by the swap of N-terminal arms of the two monomers (Figure 4) [83]. In contrast, biochemical experiments indicated that the enzyme is a monomer in solution, even though it was suggested that in the crowded lumen under certain conditions it can occur as a dimer. CAH3 monomer retains the typical α-CA fold, and the geometry of the active site is well preserved; however, the protein surface is more hydrophobic with respect to that of human isoforms, allowing the interaction of the enzyme with the thylakoid membrane [83]. From the structural comparison with hCA II, it emerges that some sequence deletions are present in CAH3 (Figure 3). Among these, the deletion corresponding to the hCA II region 130–139 involves the rim of the catalytic cavity and is responsible for the narrowing of the active site cavity of CAH3 compared to that of the human enzyme.
Dunaliella salina is another unicellular green alga, capable to survive in very variable salinity conditions, from freshwater to hyper-saline lakes, such as the Dead Sea. This alga possesses two extracellular α-CAs, i.e. dCAI and dCAII, which have been proposed to be involved under limiting CO2 conditions in the supply of CO2 to the cells, obtaining it from bicarbonate [86,87]. dCAI is an unusual internally duplicated 60 kDa protein, consisting of two 52% identical α-CA domains, whereas dCAII is a single domain protein exhibiting 55% sequence identity to each dCAI domain [88]. The two proteins exhibit comparable catalytic activity and retain an active conformation over a large range of salinities [88,89]. The crystal structure of dCAII has been solved, highlighting that the protein crystallized as a dimer, with the two monomers related by a non-crystallographic two-fold axis of symmetry (Figure 5). The global fold of each monomer, characterized by a central ten-stranded β-sheet, is highly conserved although presenting some peculiar features [89]. Indeed, a comparison with other α-CAs revealed that in the dCAII structure there is a higher content of helical structure and a reduced amount of β-strand one. Moreover, the dCAII sequence shows several insertions and deletions with respect to α-CAs and the proton-shuttle His64 is missing (Figure 3) [89]. Interestingly, the solvent-accessible surface of dCAII was characterized by a reduced number of basic residues, mainly lysines, with respect to the other α-CAs, thus leading to a predominantly negative electrostatic potential surface and to a decrease in the surface hydrophobic character. It was hypothesized that the preferentially negative electrostatic potential surface could enhance protein stability and solubility in high salt concentrations. These properties are in common with the other halophilic proteins previously reported [90], although dCAII differs in its ability to maintain solubility, enzymatic activity, and correct folding even at low salt concentrations [89].
Nannochloropsis oceanica is a unicellular picoplanktonic alga that represents an emerging model for research on photosynthesis and algae biology [91,92,93]. Differently from C. reinhardtii, which possesses the transpyrenoidal thylakoids where CAH3, one of the main players of CCM, is located (Figure 1) [64,75], N. oceanica lacks a pyrenoid, thus suggesting that the CCM components of this alga have a different spatial configuration. In agreement with this hypothesis, recent experimental evidence indicated NoCAH1, an α-CA localized in the lumen of the epiplastid ER, as an essential component of the CCM in N. oceanica [6,36]. In particular, the proposed model assumes that bicarbonate transporters pump HCO3 into the cytoplasm and then into the ER lumen, where NoCAH1 accumulates. Thus, the protein catalyzes the formation of CO2 that either diffuses into the chloroplast stroma to be fixed by Rubisco or escapes from the cell (Figure 6) [36]. Accordingly, with this model, NoCAH1 expression was regulated by the concentration of external Ci at both the protein and transcript levels [36].
Moreover, Gracilariopsis chorda, an agar-producing multicellular marine red algal species, contains four α-CAs, namely GcCAHα1, GcCAHα2, GcCAHα3 and GcCAHα4 [94]. In silico analysis of these proteins using different prediction tools revealed in all of them an N-terminal sequence with a high degree of hydrophobicity that could function as a leader sequence for their targeting to the ER. In agreement with this hypothesis, when heterologously expressed in protoplasts of A. thaliana leaf cells, the four GcCAHαs were localized in the ER and two of them (i.e. GcCAHα2 and GcCAHα4) were further targeted to the vacuole [94]. However, it cannot be excluded that in G. chorda GcCAHα1 and GcCAHα3 are initially targeted to the ER to be subsequently transported to other organelles, such as the chloroplast, and that this does not happen in Arabidopsis due to the difference in the mechanisms of trafficking from the endomembrane compartments to chloroplasts between red algae and Arabidopsis [94]. As it concerns the physiological roles of the G. chorda α-CAs, even if their involvement in the red alga CCM could be hypothesized, further studies are absolutely needed to clarify this point.

4. Cyanobacteria

Cyanobacteria are a very large group of photosynthetic bacteria, diffused in various habitats [95]. They appeared on the earth at least 3.5 billion years ago [96] and during their evolution they have been subjected to profound mutations due to deep changes in the gaseous composition of the earth’s environment, initially containing a high CO2 content and low O2 levels [33,35]. The conversion of the early oxygen-poor reducing atmosphere into an oxidizing one, with a drastic reduction in CO2 concentration, pushed cyanobacteria to develop alternative mechanisms for efficiently acquiring Ci for photosynthesis. In particular, they developed a very efficient photosynthetic CCM, which allowed them to survive at low CO2 concentrations [35,42]. The carboxysome is an essential part of the cyanobacterial CCM, unlike algae in which the pyrenoid is instead present (Figure 7) [78,97]. Cyanobacteria CAs play a central role in this mechanism [42] and up to now, α-, β- and γ-CAs [42,98] have been identified in these organisms. β- and γ-CAs are predominant, while α-CAs are less widespread [99].
A gene encoding an α-CA (EcaA, external carbonic anhydrase alpha class) was initially isolated from the cyanobacteria Anabaena sp. strain PCC7120 and Synechococcus elongatus PCC7942 [41]. Anabaena EcaA is a protein of approximately 29 kDa which shows a significant amino acid sequence homology with several human α-CAs, such as hCA I and hCA II, including conservation of most active site residues required for Zn2+ binding and catalytic activity (Figure 8) [41,98]. A twin-arginine translocation (Tat) signal peptide is present in the N-terminal region of this protein and is responsible for protein transport to the outside of the cell [98,100]. Indeed, immunogold localization studies with polyclonal antisera directed against Anabaena EcaA showed that this CA has an extracellular location and is associated with the cell wall, periplasmic space or cytoplasmic membrane [41]. Expression of Anabaena EcaA is regulated by CO2 concentration in the growth medium, being highest in cells grown at elevated CO2 levels (1% CO2 in air), whereas cells transferred in medium with low levels of CO2 (0.01%) undergo a gradual decline in protein expression [41].
Synechococcus EcaA is a 26 KDa protein with high sequence homology with both human α-CAs and Anabaena EcaA (Figure 8), and as for the latter, it contains at the N-terminus many positively charged residues indicative of the presence of the signal peptide for membrane targeting [41]. Moreover, protein expression in this case is also influenced by CO2 concentration, with high levels of the enzyme occurring when cells are grown at elevated concentrations of CO2 [41].
Despite EcaA proteins being discovered more than twenty years ago, their CA activity was not confirmed for a long time. In 2009, Wang and coworkers [101] reported a relatively high activity of the Anabaena EcaA protein heterologously expressed in E. coli, whereas only recently clear evidence of the in vitro CA activity of Synechococcus enzyme was obtained [102]. Interestingly, differently from Anabaena EcaA, the Synechococcus EcaA contains a disulphide bond that is important for the enzyme activity, as already reported for other CAs belonging to the α-class. Indeed, the enzyme shows high specific CA activity only when it is expressed in heterologous bacterial systems which support the formation of disulfide bonds, independently from the presence or absence of the leader peptide at the N-terminus. Accordingly, the enzyme activity of recombinant Synechococcus EcaA expressed in bacterial systems that do not support optimal disulfide bond formation can be restored by the addition of a thiol-oxidizing agent [102].
The functional role of the EcaA proteins in cyanobacterial photosynthesis was also investigated by generating a deletion mutant of the Synechococcus protein. Surprisingly, both the CA activity, measured at the cell surface, and the growth rate of the mutant were comparable to those of the WT cells [41,103], suggesting that EcaA does not have a significant role in the CCM. Thus, due to its capability to bind CO2 and HCO3 and to its cell-surface location, it was hypothesized that this protein could play a role as a sensor, detecting or signaling changes in the level of CO2 in the environment [40,41,42,98]. Additional studies are necessary to clarify this point.
Further, EcaA enzymes have been found in other cyanobacteria, such as the alkalophilic cyanobacteria Rhabdoderma lineare and Microcoleus chthonoplates, and the marine cyanobacterium Cyanothece sp. ATCC51142 [104,105,106]. Cyanothece EcaA is a highly active CA, whose enzymatic activity was confirmed both in vivo and in vitro, using intact cells and a recombinant protein expressed in a heterologous host, respectively [106]. Interestingly, even if Cyanothece EcaA contains two cysteine residues (Cys55 and Cys209), unlike Synechococcus EcaA, its enzymatic activity does not show a redox regulation, indicating the absence of the disulfide bond [102]. Inhibition studies using specific α-CA inhibitors with different membrane permeability suggested that Cyanothece EcaA is not associated with the outer membrane, since the rate of CO2 hydration was highly reduced by EZA, which is characterized by high permeability through cell membranes but not by AZA, which cannot penetrate cells [106]. Further studies, carried out in E. coli cells with several recombinant forms of the Cyanothece EcaA protein (with and without the leader peptide), confirmed the transport of recombinant proteins containing the leader peptide across the inner cytoplasmic membrane in E. coli cells, as also directly confirmed by immunofluorescent microscopy [107].
Finally, EcaAs isolated from R. lineare and M. chthonoplates are two extracellular α-CAs, localized in the glycocalyx membrane [105,108], probably involved in cell survival under extreme conditions of soda lakes. Indeed, it has been suggested that these enzymes may preserve the intracellular Ci pool for photoautotrophic assimilation, preventing CO2 leakage from the cell through its conversion into bicarbonate [99,109].

5. Conclusions and Future Perspectives

In this review, we have provided a comprehensive overview of experimental data existing on α-CAs present in photosynthetic organisms, highlighting that these enzymes can have different cellular localization and adopt many physiological roles. In higher plants, these enzymes have been little studied; however, available data indicate their involvement in both supplying CO2 at the active site of Rubisco and determining proton concentration in PSII membranes. On the contrary, a greater number of studies are available on the algal and cyanobacterial α-CAs, which have been demonstrated to be actively involved in CCM, photosynthetic reactions and detecting or signaling changes in the environment’s CO2 level. Even the biochemical and structural studies are limited in number and do not allow us to rationalize the presence of structural features associated with particular cellular localizations or physiological functions. From this scenario, it emerges that many studies are still needed on α-CAs from photosynthetic organisms in order to provide a clear structure–function relationship and to understand their correlation with the physiological functions exercised by the other classes of CAs present in the same organisms.

Author Contributions

Conceptualization, A.D.F. and K.D.; writing—original draft preparation, E.L., A.D.F., V.A., S.M.M., G.D.S. and K.D.; writing—review and editing, E.L., A.D.F., V.A., S.M.M., G.D.S. and K.D.; funding acquisition, G.D.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by MUR (grant FISR2019_04819 BacCAD).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We thank Florinda Pignatiello for administrative support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Monti, D.M.; De Simone, G.; Langella, E.; Supuran, C.T.; Di Fiore, A.; Monti, S.M. Insights into the Role of Reactive Sulfhydryl Groups of Carbonic Anhydrase III and VII during Oxidative Damage. J. Enzym. Inhib. Med. Chem. 2017, 32, 5–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Di Fiore, A.; Monti, D.M.; Scaloni, A.; De Simone, G.; Monti, S.M. Protective Role of Carbonic Anhydrases III and VII in Cellular Defense Mechanisms upon Redox Unbalance. Oxid. Med. Cell. Longev. 2018, 2018, 1–9. [Google Scholar] [CrossRef] [PubMed]
  3. Buonanno, M.; Langella, E.; Zambrano, N.; Succoio, M.; Sasso, E.; Alterio, V.; Di Fiore, A.; Sandomenico, A.; Supuran, C.T.; Scaloni, A.; et al. Disclosing the Interaction of Carbonic Anhydrase IX with Cullin-Associated NEDD8-Dissociated Protein 1 by Molecular Modeling and Integrated Binding Measurements. ACS Chem. Biol. 2017, 12, 1460–1465. [Google Scholar] [CrossRef] [PubMed]
  4. Buanne, P.; Renzone, G.; Monteleone, F.; Vitale, M.; Monti, S.M.; Sandomenico, A.M.; Garbi, C.; Montanaro, D.; Accardo, M.; Troncone, G.; et al. Characterization of Carbonic Anhydrase IX Interactome Reveals Proteins Assisting Its Nuclear Localization in Hypoxic Cells. J. Proteome Res. 2013, 12, 282–292. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Lionetto, M.G.; Caricato, R.; Giordano, M.E.; Schettino, T. The Complex Relationship between Metals and Carbonic Anhydrase: New Insights and Perspectives. Int. J. Mol. Sci. 2016, 17, 127. [Google Scholar] [CrossRef] [Green Version]
  6. DiMario, R.J.; Machingura, M.C.; Waldrop, G.L.; Moroney, J.V. The Many Types of Carbonic Anhydrases in Photosynthetic Organisms. Plant Sci. 2018, 268, 11–17. [Google Scholar] [CrossRef]
  7. Alterio, V.; Di Fiore, A.; D’Ambrosio, K.; Supuran, C.T.; De Simone, G. Multiple Binding Modes of Inhibitors to Carbonic Anhydrases: How to Design Specific Drugs Targeting 15 Different Isoforms? Chem. Rev. 2012, 112, 4421–4468. [Google Scholar] [CrossRef] [Green Version]
  8. DiMario, R.J.; Clayton, H.; Mukherjee, A.; Ludwig, M.; Moroney, J.V. Plant Carbonic Anhydrases: Structures, Locations, Evolution, and Physiological Roles. Mol. Plant 2017, 10, 30–46. [Google Scholar] [CrossRef] [Green Version]
  9. Monti, S.M.; Supuran, C.T.; De Simone, G. Carbonic Anhydrase IX as a Target for Designing Novel Anticancer Drugs. Curr. Med. Chem. 2012, 19, 821–830. [Google Scholar] [CrossRef]
  10. Esbaugh, A.J.; Tufts, B.L. The Structure and Function of Carbonic Anhydrase Isozymes in the Respiratory System of Vertebrates. Respir. Physiol. Neurobiol. 2006, 154, 185–198. [Google Scholar] [CrossRef]
  11. Rowlett, R.S. Structure and Catalytic Mechanism of the Beta-Carbonic Anhydrases. Biochim. Biophys. Acta 2010, 1804, 362–373. [Google Scholar] [CrossRef] [PubMed]
  12. Dathan, N.A.; Alterio, V.; Troiano, E.; Vullo, D.; Ludwig, M.; De Simone, G.; Supuran, C.T.; Monti, S.M. Biochemical Characterization of the Chloroplastic β-Carbonic Anhydrase from Flaveria Bidentis (L.) “Kuntze”. J. Enzym. Inhib. Med. Chem. 2014, 29, 500–504. [Google Scholar] [CrossRef] [PubMed]
  13. Kimber, M.; Pai, E. The Active Site Architecture of Pisum Sativum Beta-Carbonic Anhydrase Is a Mirror Image of That of Alpha-Carbonic Anhydrases. EMBO J. 2000, 19, 1407–1418. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Monti, S.M.; De Simone, G.; Dathan, N.A.; Ludwig, M.; Vullo, D.; Scozzafava, A.; Capasso, C.; Supuran, C.T. Kinetic and Anion Inhibition Studies of a β-Carbonic Anhydrase (FbiCA 1) from the C4 Plant Flaveria Bidentis. Bioorg. Med. Chem. Lett. 2013, 23, 1626–1630. [Google Scholar] [CrossRef]
  15. Parisi, G.; Perales, M.; Fornasari, M.S.; Colaneri, A.; González-Schain, N.; Gómez-Casati, D.; Zimmermann, S.; Brennicke, A.; Araya, A.; Ferry, J.G.; et al. Gamma Carbonic Anhydrases in Plant Mitochondria. Plant Mol. Biol. 2004, 55, 193–207. [Google Scholar] [CrossRef]
  16. Cox, E.H.; McLendon, G.L.; Morel, F.M.M.; Lane, T.W.; Prince, R.C.; Pickering, I.J.; George, G.N. The Active Site Structure of Thalassiosira Weissflogii Carbonic Anhydrase 1. Biochemistry 2000, 39, 12128–12130. [Google Scholar] [CrossRef]
  17. Lane, T.W.; Morel, F.M.M. Regulation of Carbonic Anhydrase Expression by Zinc, Cobalt, and Carbon Dioxide in the Marine Diatom Thalassiosira Weissflogii. Plant Physiol. 2000, 123, 345–352. [Google Scholar] [CrossRef] [Green Version]
  18. Alterio, V.; Langella, E.; De Simone, G.; Monti, S.M. Cadmium-Containing Carbonic Anhydrase CDCA1 in Marine Diatom Thalassiosira Weissflogii. Mar. Drugs 2015, 13, 1688–1697. [Google Scholar] [CrossRef] [Green Version]
  19. Angeli, A.; Buonanno, M.; Donald, W.A.; Monti, S.M.; Supuran, C.T. The Zinc—but Not Cadmium—Containing ζ-Carbonic from the Diatom Thalassiosira Weissflogii Is Potently Activated by Amines and Amino Acids. Bioorg. Chem. 2018, 80, 261–265. [Google Scholar] [CrossRef]
  20. Viparelli, F.; Monti, S.M.; Simone, G.D..; Innocenti, A.; Scozzafava, A.; Xu, Y.; Morel, F.M.M.; Supuran, C.T. Inhibition of the R1 Fragment of the Cadmium-Containing ζ-Class Carbonic Anhydrase from the Diatom Thalassiosira Weissflogii with Anions. Bioorg. Med. Chem. Lett. 2010, 20, 4745–4748. [Google Scholar] [CrossRef]
  21. Alterio, V.; Langella, E.; Buonanno, M.; Esposito, D.; Nocentini, A.; Berrino, E.; Bua, S.; Polentarutti, M.; Supuran, C.T.; Monti, S.M.; et al. Zeta-Carbonic Anhydrases Show CS2 Hydrolase Activity: A New Metabolic Carbon Acquisition Pathway in Diatoms? Comput. Struct. Biotechnol. J. 2021, 19, 3427–3436. [Google Scholar] [CrossRef] [PubMed]
  22. Del Prete, S.; Vullo, D.; Fisher, G.M.; Andrews, K.T.; Poulsen, S.A.; Capasso, C.; Supuran, C.T. Discovery of a New Family of Carbonic Anhydrases in the Malaria Pathogen Plasmodium Falciparum--the η-Carbonic Anhydrases. Bioorg. Med. Chem. Lett. 2014, 24, 4389–4396. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Kikutani, S.; Nakajima, K.; Nagasato, C.; Tsuji, Y.; Miyatake, A.; Matsuda, Y. Thylakoid Luminal Θ-Carbonic Anhydrase Critical for Growth and Photosynthesis in the Marine Diatom Phaeodactylum Tricornutum. Proc. Natl. Acad. Sci. USA 2016, 113, 9828–9833. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. 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]
  25. Nocentini, A.; Supuran, C.T.; Capasso, C. An Overview on the Recently Discovered Iota-Carbonic Anhydrases. J. Enzym. Inhib. Med. Chem. 2021, 36, 1988–1995. [Google Scholar] [CrossRef] [PubMed]
  26. Di Fiore, A.; D’Ambrosio, K.; Ayoub, J.; Alterio, V.; De Simone, G. Chapter 2—α-Carbonic Anhydrases. In Carbonic Anhydrases; Supuran, C.T., Nocentini, A., Eds.; Academic Press: Cambridge, MA, USA, 2019; pp. 19–54. ISBN 978-0-12-816476-1. [Google Scholar]
  27. D’Ambrosio, K.; Smaine, F.Z.; Carta, F.; De Simone, G.; Winum, J.Y.; Supuran, C.T. Development of Potent Carbonic Anhydrase Inhibitors Incorporating Both Sulfonamide and Sulfamide Groups. J. Med. Chem. 2012, 55, 6776–6783. [Google Scholar] [CrossRef] [PubMed]
  28. Di Fiore, A.; Truppo, E.; Supuran, C.T.; Alterio, V.; Dathan, N.; Bootorabi, F.; Parkkila, S.; Monti, S.M.; Simone, G. De Crystal Structure of the C183S/C217S Mutant of Human CA VII in Complex with Acetazolamide. Bioorg. Med. Chem. Lett. 2010, 20, 5023–5026. [Google Scholar] [CrossRef]
  29. Alterio, V.; De Simone, G.; Monti, S.M.; Scozzafava, A.; Supuran, C.T. Carbonic Anhydrase Inhibitors: Inhibition of Human, Bacterial, and Archaeal Isozymes with Benzene-1,3-Disulfonamides--Solution and Crystallographic Studies. Bioorg. Med. Chem. Lett. 2007, 17, 4201–4207. [Google Scholar] [CrossRef]
  30. Langella, E.; Buonanno, M.; Vullo, D.; Dathan, N.; Leone, M.; Supuran, C.T.; De Simone, G.; Monti, S.M. Biochemical, Biophysical and Molecular Dynamics Studies on the Proteoglycan-like Domain of Carbonic Anhydrase IX. Cell. Mol. Life Sci. 2018 7517 2018, 75, 3283–3296. [Google Scholar] [CrossRef]
  31. Moroney, J.V.; Bartlett, S.G.; Samuelsson, G. Carbonic Anhydrases in Plants and Algae. Plant. Cell Environ. 2001, 24, 141–153. [Google Scholar] [CrossRef]
  32. Wang, Y.; Stessman, D.J.; Spalding, M.H. The CO2 Concentrating Mechanism and Photosynthetic Carbon Assimilation in Limiting CO2: How Chlamydomonas Works against the Gradient. Plant J. 2015, 82, 429–448. [Google Scholar] [CrossRef] [PubMed]
  33. Tomar, V.; Sidhu, G.K.; Nogia, P.; Mehrotra, R.; Mehrotra, S. Regulatory Components of Carbon Concentrating Mechanisms in Aquatic Unicellular Photosynthetic Organisms. Plant Cell Rep. 2017, 36, 1671–1688. [Google Scholar] [CrossRef] [PubMed]
  34. Tsuji, Y.; Nakajima, K.; Matsuda, Y. Molecular Aspects of the Biophysical CO2-Concentrating Mechanism and Its Regulation in Marine Diatoms. J. Exp. Bot. 2017, 68, 3763–3772. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Badger, M.R.; Price, G.D. CO2 Concentrating Mechanisms in Cyanobacteria: Molecular Components, Their Diversity and Evolution. J. Exp. Bot. 2003, 54, 609–622. [Google Scholar] [CrossRef]
  36. Gee, C.W.; Niyogi, K.K. The Carbonic Anhydrase CAH1 Is an Essential Component of the Carbon-Concentrating Mechanism in Nannochloropsis Oceanica. Proc. Natl. Acad. Sci. USA 2017, 114, 4537–4542. [Google Scholar] [CrossRef] [Green Version]
  37. Ogren, W.L. Photorespiration: Pathways, Regulation, and Modification. Annu. Rev. Plant Physiol. 1984, 35, 415–442. [Google Scholar] [CrossRef]
  38. Bauwe, H.; Hagemann, M.; Fernie, A.R. Photorespiration: Players, Partners and Origin. Trends Plant Sci. 2010, 15, 330–336. [Google Scholar] [CrossRef]
  39. Badger, M. The Roles of Carbonic Anhydrases in Photosynthetic CO2 Concentrating Mechanisms. Photosynth. Res. 2003, 77, 83–94. [Google Scholar] [CrossRef]
  40. So, A.K.C.; Van Spall, H.G.C.; Coleman, J.R.; Espie, G.S. Catalytic Exchange of 18O from 13C18O-Labelled CO2 by Wild-Type Cells and EcaA, EcaB, and CcaA Mutants of the Cyanobacteria Synechococcus PCC7942 and Synechocystis PCC6803. Can. J. Bot. 1998, 76, 1153–1160. [Google Scholar] [CrossRef]
  41. Soltes-Rak, E.; Mulligan, M.E.; Coleman, J.R. Identification and Characterization of a Gene Encoding a Vertebrate-Type Carbonic Anhydrase in Cyanobacteria. J. Bacteriol. 1997, 179, 769. [Google Scholar] [CrossRef]
  42. Smith, K.S.; Ferry, J.G. Prokaryotic Carbonic Anhydrases. FEMS Microbiol. Rev. 2000, 24, 335–366. [Google Scholar] [CrossRef] [PubMed]
  43. Langella, E.; Alterio, V.; D’Ambrosio, K.; Cadoni, R.; Winum, J.Y.; Supuran, C.T.; Monti, S.M.; De Simone, G.; Di Fiore, A. Exploring Benzoxaborole Derivatives as Carbonic Anhydrase Inhibitors: A Structural and Computational Analysis Reveals Their Conformational Variability as a Tool to Increase Enzyme Selectivity. J. Enzym. Inhib. Med. Chem. 2019, 34, 1498–1505. [Google Scholar] [CrossRef]
  44. De Simone, G.; Langella, E.; Esposito, D.; Supuran, C.T.; Monti, S.M.; Winum, J.Y.; Alterio, V. Insights into the Binding Mode of Sulphamates and Sulphamides to HCA II: Crystallographic Studies and Binding Free Energy Calculations. J. Enzym. Inhib. Med. Chem. 2017, 32, 1002–1011. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Momayyezi, M.; McKown, A.D.; Bell, S.C.S.; Guy, R.D. Emerging Roles for Carbonic Anhydrase in Mesophyll Conductance and Photosynthesis. Plant J. 2020, 101, 831–844. [Google Scholar] [CrossRef] [Green Version]
  46. Villarejo, A.; Burén, S.; Larsson, S.; Déjardin, A.; Monné, M.; Rudhe, C.; Karlsson, J.; Jansson, S.; Lerouge, P.; Rolland, N.; et al. Evidence for a Protein Transported through the Secretory Pathway En Route to the Higher Plant Chloroplast. Nat. Cell Biol. 2005, 7, 1224–1231. [Google Scholar] [CrossRef] [PubMed]
  47. Blanco-Rivero, A.; Shutova, T.; Román, M.J.; Villarejo, A.; Martinez, F. Phosphorylation Controls the Localization and Activation of the Lumenal Carbonic Anhydrase in Chlamydomonas Reinhardtii. PLoS ONE 2012, 7, e49063. [Google Scholar] [CrossRef] [Green Version]
  48. Rudenko, N.N.; Ignatova, L.K.; Fedorchuk, T.P.; Ivanov, B.N. Carbonic Anhydrases in Photosynthetic Cells of Higher Plants. Biochemistry. (Mosc). 2015, 80, 674–687. [Google Scholar] [CrossRef]
  49. Burén, S.; Ortega-Villasante, C.; Blanco-Rivero, A.; Martínez-Bernardini, A.; Shutova, T.; Shevela, D.; Messinger, J.; Bako, L.; Villarejo, A.; Samuelsson, G. Importance of Post-Translational Modifications for Functionality of a Chloroplast-Localized Carbonic Anhydrase (CAH1) in Arabidopsis Thaliana. PLoS ONE 2011, 6, e21021. [Google Scholar] [CrossRef] [Green Version]
  50. Friso, G.; Giacomelli, L.; Ytterberg, A.J.; Peltier, J.B.; Rudella, A.; Sun, Q.; Van Wijk, K.J. In-Depth Analysis of the Thylakoid Membrane Proteome of Arabidopsis Thaliana Chloroplasts: New Proteins, New Functions, and a Plastid Proteome Database. Plant Cell 2004, 16, 478–499. [Google Scholar] [CrossRef] [Green Version]
  51. Fedorchuk, T.P.; Opanasenko, V.K.; Rudenko, N.N.; Ivanov, B.N. Bicarbonate-Induced Stimulation of Photophosphorylation in Isolated Thylakoids: Effects of Carbonic Anhydrase Inhibitors. Biol. Membr. 2018, 35, 34–41. [Google Scholar] [CrossRef]
  52. Fedorchuk, T.P.; Kireeva, I.A.; Opanasenko, V.K.; Terentyev, V.V.; Rudenko, N.N.; Borisova-Mubarakshina, M.M.; Ivanov, B.N. Alpha Carbonic Anhydrase 5 Mediates Stimulation of ATP Synthesis by Bicarbonate in Isolated Arabidopsis Thylakoids. Front. Plant Sci. 2021, 12, 891. [Google Scholar] [CrossRef]
  53. Zhurikova, E.M.; Ignatova, L.K.; Rudenko, N.N.; Mudrik, V.A.; Vetoshkina, D.V.; Ivanov, B.N. Participation of Two Carbonic Anhydrases of the Alpha Family in Photosynthetic Reactions in Arabidopsis Thaliana. Biochemistry. (Mosc). 2016, 81, 1182–1187. [Google Scholar] [CrossRef] [PubMed]
  54. Rudenko, N.N.; Ignatova, L.K.; Nadeeva-Zhurikova, E.M.; Fedorchuk, T.P.; Ivanov, B.N.; Borisova-Mubarakshina, M.M. Advances in Understanding the Physiological Role and Locations of Carbonic Anhydrases in C3 Plant Cells. Protoplasma 2021, 258, 249–262. [Google Scholar] [CrossRef] [PubMed]
  55. Rudenko, N.N.; Ivanov, B.N. Unsolved Problems of Carbonic Anhydrases Functioning in Photosynthetic Cells of Higher C3 Plants. Biochemistry. (Mosc). 2021, 86, 1243–1255. [Google Scholar] [CrossRef] [PubMed]
  56. Burén, S. Targeting and Function of CAH1-Characterization of a Novel Protein Pathway to the Plant Cell Chloroplast; Umeå University: Umeå, Sweden, 2010. [Google Scholar]
  57. Rudenko, N.N.; Vetoshkina, D.V.; Fedorchuk, T.P.; Ivanov, B.N. Effect of Light Intensity under Different Photoperiods on Expression Level of Carbonic Anhydrase Genes of the α- and β-Families in Arabidopsis Thaliana Leaves. Biochemistry. (Mosc). 2017, 82, 1025–1035. [Google Scholar] [CrossRef]
  58. Rudenko, N.N.; Fedorchuk, T.P.; Vetoshkina, D.V.; Zhurikova, E.M.; Ignatova, L.K.; Ivanov, B.N. Influence of Knockout of At4g20990 Gene Encoding α-CA4 on Photosystem II Light-Harvesting Antenna in Plants Grown under Different Light Intensities and Day Lengths. Protoplasma 2018, 255, 69–78. [Google Scholar] [CrossRef]
  59. Ignatova, L.; Zhurikova, E.; Ivanov, B. The Presence of the Low Molecular Mass Carbonic Anhydrase in Photosystem II of C3 Higher Plants. J. Plant Physiol. 2019, 232, 94–99. [Google Scholar] [CrossRef]
  60. Makita, Y.; Shimada, S.; Kawashima, M.; Kondou-Kuriyama, T.; Toyoda, T.; Matsui, M. MOROKOSHI: Transcriptome Database in Sorghum Bicolor. Plant Cell Physiol. 2015, 56, e6. [Google Scholar] [CrossRef] [Green Version]
  61. Tang, H.; Krishnakumar, V.; Bidwell, S.; Rosen, B.; Chan, A.; Zhou, S.; Gentzbittel, L.; Childs, K.L.; Yandell, M.; Gundlach, H.; et al. An Improved Genome Release (Version Mt4.0) for the Model Legume Medicago Truncatula. BMC Genom. 2014, 15, 1–14. [Google Scholar] [CrossRef] [Green Version]
  62. Moroney, J.V.; Ma, Y.; Frey, W.D.; Fusilier, K.A.; Pham, T.T.; Simms, T.A.; DiMario, R.J.; Yang, J.; Mukherjee, B. The Carbonic Anhydrase Isoforms of Chlamydomonas Reinhardtii: Intracellular Location, Expression, and Physiological Roles. Photosynth. Res. 2011, 109, 133–149. [Google Scholar] [CrossRef]
  63. Aspatwar, A.; Haapanen, S.; Parkkila, S. An Update on the Metabolic Roles of Carbonic Anhydrases in the Model Alga Chlamydomonas Reinhardtii. Metabolites 2018, 8, 22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Tachibana, M.; Allen, A.E.; Kikutani, S.; Endo, Y.; Bowler, C.; Matsuda, Y. Localization of Putative Carbonic Anhydrases in Two Marine Diatoms, Phaeodactylum Tricornutum and Thalassiosira Pseudonana. Photosynth. Res. 2011, 109, 205–221. [Google Scholar] [CrossRef] [PubMed]
  65. Samukawa, M.; Shen, C.; Hopkinson, B.M.; Matsuda, Y. Localization of Putative Carbonic Anhydrases in the Marine Diatom, Thalassiosira Pseudonana. Photosynth. Res. 2014, 121, 235–249. [Google Scholar] [CrossRef] [PubMed]
  66. Coleman, J.R.; Berry, J.A.; Togasaki, R.K.; Grossman, A.R. Identification of Extracellular Carbonic Anhydrase of Chlamydomonas Reinhardtii. Plant Physiol. 1984, 76, 472–477. [Google Scholar] [CrossRef] [Green Version]
  67. Yang, S.Y.; Tsuzuki, M.; Miyachi, S. Carbonic Anhydrase of Chlamydomonas: Purification and Studies on Its Induction Using Antiserum against Chlamydomonas Carbonic Anhydrase. Plant Cell Physiol. 1985, 26, 25–34. [Google Scholar] [CrossRef]
  68. Fukuzawa, H.; Fujiwara, S.; Yamamoto, Y.; Dionisio-Sese, M.L.; Miyachi, S. CDNA Cloning, Sequence, and Expression of Carbonic Anhydrase in Chlamydomonas Reinhardtii: Regulation by Environmental CO2 Concentration. Proc. Natl. Acad. Sci. USA 1990, 87, 4383–4387. [Google Scholar] [CrossRef] [Green Version]
  69. Fujiwara, S.; Fukuzawa, H.; Tachiki, A.; Miyachi, S. Structure and Differential Expression of Two Genes Encoding Carbonic Anhydrase in Chlamydomonas Reinhardtii. Proc. Natl. Acad. Sci. USA 1990, 87, 9779–9783. [Google Scholar] [CrossRef] [Green Version]
  70. Ishida, S.; Muto, S.; Miyachi, S. Structural Analysis of Periplasmic Carbonic Anhydrase 1 of Chlamydomonas Reinhardtii. Eur. J. Biochem. 1993, 214, 9–16. [Google Scholar] [CrossRef]
  71. Shutova, T.; Kenneweg, H.; Buchta, J.; Nikitina, J.; Terentyev, V.; Chernyshov, S.; Andersson, B.; Allakhverdiev, S.I.; Klimov, V.V.; Dau, H.; et al. The Photosystem II-Associated Cah3 in Chlamydomonas Enhances the O2 Evolution Rate by Proton Removal. EMBO J. 2008, 27, 782–791. [Google Scholar] [CrossRef]
  72. Küken, A.; Sommer, F.; Yaneva-Roder, L.; Mackinder, L.C.M.; Höhne, M.; Geimer, S.; Jonikas, M.C.; Schroda, M.; Stitt, M.; Nikoloski, Z.; et al. Effects of Microcompartmentation on Flux Distribution and Metabolic Pools in Chlamydomonas Reinhardtii Chloroplasts. Elife 2018, 7, e37960. [Google Scholar] [CrossRef]
  73. Yagawa, Y.; Aizawa, K.; Yang, S.Y.; Miyachi, S. Release of Carbonic Anhydrase from the Cell Surface of Chlamydomonas Reinhardtii by Trypsin. Plant Cell Physiol. 1986, 27, 215–221. [Google Scholar] [CrossRef]
  74. Kimpel, D.L.; Togasaki, R.K.; Miyachi, S. Carbonic Anhydrase in Chlamydomonas Reinhardtii. I. Localization. Plant Cell Physiol. 1983, 24, 255–259. [Google Scholar] [CrossRef]
  75. Karlsson, J.; Ciarke, A.K.; Chen, Z.Y.; Hugghins, S.Y.; Park, Y.I.; Husic, H.D.; Moroney, J.V.; Samuelsson, G. A Novel Alpha-Type Carbonic Anhydrase Associated with the Thylakoid Membrane in Chlamydomonas Reinhardtii Is Required for Growth at Ambient CO2. EMBO J. 1998, 17, 1208–1216. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Karlsson, J.; Hiltonen, T.; Husic, H.D.; Ramazanov, Z.; Samuelsson, G. Intracellular Carbonic Anhydrase of Chlamydomonas Reinhardtii. Plant Physiol. 1995, 109, 533–539. [Google Scholar] [CrossRef] [Green Version]
  77. Badger, M.R.; Kaplan, A.; Berry, J.A. Internal Inorganic Carbon Pool of Chlamydomonas Reinhardtii. Plant Physiol. 1980, 66, 407–413. [Google Scholar] [CrossRef] [Green Version]
  78. Moroney, J.V.; Husic, H.D.; Tolbert, N.E. Effect of Carbonic Anhydrase Inhibitors on Inorganic Carbon Accumulation by Chlamydomonas Reinhardtii. Plant Physiol. 1985, 79, 177–183. [Google Scholar] [CrossRef] [Green Version]
  79. Moroney, J.V.; Ynalvez, R.A. Proposed Carbon Dioxide Concentrating Mechanism in Chlamydomonas Reinhardtii. Eukaryot. Cell 2007, 6, 1251. [Google Scholar] [CrossRef] [Green Version]
  80. Markelova, A.G.; Sinetova, M.P.; Kupriyanova, E.V.; Pronina, N.A. Distribution and Functional Role of Carbonic Anhydrase Cah3 Associated with Thylakoid Membranes in the Chloroplast and Pyrenoid of Chlamydomonas Reinhardtii. Russ. J. Plant Physiol. 2009, 56, 761–768. [Google Scholar] [CrossRef]
  81. Terentyev, V.V.; Shukshina, A.K.; Shitov, A.V. Carbonic Anhydrase CAH3 Supports the Activity of Photosystem II under Increased PH. Biochim. Biophys. acta. Bioenerg. 2019, 1860, 582–590. [Google Scholar] [CrossRef]
  82. Suzuki, K.; Yang, S.Y.; Shimizu, S.; Morishita, E.C.; Jiang, J.; Zhang, F.; Hoque, M.M.; Sato, Y.; Tsunoda, M.; Sekiguchi, T.; et al. The Unique Structure of Carbonic Anhydrase ACA1 from Chlamydomonas Reinhardtii. Acta Crystallogr. D. Biol. Crystallogr. 2011, 67, 894–901. [Google Scholar] [CrossRef]
  83. Benlloch, R.; Shevela, D.; Hainzl, T.; Grundström, C.; Shutova, T.; Messinger, J.; Samuelsson, G.; Elisabeth Sauer-Eriksson, A. Crystal Structure and Functional Characterization of Photosystem II-Associated Carbonic Anhydrase CAH3 in Chlamydomonas Reinhardtii. Plant Physiol. 2015, 167, 950–962. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Touw, W.G.; Baakman, C.; Black, J.; Te Beek, T.A.H.; Krieger, E.; Joosten, R.P.; Vriend, G. A Series of PDB-Related Databanks for Everyday Needs. Nucleic Acids Res. 2015, 43, D364–D368. [Google Scholar] [CrossRef]
  85. Kabsch, W.; Sander, C. Dictionary of Protein Secondary Structure: Pattern Recognition of Hydrogen-bonded and Geometrical Features. Biopolymers 1983, 22, 2577–2637. [Google Scholar] [CrossRef] [PubMed]
  86. Fisher, M.; Gokhman, I.; Pick, U.; Zamir, A. A Salt-Resistant Plasma Membrane Carbonic Anhydrase Is Induced by Salt in Dunaliella Salina. J. Biol. Chem. 1996, 271, 17718–17723. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Fisher, M.; Pick, U.; Zamir, A. A Salt-Induced 60-Kilodalton Plasma Membrane Protein Plays a Potential Role in the Extreme Halotolerance of the Alga Dunaliella. Plant Physiol. 1994, 106, 1359–1365. [Google Scholar] [CrossRef] [PubMed]
  88. Bageshwar, U.K.; Premkumar, L.; Gokhman, I.; Savchenko, T.; Sussman, J.L.; Zamir, A. Natural Protein Engineering: A Uniquely Salt-Tolerant, but Not Halophilic, Alpha-Type Carbonic Anhydrase from Algae Proliferating in Low- to Hyper-Saline Environments. Protein Eng. Des. Sel. 2004, 17, 191–200. [Google Scholar] [CrossRef] [Green Version]
  89. Premkumar, L.; Greenblatt, H.M.; Bageshwar, U.K.; Savchenko, T.; Gokhman, I.; Sussman, J.L.; Zamir, A. Three-Dimensional Structure of a Halotolerant Algal Carbonic Anhydrase Predicts Halotolerance of a Mammalian Homolog. Proc. Natl. Acad. Sci. USA 2005, 102, 7493–7498. [Google Scholar] [CrossRef] [Green Version]
  90. Bieger, B.; Essen, L.O.; Oesterhelt, D. Crystal Structure of Halophilic Dodecin: A Novel, Dodecameric Flavin Binding Protein from Halobacterium Salinarum. Structure 2003, 11, 375–385. [Google Scholar] [CrossRef] [Green Version]
  91. Moog, D.; Stork, S.; Reislöhner, S.; Grosche, C.; Maier, U.G. In Vivo Localization Studies in the Stramenopile Alga Nannochloropsis Oceanica. Protist 2015, 166, 161–171. [Google Scholar] [CrossRef]
  92. Murakami, R.; Hashimoto, H. Unusual Nuclear Division in Nannochloropsis Oculata (Eustigmatophyceae, Heterokonta) Which May Ensure Faithful Transmission of Secondary Plastids. Protist 2009, 160, 41–49. [Google Scholar] [CrossRef]
  93. Cavalier-Smith, T.; Van Der Giezen, M.; Martin, W.; Wilkins, A.; Doolittle, W.F.; Allen, J.F.; Leaver, C.J.; Whatley, F.R.; Howe, C.J. Genomic Reduction and Evolution of Novel Genetic Membranes and Protein-Targeting Machinery in Eukaryote-Eukaryote Chimaeras (Meta-Algae). Philos. Trans. R. Soc. Lond. B. Biol. Sci. 2003, 358, 109–134. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Razzak, M.A.; Lee, J.M.; Lee, D.W.; Kim, J.H.; Yoon, H.S.; Hwang, I. Expression of Seven Carbonic Anhydrases in Red Alga Gracilariopsis Chorda and Their Subcellular Localization in a Heterologous System, Arabidopsis Thaliana. Plant Cell Rep. 2019, 38, 147–159. [Google Scholar] [CrossRef] [PubMed]
  95. Whitton, B.A.; Potts, M. Introduction to the Cyanobacteria. In Ecology of Cyanobacteria II; Whitton, B., Ed.; Springer: Dordrecht, The Netherlands, 2012; pp. 1–13. ISBN 9789400738553. [Google Scholar]
  96. Schopf, J.W.; Packer, B.M. Early Archean (3.3-Billion to 3.5-Billion-Year-Old) Microfossils from Warrawoona Group, Australia. Science 1987, 237, 70–73. [Google Scholar] [CrossRef] [PubMed]
  97. Cannon, G.C.; Bradburne, C.E.; Aldrich, H.C.; Baker, S.H.; Heinhorst, S.; Shively, J.M. Microcompartments in Prokaryotes: Carboxysomes and Related Polyhedra. Appl. Environ. Microbiol. 2001, 67, 5351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. So, A.K.C.; Espie, G.S. Cyanobacterial Carbonic Anhydrases. Can. J. Bot. 2005, 83, 721–734. [Google Scholar] [CrossRef]
  99. Kupriyanova, E.V.; Sinetova, M.A.; Cho, S.M.; Park, Y.I.; Markelova, A.G.; Los, D.A.; Pronina, N.A. Specific Features of the System of Carbonic Anhydrases of Alkaliphilic Cyanobacteria. Russ. J. Plant Physiol. 2013, 60, 465–471. [Google Scholar] [CrossRef]
  100. Berks, B.C.; Sargent, F.; Palmer, T. The Tat Protein Export Pathway. Mol. Microbiol. 2000, 35, 260–274. [Google Scholar] [CrossRef]
  101. Wang, D.; Li, Q.; Li, W.; Xing, J.; Su, Z. Improvement of Succinate Production by Overexpression of a Cyanobacterial Carbonic Anhydrase in Escherichia Coli. Enzym. Microb. Technol. 2009, 45, 491–497. [Google Scholar] [CrossRef]
  102. Kupriyanova, E.V.; Sinetova, M.A.; Bedbenov, V.S.; Pronina, N.A.; Los, D.A. Putative Extracellular α-Class Carbonic Anhydrase, EcaA, of Synechococcus Elongatus PCC 7942 Is an Active Enzyme: A Sequel to an Old Story. Microbiology 2018, 164, 576–586. [Google Scholar] [CrossRef]
  103. So, A.K.C.; Espie, G.S. Cloning, Characterization and Expression of Carbonic Anhydrase from the Cyanobacterium Synechocystis PCC6803. Plant Mol. Biol. 1998, 37, 205–215. [Google Scholar] [CrossRef]
  104. Kupriyanova, E.V.; Lebedeva, N.V.; Dudoladova, M.V.; Gerasimenko, L.M.; Alekseeva, S.G.; Pronina, N.A.; Zavarzin, G.A. Carbonic Anhydrase Activity of Alkalophilic Cyanobacteria from Soda Lakes. Russ. J. Plant Physiol. 2003, 50, 532–539. [Google Scholar] [CrossRef]
  105. Kupriyanova, E.V.; Markelova, A.G.; Lebedeva, N.V.; Gerasimenko, L.M.; Zavarzin, G.A.; Pronina, N.A. Carbonic Anhydrase of the Alkaliphilic Cyanobacterium Microcoleus Chthonoplastes. Microbiology 2004, 73, 255–259. [Google Scholar] [CrossRef]
  106. Kupriyanova, E.V.; Sinetova, M.A.; Mironov, K.S.; Novikova, G.V.; Dykman, L.A.; Rodionova, M.V.; Gabrielyan, D.A.; Los, D.A. Highly Active Extracellular α-Class Carbonic Anhydrase of Cyanothece Sp. ATCC 51142. Biochimie 2019, 160, 200–209. [Google Scholar] [CrossRef] [PubMed]
  107. Kupriyanova, E.V.; Sinetova, M.A.; Leusenko, A.V.; Voronkov, A.S.; Los, D.A. A Leader Peptide of the Extracellular Cyanobacterial Carbonic Anhydrase Ensures the Efficient Secretion of Recombinant Proteins in Escherichia Coli. J. Biotechnol. 2022, 344, 11–23. [Google Scholar] [CrossRef] [PubMed]
  108. Dudoladova, M.V.; Markelova, A.G.; Lebedeva, N.V.; Pronina, N.A. Compartmentation of α- and β-Carbonic Anhydrases in Cells of Halo- and Alkalophilic Cyanobacteria Rhabdoderma Lineare. Russ. J. Plant Physiol. 2004, 51, 806–814. [Google Scholar] [CrossRef]
  109. Dudoladova, M.V.; Kupriyanova, E.V.; Markelova, A.G.; Sinetova, M.P.; Allakhverdiev, S.I.; Pronina, N.A. The Thylakoid Carbonic Anhydrase Associated with Photosystem II Is the Component of Inorganic Carbon Accumulating System in Cells of Halo- and Alkaliphilic Cyanobacterium Rhabdoderma Lineare. Biochim. Biophys. Acta-Bioenerg. 2007, 1767, 616–623. [Google Scholar] [CrossRef]
Figure 1. Representation of CCM in C. reinhardtii showing the subcellular localization of the different CAs involved. CAH1, CAH2, and CAH3 belong to the α-class and are shown in red, while CAH7, CAH8, and CAH9 are β-CAs and are colored in black. CAH7 localization is not fully demonstrated. Pink rectangles represent bicarbonate transporters whereas blue circles symbolize low CO2-inducible proteins (β-CAs).
Figure 1. Representation of CCM in C. reinhardtii showing the subcellular localization of the different CAs involved. CAH1, CAH2, and CAH3 belong to the α-class and are shown in red, while CAH7, CAH8, and CAH9 are β-CAs and are colored in black. CAH7 localization is not fully demonstrated. Pink rectangles represent bicarbonate transporters whereas blue circles symbolize low CO2-inducible proteins (β-CAs).
Ijms 23 12045 g001
Figure 2. Cartoon representation of the CAH1 dimer. For both monomers, the big subunits are colored in magenta, while the small subunits in cyan. Disulfide bonds between the two monomers and between big and small subunits are reported in yellow. The zinc ions and their coordinating histidines are also shown in green.
Figure 2. Cartoon representation of the CAH1 dimer. For both monomers, the big subunits are colored in magenta, while the small subunits in cyan. Disulfide bonds between the two monomers and between big and small subunits are reported in yellow. The zinc ions and their coordinating histidines are also shown in green.
Ijms 23 12045 g002
Figure 3. Structure-based sequence alignment of hCA II with the green algal α-CAs with known structure: CAH1 and CAH3 from C. reinhardtii and dCAII from D. salina. Histidine proton shuttles are represented in bold, while zinc coordinating histidines and gatekeeper residues Thr199 and Glu106 (hCA II numbering) are indicated with (*) and (:), respectively. For CAH1, residues of the small subunit (345–377) are underlined and ♦ indicates the break between small and big subunit (residues 21–297). The α-helix regions are colored in red and β-strand regions in green. Secondary structural elements are calculated with the DSSP program [84,85].
Figure 3. Structure-based sequence alignment of hCA II with the green algal α-CAs with known structure: CAH1 and CAH3 from C. reinhardtii and dCAII from D. salina. Histidine proton shuttles are represented in bold, while zinc coordinating histidines and gatekeeper residues Thr199 and Glu106 (hCA II numbering) are indicated with (*) and (:), respectively. For CAH1, residues of the small subunit (345–377) are underlined and ♦ indicates the break between small and big subunit (residues 21–297). The α-helix regions are colored in red and β-strand regions in green. Secondary structural elements are calculated with the DSSP program [84,85].
Ijms 23 12045 g003
Figure 4. Cartoon representation of the CAH3 dimer. One subunit is colored in orange and the other one in yellow; the zinc ions and their coordinating histidines are also depicted.
Figure 4. Cartoon representation of the CAH3 dimer. One subunit is colored in orange and the other one in yellow; the zinc ions and their coordinating histidines are also depicted.
Ijms 23 12045 g004
Figure 5. Cartoon representation of the algal dCAII dimer. One subunit is colored in cyan and the other one in green; the zinc ion and its coordinating histidines are also depicted.
Figure 5. Cartoon representation of the algal dCAII dimer. One subunit is colored in cyan and the other one in green; the zinc ion and its coordinating histidines are also depicted.
Ijms 23 12045 g005
Figure 6. Proposed model for the N. oceanica CCM. The chloropast is separated from the cytoplasm by four membranes, the outermost of which, called the chloroplast ER (cER) membrane, is connected to the nuclear envelope and the ER membrane. Bicarbonate transporters push bicarbonate into the cytoplasm and then into the ER lumen, where NoCAH1 is localized. Here, this enzyme catalyzes the formation of CO2, which can either diffuse into the chloroplast stroma to be fixed by Rubisco in the Calvin–Benson–Bassham (CCB) cycle, or escape from the cell (dark blue and light blue rectangles represent bicarbonate transporters).
Figure 6. Proposed model for the N. oceanica CCM. The chloropast is separated from the cytoplasm by four membranes, the outermost of which, called the chloroplast ER (cER) membrane, is connected to the nuclear envelope and the ER membrane. Bicarbonate transporters push bicarbonate into the cytoplasm and then into the ER lumen, where NoCAH1 is localized. Here, this enzyme catalyzes the formation of CO2, which can either diffuse into the chloroplast stroma to be fixed by Rubisco in the Calvin–Benson–Bassham (CCB) cycle, or escape from the cell (dark blue and light blue rectangles represent bicarbonate transporters).
Ijms 23 12045 g006
Figure 7. Schematic picture of cyanobacterium cell, highlighting the presence of the carboxysome and thylakoid membranes within the cytosol. Subcellular localization of CAs is also shown: EcaA (α-CA) is extracellular, EcaB (β-CA) is primarily located in the thylakoid membrane with a smaller component associated with the plasma membrane, and β- and γ-CAs are localized in the carboxysome.
Figure 7. Schematic picture of cyanobacterium cell, highlighting the presence of the carboxysome and thylakoid membranes within the cytosol. Subcellular localization of CAs is also shown: EcaA (α-CA) is extracellular, EcaB (β-CA) is primarily located in the thylakoid membrane with a smaller component associated with the plasma membrane, and β- and γ-CAs are localized in the carboxysome.
Ijms 23 12045 g007
Figure 8. Multiple sequence alignment of Anabaena EcaA (aEcaA), Synechococcus EcaA (sEcaA), hCA I and hCA II. Conserved residues are indicated with an asterisk (*), while (:) and (.) indicate conservative and semi-conservative substitutions, respectively. Histidine residues involved in zinc coordination, Thr199 and Glu106 (hCA II numbering) are highlighted in yellow. Tat-signal peptides are underlined.
Figure 8. Multiple sequence alignment of Anabaena EcaA (aEcaA), Synechococcus EcaA (sEcaA), hCA I and hCA II. Conserved residues are indicated with an asterisk (*), while (:) and (.) indicate conservative and semi-conservative substitutions, respectively. Histidine residues involved in zinc coordination, Thr199 and Glu106 (hCA II numbering) are highlighted in yellow. Tat-signal peptides are underlined.
Ijms 23 12045 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Langella, E.; Di Fiore, A.; Alterio, V.; Monti, S.M.; De Simone, G.; D’Ambrosio, K. α-CAs from Photosynthetic Organisms. Int. J. Mol. Sci. 2022, 23, 12045. https://doi.org/10.3390/ijms231912045

AMA Style

Langella E, Di Fiore A, Alterio V, Monti SM, De Simone G, D’Ambrosio K. α-CAs from Photosynthetic Organisms. International Journal of Molecular Sciences. 2022; 23(19):12045. https://doi.org/10.3390/ijms231912045

Chicago/Turabian Style

Langella, Emma, Anna Di Fiore, Vincenzo Alterio, Simona Maria Monti, Giuseppina De Simone, and Katia D’Ambrosio. 2022. "α-CAs from Photosynthetic Organisms" International Journal of Molecular Sciences 23, no. 19: 12045. https://doi.org/10.3390/ijms231912045

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