Next Article in Journal
Binding of CML-Modified as Well as Heat-Glycated β-lactoglobulin to Receptors for AGEs Is Determined by Charge and Hydrophobicity
Next Article in Special Issue
Changes in Expression Level of OsHKT1;5 Alters Activity of Membrane Transporters Involved in K+ and Ca2+ Acquisition and Homeostasis in Salinized Rice Roots
Previous Article in Journal
Neuropathophysiology, Genetic Profile, and Clinical Manifestation of Mucolipidosis IV—A Review and Case Series
Previous Article in Special Issue
Function and Regulation of Ammonium Transporters in Plants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Diverse Physiological Functions of Cation Proton Antiporters across Bacteria and Plant Cells

Department of Biomolecular Engineering, Graduate School of Engineering, Tohoku University, Aobayama 6-6-07, Sendai 980-8579, Japan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2020, 21(12), 4566; https://doi.org/10.3390/ijms21124566
Submission received: 1 June 2020 / Revised: 22 June 2020 / Accepted: 24 June 2020 / Published: 26 June 2020
(This article belongs to the Special Issue Ion Transport and Homeostasis in Plants)

Abstract

:
Membrane intrinsic transport systems play an important role in maintaining ion and pH homeostasis and forming the proton motive force in the cytoplasm and cell organelles. In most organisms, cation/proton antiporters (CPAs) mediate the exchange of K+, Na+ and Ca2+ for H+ across the membrane in response to a variety of environmental stimuli. The tertiary structure of the ion selective filter and the regulatory domains of Escherichia coli CPAs have been determined and a molecular mechanism of cation exchange has been proposed. Due to symbiogenesis, CPAs localized in mitochondria and chloroplasts of eukaryotic cells resemble prokaryotic CPAs. CPAs primarily contribute to keeping cytoplasmic Na+ concentrations low and controlling pH, which promotes the detoxification of electrophiles and formation of proton motive force across the membrane. CPAs in cyanobacteria and chloroplasts are regulators of photosynthesis and are essential for adaptation to high light or osmotic stress. CPAs in organellar membranes and in the plasma membrane also participate in various intracellular signal transduction pathways. This review discusses recent advances in our understanding of the role of CPAs in cyanobacteria and plant cells.

1. Introduction

The flux of cations and protons across membranes is a key process in maintaining intracellular ion homeostasis in various environments. A large number of ion transporters in the plasma membrane and in organellar membranes are involved in the regulation of ion translocation. They mediate the flux of potassium ions (K+), sodium ions (Na+) and protons (H+), common cations in cells and in the environment. The concentration of K+ in cells is estimated to be around 0.1-0.2 M, and it is the most prevalent cation in living cells [1,2]. K+ is required for many biological processes such as the formation of the membrane potential, osmotic adjustment and turgor pressure [3,4,5,6]. K+ is also needed for the proper activation of certain enzymes, for instance pyruvate kinase [7]. In contrast, the Na+ concentration in cells is usually only 5-14 mM. Excess Na+ often has negative effects on terrestrial plants and some bacteria [8,9,10,11] since a Na+/K+-ATPase is missing in bacteria and plants [12]. Excess Na+ induces dehydration and reduces the intracellular K+ content, resulting in fatal cell damage.
The exchange of K+ or Na+ with H+ across cell membranes and organelles is coordinated by a variety of ion transporters. Among them, the cation/proton antiporter (CPA) superfamily is one of the larger groups in the superfamily of ion transporters in a variety of organisms [13]. Transport direction and localization of CPAs are diverse and CPAs therefore play a range of physiological roles. For example, some CPAs respond to salt stress by exporting excess cations from the cytoplasm into the extracellular space or into organelles [14,15,16]. Some CPAs are involved in the transport of proteins into endosomal compartments [17,18].
Plants contain a large number of conserved CPAs [19], which are localized in the plasma membrane and in the membranes of organelles such as vacuoles, endosomes and chloroplasts. Because mitochondria and chloroplasts evolved from prokaryotes through symbiogenesis, understanding the physiological role of prokaryotic CPAs is important to elucidate their role in plants. In this study, we focus on the fundamental structure and physiological roles of CPAs in Escherichia coli, cyanobacteria and plant cells to evaluate their similarities and differences.

2. Structure of CPAs

While numerous studies have shown the physiological importance of cation/H+ antiporters, their structure and transport activity remain poorly understood. Direct measurement of cation/H+ exchange activity by CPAs is usually conducted using inverted membranes of E. coli, organelles of eukaryotic cells or liposomes [15,20,21,22,23] (Figure 1A). Fundamental understanding of CPAs in living cells comes from crystallography studies of the E. coli transporter NhaA (henceforth NhaA). NhaA, an electrogenic pH-dependent Na+(Li+)/H+ antiporter [24], is a member of the CPA superfamily, from the NHA clade of CPA2 [13]. To date, two different crystal structures of NhaA have been reported, one in monomeric [25] and one in dimeric configuration [26]. Both structures have been determined under low pH conditions (at pH 4 and 3.5), and therefore they correspond to the inactive state of NhaA [25,26]. NhaA is composed of 12 transmembrane segments (TMs) with N and C terminus located in the cytoplasm [25]. The monomeric crystal structure of NhaA reveals a unique structural conformation consisting of TM4 and TM11 [25], designated as the NhaA fold [27]. These two transmembrane regions contain the ion-binding site near their intersection [27,28]. The NhaA fold is critical for antiporter activity by creating an electrostatic balance in close proximity to the substrate binding site [25,28]. The Na+/Li+ binding site is composed of two essential aspartic acid residues (D163 and D164) in TM5, together with D133 and T132 in TM4 (Figure 1B) [25], which was confirmed by isothermal calorimeter experiments [29]. The crystal structure also reveals a putative ‘pH sensor’ region, a cluster of ionizable residues located at the opening of the cytoplasmic funnel (N-terminal region of TM9) [25]. Mutations affecting amino acid residues in the pH sensor region and in the TM8-9 loop significantly change the pH profile of the antiporter. Increasing pH induces conformational changes (resulting in movement of the TM4 and TM11 assembly) that expose the Na+ binding site and eventually activate NhaA [28].
The stoichiometry of NhaA is 1Na+/2H+ but the exact mechanisms of the antiport activity is still under debate, due to unavailability of an active-state structure. In particular, the identity of the second H+ donor remains unclear. Both D163 and D164 have been proposed as the first and second H+ donors in the NhaA antiporter cycle [25]. However, the second, dimeric NhaA crystal structure reassigns TM10, which changes the position of K300 to be within salt bridge distance to D163, which suggests an alternative hypothesis of K300 as the second H+ donor (Figure 1C) [26]. Furthermore, the constant pH molecular dynamics simulations (CpHMD) suggest that Na+ binding to D163 disrupts the salt bridge between D163 and K300, destabilizes the charged state of K300 and sparks the discharge of H+ [30]. Consequently, CpHMD analysis supports the proposed mechanism of K300 as the second H+ donor [30]. Interestingly, a recent report shows that replacement of both salt bridge residues (D163N/K300Q) does not abolish electrogenic transport activity despite the absence of a salt bridge near the ion binding site. The authors concluded that NhaA might have several alternative mechanisms to ensure the electrogenicity of the Na+/H+ exchange and that NhaA is able to utilize an alternative second H+ donor when the main second H+ donor is unavailable [31]. Therefore, efforts to solve the NhaA active structure are still ongoing.
Regardless of differences in amino acid sequence and length, some other members of CPA have a similar structural conformation to NhaA, such as Thermus thermophilus NapA (TtNapA), Neisseria meningitidis ABST (NmABST), Pyrococcus abyssi NhaP (PaNhaP), and Methanocaldococcus jannaschii NhaP1 (MjNhaP1) [27]. A broad phylogenic analysis of 6000 CPAs found a common CPA motif, consisting of eight non-linear amino acids that might be responsible for electrogenicity and ion selectivity of the CPA superfamily [32]. The CPA motif is located in the vicinity of the ion binding site [32]. The electrogenicity of the transport activity of an electrogenic CPA is usually characterized by triplet residues (two acidic residues corresponding to D163 and D164 at NhaA and a basic residue corresponding to K300 at NhaA) [32].

3. Physiological Functions of CPAs in E. coli

3.1. CPAs as Na+ Extrusion Systems

Excess Na+ has detrimental effects on E. coli growth. E. coli contains eight CPAs (Table 1 and, Figure 2A). Of those, three Na+ extrusion systems, NhaA, NhaB and ChaA, help to maintain a low concentration of Na+ in the cytoplasm. NhaB (formerly AntB) was discovered by analyzing the remaining antiporter activity in an nhaA deletion mutant [33]. NhaA and NhaB share relatively low homology of their amino acid sequence (22.4% identity and 35.6% similarity) [34]. NhaB has distinct characteristics from NhaA in terms of weak pH-dependency and lower Li+ affinity (approximately 15 times lower than NhaA) [33]. The third system, ChaA, was initially thought to be a Ca2+/H+ antiporter in E. coli. However, a follow-up study shows that ChaA has Na+ extrusion activity that is able to complement E. coli EP432 (nhaA nhaB) [35]. An E. coli mutant lacking all three of these Na+ extrusion systems lacks most Na+ extrusion ability, evidenced by its sensitivity to 0.1 M NaCl at pH 8.5 [36]. NhaA, NhaB and ChaA prevent Na+ toxicity under physiological conditions.
Why does E. coli have multiple Na+ extrusion systems (Figure 2A)? The answer probably lies in the different optimal pH conditions for each transporter. NhaA is strongly pH dependent and is only active at pH between 6.5 and 8.5, NhaB functions at pH below 8 and ChaA is active above pH 8 [36]. The mechanism leading to pH activation of NhaB and ChaA remains to be determined. Another characteristic differing between NhaA, NhaB and ChaA is their substrate specificity. NhaA and NhaB transport Li+ in addition to Na+, which promotes Li+ detoxification [37]. ChaA has a broader range of substrates: Na+, Ca+, and K+. Although ChaA has Ca+/H+ antiport capacity, ChaA does not play a role in Ca2+ extrusion in vivo [38], but rather is involved in protection against K+ salinity stress [39].

3.2. CPAs as pH Regulators at Alkaline pH

The extrusion of Na+ from cells via CPAs is accompanied by H+ uptake, which results in cytoplasmic acidification. E. coli is a neutralophilic bacterium which needs to maintain an intracellular pH of around 7.4–7.8 over a wide range of external pH from 5.0 to 9.0 [40]. At alkaline pH conditions, pH homeostasis can be achieved through the electrogenic nature of NhaA and NhaB transport. NhaA transports 2H+ per 1Na+ exported [41], whereas NhaB transports 3H+ per 2Na+ exported [42]. Therefore, H+ accumulation leads to an acidification of the cytoplasm at alkaline pH in the presence of Na+. Besides those three Na+/H+ transporters, two other transport systems also assist in maintaining pH homeostasis in E. coli: MdtM and MdfA. Both transporters are multidrug-resistance transporters (Mdr transporter), that mainly function as multi-drug/H+ antiporters but also have cation/H+ antiport activity. Lewison et al. [43] found that MdfA also mediates Na+(K+)/H+ antiport, which contributes to alkali tolerance in E. coli at extremely high pH (9.5–10). In addition, MdtM, an MdfA homolog, also exhibits Na+ (K+, Rb+, Li+)/H+ antiport capacity [44]. Both MdtM and MdfA are able to extend the pH tolerance of E. coli up to pH 10 in the presence of Na+ or K+. Since ChaA, MdtM and MdfA have K+/H+ antiport capacity, they are able to conduct H+ influx at higher alkaline pH conditions than Na+(Li+)-specific antiporters (NhaA and NhaB). Their ability to utilize K+ as a driving force is a critical advantage in adaptation to extreme alkaline conditions, since E. coli accumulates high K+ concentration intracellularly [45].

3.3. CPAs as Antimicrobial Stress Mechanisms

Another physiological function of CPAs in E. coli is antimicrobial resistance. This function can be further divided into two categories: direct and indirect. CPAs, like MdtM and MdfA, directly contribute to antimicrobial resistance by actively extruding a broad range of toxic compounds such as antibiotics, mutagen and xenobiotics: chloramphenicol, ethidium bromide, various quaternary ammonium compounds and unconjugated bile salt by MtdM [46,47,48], zwitterionic lipophilic compounds, certain aminoglycosides and fluoroquinolones by MdfA [49].
E. coli takes up K+ through Kup, Kdp, TrkG and TrkH [50], but effluxes K+ via KefB and KefC under electrophilic stress conditions. KefB and KefC indirectly contribute to electrophilic stress tolerance as they transport H+ in exchange with K+ to acidify the cytoplasm which in turn alleviates the stress [51,52]. They are glutathione-gated K+ efflux systems in E. coli; each efflux system requires an ancillary protein: KefG (formerly YabF) for KefB and KefF (formerly YheR) for KefC, respectively, to stabilize their conformation and achieve maximum activity [53]. KefC consists of a transmembrane domain and a cytosolic KTN (K+ transport nucleotide binding) domain. The KTN domain is involved in dimer formation, and binding of glutathione to its cleft inhibits KefC activity [54,55]. When cells are exposed to electrophiles (e.g., methylglyoxal, N-ethylmaleimide and chlorodinitrobenzene), these react with glutathione to form glutathione conjugates which in turn activate KefB and KefC-mediated K+/H+ antiport transport activity [52].
YcgO, also known as CvrA, contributes to cell volume regulation under low osmolarity conditions [56]. YcgO functions as K+ efflux system in a mutant lacking PtsN (EIIANtr) [57]. YgcO activity is regulated by the phosphorylation state of PtsN, a member of the nitrogen phosphor transferase system (PTSNtr). It is therefore possible that certain stresses, that shift the balance toward phospho-PtsN, activate K+ efflux via YgcO and induce growth cessation [57].

4. Physiological Function of CPAs in Cyanobacteria

Cyanobacteria, the ancestors of plant chloroplasts, need to control the Na+ content in their cytoplasm. Salt stress decreases cell volume and stimulates the inactivation of photosystem I (PSI) and photosystem II (PSII) [58]. Na+/H+ antiporters are thought to be the conduit of Na+ influx and efflux in cyanobacteria. Synechocystis sp. PCC6803 (hereafter Synechocystis), the model strain of cyanobacteria, possesses six Na+/H+ antiporters (NhaS1-NhaS6) (Table 2 and Figure 2B) [16,20,59,60,61,62]. Among them, only NhaS1 and NhaS3 were shown to be functional Na+/H+ antiporters when expressed in E. coli [20,59], whereas the activity of other antiporters remains elusive. A triple mutant of NhaS1/4/5 and a single mutant of NhaS1/2/4/5 do not show any significant difference compared with the wild type, suggesting that these transporters are not involved in salinity stress adaptation [16,60]. On the other hand, ΔnhaS3 has a salt-sensitive phenotype, suggesting that NhaS3 confers salt tolerance [16,20]. NhaS3 transports Na+ from the cytosol to the thylakoid lumen to keep concentrations of Na+ low in Synechocystis [20]. Freshwater cyanobacterium, Synechococcus sp. strain PCC7942 (hereafter Synechococcus) possesses seven Na+/H+ antiporters (Table 2). Synechococcus NhaS3 alleviates salinity stress in a pH-dependent manner [63], similar to the role of Synechocystis NhaS3. Loss of nhaS2nhaS2) reduces growth in media containing low Na+ concentrations [61] and deletion of nhaS4 also results in hypersensitivity to low Na+ conditions [16]. While ΔnhaS2 requires 250 mM NaCl in the medium to reach growth that is similar to the wild type, ΔnhaS4 only requires 20 mM NaCl. NhaS2 and NhaS4 act as Na+ influx transporters with different affinities for Na+.
In eukaryotes, Ca2+ is an essential cation, however, the physiological role of Ca2+ in prokaryotes is largely unknown. Excess Ca2+ concentrations in the cytoplasm are toxic for both eukaryotes and prokaryotes. Synechocystis possesses a Ca2+/H+ antiporter in the plasma membrane, SynCax [64]. Deletion of synCax (Δsyncax) reduces Ca2+ efflux activity. In addition, the mutant has higher intracellular Na+ content and reduced growth in high salt medium. SynCax-mediated Ca2+ efflux across the cell membrane is related to adaptation to salinity stress.

5. Physiological Functions of CPAs in Plants

5.1. CPAs in the Plasma Membrane Are Required for Salt Extrusion and Nutrient Uptake

Plant CPAs are classified into eight Na+/H+ exchangers (NHX), 28 cation/H+ exchangers (CHX), six K+ efflux antiporters (KEA) and two sodium hydrogen antiporter (NHAD)-type carriers and one Ca2+/H+ exchanger (Table 3 and Figure 2C) [19]. In plant cells, CPAs localized in the cell membrane expel excess cations from the cell or replenish cations into the cell. Arabidopsis thaliana possesses a salt extrusion system of Na+/H+ antiporters, such as SOS1 (NHX7) [14]. The SOS (Salt Overly Sensitive) pathway consists of SOS1, SOS2, a CBL-interacting protein kinase (also called CIPK24), and SOS3, a calcineurin B-like protein (also called CBL4) [65]. When plant cells experience salt stress, the intracellular Ca2+ levels transiently increase. The binding of Ca2+ to SOS3 activates SOS2, resulting in phosphorylation of SOS1 and Na+ extrusion from the cytosol. Arabidopsis ∆sos1 accumulates Na+ in the cell, which inhibits the K+ uptake channel AKT1, leading to a decrease in the intracellular K+ concentration [66]. The Ca2+ for SOS activation may be replenished through a Ca2+ influx channel regulated by a glucuronosyltransferase (MOCA1) for glycosyl inositol phosphorylceramide (GIPC) sphingolipids in the plasma membrane [67]. NHX8 is also localized to the plasma membrane but is not involved in NaCl tolerance [68]. Growth of Arabidopsis ∆nhx8 is strongly reduced in the presence of 10 mM LiCl. This Li+ sensitivity is also seen in ∆sos1, and overexpression of NHX8 rescues the Li+ sensitivity of ∆sos1. Under high Li+ conditions, plants use SOS1 and NHX8 to extrude Li+ out of the cells.
Both CHX13 and CHX14 function as K+/H+ antiporters in the cell membrane, but their roles are different [69,70]. CHX13 has K+ uptake activity and therefore ∆chx13 shows reduced growth and pale yellow leaves when grown on low K+ medium. On the other hand, CHX14 has K+ efflux activity and ∆chx14 shows reduced root elongation and increased K+ content in root cells when exposed to excess K+.

5.2. Vacuolar CPAs Are Important for Cell Turgor Pressure and Storage of Intracellular K+

NHX1–NHX4 are localized in the vacuolar membrane and are involved in the regulation of cell turgor pressure, intracellular K+ accumulation and cytoplasmic Na+ homeostasis. NHX1 and NHX2 are specific for K+ over Na+ in their cation transport activity [71,72]. An Arabidopsis ∆nhx1 mutant shows reduced growth [71], while ∆nhx2 exhibits the same growth phenotype as the wild type. However, growth of the ∆nhx1nhx2 double mutant is more reduced than that of the ∆nhx1 single mutant. The ∆nhx1nhx2 double mutant also has smaller cells and immature development of embryos and stamen. NHX1-mediated K+ and Na+ transport activity is regulated by calmodulin (CaM)-like protein 15 (CAM15) [73]. CAM15 binds to the C-terminus of NHX1 inside the vacuole and significantly reduces Na+ transport activity of NHX1. CAM15 binds to NHX1 in a pH-dependent manner and fails to bind NHX1 under alkaline conditions. When plants are exposed to salt stress, the pH of the vacuole lumen increases [74]. Under non-stress conditions, NHX1 interacting with CAM15 may lead to swelling of the cell due to the influx of K+ into the vacuole. When the plant is exposed to salt stress, CAM15 separates from NHX1 and then NHX1 promotes Na+ transport into the vacuole.
NHX3 and NHX4 function differs from NHX1 and NHX2. NHX3 has a role in K+ accumulation in cells [75,76] and the expression of Arabidopsis NHX3 in sugar beets increases salt stress tolerance [75]. NHX3-expressing sugar beets contain lower Na+ and higher K+ concentrations. In the same plants, K+/H+ antiporter activity in vacuolar membranes increases, but Na+/H+ antiporter activity does not. In Arabidopsis, NHX3 is expressed in germinating seeds and reproductive tissues [76]. Arabidopsis ∆nhx3 is sensitive to low K+ concentrations during germination, therefore NHX3 is likely involved in K+ uptake during seed germination. Tonoplast-localized NHX4 has Na+ transport activity, and translocates Na+ from the vacuole into the cytoplasm [77].

5.3. Regulation of Photosynthesis by Multiple CPAs

In chloroplasts, the light reaction leads to proton accumulation in the thylakoid lumen, resulting in the generation of a pH gradient component (∆pH) and membrane potential (∆Ψ). ATP production is driven by H+-ATPase using the proton motive force, which is the sum of ∆pH and ∆Ψ. Excessive reduction elicited by high light energy causes overproduction of reactive oxygen species, which hamper the photochemical system. To avoid this damage, non-photochemical quenching (NPQ) dissipates the collected light energy [78]. NPQ is triggered by acidification of the thylakoid lumen. Under low light conditions, NPQ is inhibited. KEA3 is a thylakoid lumen-localized K+/H+ antiporter. KEA3-mediated proton translocation from lumen to stroma inhibits NPQ [21,79,80,81,82,83]. NPQ is strongly suppressed in transgenic Arabidopsis with enhanced KEA3 transport activity [21,82]. Under constant light conditions the NPQ level of Arabidopsis ∆kea3 is similar to the wild type, but the NPQ level of ∆kea3 is increased under fluctuating light conditions [79,80]. KEA3 has a KTN domain in its C-terminal region and overexpression of a splicing variant without this domain represses NPQ. Interestingly, hyperosmotic stress does not lead to increased cytosolic Ca2+ concentrations in ∆kea3, suggesting that KEA3 is involved in Ca2+ release from the thylakoid lumen [84].
KEA1 and KEA2 function as regulators of photosynthesis in the envelope membrane of plastids [79,85]. While ∆kea1 and ∆kea2 do not show any changes in growth and photosynthesis, the double mutant ∆kea1kea2 shows impaired growth with pale leaves and reduced NPQ in high light. KEA1 and KEA2 contribute to maintaining K+ and H+ homeostasis in the chloroplasts, leading to correct induction of NPQ. NHD1, a sodium hydrogen antiporter (NHAD)-type carrier, also localizes to the chloroplast envelope [86]. NHD1 excludes Na+ from the chloroplast into the cytosol, protecting photosynthetic activity from salinity stress. In Fraveria bidentis, NHD1 establishes a Na+ gradient across the chloroplast envelope to activate Na+-dependent pyruvate uptake into the chloroplast via bile acid:sodium symporter family protein 2 (BASS2) [87].
The thylakoid lumen in the chloroplast is known to function as a Ca2+ store for stress responses [88]. In addition, the oxygen evolving complex (OEC) in the thylakoid membrane requires Ca2+. Ca2+ transport across the thylakoid membrane is crucial for chloroplasts, however the transport mechanism is unknown. The chloroplast-localized Ca2+/H+ antiporter CCHA1 is a potential candidate for this Ca2+ transport system into the thylakoid lumen [89,90]. The ccha1 mutation in Arabidopsis decreases photosynthetic activity and oxygen generation rate and alters the cytoplasmic Ca2+ concentration and the pH in guard cells. CCHA1 likely regulates photosynthetic activity and stomatal aperture by exchanging Ca2+ and H+ across the thylakoid membrane.

5.4. Role of Endosome-Localized CPAs

Various types of CPAs are localized in endosomal membranes, functioning in regulating the internal ion concentration and pH. NHX5 and NHX6 participate in control of the endosomal ion environment [18,91,92]. The acidic pH in the trans-Golgi network (TGN), prevacuolar compartment (PVC) and vacuolar sorting receptors (VSRs) in ∆nhx5∆nhx6 suggests that NHX5 and NHX6 transport K+ into endosomes and release H+ into the cytoplasm [91]. NHX6 is expressed in primary root tips and lateral root primordia. Delayed growth of lateral root primordia occurs in ∆nhx5∆nhx6 [92]. Moreover, transport activity of the auxin efflux PIN FORMED (PIN) family, which is required for correct auxin distribution, is inhibited in the plasma membrane of ∆nhx5∆nhx6. NHX5 and NHX6 therefore promote salt extrusion and auxin distribution.
CHX17 is localized in the PVC [93], and its C-terminal region is important for this localization [94]. Expression of CHX17 is enhanced by salt stress, K+ starvation, low pH and ABA treatment. Salt stress decreases K+ content in ∆chx17. CHX17/18/19 localize to the PVC, and CHX20 localizes to both PVC and vacuolar membrane [17]. CHX17 and CHX20 are involved in protein sorting. CHX23 localizes to the endoplasmic reticulum (ER) in pollen tubes [95]. Simultaneous deletion of CHX23 and plasma membrane-localized CHX21 reduces pollen tube guidance to ovules.
KEA4/5/6 is expressed in the Golgi and in endosomes in all plant tissues [96,97]. A triple mutant of ∆kea4/5/6 has reduced growth, and is highly sensitive to K+ starvation and salinity stress [96]. The ∆kea4/5/6 triple mutant has acidic pH in organelles, similar to ∆nhx5 and ∆nhx6. Expression of NHX5 or NHX6 rescues the salt stress sensitivity of ∆kea4/5/6, whereas expression of the plasma membrane transporters NHX1 and NHX2 cannot rescue the mutant. The function of KEA4/5/6 is similar to that of NHX5/6 [97].

6. Conclusions

Numerous reports show the structure and function of CPAs and physiological importance of the exchange of cations and H+ across membranes in living cells. The transport of cations and H+ by CPAs directly controls Na+/K+ and proton concentrations in the cytosol, and indirectly supports photosynthesis, protein processing and resistance against toxins. Further studies are needed to fully understand the transport mechanism of CPAs and their physiological role in cellular signaling pathways that contribute to improved plant tolerance to abiotic and biotic stress.

Author Contributions

E.T., M.T. and N.U. wrote the manuscript. All authors have read and agree to the published version of the manuscript.

Funding

This work was supported by JSPS KAKENHI Grant Number (16H06558, 18H03762, 19H02880, 19K22264 and 19J10317).

Acknowledgments

We thank Anke Reinders for critical reading of the manuscript.

Conflicts of Interest

The authors declare that there is no conflict of interest.

Abbreviations

CPAcation/proton antiporter
CpHMDconstant pH molecular dynamics simulations
KTNK+ transport nucleotide binding
NHXNa+/H+ exchanger
CHXcation/H+ exchanger
KEAK+ efflux antiporters
NPQnon-photochemical quenching
OECoxygen evolving complex
TGNtrans-Golgi network
PVCprevacuolar compartment
VSRvacuolar sorting receptor
PINPIN FORMED
ERendoplasmic reticulum

References

  1. Meury, J.; Kepes, A. The regulation of potassium fluxes for the adjustment and maintenance of potassium levels in Escherichia coli. Eur. J. Biochem. 1981, 119, 165–170. [Google Scholar] [CrossRef]
  2. Dreyer, I.; Uozumi, N. Potassium channels in plant cells. FEBS J. 2011, 278, 4293–4303. [Google Scholar] [CrossRef]
  3. Etherton, B.; Higinbotham, N. Transmembrane potential measurements of cells of higher plants as related to salt uptake. Science 1960, 131, 409–410. [Google Scholar] [CrossRef]
  4. Cheeseman, J.M.; Hanson, J.B. Mathematical analysis of the dependence of cell potential on external potassium in corn roots. Plant Physiol. 1979, 63, 1–4. [Google Scholar] [CrossRef] [Green Version]
  5. Maathuis, F.J.M.; Sanders, D. Energization of potassium uptake in Arabidopsis thaliana. Planta 1993, 191, 302–307. [Google Scholar] [CrossRef]
  6. Hasanuzzaman, M.; Bhuyan, M.H.M.B.; Nahar, K.; Hossain, M.S.; Al Mahmud, J.; Hossen, M.S.; Masud, A.A.C.; Moumita; Fujita, M. Potassium: A vital regulator of plant responses and tolerance to abiotic stresses. Agronomy 2018, 8, 31. [Google Scholar] [CrossRef] [Green Version]
  7. Gohara, D.W.; Di Cera, E. Molecular mechanisms of enzyme activation by monovalent cations. J. Biol. Chem. 2016, 291, 20840–20848. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Kronzucker, H.J.; Coskun, D.; Schulze, L.M.; Wong, J.R.; Britto, D.T. Sodium as nutrient and toxicant. Plant Soil 2013, 369, 1–23. [Google Scholar] [CrossRef] [Green Version]
  9. Isayenkov, S.V.; Maathuis, F.J.M. Plant salinity stress: Many unanswered questions remain. Front. Plant Sci. 2019, 10, 80. [Google Scholar] [CrossRef] [Green Version]
  10. Uozumi, N. Uniquely evolved plant ion channels. FEBS J. 2011, 278, 4261. [Google Scholar] [CrossRef]
  11. Lo, C.J.; Leake, M.C.; Berry, R.M. Fluorescence measurement of intracellular sodium concentration in single Escherichia coli cells. Biophys. J. 2006, 90, 357–365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Uozumi, N.; Dreyer, I. Structure-function correlates in plant ion channels. In Comprehensive Biophysics; Academic Press: Cambridge, MA, USA, 2012; Volume 6, pp. 234–245. [Google Scholar]
  13. Brett, C.L.; Donowitz, M.; Rao, R. Evolutionary origins of eukaryotic sodium/proton exchangers. Am. J. Physiol. Cell Physiol. 2005, 288, C223–C229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Wu, S.J.; Ding, L.; Zhu, J.K. SOS1, a genetic locus essential for salt tolerance and potassium acquisition. Plant Cell 1996, 8, 617–627. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Schuldiner, S.; Fishkes, H. Sodium-proton antiport in isolated membrane vesicles of Escherichia coli. Biochemistry 1978, 17, 706–711. [Google Scholar] [CrossRef] [PubMed]
  16. Wang, H.L.; Postier, B.L.; Burnap, R.L. Polymerase chain reaction-based mutageneses identify key transporters belonging to multigene families involved in Na+ and ph homeostasis of Synechocystis sp. PCC 6803. Mol. Microbiol. 2002, 44, 1493–1506. [Google Scholar] [CrossRef] [Green Version]
  17. Chanroj, S.; Lu, Y.; Padmanaban, S.; Nanatani, K.; Uozumi, N.; Rao, R.; Sze, H. Plant-specific cation/H+ exchanger 17 and its homologs are endomembrane K+ transporters with roles in protein sorting. J. Biol. Chem. 2011, 286, 33931–33941. [Google Scholar] [CrossRef] [Green Version]
  18. Bassil, E.; Ohto, M.A.; Esumi, T.; Tajima, H.; Zhu, Z.; Cagnac, O.; Belmonte, M.; Peleg, Z.; Yamaguchi, T.; Blumwald, E. The Arabidopsis intracellular Na+ /H+ antiporters NHX5 and NHX6 are endosome associated and necessary for plant growth and development. Plant Cell 2011, 23, 224–239. [Google Scholar] [CrossRef] [Green Version]
  19. Chanroj, S.; Wang, G.; Venema, K.; Zhang, M.W.; Delwiche, C.F.; Sze, H. Conserved and diversified gene families of monovalent cation/H+ antiporters from algae to flowering plants. Front. Plant Sci. 2012, 3, 25. [Google Scholar] [CrossRef] [Green Version]
  20. Tsunekawa, K.; Shijuku, T.; Hayashimoto, M.; Kojima, Y.; Onai, K.; Morishita, M.; Ishiura, M.; Kuroda, T.; Nakamura, T.; Kobayashi, H.; et al. Identification and characterization of the Na+/H+ antiporter NhaS3 from the thylakoid membrane of Synechocystis sp. PCC 6803. J. Biol. Chem. 2009, 284, 16513–16521. [Google Scholar] [CrossRef] [Green Version]
  21. Tsujii, M.; Kera, K.; Hamamoto, S.; Kuromori, T.; Shikanai, T.; Uozumi, N. Evidence for potassium transport activity of Arabidopsis KEA1-KEA6. Sci. Rep. 2019, 9, 10040. [Google Scholar] [CrossRef]
  22. Venema, K.; Quintero, F.J.; Pardo, J.M.; Donaire, J.P. The Arabidopsis Na+/H+ exchanger AtNHX1 catalyzes low affinity Na+ and K+ transport in reconstituted liposomes. J. Biol. Chem. 2002, 277, 2413–2418. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Aranda-Sicilia, M.N.; Cagnac, O.; Chanroj, S.; Sze, H.; Rodríguez-Rosales, M.P.; Venema, K. Arabidopsis KEA2, a homolog of bacterial KefC, encodes a K+/H+ antiporter with a chloroplast transit peptide. Biochim. Biophys. Acta Biomembr. 2012, 1818, 2362–2371. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Taglicht, D.; Padan, E.; Schuldiner, S. Overproduction and purification of a functional Na+/H+ antiporter coded by nhaA (ant) from Escherichia coli. J. Biol. Chem. 1991, 266, 11289–11294. [Google Scholar] [PubMed]
  25. Hunte, C.; Screpanti, E.; Venturi, M.; Rimon, A.; Padan, E.; Michel, H. Structure of a Na+/H+ antiporter and insights into mechanism of action and regulation by pH. Nature 2005, 435, 1197–1202. [Google Scholar] [CrossRef] [PubMed]
  26. Lee, C.; Yashiro, S.; Dotson, D.L.; Uzdavinys, P.; Iwata, S.; Sansom, M.S.P.; von Ballmoos, C.; Beckstein, O.; Drew, D.; Cameron, A.D. Crystal structure of the sodium-proton antiporter NhaA dimer and new mechanistic insights. J. Gen. Physiol. 2014, 144, 529–544. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Padan, E.; Michel, H. NhaA: A unique structural fold of secondary active transporters. Isr. J. Chem. 2015, 55, 1233–1239. [Google Scholar] [CrossRef]
  28. Padan, E.; Kozachkov, L.; Herz, K.; Rimon, A. NhaA crystal structure: Functional-structural insights. J. Exp. Biol. 2009, 212, 1593–1603. [Google Scholar] [CrossRef] [Green Version]
  29. Maes, M.; Rimon, A.; Kozachkov-Magrisso, L.; Friedler, A.; Padan, E. Revealing the ligand binding site of NhaA Na+/H+ antiporter and its pH dependence. J. Biol. Chem. 2012, 287, 38150–38157. [Google Scholar] [CrossRef] [Green Version]
  30. Huang, Y.; Chen, W.; Dotson, D.L.; Beckstein, O.; Shen, J. Mechanism of pH-dependent activation of the sodium-proton antiporter NhaA. Nat. Commun. 2016, 7, 12940. [Google Scholar] [CrossRef]
  31. Patiño-Ruiz, M.; Dwivedi, M.; Cǎlinescu, O.; Karabel, M.; Padan, E.; Fendler, K. Replacement of Lys-300 with a glutamine in the NhaA Na+/H+ antiporter of Escherichia coli yields a functional electrogenic transporter. J. Biol. Chem. 2019, 294, 246–256. [Google Scholar] [CrossRef] [Green Version]
  32. Masrati, G.; Dwivedi, M.; Rimon, A.; Gluck-Margolin, Y.; Kessel, A.; Ashkenazy, H.; Mayrose, I.; Padan, E.; Ben-Tal, N. Broad phylogenetic analysis of cation/proton antiporters reveals transport determinants. Nat. Commun. 2018, 9, 4205. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Padan, E.; Maisler, N.; Taglicht, D.; Karpel, R.; Schuldiner, S. Deletion of ant in Escherichia coli reveals its function in adaptation to high salinity and an alternative Na+/H+ antiporter system(s). J. Biol. Chem. 1989, 264, 20297–20302. [Google Scholar] [PubMed]
  34. Madeira, F.; Park, Y.M.; Lee, J.; Buso, N.; Gur, T.; Madhusoodanan, N.; Basutkar, P.; Tivey, A.R.N.; Potter, S.C.; Finn, R.D.; et al. The EMBL-EBI search and sequence analysis tools APIs in 2019. Nucleic Acids Res. 2019, 47, W636–W641. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Ivey, D.M.; Guffanti, A.A.; Zemsky, J.; Pinner, E.; Karpel, R.; Padan, E.; Schuldiner, S.; Krulwich, T.A. Cloning and characterization of a putative Ca2+/H+ antiporter gene from Escherichia coli upon functional complementation of Na+/H+ antiporter- deficient strains by the overexpressed gene. J. Biol. Chem. 1993, 268, 11296–11303. [Google Scholar] [PubMed]
  36. Sakuma, T.; Yamada, N.; Saito, H.; Kakegawa, T.; Kobayashi, H. pH dependence of the function of sodium ion extrusion systems in Escherichia coli. Biochim. Biophys. Acta Bioenerg. 1998, 1363, 231–237. [Google Scholar] [CrossRef] [Green Version]
  37. Inaba, K.; Kuroda, T.; Shimamoto, T.; Kayahara, T.; Tsuda, M.; Tsuchiya, T. Lithium toxicity and Na+(Li+)/H+ antiporter in Escherichia coli. Biol. Pharm. Bull. 1994, 17, 395–398. [Google Scholar] [CrossRef] [Green Version]
  38. Naseem, R.; Holland, I.B.; Jacq, A.; Wann, K.T.; Campbell, A.K. pH and monovalent cations regulate cytosolic free Ca2+ in E. coli. Biochim. Biophys. Acta Biomembr. 2008, 1778, 1415–1422. [Google Scholar] [CrossRef]
  39. Radchenko, M.V.; Tanaka, K.; Waditee, R.; Oshimi, S.; Matsuzaki, Y.; Fukuhara, M.; Kobayashi, H.; Takabe, T.; Nakamura, T. Potassium/proton antiport system of Escherichia coli. J. Biol. Chem. 2006, 281, 19822–19829. [Google Scholar] [CrossRef] [Green Version]
  40. Slonczewski, J.L.; Fujisawa, M.; Dopson, M.; Krulwich, T.A. Cytoplasmic pH measurement and homeostasis in bacteria and archaea. Adv. Microb. Physiol. 2009, 55, 1–79. [Google Scholar]
  41. Taglicht, D.; Padan, E.; Schuldiner, S. Proton-sodium stoichiometry of NhaA, an electrogenic antiporter from Escherichia coli. J. Biol. Chem. 1993, 268, 5382–5387. [Google Scholar]
  42. Pinner, E.; Padan, E.; Schuldiner, S. Kinetic properties of NhaB, a Na+/H+ antiporter from Escherichia coli. J. Biol. Chem. 1994, 269, 26274–26279. [Google Scholar]
  43. Lewinson, O.; Padan, E.; Bibi, E. Alkalitolerance: A biological function for a multidrug transporter in pH homeostasis. Proc. Natl. Acad. Sci. USA 2004, 101, 14073–14078. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Holdsworth, S.R.; Law, C.J. Multidrug resistance protein MdtM adds to the repertoire of antiporters involved in alkaline pH homeostasis in Escherichia coli. BMC Microbiol. 2013, 13, 1987. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Padan, E.; Bibi, E.; Ito, M.; Krulwich, T.A. Alkaline pH homeostasis in bacteria: New insights. Biochim. Biophys. Acta Biomembr. 2005, 1717, 67–88. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Holdsworth, S.R.; Law, C.J. Functional and biochemical characterisation of the Escherichia coli major facilitator superfamily multidrug transporter MdtM. Biochimie 2012, 94, 1334–1346. [Google Scholar] [CrossRef] [PubMed]
  47. Holdsworth, S.R.; Law, C.J. The major facilitator superfamily transporter MdtM contributes to the intrinsic resistance of Escherichia coli to quaternary ammonium compounds. J. Antimicrob. Chemother. 2013, 68, 831–839. [Google Scholar] [CrossRef] [Green Version]
  48. Paul, S.; Alegre, K.O.; Holdsworth, S.R.; Rice, M.; Brown, J.A.; Mcveigh, P.; Kelly, S.M.; Law, C.J. A single-component multidrug transporter of the major facilitator superfamily is part of a network that protects Escherichia coli from bile salt stress. Mol. Microbiol. 2014, 92, 872–884. [Google Scholar] [CrossRef] [Green Version]
  49. Edgar, R.; Bibi, E. MdfA, an Escherichia coli multidrug resistance protein with an extraordinarily broad spectrum of drug recognition. J. Bacteriol. 1997, 179, 2274–2280. [Google Scholar] [CrossRef] [Green Version]
  50. Tanudjaja, E.; Hoshi, N.; Su, Y.-H.; Hamamoto, S.; Uozumi, N. Kup-mediated Cs+ uptake and Kdp-driven K+ uptake coordinate to promote cell growth during excess Cs+ conditions in Escherichia coli. Sci. Rep. 2017, 7, 2122. [Google Scholar] [CrossRef] [Green Version]
  51. Ferguson, G.P.; Nikolaev, Y.; McLaggan, D.; Maclean, M.; Booth, I.R. Survival during exposure to the electrophilic reagent N-ethylmaleimide in Escherichia coli: Role of KefB and KefC potassium channels. J. Bacteriol. 1997, 179, 1007–1012. [Google Scholar] [CrossRef] [Green Version]
  52. Ferguson, G.P.; McLaggan, D.; Booth, I.R. Potassium channel activation by glutathione-S-conjugates in Escherichia coli: Protection against methylglyoxal is mediated by cytoplasmic acidification. Mol. Microbiol. 1995, 17, 1025–1033. [Google Scholar] [CrossRef]
  53. Miller, S.; Ness, L.S.; Wood, C.M.; Fox, B.C.; Booth, I.R. Identification of an ancillary protein, YabF, required for activity of the KefC glutathione-gated potassium efflux system in Escherichia coli. J. Bacteriol. 2000, 182, 6536–6540. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Roosild, T.P.; Castronovo, S.; Healy, J.; Miller, S.; Pliotas, C.; Rasmussen, T.; Bartlett, W.; Conway, S.J.; Booth, I.R. Mechanism of ligand-gated potassium efflux in bacterial pathogens. Proc. Natl. Acad. Sci. USA 2010, 107, 19784–19789. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Roosild, T.P.; Castronovo, S.; Miller, S.; Li, C.; Rasmussen, T.; Bartlett, W.; Gunasekera, B.; Choe, S.; Booth, I.R. KTN (RCK) domains regulate K+ channels and transporters by controlling the dimer-hinge conformation. Structure 2009, 17, 893–903. [Google Scholar] [CrossRef] [PubMed]
  56. Verkhovskaya, M.L.L.; Barquera, B.; Wikström, M. Deletion of one of two Escherichia coli genes encoding putative Na+/H+ exchangers (ycgO) perturbs cytoplasmic alkali cation balance at low osmolarity. Microbiology 2001, 147, 3005–3013. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Sharma, R.; Shimada, T.; Mishra, V.K.; Upreti, S.; Sardesai, A.A. Growth inhibition by external potassium of Escherichia coli lacking PtsN (EIIANtr) is caused by potassium limitation mediated by YcgO. J. Bacteriol. 2016, 198, 1868–1882. [Google Scholar] [CrossRef] [Green Version]
  58. Allakhverdiev, S.I.; Murata, N. Salt stress inhibits photosystems II and I in cyanobacteria. Photosynth. Res. 2008, 98, 529–539. [Google Scholar] [CrossRef]
  59. Inaba, M.; Sakamoto, A.; Murata, N. Functional expression in Escherichia coli of low-affinity and high-affinity Na+(Li+)/H+ antiporters of Synechocystis. J. Bacteriol. 2001, 183, 1376–1384. [Google Scholar] [CrossRef] [Green Version]
  60. Elanskaya, I.V.; Karandashova, I.V.; Bogachev, A.V.; Hagemann, M. Functional analysis of the Na+/H+ antiporter encoding genes of the cyanobacterium Synechocystis PCC 6803. Biochemistry (Mosc.) 2002, 67, 432–440. [Google Scholar] [CrossRef]
  61. Mikkat, S.; Milkowski, C.; Hagemann, M. The gene sll0273 of the cyanobacterium Synechocystis sp. strain PCC6803 encodes a protein essential for growth at low Na+/K.+ ratios. Plant Cell Environ. 2000, 23, 549–559. [Google Scholar] [CrossRef]
  62. Hamada, A.; Hibino, T.; Nakamura, T.; Takabe, T. Na+/H+ antiporter from Synechocystis species PCC 6803, homologous to SOS1, contains an aspartic residue and long c-terminal tail important for the carrier activity. Plant Physiol. 2001, 125, 437–446. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Billini, M.; Stamatakis, K.; Sophianopoulou, V. Two members of a network of putative Na+/H+ antiporters are involved in salt and pH tolerance of the freshwater cyanobacterium Synechococcus elongatus. J. Bacteriol. 2008, 190, 6318–6329. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Waditee, R.; Hossain, G.S.; Tanaka, Y.; Nakamura, T.; Shikata, M.; Takano, J.; Takabe, T.; Takabe, T. Isolation and functional characterization of Ca2+/H+ antiporters from cyanobacteria. J. Biol. Chem. 2004, 279, 4330–4338. [Google Scholar] [CrossRef] [Green Version]
  65. Qiu, Q.S.; Guo, Y.; Dietrich, M.A.; Schumaker, K.S.; Zhu, J.K. Regulation of SOS1, a plasma membrane Na+/H+ exchanger in Arabidopsis thaliana, by SOS2 and SOS3. Proc. Natl. Acad. Sci. USA 2002, 99, 8436–8441. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Qi, Z.; Spalding, E.P. Protection of plasma membrane K+ transport by the salt overly sensitive1 Na+/H+ antiporter during salinity stress. Plant Physiol. 2004, 136, 2548–2555. [Google Scholar] [CrossRef] [Green Version]
  67. Jiang, Z.; Zhou, X.; Tao, M.; Yuan, F.; Liu, L.; Wu, F.; Wu, X.; Xiang, Y.; Niu, Y.; Liu, F.; et al. Plant cell-surface GIPC sphingolipids sense salt to trigger Ca2+ influx. Nature 2019, 572, 341–346. [Google Scholar] [CrossRef]
  68. An, R.; Chen, Q.J.; Chai, M.F.; Lu, P.L.; Su, Z.; Qin, Z.X.; Chen, J.; Wang, X.C. AtNHX8, a member of the monovalent cation:proton antiporter-1 family in Arabidopsis thaliana, encodes a putative Li+/H+ antiporter. Plant J. 2007, 49, 718–728. [Google Scholar] [CrossRef]
  69. Zhao, J.; Cheng, N.H.; Motes, C.M.; Blancaflor, E.B.; Moore, M.; Gonzales, N.; Padmanaban, S.; Sze, H.; Ward, J.M.; Hirschi, K.D. AtCHX13 is a plasma membrane K+ transporter. Plant Physiol. 2008, 148, 796–807. [Google Scholar] [CrossRef] [Green Version]
  70. Zhao, J.; Li, P.; Motes, C.M.; Park, S.; Hirschi, K.D. CHX14 is a plasma membrane K-efflux transporter that regulates K+ redistribution in Arabidopsis thaliana. Plant Cell Environ. 2015, 38, 2223–2238. [Google Scholar] [CrossRef]
  71. Bassil, E.; Tajima, H.; Liang, Y.C.; Ohto, M.; Ushijima, K.; Nakano, R.; Esumi, T.; Coku, A.; Belmonte, M.; Blumwald, E. The arabidopsis Na+/H+ antiporters NHX1 and NHX2 control vacuolar pH and K+ homeostasis to regulate growth, flower development, and reproduction. Plant Cell 2011, 23, 3482–3497. [Google Scholar] [CrossRef] [Green Version]
  72. Barragán, V.; Leidi, E.O.; Andrés, Z.; Rubio, L.; de Luca, A.; Fernández, J.A.; Cubero, B.; Pardo, J.M. Ion exchangers NHX1 and NHX2 mediate active potassium uptake into vacuoles to regulate cell turgor and stomatal function in Arabidopsis. Plant Cell 2012, 24, 1127–1142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Yamaguchi, T.; Aharon, G.S.; Sottosanto, J.B.; Blumwald, E. Vacuolar Na+/H+ antiporter cation selectivity is regulated by calmodulin from within the vacuole in a Ca2+- and pH-dependent manner. Proc. Natl. Acad. Sci. USA. 2005, 102, 16107–16112. [Google Scholar] [CrossRef] [Green Version]
  74. Gruwel, M.L.H.; Rauw, V.L.; Loewen, M.; Abrams, S.R. Effects of sodium chloride on plant cells; A 31P and 23Na NMR system to study salt tolerance. Plant Sci. 2001, 160, 785–794. [Google Scholar] [CrossRef]
  75. Liu, H.; Wang, Q.; Yu, M.; Zhang, Y.; Wu, Y.; Zhang, H. Transgenic salt-tolerant sugar beet (Beta vulgaris L.) constitutively expressing an Arabidopsis thaliana vacuolar Na+/H+ antiporter gene, AtNHX3, accumulates more soluble sugar but less salt in storage roots. Plant Cell Environ. 2008, 31, 1325–1334. [Google Scholar] [CrossRef] [PubMed]
  76. Liu, H.; Tang, R.; Zhang, Y.; Wang, C.; Lv, Q.; Gao, X.; Li, W.B.; Zhang, H. AtNHX3 is a vacuolar K+/H+ antiporter required for low-potassium tolerance in Arabidopsis thaliana. Plant Cell Environ. 2010, 33, 1989–1999. [Google Scholar] [CrossRef] [PubMed]
  77. Li, H.T.; Liu, H.; Gao, X.S.; Zhang, H. Knock-out of Arabidopsis AtNHX4 gene enhances tolerance to salt stress. Biochem. Biophys. Res. Commun. 2009, 382, 637–641. [Google Scholar] [CrossRef]
  78. Müller, P.; Li, X.P.; Niyogi, K.K. Non-photochemical quenching. A response to excess light energy. Plant Physiol. 2001, 125, 1558–1566. [Google Scholar] [CrossRef] [Green Version]
  79. Kunz, H.-H.; Gierth, M.; Herdean, A.; Satoh-Cruz, M.; Kramer, D.M.; Spetea, C.; Schroeder, J.I. Plastidial transporters KEA1, -2, and -3 are essential for chloroplast osmoregulation, integrity, and pH regulation in Arabidopsis. Proc. Natl. Acad. Sci. USA 2014, 111, 7480–7845. [Google Scholar] [CrossRef] [Green Version]
  80. Armbruster, U.; Carrillo, L.R.; Venema, K.; Pavlovic, L.; Schmidtmann, E.; Kornfeld, A.; Jahns, P.; Berry, J.A.; Kramer, D.M.; Jonikas, M.C. Ion antiport accelerates photosynthetic acclimation in fluctuating light environments. Nat. Commun. 2014, 5, 5439. [Google Scholar] [CrossRef] [Green Version]
  81. Armbruster, U.; Leonelli, L.; Galvis, V.C.; Strand, D.; Quinn, E.H.; Jonikas, M.C.; Niyogi, K.K. Regulation and levels of the thylakoid K+/H+ antiporter KEA3 shape the dynamic response of photosynthesis in fluctuating light. Plant Cell Physiol. 2016, 57, 1557–1567. [Google Scholar] [CrossRef] [Green Version]
  82. Wang, C.; Yamamoto, H.; Narumiya, F.; Munekage, Y.N.; Finazzi, G.; Szabo, I.; Shikanai, T. Fine-tuned regulation of the K+/H+ antiporter KEA3 is required to optimize photosynthesis during induction. Plant J. 2017, 89, 540–553. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Wang, C.; Shikanai, T. Modification of Activity of the Thylakoid H+/K+ antiporter KEA3 disturbs ∆pH-dependent regulation of photosynthesis. Plant Physiol. 2019, 181, 762–773. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Stephan, A.B.; Kunz, H.-H.; Yang, E.; Schroeder, J.I. Rapid hyperosmotic-induced Ca2+ responses in Arabidopsis thaliana exhibit sensory potentiation and involvement of plastidial KEA transporters. Proc. Natl. Acad. Sci. USA 2016, 113, E5242–E5249. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Aranda-Sicilia, M.N.; Aboukila, A.; Armbruster, U.; Cagnac, O.; Schumann, T.; Kunz, H.-H.; Jahns, P.; Rodríguez-Rosales, M.P.; Sze, H.; Venema, K. Envelope K+/H+ antiporters AtKEA1 and AtKEA2 function in plastid development. Plant Physiol. 2016, 172, 441–449. [Google Scholar] [CrossRef] [Green Version]
  86. Müller, M.; Kunz, H.H.; Schroeder, J.I.; Kemp, G.; Young, H.S.; Ekkehard Neuhaus, H. Decreased capacity for sodium export out of Arabidopsis chloroplasts impairs salt tolerance, photosynthesis and plant performance. Plant J. 2014, 78, 646–658. [Google Scholar] [CrossRef] [Green Version]
  87. Furumoto, T.; Yamaguchi, T.; Ohshima-Ichie, Y.; Nakamura, M.; Tsuchida-Iwata, Y.; Shimamura, M.; Ohnishi, J.; Hata, S.; Gowik, U.; Westhoff, P.; et al. A plastidial sodium-dependent pyruvate transporter. Nature 2011, 476, 472–475. [Google Scholar] [CrossRef] [Green Version]
  88. Nomura, H.; Shiina, T. Calcium signaling in plant endosymbiotic organelles: Mechanism and role in physiology. Mol. Plant 2014, 7, 1094–1104. [Google Scholar] [CrossRef] [Green Version]
  89. Wang, C.; Xu, W.; Jin, H.; Zhang, T.; Lai, J.; Zhou, X.; Zhang, S.; Liu, S.; Duan, X.; Wang, H.; et al. A Putative Chloroplast-Localized Ca2+/H+ Antiporter CCHA1 Is involved in Calcium and pH homeostasis and required for PSII function in arabidopsis. Mol. Plant 2016, 9, 1183–1196. [Google Scholar] [CrossRef] [Green Version]
  90. Schneider, A.; Steinberger, I.; Herdean, A.; Gandini, C.; Eisenhut, M.; Kurz, S.; Morper, A.; Hoecker, N.; Rühle, T.; Labs, M.; et al. The evolutionarily conserved protein PHOTOSYNTHESIS AFFECTED MUTANT71 is required for efficient manganese uptake at the thylakoid membrane in Arabidopsis. Plant Cell 2016, 28, 892–910. [Google Scholar] [CrossRef]
  91. Reguera, M.; Bassil, E.; Tajima, H.; Wimmer, M.; Chanoca, A.; Otegui, M.S.; Paris, N.; Blumwald, E. pH regulation by NHX-type antiporters is required for receptor-mediated protein trafficking to the vacuole in Arabidopsis. Plant Cell 2015, 27, 1200–1217. [Google Scholar] [CrossRef] [Green Version]
  92. Dragwidge, J.M.; Ford, B.A.; Ashnest, J.R.; Das, P.; Gendall, A.R. Two endosomal NHX-Type Na+/H+ antiporters are involved in auxin-mediated development in Arabidopsis Thaliana. Plant Cell Physiol. 2018, 59, 1660–1669. [Google Scholar] [CrossRef] [PubMed]
  93. Cellier, F.; Conéjéro, G.; Ricaud, L.; Doan, T.L.; Lepetit, M.; Gosti, F.; Casse, F. Characterization of AtCHX17, a member of the cation/H+ exchangers, CHX family, from Arabidopsis thaliana suggests a role in K+ homeostasis. Plant J. 2004, 39, 834–846. [Google Scholar] [CrossRef] [PubMed]
  94. Chanroj, S.; Padmanaban, S.; Czerny, D.D.; Jauh, G.Y.; Sze, H. K+ transporter AtCHX17 with its hydrophilic C tail localizes to membranes of the secretory/endocytic system: Role in reproduction and seed set. Mol. Plant 2013, 6, 1226–1246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Lu, Y.; Chanroj, S.; Zulkifli, L.; Johnson, M.A.; Uozumi, N.; Cheung, A.; Sze, H. Pollen tubes lacking a pair of K+ transporters fail to target ovules in Arabidopsis. Plant Cell 2011, 23, 81–93. [Google Scholar] [CrossRef] [Green Version]
  96. Zhu, X.; Pan, T.; Zhang, X.; Fan, L.; Quintero, F.J.; Zhao, H.; Su, X.; Li, X.; Villalta, I.; Mendoza, I.; et al. K+ efflux antiporters 4, 5, and 6 mediate pH and K+ homeostasis in endomembrane compartments. Plant Physiol. 2018, 178, 1657–1678. [Google Scholar] [CrossRef] [Green Version]
  97. Wang, Y.; Tang, R.J.; Yang, X.; Zheng, X.; Shao, Q.; Tang, Q.L.; Fu, A.; Luan, S. Golgi-localized cation/proton exchangers regulate ionic homeostasis and skotomorphogenesis in Arabidopsis. Plant Cell Environ. 2019, 42, 673–687. [Google Scholar] [CrossRef]
Figure 1. Structure and molecular function of NhaA. (A) Typical profile of the proton movement across an inverted membrane vesicle. Na+/H+ antiporter activity is monitored by dequenching of a pH sensitive fluorescent probe like acridine orange. (B) Three-dimensional structure model of NhaA, the Na+/H+ antiporter of E. coli (PDB 4AU5 [26]), was made by using PyMOL, version 2.3.5, Schrodinger LLC. Overview (left) and ion biding sites (right). The numbers in the left figure correspond to the numbers of the transmembrane domains. (C) Proposed mechanism of Na+ and H+ translocation by NhaA [26], D163 and K300 form a salt bridge (dotted line). When Na+ binds to D163 and D164, the salt bridges are disrupted, and two protons are released from D163 and K300.
Figure 1. Structure and molecular function of NhaA. (A) Typical profile of the proton movement across an inverted membrane vesicle. Na+/H+ antiporter activity is monitored by dequenching of a pH sensitive fluorescent probe like acridine orange. (B) Three-dimensional structure model of NhaA, the Na+/H+ antiporter of E. coli (PDB 4AU5 [26]), was made by using PyMOL, version 2.3.5, Schrodinger LLC. Overview (left) and ion biding sites (right). The numbers in the left figure correspond to the numbers of the transmembrane domains. (C) Proposed mechanism of Na+ and H+ translocation by NhaA [26], D163 and K300 form a salt bridge (dotted line). When Na+ binds to D163 and D164, the salt bridges are disrupted, and two protons are released from D163 and K300.
Ijms 21 04566 g001
Figure 2. Localization of CPAs (cation/proton antiporters) in Escherichia coli (A), cyanobacteria (B) and Arabidopsis thaliana (C).
Figure 2. Localization of CPAs (cation/proton antiporters) in Escherichia coli (A), cyanobacteria (B) and Arabidopsis thaliana (C).
Ijms 21 04566 g002
Table 1. CPAs in Escherichia coli.
Table 1. CPAs in Escherichia coli.
TransporterActivityPhysiological FunctionFamilyReference
NhaANa+(Li+)/H+ antiporterNa+ extrusion, alkaline pH homeostasisNhaA(Taglicht et al., 1991)
NhaBNa+(Li+)/H+ antiporterNa+ extrusion, alkaline pH homeostasisNhaB(Pinner et al., 1993,
Shimamoto et al., 1994)
ChaANa+(Ca+, K+)/H+ antiporterNa+ extrusion, alkaline pH homeostasis, KCl salinity toleranceCaCA(Ivey et al., 1993,
Radchenko et al., 2006)
KefBglutathione-regulated K+ efflux (K+/H+ antiporter)electrophilic stress toleranceCPA2(Ferguson et al., 1997,
Elmore et al., 1997)
KefCglutathione-regulated K+ efflux (K+/H+ antiporter)electrophilic stress toleranceCPA2(Ferguson et al., 1997,
Elmore et al., 1997)
MdfAdrugs (Na+,K+)/H+ antiporterantimicrobial resistance, alkaline pH homeostasisDHA1(Lewinson et al., 2004,
Lewinson et al., 2003)
MdtMdrugs (Na+,K+, Rb+, Li+)/H+ antiporterantimicrobial resistance, alkaline pH homeostasisDHA1(Holdsworth et al., 2013)
YcgOK+/H+ antiportergrowth at low osmolarity CPA1(Sharma et al., 2016,
Verkhovskaya et al., 2001)
Table 2. CPAs in cyanobacteria.
Table 2. CPAs in cyanobacteria.
Prokaryote IsoformMembrane LocationActivityPhysiological FunctionFamilyReference
NhaS1 (Syncechocystis sp. pcc6803)-Na+ (Li+)/H+ antiporter-CPA1(Inaba et al., 2001, Hamada et al., 2001, Waditee et al., 2006)
NhaS2 (Syncehocystis sp. pcc6803)--Na+ influxCPA1(Mikkat et al., 2000, Wang et al., 2002)
NhaS3 (Syncehocystis sp. pcc6803)Thylakoid membraneNa+/H+ antiporterNa+ homeostasis, adaptation to osmotic stressCPA2(Wang et al., 2002, Tsunekawa et al., 2009)
NhaS4 (Syncehocystis sp. pcc6803)--Na+ influxCPA2(Wang et al., 2002)
SynCAX (Syncehocystis sp. pcc6803)-Ca2+/H+ antiporteralkaline homeostasis, adaptation to salt stress-(Waditee et al., 2004)
NhaS2 (Syncechococcus sp. strain PCC 7942)--Na+ influxCPA1(Billini et al., 2008)
NhaS3 (Syncechococcus sp. strain PCC 7942)-Na+/H+ antiporterNa+ homeostasisCPA2(Billini et al., 2008)
Table 3. CPAs in Arabidopsis thaliana.
Table 3. CPAs in Arabidopsis thaliana.
Plant IsoformMembrane LocationActivityPhysiological FunctionFamilyReference
NHX1, NHX2VacuoleK+(Na+)/H+ antiportermaintaining cell turgor pressure, adaptation to excess K+ conditions, floweringCPA1(Shi and Zhu, 2002, Bassil et al., 2011, Barragan et al., 2012)
NHX3VacuoleK+(Na+)/H+ antiporteradaptation to low K+ conditionsCPA1(Liu et al., 2008, Liu et al., 2010)
NHX4VacuoleNa+/H+ antiporterNa+ homeostasisCPA1(Li et al., 2009)
NHX5, NHX6Golgi, Trans-golgi network, prevacuolar compartment-protein sorting to vacuole, adaptation to salt stress, root extention, redistribution of auxinCPA1(Bassil et al., 2011, Reguela et al., 2015, Dragwidge et al., 2018)
NHX7 (SOS1)Plasma membraneNa+/H+ antiporterNa+ extrusion, maintaining intracellular K+ contentCPA1(Wu et al., 1996, Shi et al., 2002, Qiu et al., 2002, Qi and Spalding, 2004, Ariga et al., 2013)
NHX8Plasma membrane(Li+)/H+ antiporterLi+ extrusionCPA1(An et al., 2007)
CHX13Plasma membraneK+/H+ antiporterK+ uptakeCPA2(Zhao et al., 2008)
CHX14Plasma membraneK+/H+ antiporterK+ extrusionCPA2(Zhao et al., 2015)
CHX17Prevacuolar compartmentK+/H+ antiporterprotein sorting to vacuole, K+ homeostasis, seed developmentCPA2(Cellier et al., 2004, Chanroj et al., 2011, Chanroj et al., 2013)
CHX18, CHX19Prevacuolar compartmentK+/H+ antiporter-CPA2(Chanroj et al., 2011)
CHX20Prevacuolar compartment, ERK+/H+ antiporterprotein sorting to vacuoleCPA2(Chanroj et al., 2011)
CHX21Plasma membrane-adaptation to high Na+ (K+) conditions, pollen tube growthCPA2(Hall et al., 2006, Evans et al., 2011)
CHX23ERK+/H+ antiporterpollen tube growthCPA2(Lu et al., 2011, Evans et al., 2011)
KEA1, KEA2Plastid envelopeK+/H+ antiporterchloroplast development, adaptation to hyper-osmotic stressCPA2(Kunz et al., 2014, Aranda Sicilia et al., 2016, Stephan et al., 2016, Tsujii et al., 2019)
KEA3Plastid thylakoid membraneK+/H+ antiporterfine-tuning of photosynthesis, chloroplast development, adaptation to hyper-osmotic stressCPA2(Kunz et al., 2014, Armbruster et al., 2014, 2016, Stephan et al., 2016, Wang et al., 2017, 2019, Tsujii et al., 2019)
KEA4, KEA5, KEA6Golgi, Trans-golgi network, prevacuolar compartmentK+/H+ antiportermaintaining ion homeostasis in low K+ conditions, high K+(Na+) stress, pH regulation in vacuoleCPA2(Zhu et al., 2018, Wang et al., 2018, Tsujii et al., 2019)
NHD1Plastid envelopeNa+/H+ antiporterNa+ extrusion from chloroplastNhaD(Maria et al., 2014)
CCHA1 (PAM71)Plastid thylakoid membrane-Photosynthetic regulation, regulation of cytosolic pH in guard cells-(Wang et al., 2016, Schneider et al., 2016)

Share and Cite

MDPI and ACS Style

Tsujii, M.; Tanudjaja, E.; Uozumi, N. Diverse Physiological Functions of Cation Proton Antiporters across Bacteria and Plant Cells. Int. J. Mol. Sci. 2020, 21, 4566. https://doi.org/10.3390/ijms21124566

AMA Style

Tsujii M, Tanudjaja E, Uozumi N. Diverse Physiological Functions of Cation Proton Antiporters across Bacteria and Plant Cells. International Journal of Molecular Sciences. 2020; 21(12):4566. https://doi.org/10.3390/ijms21124566

Chicago/Turabian Style

Tsujii, Masaru, Ellen Tanudjaja, and Nobuyuki Uozumi. 2020. "Diverse Physiological Functions of Cation Proton Antiporters across Bacteria and Plant Cells" International Journal of Molecular Sciences 21, no. 12: 4566. https://doi.org/10.3390/ijms21124566

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