Hostname: page-component-7479d7b7d-jwnkl Total loading time: 0 Render date: 2024-07-08T17:09:57.632Z Has data issue: false hasContentIssue false

Energizing the plasmalemma of marine photosynthetic organisms: the role of primary active transport

Published online by Cambridge University Press:  13 April 2020

John A. Raven*
Affiliation:
Division of Plant Science, University of Dundee at the James Hutton Institute, Invergowrie, DundeeDD2 5DA, UK Climate Change Cluster, University of Technology Sydney, Ultimo, NSW2007, Australia School of Biological Sciences, University of Western Australia, 35 Stirling Highway, Crawley, WA6009, Australia
John Beardall
Affiliation:
School of Biological Sciences, Monash University, Clayton, VIC3800, Australia State Key Laboratory of Marine Environmental Science & College of Ocean and Earth Sciences, Xiamen University, Xiamen361005, China
*
Author for correspondence: John A. Raven, E-mail: [email protected]
Rights & Permissions [Opens in a new window]

Abstract

Generation of ion electrochemical potential differences by primary active transport can involve energy inputs from light, from exergonic redox reactions and from exergonic ATP hydrolysis. These electrochemical potential differences are important for homoeostasis, for signalling, and for energizing nutrient influx. The three main ions involved are H+, Na+ (efflux) and Cl (influx). In prokaryotes, fluxes of all three of these ions are energized by ion-pumping rhodopsins, with one archaeal rhodopsin pumping H+into the cells; among eukaryotes there is also an H+ influx rhodopsin in Acetabularia and (probably) H+ efflux in diatoms. Bacteriochlorophyll-based photoreactions export H+ from the cytosol in some anoxygenic photosynthetic bacteria, but chlorophyll-based photoreactions in marine cyanobacteria do not lead to export of H+. Exergonic redox reactions export H+ and Na+ in photosynthetic bacteria, and possibly H+ in eukaryotic algae. P-type H+- and/or Na+-ATPases occur in almost all of the photosynthetic marine organisms examined. P-type H+-efflux ATPases occur in charophycean marine algae and flowering plants whereas P-type Na+-ATPases predominate in other marine green algae and non-green algae, possibly with H+-ATPases in some cases. An F-type Cl-ATPase is known to occur in Acetabularia. Some assignments, on the basis of genomic evidence, of P-type ATPases to H+ or Na+ as the pumped ion are inconclusive.

Type
Review
Copyright
Copyright © Marine Biological Association of the United Kingdom 2020

Introduction

Transport of solutes and water across the plasmalemma is vital for homoeostasis and growth of all organisms. For many of these solutes, and perhaps water in some cases, the movement of the ions and molecules is in the direction opposite to that dictated by the free energy difference across the plasmalemma for that ion or compound: this is the strict definition of active transport. For neutral solutes and water, the free energy difference is the difference in chemical activity (concentration times activity coefficient). For ions there is the additional component of electrical potential difference in the overall free energy difference. Active transport necessitates an input of free energy per molecule or ion transported in excess of the free energy difference per molecule or ion. We follow here the distinction made by Mitchell (Reference Mitchell1979) between primary and secondary active transport (see also Saier, Reference Saier2000; Saier et al., Reference Saier, Reddy, Tsu, Ahmed, Li and Moreno-Hagelsieb2015). In primary active transport the energy input to the transporter is an exergonic scalar chemical reaction (e.g. oxidation of NAD(P)H by O2; hydrolysis of ATP to produce ADP and Pi) or photons absorbed by a chromophore component of the transporter (e.g. the retinal component of ion-pumping rhodopsins). In secondary active transport the energy input to the transporter is from the energetically downhill transmembrane flux of a driving ion or ions, with re-energization by primary active transport of those driving ions.

The three major ions involved in primary active transport at the plasmalemma of Archaea, Bacteria and Eukarya are H+, Na+ and Cl. Table 1 shows what is known of the distribution of these pumps among the three higher taxa and the immediate energy sources. In marine photosynthetic (including photoheterotrophic) organisms there are examples of all of the pumps in Table 1 located in the plasmalemma. It is these ions that are the focus of the subsequent discussion in this paper. Other important ions subject to primary active transport are Ca2+ and Mg2+, transported out of cells by P-type ATPases, that also pump, in various organisms, Na+ out:K+ in, K+ in, Na+ out, H+ out and H+ out:K+ in (Kuhlbrandt, Reference Kuhlbrandt2004; Bublitz et al., Reference Bublitz, Morth and Nissen2011; Chan et al., Reference Chan, Reyes-Prieto and Bhattacharya2011, Reference Chan, Zäuner, Wheeler, Grossman, Prochnik, Blouin, Zhuang, Benning, Berg, Yarish, Ekiksen, Klein, Lin, Levine, Brawley and Bhattachaya2012; Palmgren & Nissen, Reference Palmgren and Nissen2011; Søndergaard & Pedersen, Reference Søndergaard and Pedersen2015; Palmgren et al., Reference Palmgren, Sorenson, Hallström, Säll and Broberg2020). As well as variants pumping H+, Na+, K+, Mg2+ and Ca2+ (topological type II: Thever & Saier, Reference Thever and Saier2009, in prokaryotes and eukaryotes), there are also P-type ATPases (topological type I, Thever & Saier, Reference Thever and Saier2009, not considered further here) that pump out metal cations such as Cu(I), Ag(I), Zn(II), Cd(II) and Pb(II) (Kuhlbrandt, Reference Kuhlbrandt2004; Thever & Saier, Reference Thever and Saier2009; Bublitz et al., Reference Bublitz, Morth and Nissen2011; Palmgren & Nissen, Reference Palmgren and Nissen2011; Søndergaard & Pedersen, Reference Søndergaard and Pedersen2015). There is also an F-type Cl influx ATPase in the plasmalemma of some marine algae, as well as P-type Cl-ATPases as found in metazoans (Gerencser & Zhang, Reference Gerencser and Zhang2003; Raven, Reference Raven2017). There are also ATP (adenosine triphosphate) binding cassette proteins (sometimes referred to as ABC transporters) involved in active solute influx at the plasmalemma, although there is little evidence of these in marine photosynthetic organisms (Chan et al., Reference Chan, Reyes-Prieto and Bhattacharya2011, Reference Chan, Zäuner, Wheeler, Grossman, Prochnik, Blouin, Zhuang, Benning, Berg, Yarish, Ekiksen, Klein, Lin, Levine, Brawley and Bhattachaya2012; but see Badger & Price, Reference Badger and Price2003 for HCO3 accumulation in freshwater and marine cyanobacteria).

Table 1. Energy sources for primary active transporters of H+, Na+ and Cl in the Archaea, Bacteria and Eukarya

See also Kuhlbrandt (Reference Kuhlbrandt2004), Bublitz et al. (Reference Bublitz, Morth and Nissen2011), Palmgren & Nissen (Reference Palmgren and Nissen2011), Søndergaard & Pedersen (Reference Søndergaard and Pedersen2015).

Possible original functions of primary active H+, Na+ and Cl pumps at the plasmalemma

An argument for the early origin and functions of primary active H+ pumps (ATP-driven; light-driven via rhodopsins) is intracellular acid-base regulation in early cells related, for example, to fermentation of neutral substrates (e.g. sugars) to organic acids (e.g. lactic acid) (Raven & Smith, Reference Raven and Smith1981, Reference Raven and Smith1982) in anoxic environments. Occurrence of the redox-driven or light-driven H+ pumps and ATP-driven H+ pumps in the plasmalemma of the same cell permits the use of redox or light energy to phosphorylate ADP to ATP, granted appropriate free energy differences and stoichiometries (Raven & Smith, Reference Raven and Smith1981, Reference Raven and Smith1982). Early chemolithotrophy could generate the H+ free energy difference across the plasmalemma, and thence phosphorylate ADP (Russell & Hall, Reference Russell and Hall1997; Martin & Russell, Reference Martin and Russell2007; Mulkidjanian et al., Reference Mulkidjanian, Dibrov and Galperin2008; Duchuzeau et al., Reference Duchuzeau, Schoepp-Cothenet, Baymann, Russell and Nitscke2014; but see Jackson, Reference Jackson2016). In seawater at the present pH, and with intracellular (cytosolic) pH about 0.5 units lower than seawater (Raven & Smith, Reference Raven and Smith1981, Reference Raven and Smith1982), the electrical potential difference across the membrane must be more negative than is found in most marine eukaryotes to definitely need active H+ efflux (see below). The requirement for intracellular pH regulation in oxygenic photolithotrophic marine cells involves several metabolic reactions assimilating inorganic nutrient solutes and is shown in Table 2. Volume regulation of wall-less marine cells by active Na+ efflux would, in the case of primary active H+ transport, involve H+:Na+ antiport (Raven & Smith, Reference Raven and Smith1982; Katz et al., Reference Katz, Bental, Degani and Avron1991; Gimmler, Reference Gimmler2000).

Table 2. Intracellular H+ production and consumption in redox reactions, and calcification, in marine phytoplankton

Elemental atomic ratios based on the Redfield ratio, with a upper limit on reduced S from Ksionzek et al. (Reference Ksionzek, Lechtenfeld, McAllister, Schmitt-Kopplin, Geuer, Geibert and Koch2016) since values are for total S and not all S in the cell is reduced to the −SH level. For intracellular calcification in coccolithophores (Taylor et al., Reference Taylor, Brownlee and Wheeler2017) and some dinoflagellates (Van de Waal et al., Reference Van de Waal, John, Ziveri, Hoins, Sluija and Röst2013), a particulate inorganic C: particulate organic C ratio of 1.0 is assumed. Other references for H+:N are assimilated from Brewer & Goldman (Reference Brewer and Goldman1976), Smith & Raven (Reference Smith and Raven1979) and Raven (Reference Raven2013).

Well characterized in the, mainly freshwater, green algal macrophytes of the Characeae (Walker et al., Reference Walker, Smith and Cathers1980), localized active efflux of H+ across the plasmalemma causes a localized acidification in the cell wall and diffusion boundary layer. This shifts the CO2:HCO3 equilibrium in favour of CO2, increasing the rate of uncatalysed HCO3 conversion to CO2 and thereby improving cellular CO2 supply (Raven & Beardall, Reference Raven and Beardall2016). A similar mechanism is thought to occur in many marine macrophytes as a mechanism of using external HCO3 (Raven & Hurd, Reference Raven and Hurd2012). However, the evidence for this in marine macroalgae is indirect and based on effects of buffers on HCO3 use, inhibition of external carbonic anhydrase using membrane-impermeant inhibitors or a lack of effects of inhibitors of HCO3 transport (Raven & Hurd, Reference Raven and Hurd2012; Raven & Beardall, Reference Raven and Beardall2016). It is considered unlikely that this process occurs in microalgae, however, as their size would mean they have a thinner diffusive boundary layer and greater proton leakage (Flynn et al., Reference Flynn, Blackford, Baird, Raven, Clark, Beardall, Brownlee, Fabian and Wheeler2012; Raven & Beardall, Reference Raven and Beardall2016).

For primary active Na+ efflux at the plasmalemma of wall-less cells, the higher intracellular K+:Na+ ratio in the cytosol than in a high-salinity medium such as seawater, driven by the primary active Na+ efflux, could be involved in cell volume regulation, countering the Donnan effect of negative charge in the cytosol (Raven & Smith, Reference Raven and Smith1982). The high K+:Na+ in the cytosol is attributed by Dibrova et al. (Reference Dibrova, Galperin, Koonin and Mulkidjanian2015) to the origin of life at inland hot springs with high K+:Na+, resulting in a requirement for high (about 100 mol m−3) K+ for activity of many enzymes. When life invaded the much more widespread freshwater and marine habitats with low K+:Na+ ratios, active Na+ efflux maintained high internal K+:Na+ ratios despite a finite Na+ permeability of the plasmalemma and, in some bacteria, significant energy storage as a Na+ electrochemical potential gradient (Na+ motive force) (Skulachev, Reference Skulachev1984; Skulachev, Reference Skulachev1989; Mulkidjanian et al., Reference Mulkidjanian, Dibrov and Galperin2008; Dibrova et al., Reference Dibrova, Galperin, Koonin and Mulkidjanian2015). Primary active Na+ efflux in the marine species of the wall-less chlorophycean Dunaliella spp. is involved in osmoregulation with varying external osmolarities (Ehrenfeld & Cousin, Reference Ehrenfeld and Cousin1984; Gimmler, Reference Gimmler2000; Popova et al., Reference Popova, Shumkova, Andreev and Balnokin2005; Popova & Balnokin, Reference Popova and Balnokin2013). The Na+ electrochemical potential gradient is widely used in marine (and some other) photosynthetic organisms to energize the influx of nutrients and osmolytes by Na+ co-transport (Raven, Reference Raven1984; Chan et al., Reference Chan, Reyes-Prieto and Bhattacharya2011, Reference Chan, Zäuner, Wheeler, Grossman, Prochnik, Blouin, Zhuang, Benning, Berg, Yarish, Ekiksen, Klein, Lin, Levine, Brawley and Bhattachaya2012). A further function of the Na+ electrochemical potential difference is in the action potentials in marine diatoms (Taylor, Reference Taylor2009; Helliwell et al., Reference Helliwell, Chrachri, Koester, Wharam, Verret, Taylor, Wheeler and Brownlee2019) and, from genomic evidence, the marine phaeophycean Ectocarpus and the marine prasinophyceans Micromonas and Ostreococcus, as well as freshwater chlorophyceans (Fux et al., Reference Fux, Mehta, Moffat and Spafford2018). Recovery of the resting state after an action potential requires active Na+ efflux, either by primary active Na+ efflux, or by Na+:H+ antiport following primary active H+ efflux.

Primary active transport of Cl at the plasmalemma of walled marine ulvophycean algal cells is involved in turgor generation, and in action potentials (Bisson et al., Reference Bisson, Beilby and Shepherd2006; Raven, Reference Raven2017). At least the turgor generation role could be performed by Cl influx coupled to H+ or Na+ influx and primary active H+ or Na+ efflux. The roles of Cl as an essential micronutrient in photosynthesis (Raven, Reference Raven2017) can be satisfied by passive Cl distribution between seawater and the cytosol even with an inside-negative electrical potential of 150 mV, but not the role established in terrestrial flowering plants as a beneficial nutrient (Raven, Reference Raven2017).

Not considered in detail here is active Ca2+ efflux at the plasmalemma that maintains the very low free Ca2+ concentration in the cytosol (100–200 μmol m−3) used in preventing damage to proteins and in signalling (Roberts et al., Reference Roberts, Illot and Brownlee1994; Thompson et al., Reference Thompson, Callow, Callow, Wheeler, Taylor and Brownlee2007; Wheeler et al., Reference Wheeler, Helliwell and Brownlee2019). Ca2+-H+ antiporters and Ca2+ P-ATPases that are involved in Ca2+ efflux from the cytosol are known, on genomic evidence, to be very widespread among algae (Emery et al., Reference Emery, Whelan, Hirschi and Pittman2012; Palmgren et al., Reference Palmgren, Sorenson, Hallström, Säll and Broberg2020).

Energetics of primary active transport

An important aspect of analysing primary active ion transport at the plasmalemma is the electrochemical potential difference for ions across that membrane. The electrochemical potential difference is defined by equation (1) (Mitchell, Reference Mitchell1979; Raven, Reference Raven1984; Nobel, Reference Nobel2009; Nichols & Ferguson, Reference Nichols and Ferguson2013).

(1)$$\Delta \lpar \bar{\mu }\rpar _{\,j^{ +{/}-}{\rm NP}} = zF\Psi _{{\rm NP}} + {\rm RT}\ln \displaystyle{{{\lsqb {\,j^{ +{/}-}} \rsqb }_{\rm N}} \over {{\lsqb {\,j^{ +{/}-}} \rsqb }_{\rm P}}}$$

where $\Delta \lpar \bar{\mu }\rpar _{j^{ +{/}-}{\rm NP}}$ = electrochemical potential of ion j +/− in phase N relative to that in phase P; N = electrically negative phase (cytosol for the plasmalemma); P = electrically positive phase (aqueous medium for the plasmalemma); j +/− = ion under consideration, either a cation (j +) or anion (j ); z = numerical charge on the ion (+1 for H+ or Na+, +2 for Ca2+, −1 for Cl); F = Faraday constant (96,485 Joule V−1 mol−1); Ψ NP = electrical potential of phase N relative to phase P; R = gas constant (8.314 Joule mol−1 °K); T = temperature (°K); ln = natural logarithm; [j +]N = concentration of j + in phase N (mol m−3); [j ]P = concentration of j in phase P (mol m−3).

The sign of the electrochemical potential difference defines the direction of active transport of the ion, i.e. in the energetically uphill direction, requiring an input of energy from coupling to photons, exergonic redox reactions, or ATP conversion to ADP and phosphate in primary active transport, or coupling to exergonic ion fluxes in secondary active transport. The magnitude of the electrochemical potential difference defines the minimum energy input to primary or secondary active transport. As will be seen in the rest of the paper, Na+ invariably, and H+ very widely, are actively transported by marine photosynthetic organisms from the cytosol to the medium. If only one of these ion species, e.g. H+, is subject to primary active transport, then antiport coupling exergonic H+ re-entry to secondary active Na+ efflux must have a H+:Na+ stoichiometry consistent with the antiport being overall exergonic.

Hereinafter, the electrical potential of the cytosol (N phase) relative to the outside medium (P phase) is represented as Ψ CO.

Organisms with ion-pumping rhodopsins

Ion-pumping rhodopsins are light energy transducers that bring about active ion transport as the sole product of the photochemical reaction usable in cell metabolism (Oesterhelt & Stoeckenius, Reference Oesterhelt and Stoeckenius1973). The photoreaction and ion transport occurs using a single retinol-binding opsin protein, and brings about positive charge (H+ or Na+) flux from the N to the P side of the membrane, or negative charge (Cl) flux from the P to the N side of the membrane, where ‘N’ and ‘P’ are as defined by Mitchell (Reference Mitchell1979). For the plasmalemma, the cytosol is the N side and the outside medium is the P side (see equation (1)).

Some marine bacteria have rhodopsins that pump H+ or Na+ out of the cells, or Cl into the cells; some marine Archaea have been shown to have H+ efflux or Cl influx rhodopsins, but there are no reports of Na+ efflux rhodopsins in Archaea (Oesterhelt & Stoeckenius, Reference Oesterhelt and Stoeckenius1973; Bejá & Lanyi, Reference Bejá and Lanyi2014; Yoshizawa et al., Reference Yoshizawa, Kumuga, Kim, Ogawa, Hayashi, Iwasaki, DeLong and Kogure2014; Larkum et al., Reference Larkum, Ritchie and Raven2018). Unexpectedly, there is an inwardly directed H+ pump driven by xenorhodopsin in Nanosalina (Shevchenko et al., Reference Shevchenko, Mager, Kovalev, Polovinkin, Alekseev, Juettner, Chizhov, Bamann, Vavourakis, Ghai, Gushchin, Borshchevskiy, Rogachev, Melnikov, Popov, Balandin, Rodriguez-Valera, Manstein, Bueldt, Bamberg and Gordelly2017); the significance of this energized, but energetically downhill, flux is not clear. Some marine Archaea and bacteria have autotrophic CO2 assimilation pathways, but none of these are energized entirely by ion-pumping rhodopsins despite the possibility of such energization (Raven & Smith, Reference Raven and Smith1981; Raven, Reference Raven2009a, Reference Raven2009b; Berg et al., Reference Berg, Kockelhorn, Ramos-Vera, Say, Zarzycki, Hügler, Alber and Fuchs2010; Bejá & Lanyi, Reference Bejá and Lanyi2014; Larkum et al., Reference Larkum, Ritchie and Raven2018). A marine aerobic anoxygenic photoheterotrophic bacterium, with bacteriochlorophylls, also has an H+-pumping rhodopsin, presumably in the plasmalemma (Larkum et al., Reference Larkum, Ritchie and Raven2018). There seem to be no reports of ion-pumping rhodopsins in marine aerobic or anaerobic anoxygenic photosynthetic bacteria, or in marine photosynthetic cyanobacteria, although a freshwater/terrestrial cyanobacterium (Gloeobacter) has a H+ efflux rhodopsin, with respiratory and photosynthetic H+ pumps, in the plasmalemma (Larkum et al., Reference Larkum, Ritchie and Raven2018). In an analysis of the energetics of phototrophic marine picoplankton, ion-pumping rhodopsins are major energy-transducing pigments in the ocean in organisms that are small enough to pass through a 2 μm filter (Gómez-Consarnau et al., Reference Gómez-Consarnau, Raven, Levin, Cutter, Wang, Seegers, Arístegui, Fuhrman J, Gasel and Sañudo-Wilhelmy2019; see also Kirchman & Hanson, Reference Kirchman and Hanson2013).

Ion-pumping rhodopsins also occur in eukaryotes. Among marine photosynthetic eukaryotes, the best-characterized case of ion-pumping rhodopsin in the plasmalemma is in Acetabularia, which contains an H+-pumping rhodopsin. Paradoxically, this rhodopsin moves H+into the cytosol, rather than the expected direction of acting as a H+ efflux pump (Raven, Reference Raven2009a; Wada et al., Reference Wada, Shimono, Kikukawa, Hato, Shinya, Ki, Kimura-Somega, Shirouza, Taogami, Miyachi, Juing, Kamo and Yokoyama2011; Bejá & Lanyi, Reference Bejá and Lanyi2014; Tamogami et al., Reference Tamogami, Kikuyawa, Wada, Demura, Kimura-Someya, Shirouzo, Yokayama, Miauchi, Simono and Kamo2017; Larkum et al., Reference Larkum, Ritchie and Raven2018), i.e. in the same direction as the inwardly directed H+ pump driven by xenorhodopsin in the bacterium Nanosalina (Shevchenko et al., Reference Shevchenko, Mager, Kovalev, Polovinkin, Alekseev, Juettner, Chizhov, Bamann, Vavourakis, Ghai, Gushchin, Borshchevskiy, Rogachev, Melnikov, Popov, Balandin, Rodriguez-Valera, Manstein, Bueldt, Bamberg and Gordelly2017). The Cl ATPase in the plasmalemma of Acetabularia generates a Ψ CO of −180 mV, inside negative, in the light that, with a probable cytosol pH between 7 and 8, means an inwardly directed H+ electrochemical difference. This means that the H+-transporting rhodopsin ‘pumps’ H+ downhill. Other cases, where it is thought that the H+-pumping rhodopsin is in the plasmalemma and functions in H+ efflux, are found in some diatoms where it is thought to be an Fe-sparing way of energizing the plasmalemma (Raven, Reference Raven2009a; Slamovits et al., Reference Slamovits, Okamoto, Burri, James and Keeling2011; Marchetti et al., Reference Marchetti, Corlett, Hoplinson, Ellis and Cassar2015; Cohen et al., Reference Cohen, Ellis, Lampe, McNair, Twining, Maldonado, Brzezinski, Kizminov, Thamatrakoln, Till, Bruland, Sunda, Bangu and Marchetti2017; Larkum et al., Reference Larkum, Ritchie and Raven2018). In at least one case (the dinoflagellate Noctiluca which lacks oxygenic photosynthesis, except by symbiosis) the H+-pumping rhodopsin is found in a digestive vacuole; H+-pumping rhodopsins also occur in dinoflagellates with oxygenic photosynthesis (Slamovits et al., Reference Slamovits, Okamoto, Burri, James and Keeling2011; Vader et al., Reference Vader, Laughinghouse IV, Griffiths, Jakobsen and Gabrielsen2018).

It is clear that there is a significant variation in the phylogenetic analysis, and the richness of relevant data available among taxa. This is also the case for data considered below for organisms with bacteriochlorophylls and chlorophylls.

Organisms with bacteriochlorophylls

The anoxygenic photosynthetic bacteria function as photolithotrophs with reductants other than water in anaerobic environments or photoheterotrophically by assimilating CO2 with an organic reductant more oxidizing than H2 in anaerobic environments. In both of these cases autotrophic CO2 assimilation pathways are used (Larkum et al., Reference Larkum, Ritchie and Raven2018). These photolithotrophic organisms occupy geographically limited benthic marine anoxic illuminated habitats. The other marine bacteria with bacteriochlorophyll are the planktonic aerobic anoxygenic bacteria (Kolber et al., Reference Kolber, Gerald, Lang, Beatty, Blankenship, VanDover, Vetriani, Ratheber and Falkowski2001) that lack autotrophic CO2 assimilation pathways, and so are photoheterotrophs (Larkum et al., Reference Larkum, Ritchie and Raven2018; Gómez-Consarnau et al., Reference Gómez-Consarnau, Raven, Levin, Cutter, Wang, Seegers, Arístegui, Fuhrman J, Gasel and Sañudo-Wilhelmy2019). Bacteriochlorophylls are less significant in terms of photon absorption than ion-pumping rhodopsins or than chlorophylls as energy-transducing pigments in the ocean in organisms capable of passing through a 2 μm filter (Gómez-Consarnau et al., Reference Gómez-Consarnau, Raven, Levin, Cutter, Wang, Seegers, Arístegui, Fuhrman J, Gasel and Sañudo-Wilhelmy2019; see also Kirchman & Hanson, Reference Kirchman and Hanson2013).

Bacteriochlorophylls, like chlorophylls but unlike ion-pumping rhodopsins, are light energy transducers that indirectly bring about active transport of H+. The primary photochemistry moves an electron from a high potential electron donor on the P side of the membrane (sensu Mitchell, Reference Mitchell1979) to a low potential acceptor on the N side of the membrane (sensu Mitchell, Reference Mitchell1979) (equation (1)). H+ active transport from the P to the N side of the membrane is a result of secondary, thermodynamically downhill, redox reactions.

The freshwater and marine Chlorobi and Chloroflexi and, in inland hot springs, Acidobacteria, are photolithotrophic bacteria with light-harvesting bacteriochlorophyll c in chlorosomes. These organisms have no intracellular membranes, and have either Type 1 (Chlorobi, Acidobacteria) or Type 2 (Chloroflexi) reaction centres and associated redox catalysts, and the (photo-)redox H+ pumps and H+ electrochemical difference-driven FOF1 ATP synthases, are in their plasmalemma (Adams et al., Reference Adams, Cadby, Robinson, Tsukatani, Tank, Wen, Blankenship, Bryant and Hunter2013; Larkum et al., Reference Larkum, Ritchie and Raven2018). This plasmalemma location of photochemistry is also the case for the firmicute Heliobacterium without chlorosomes and with Type 1 reaction centres. H+ efflux across the plasmalemma is driven by light energy in the photoperiod, and by non-photochemical redox reactions in the scotophase. The H+ free energy difference across the plasmalemma in the light is large enough to synthesize ATP, granted the H+:ATP ratio of the bacterial FOF1 ATP synthase (Larkum et al., Reference Larkum, Ritchie and Raven2018).

The marine photosynthetic Proteobacteria with Type 2 reaction centres are the anaerobic photolithotrophic purple sulphur bacteria, and the anaerobic or aerobic photoheterotrophic purple non-sulphur bacteria. These organisms have their Type 2 reaction centres in plasmalemma invaginations and/or intracellular vesicles or flattened thylakoids (Larkum et al., Reference Larkum, Ritchie and Raven2018). These structural features mean that most, or all, of the photochemistry and associated H+ pumping is in membranes other than the plasmalemma that directly exchanges solutes with the bulk medium. It is, however, unclear what primary ion pumps occur in those parts of the plasmalemma that are not invaginated, and so are involved in nutrient uptake.

The marine anaerobic anoxygenic photosynthetic bacteria of the Chlorobi, Chloroflexi and Proteobacteria make a very small contribution (less than 0.1%) to global marine primary productivity (Johnson et al., Reference Johnson, Wolfe-Simon, Pearson and Knoll2009; Raven, Reference Raven2009b). Photons absorbed by marine aerobic anoxygenic photoheterotrophic bacteria probably make a larger contribution (0.5–5%) to global marine primary productivity than the anaerobic anoxygenic photolithotrophic bacteria (Kolber et al., Reference Kolber, VanDover, Niederman and Falkowski2000, Reference Kolber, Gerald, Lang, Beatty, Blankenship, VanDover, Vetriani, Ratheber and Falkowski2001; Goericke, Reference Goericke2002; Johnson et al., Reference Johnson, Wolfe-Simon, Pearson and Knoll2009; Raven, Reference Raven2009b; Kirchman & Hanson, Reference Kirchman and Hanson2013; Gómez-Consarnau et al., Reference Gómez-Consarnau, Raven, Levin, Cutter, Wang, Seegers, Arístegui, Fuhrman J, Gasel and Sañudo-Wilhelmy2019).

Organisms with chlorophylls

Cyanobacteria

Some phylogenetic evidence is consistent with a freshwater/terrestrial origin of cyanobacteria (Blank & Sánchez-Baracaldo, Reference Blank and Sánchez-Baracaldo2010; Blank, Reference Blank2013b; Sánchez-Baracaldo et al., Reference Sánchez-Baracaldo, Ridgwell and Raven2014). Data from freshwater cyanobacteria show that respiratory and photosynthetic redox and proton pumping reactions, and the associated CFOCF1 ATP synthase, occur in thylakoids. The exception is in Gloeobacter, where there are no thylakoids and photosynthesis and respiration, and ion-pumping rhodopsin, as well as nutrient transporters, occur in the plasmalemma (Mullineaux, Reference Mullineaux2014; Lea-Smith et al., Reference Lea-Smith, Bombelli, Vasudevan and Howe2016). The plasmalemma of thylakoid-containing freshwater Cyanobacteria has vanadate-sensitive (presumably ATP-driven) H+ efflux (Scherer & Böger, Reference Scherer and Böger1984; Kaplan et al., Reference Kaplan, Scherer and Lerner1989; Schultze et al., Reference Schultze, Forberich, Rexroth, Dyczmons, Roegner and Appel2009), and can oxidize NAD(P)H and reduce O2; and contains plastoquinone (PQ) and ATP synthase (Mullineaux, Reference Mullineaux2014; Lea-Smith et al., Reference Lea-Smith, Bombelli, Vasudevan and Howe2016), and moves protons out of the cell (Scherer et al., Reference Scherer, Stürzl and Böger1984). While this is consistent with primary active H+ efflux at the plasmalemma in freshwater cyanobacteria, the work of Ritchie (Reference Ritchie1992) shows that there is an electrogenic Na+ efflux pump at the plasmalemma of Synechococcus R2, and that the H+ efflux involves a Na+:H+ antiporter. This Na+ active efflux occurs over a wide range of external pH (5–10), K+ (0.1–300 mol m−3) and Na+ (0.1–300 mol m−3). The halophilic, alkalophilic cyanobacterium Aphanothece halophytica has a Na+-dependent FOF1 ATP synthase which, working as an ATPase in the plasmalemma, could act in active Na+ efflux (Wiangno et al., Reference Wiangno, Raksajit and Incharoensakdi2007; Soontharapirakkul & Incharoensakdi, Reference Soontharapirakkul and Incharoensakdi2010; Soontharapirakkul et al., Reference Soontharapirakkul, Promden, Yamada, Kageyama, Incharoensakdi, Iwamoto-Kihara and Takabe2011; see prediction by Mitchell, Reference Mitchell1979). Gabbay-Azaria et al. (2000) demonstrated cytochrome oxidase activity in the plasmalemma of the marine cyanobacterium Spirulina subsalsa (now Arthrospira subsalsa) that could be involved in active H+ efflux (Gabbay-Azaria et al., Reference Gabbay-Azaria, Schonfeld, Tel-Or, Messinger and Tel-Or1992). Bergman et al. (Reference Bergman, Siddiqui, Carpenter and Pescher1993) found cytochrome oxidase in the plasmalemma as well as the thylakoid membranes of the marine diazotrophic cyanobacterium Trichodesmium thiebaultii. A P-type Ca2+ ATPase has been found in a marine cyanobacterium that can excavate solid CaCO3 (Garcia-Pichel et al., Reference Garcia-Pichel, Ramirez-Reinat and Guo2010).

Na+ symport of HCO3 is one of the means of concentrating inorganic C in freshwater cyanobacterial cells, although these organisms also have other HCO3 transporters, one of which is an ABC transporter (Omata et al., Reference Omata, Price, Badger, Okamura, Gohta and Ogawa1999; Badger & Price, Reference Badger and Price2003). Freshwater cyanobacteria also use ABC transporters for NO3 (Maeda & Omata, Reference Maeda and Omata1997). However, the HCO3 and NO3 transporters in marine cyanobacteria are not ABC transporters such as occur in their freshwater counterparts (Wang et al., Reference Wang, Li and Post2000; Badger & Price, Reference Badger and Price2003; Maeda et al., Reference Maeda, Murakami, Ito, Tanaka and Omata2015).

Eukaryotic algae: relation to the supergroups in the New Tree of Life

Almost all eukaryotic photolithotrophs have a photosynthetic apparatus most of whose core genes were derived from endosymbiosis of a freshwater β-cyanobacterium with a non-photosynthetic unicellular eukaryote to comprise the Archaeplastida (Burki et al., Reference Burki, Roger, Brown and Simpson2020; Palmgren et al., Reference Palmgren, Sorenson, Hallström, Säll and Broberg2020). ‘Green’ Archaeplastida with chlorophyll b produced, by endosymbiosis in an Excavate cell, the secondarily photosynthetic Euglenophyta and, by endosymbiosis in a cell of the Rhizaria in the TSAR (Telonemia, Stramenopila, Alveolata, Rhizaria) supergroup, the secondarily photosynthetic Chlorarachniophyta. In a further set of secondary endosymbioses, ‘Red’ Archaeplastida with phycobilins produced, in cells of the Cryptista, the phycobilin-containing Cryptophyta and, with the loss of light-harvesting phyobilins, in cells of the Haptista to produce the Haptophyta, and in cells of the TSAR supergroup to produce the chromerids and dinoflagellates in the Alveolata, and the Ochrista in the Stramenopila. A very small minority of photolithotrophic eukaryotes (the genus Paulinella) arose by endosymbiosis of an α-cyanobacterium in a rhizarian (TSAR supergroup) (Nowack, Reference Nowack2014).

Eukaryotic algae: Archaeplastida (Glaucophyta, Rhodophyta, Chlorophyta, algal members of the Streptophyta)

The primary endosymbiosis of a β-cyanobacterium in an aerobic eukaryote that gave rise to chloroplasts of the Archaeplastida involved a cyanobacterium whose closest living relative is the freshwater Gloeomargarita lithophora (Brasier, Reference Brasier2013; Blank, Reference Blank2013a; Lewis, Reference Lewis2017; Ponce-Toledo et al., Reference Ponce-Toledo, Deschamps, Lôpez-García, Zivanovic, Benzenara and Moreira2017; Sánchez-Baracaldo et al., Reference Sánchez-Baracaldo, Raven, Pisani and Knoll2017). The basal extant Archaeplastida are the freshwater Glaucophyta; the basal Rhodophyta, the Cyanidiophyceae, are non-marine, but their habitat of acid hot springs cannot be described as freshwater (Dittami et al., Reference Dittami, Heesch, Olsen and Collén2017). There have been numerous freshwater–marine transitions in the Rhodophyta and, especially, in the Chlorophyta, but less so in the predominantly freshwater algal Streptophyta (Dittami et al., Reference Dittami, Heesch, Olsen and Collén2017). These habitat variations make it difficult to predict which primary active ion transporters occur in the plasmalemma of these organisms.

In the Rhodophyta, the non-marine (acid hot spring) Cyanidium caldarium, Cyanidioschyzon merolae and Galdieria sulphuraria (Cyanidiophyceae) have P-type H+-ATPases but not Na+-ATPases (Ohta et al., Reference Ohta, Shirakawa, Uchida, Yoshida, Matuo and Ehami1997; Lee et al., Reference Lee, Ghosh and Saier2017). Two Na+-ATPases (PyKPA1 and PyKPA2) have been characterized in the marine bangiophycean Porphyra yezoensis (now Pyropia yezoensis) (Barrero-Gil et al., Reference Barrero-Gil, Garciadeblás and Benito2005; Uji et al., Reference Uji, Hirata, Mikami, Mizuto and Saga2012a, Reference Uji, Monma, Mizuta and Saga2012b; see also Chan et al., Reference Chan, Zäuner, Wheeler, Grossman, Prochnik, Blouin, Zhuang, Benning, Berg, Yarish, Ekiksen, Klein, Lin, Levine, Brawley and Bhattachaya2012; Kishimoto et al., Reference Kishimoto, Shimajiri, Oshima, Hase, Mikami and Akama2013). This organism also has two plasmalemma Na+/H+ antiporters (Uji et al., Reference Uji, Monma, Mizuta and Saga2012b). An Na+:H+ antiporter has also been described in Pyropia haitanensis by Chen et al. (Reference Chen, Wang, Xu, Xu, Ji, Chen and Xie2019), who assume, on the basis of vanadate inhibition, that there is a H+-ATPase at the plasmalemma of this alga. However, vanadate inhibition is the case for all P and F ATPases (Müller et al., Reference Müller, Jensen and Taiz1999; Araki & González, Reference Araki and González1998; Kuhlbrandt, Reference Kuhlbrandt2004; Hong & Pedersen, Reference Hong and Pedersen2008; Pedersen et al., Reference Pedersen, Axelsen, Harper and Palmgren2012), but to a much smaller extent for V ATPases (Müller et al., Reference Müller, Irkens-Kiesecker, Rubinstein and Taiz1996; Araki & González, Reference Araki and González1998). Accordingly, vanadate is a general inhibitor of H+, Na+ and Ca2+ P-ATPases, and is not specific for H+ ATPases. Reed et al. (Reference Reed, Collins and Russell1981) and Reed & Collins (Reference Reed and Collins1981) examined the energetics of ion transport in Porphyra purpurea over a wide range of hypo- and hyper-saline (relative to seawater) media. Under all salinities (1/16 seawater to 3× seawater) there is active Na+ efflux (Reed et al., Reference Reed, Collins and Russell1981). It should be noted that the measurement of Ψ CO in the essentially non-vacuolate Porphyra purpurea cells by Reed and co-workers was determined using the distribution of the lipophilic cation TPMP+; there are reservations about the use of this method of measuring transplasmalemma electrical potential differences in eukaryotes (Ritchie, Reference Ritchie1982, Reference Ritchie1984). The floridiophycean Chondrus crispus has a P-type Na+-ATPase but no H+-ATPase (Lee et al., Reference Lee, Ghosh and Saier2017). There seem to be no relevant data for freshwater red algae.

In the Chlorophyta, the microalgal Prasinophyceae occur in marine and also freshwater habitats (Tragin & Vaulot, Reference Tragin and Vaulot2018; Del Cortona et al., Reference Del Cortona, Jackson, Bucchini, Van Bel, D'hondt, Škaloud, Delwiche, Knoll, Raven, Verbruggen, Vandepoel, De Clerck and Leliaert2020). There is functional evidence for the presence in the plasmalemma of the marine Tetraselmis (=Platymonas) viridis of an ATP-driven Na+ pump (Balnokin & Popova, Reference Balnokin and Popova1994; Balnokin et al., Reference Balnokin, Popova and Gimmler1997, Reference Balnokin, Popova and Andreev1999, Reference Balnokin, Popova, Pagis and Andreev2004; Gimmler, Reference Gimmler2000; Pagis et al., Reference Pagis, Popova, Andreev and Balnokin2001, Reference Pagis, Popova, Andreev and Balnokin2003; Popova & Balnokin, Reference Popova and Balnokin2013) and an ATP-driven H+ pump (Popova & Balnokin, Reference Popova and Balnokin1992; Gimmler, Reference Gimmler2000; Pagis et al., Reference Pagis, Popova, Andreev and Balnokin2003). There is genomic evidence of a P-type Na+-ATPase in the marine Ostreococcus tauri (Rodríguez-Navarro & Benito, Reference Rodríguez-Navarro and Benito2010). Gimmler (Reference Gimmler2000) provides an energetic background to infer the presence of active H+ and Na+ efflux in Tetraselmis viridis, although there are no data from that organism on the electrical potential difference across the plasmalemma, or Na+ and H+ concentrations in the cytoplasm.

The Chlorophyta: Chlorophyceae are mainly freshwater but with some marine representatives (Tragin & Vaulot, Reference Tragin and Vaulot2018; Del Cortona et al., Reference Del Cortona, Jackson, Bucchini, Van Bel, D'hondt, Škaloud, Delwiche, Knoll, Raven, Verbruggen, Vandepoel, De Clerck and Leliaert2020), the P-type ATPases of the genus Dunaliella has been the subject of considerable investigation (Wolf et al., Reference Wolf, Slayman and Gradmann1995; Weiss & Pick, Reference Weiss and Pick1996; Popova et al., Reference Popova, Belyaev, Shuvalov, Yurchenko, Matalin, Khramov, Orlova and Balnokin2018). Sequences encoding P-type H+-ATPases have been found in the genomes of the halophilic Dunaliella bioculata (Smahel et al., Reference Smahel, Hamann and Gradmann1990; Wolf et al., Reference Wolf, Slayman and Gradmann1995), Dunaliella salina (Weiss & Pick, Reference Weiss and Pick1996; Katz et al., Reference Katz, Waridel, Shevchenko and Pick2007), Dunaliella bioculata (Bertucci et al., Reference Bertucci, Tambutté, Tambutté, Allemand and Zoccola2010) and Dunaliella tertiolecta (Popova et al., Reference Popova, Belyaev, Shuvalov, Yurchenko, Matalin, Khramov, Orlova and Balnokin2018), as well as the acidophilic Dunaliella acidophila (Weiss & Pick, Reference Weiss and Pick1996; Bertucci et al., Reference Bertucci, Tambutté, Tambutté, Allemand and Zoccola2010). While Popova et al. (Reference Popova, Belyaev, Shuvalov, Yurchenko, Matalin, Khramov, Orlova and Balnokin2018) could not find a P-type Na+-ATPase in the genome of Dunaliella tertiolecta, Popova et al. (Reference Popova, Shumkova, Andreev and Balnokin2005; see also Shumkova et al., Reference Shumkova, Popova and Balnokin2000) functionally identified an electrogenic Na+-translocating ATPase in Dunaliella maritima using inside-out plasmalemma vesicles (see also Popova & Balnokin, Reference Popova and Balnokin2013). The possibility of a primary H+ pump with H+:Na+ antiport was ruled out by the observation of stimulation, rather than inhibition, of Na+ transport by the uncoupler CCCP (m-chlorophenyl carbonyl cyanide phenylhydrazone). Popova et al. (Reference Popova, Shumkova, Andreev and Balnokin2005) did not attempt to identify whether the catalyst of Na+ transport was a P-type ATPase. Further investigation is needed into the possibility of direct redox energization of Na+ efflux from Dunaliella salina (Katz & Pick, Reference Katz and Pick2001). The freshwater Chlamydomonas reinhardtii has a P-type Na+-ATPase (Barrero-Gil et al., Reference Barrero-Gil, Garciadeblás and Benito2005; Rodríguez-Navarro & Benito, Reference Rodríguez-Navarro and Benito2010), as well as a P-type H+-ATPase (Campbell et al., Reference Campbell, Coble, Cohen, Ch'ng, Russo, Long and Armbrust2001; Barrero-Gil et al., Reference Barrero-Gil, Garciadeblás and Benito2005; Bertucci et al., Reference Bertucci, Tambutté, Tambutté, Allemand and Zoccola2010).

An important aspect of the functioning of primary active transport of H+ and Na+ in driving symport, antiport and uniport of other solutes by Dunaliella is the free energy difference across the plasmalemma. There are several reports of intracellular Na+ concentrations (Katz & Avron, Reference Katz and Avron1985; Bental et al., Reference Bental, Degani and Avron1986; Pick et al., Reference Pick, Kanni and Avron1986; Wegmann, Reference Wegmann1986; Gimmler, Reference Gimmler2000), cytoplasmic pH (Burns & Beardall, Reference Burns and Beardall1987; Gimmler et al., Reference Gimmler, Kugel, Leibfritz and Mayer1988; Ginzburg et al., Reference Ginzburg, Ratcliffe and Southon1988; Gimmler, Reference Gimmler2000) and Ψ CO (Gimmler & Greenway, Reference Gimmler and Greenway1983; Oren-Shamir et al., Reference Oren-Shamir, Pick and Avron1990; see the critique by Ritchie, Reference Ritchie1982, Reference Ritchie1984) in a variety of halophilic Dunaliella species. Ψ CO of Dunaliella acidophila has been measured using the preferable microelectrodes method (Remis et al., Reference Remis, Simonis and Gimmler1992), but not so far for halophilic Dunaliella species. The thermodynamic analysis by Gimmler (Reference Gimmler2000) dealt with Dunaliella salina and suggested active efflux of both H+ and Na+, and pre-dates the discovery of the electrogenic Na+ ATPase in Dunaliella maritima (Popova et al., Reference Popova, Shumkova, Andreev and Balnokin2005; see also Shumkova et al., Reference Shumkova, Popova and Balnokin2000). Khramov et al. (Reference Khramov, Matalin, Karpichev, Balnokin and Popova2019) showed increased transcript abundance of a putative H+ P-ATPase under hypersaline conditions in Dunaliella maritima.

As with the Chlorophyceae, the Chlorophyta: Trebouxiophyceae are mainly freshwater but with some marine members (Tragin & Vaulot, Reference Tragin and Vaulot2018; Del Cortona et al., Reference Del Cortona, Jackson, Bucchini, Van Bel, D'hondt, Škaloud, Delwiche, Knoll, Raven, Verbruggen, Vandepoel, De Clerck and Leliaert2020). There is genomic evidence of both Na+-ATPase and H+-ATPases in the halotolerant Picochlorum sp. (Foflonker et al., Reference Foflonker, Price, Qiu, Palenik, Wang and Bhattacharya2014). The only contribution to a thermodynamic analysis of the need for active H+ and/or Na+ efflux is the work of Bock et al. (Reference Bock, Jacob, Kirst, Leibfritz and Mayer1996) who, using 31P NMR, found a cytoplasmic pH of 7.8 in the high intertidal–supralittoral macroalga Prasiola crispa. For a freshwater Chlorella sp. there is genomic evidence of a Na+ -ATPase (Uji et al., Reference Uji, Hirata, Mikami, Mizuto and Saga2012a), and functional evidence consistent with a H+-ATPase (Komor et al., Reference Komor, Cho, Schricker and Schobert1989).

The earliest Chlorophyta: Ulvophyceae fossils are from marine sediments, and most extant ulvophyceans are marine, exceptions being the freshwater Dichotomosiphon and some species of Cladophora and the terrestrial Trentepohliales (Del Cortona et al., Reference Del Cortona, Jackson, Bucchini, Van Bel, D'hondt, Škaloud, Delwiche, Knoll, Raven, Verbruggen, Vandepoel, De Clerck and Leliaert2020). Blount & Levedahl (Reference Blount and Levedahl1960) found active electrogenic Na+ efflux and Cl influx in the marine gametophyte Halicystis ovalis phase of the marine ulvophycean Derbesia marina. There is genomic evidence for a P-type Na+ ATPase of the marine ulvophycean Flabellia petiolata (formerly Udotea petiolata) (Rodríguez-Navarro & Benito, Reference Rodríguez-Navarro and Benito2010), and for a number of P-type ATPases in Ulva (Zhang et al., Reference Zhang, Ye, Liang, Mou, Fan, Xu, Xu and Zhuang2012; De Clerck et al., Reference De Clerck2018).

Blinks (Reference Blinks1940, Reference Blinks1949) showed that Halicystis ovalis (the giant-celled coenocytic gametophyte phase of Derbesia marina) had a potential difference between the vacuole and seawater medium (Ψ VO) of −75 to −80 mV; for Halicystis osterhoutii (the gametophyte phase of Derbesia osterhoutii) the value is −65 to −70 mV. Replacing external Cl with NO3 (or SO42−, or a number of organic anions) caused the potential difference to become negligible, or positive. This is consistent with the short-circuit current experiments of Blount & Levedahl (Reference Blount and Levedahl1960) showing the occurrence of an electrogenic Cl influx pump in Halicystis ovalis. Graves & Gutknecht (Reference Graves and Gutknecht1976, Reference Graves and Gutknecht1977a, Reference Graves and Gutknecht1977b) examined ion concentrations and fluxes, and the effects of decreased external Cl and low temperatures and of clamped Ψ VO, in the gametophyte Halicystis parvula phase of Derbesia tenuissima, and also concluded that there is an electrogenic Cl influx pump. Graves & Gutknecht (Reference Graves and Gutknecht1976, Reference Graves and Gutknecht1977a, Reference Graves and Gutknecht1977b) infer that the Cl pump is at the plasmalemma, as is the case for Acetabularia, discussed below.

Saddler (Reference Saddler1970a, Reference Saddler1970b) measured ion (K+, Na+, Cl) fluxes between uninucleate giant-celled marine Acetabularia and the seawater medium, and electrical properties of the cell, concluding that there was an electrogenic Cl influx pump. The occurrence of active electrogenic Cl influx was confirmed by Mummert & Gradmann (Reference Mummert and Gradmann1976) and by Gradmann et al. (Reference Gradmann, Tittor and Goldfarb1982). The Ψ CO in the light is −170 mV (Saddler, Reference Saddler1970a) to −180 mV (Amtmann & Gradmann, Reference Amtmann and Gradmann1994). Saddler (Reference Saddler1970a, Reference Saddler1970b) suggested that cytoplasmic Cl in Acetabularia is the same (500 mol m−3) as that in seawater, although this is not consistent with observed inhibitory effects of Cl on metabolism (Raven, Reference Raven2017). Using the value of 500 mol m−3 for cytoplasmic Cl, the electrochemical potential gradient for Cl is 17–18 kJ mol−1, driving Cl from the cytoplasm to the medium (Saddler, Reference Saddler1970a, Reference Saddler1970b). Mummert & Gradmann (Reference Mummert and Gradmann1991a) showed the role of Cl in the action potential of Acetabularia, and also (Mummert & Gradmann, Reference Mummert and Gradmann1991a, Reference Mummert and Gradmann1991b) discovered the role of vesicular transport of ions across the cytosol to the vacuole. The Cl influx is driven by a Cl-ATPase (Gradmann et al., Reference Gradmann, Tittor and Goldfarb1982; Goldfarb & Gradmann, Reference Goldfarb and Gradmann1983; Ohhashi et al., Reference Ohhashi, Katsu and Ikeda1992) inhibited by vanadate (Smahel et al., Reference Smahel, Hamann and Gradmann1992; see discussion of vanadate inhibition of primary active ion-pumping ATPase under ‘Rhodophyta’ above). Goldfarb et al. (Reference Goldfarb, Sanders and Gradmann1984) reported that the Cl pump can be reversed, coupled to net ATP synthesis (as in the prediction of Mitchell, Reference Mitchell1979). Ikeda et al. (Reference Ikeda, Schmid and Oesterhelt1990a, Reference Ikeda, Schmid and Oesterhelt1990b) showed that the Cl-ATPase was not identical with the (C)F ATPase of the FOF1/CFOCF1 ATP synthase, but Ikeda et al. (Reference Ikeda, Kadowaki, Ikeda, Moritani and Kanazawa1997) demonstrated the interchangeability of the b subunit of the Acetabularia Cl-ATPase and the β-subunit of the Escherichia coli F-ATPase. Finally, Moritani et al. (Reference Moritani, Ohnashi, Kadowaki, Tagaya, Lottspeich, Oesterhelt and Ikeda1997) determined the primary structure of the b subunit of the Acetabularia Cl-ATPase. There seems to have been no further work on this Cl-ATPase other than studies on action potentials (Raven, Reference Raven2017).

For the major ions commonly subject to primary active transport, Na+ and H+, the cytoplasmic Na+ concentration in Acetabularia is 60 mol m−3 (Amtmann & Gradmann, Reference Amtmann and Gradmann1994), and the cytoplasmic pH is estimated at 8.0–8.4 using pH indicators, and pH 7.6–7.7 from the pH at which isolated intact chloroplasts exhibit their highest rate of photosynthesis (Dodd & Bidwell, Reference Dodd and Bidwell1971). If external Na+ is 450 mol m−3, the internal:external Na+ concentration difference is 0.133, the electrochemical potential difference for Na+ is 22 kJ mol−1 and will drive Na+ into the cell. For H+, the indicator dye-measured cytoplasmic pH of 8.0–8.2 (Dodd & Bidwell, Reference Dodd and Bidwell1971) is essentially identical to that of seawater, so the electrochemical potential difference for H+ is 17 kJ mol−1, driving H+ into the cell. No evidence is available for the occurrence of primary active transport processes at the plasmalemma of Acetabularia other than the inward Cl pump and the inward H+ pump. If secondary active efflux of Na+ is driven by Cl symport, granted the electrochemical potentials for the two ions calculated above, the Cl:Na+ ratio must be not less than 2. Active H+ efflux cannot be driven by the inwardly directed H+-pumping rhodopsin. Since the calculated electrochemical potential differences for H+ and Cl are equal and opposite, net H+ efflux driven by Cl efflux requires a ratio of Cl:H+ > 1.0.

Other large-celled ulvophyceans include the Siphonocladales and Cladophorales, mainly marine and comprising one or more coenocytic cells. In the cases examined, the vacuole-positive electrical potential difference across the tonoplast is much greater than in other vacuolated organisms, in some cases making the inside-positive tonoplast–medium electrical potential difference greater than the inside-negative cytosol–medium electrical potential difference (Hope & Walker, Reference Hope and Walker1975). There are data for the giant-celled marine Chaetomorpha darwinii (now Chaetomorpha coliforme) on the Ψ CO of −72 mV, cytoplasmic Na+ concentration (25 mol m−3) and cytoplasmic pH (pH 8.0–8.3 in the light; pH 7.5−7.8 in the dark), with external Na+ 500 mol m−3 and external pH 8.0. The free energy difference across the plasmalemma, cytosol relative to medium, for H+ is −7.6 kJ mol−1, driving H+ into the cell, and −12.6 kJ mol−1 for Na+, also driving Na+ into the cell (Dodd et al., Reference Dodd, Pitman and West1966; Findlay et al., Reference Findlay, Hope, Pitman, Smith and Walker1971; Raven and Smith, Reference Raven and Smith1980), in the light. No data are available for electrical potentials, or intracellular ion concentrations, in the dark. Less information is available for the closely related large-celled alga, i.e. species of Valonia, Valoniopsis and Ventricaria (Bisson et al., Reference Bisson, Beilby and Shepherd2006), but the available information is consistent with a situation similar to that in Chaetomorpha darwinii.

Finally among the Ulvophyceae, there are important data on Ulva spp., now incorporating the genera Ulva and Enteromorpha (Hayden et al., Reference Hayden, Blomster, Maggs, Silva, Stanhope and Walland2003). Ulva is multicellular, with small cells of which half or less of the volume is taken up by a vacuole. Ritchie (Reference Ritchie1985) examined the energetics of ion transport in Enteromorpha (=Ulva) intestinalis. The cytoplasmic pH was not measured in that study so Ritchie (Reference Ritchie1985) used a value of pH 7.3 from a wide range of studies on (mostly non-marine) cyanobacteria, eukaryotic algae and plants, with an external pH of 8.0. With the measured Ψ CO of −54 ± 5 mV in seawater in the light, and −30 ± 5 mV in the dark, the proton electrochemical potential difference across the plasmalemma is not significantly different from zero (Ritchie, Reference Ritchie1985). For Na+, the ion most likely to be subject to primary active efflux in Ulva, the electrochemical potential difference (cytosol relative to medium) across the plasmalemma is −15.5 kJ mol−1 in the light and −13.1 kJ mol−1 in the dark (Ritchie, Reference Ritchie1985).

There are also data for the closely related Ulva lactuca. Reed & Collins (Reference Reed and Collins1981) used the distribution of the lipid-soluble cation TPMP+ to measure Ψ CO for Ulva lactuca cells in seawater; they found values of −54 ± 1.8 mV in the light and −44 ± 3.5 mV in the dark. Ritchie (Reference Ritchie1988), using microelectrodes, showed that Ψ CO of U. lactuca in seawater was −39 ± 1.1 mV in the light and −25 ± 1.9 mV in the dark. In view of the comments by Ritchie (Reference Ritchie1982, Reference Ritchie1984) on problems with the use of lipid-soluble cations to estimate the electrochemical potential difference between cells and the medium in eukaryotes, the following calculations use the electrical potential difference values of Ritchie (Reference Ritchie1988). Thus, using intracellular and extracellular Na+ concentrations, we calculate the electrochemical potential difference for Na+ (cytosol relative to medium) across the plasmalemma to be −13.8 kJ mol−1 in the light and −12.2 kJ mol−1 in the dark. Ritchie (Reference Ritchie1988) does not give values for the electrochemical potential differences for H+ across the plasmalemma, but with the assumption made by Ritchie (Reference Ritchie1985) for cytoplasmic pH (7.3), the H+ electrochemical potential difference (cytosol relative to medium) across the plasmalemma is +0.30 kJ mol−1 in the light and +1.65 kJ mol−1 in the dark. The choice of using the cytoplasmic pH value of Ritchie (Reference Ritchie1988) is supported by Lundberg et al. (Reference Lundberg, Welch, Jensén and Vogel1989) who, using 31P NMR, found a cytoplasmic pH of 7.2 in U. lactuca.

The algal members of the Streptophyta are the Charophyceae sensu lato (Del Cortona et al., Reference Del Cortona, Jackson, Bucchini, Van Bel, D'hondt, Škaloud, Delwiche, Knoll, Raven, Verbruggen, Vandepoel, De Clerck and Leliaert2020). The most morphologically complex of the Charophyceae are the Charales with most of the thallus volume occupied by giant cells (Beilby, Reference Beilby2015; Nishiyama et al., Reference Nishiyama2018). Most of the Charales are freshwater, although some species occur in brackish waters (e.g. in the Baltic) and Lamprothamnium spp. grows in coastal lagoons with very large changes in salinity, with the highest values twice that of seawater (Beilby, Reference Beilby2015). Energization of transport at the plasmalemma of Lamprothamnium, like that of freshwater Charales, involves a P-type H+ efflux ATPase, with active Na+ efflux driven by H+ antiport and active Cl influx driven by H+ symport (Beilby, Reference Beilby2015). The electrochemical potential difference across the plasmalemma for H+, with external pH of 8 and cytoplasmic pH of 7.7 and a ΨCO of −160 mV (Kirst & Bisson, Reference Kirst and Bisson1982) at 25°C, is −13.7 kJ mol−1. For the major ions the electrochemical potential differences, cytosol relative to medium, are −24.2 kJ mol−1 (Na+), −3.14 kJ mol−1 (K+) and +9.27 kJ mol−1 (Cl) (Kirst & Bisson, Reference Kirst and Bisson1982). The primary active efflux of H+ energizes, directly or indirectly, the active efflux of Na+ and K+ and active influx of Cl.

Eukaryotic algae: organisms with secondary and tertiary chloroplast endosymbiosis

The earliest fossils of the diatoms (Bacillariophyceae sensu lato; Ochrophyta) are from marine habitats, with later invasion of fresh waters (Falkowski et al., Reference Falkowski, Katz, Knoll, Quigg, Raven, Schofield and Taylor2004; Siver et al., Reference Siver, Velez, Clivetti and Binda2018). There is functional evidence of a plasmalemma Na+-ATPase in the colourless marine diatom Nitzschia alba (Bhattacharya & Volcani, Reference Bhattacharya and Volcani1980). The work of Flynn et al. (Reference Flynn, Öpik and Syrett1987) on plasmalemma vesicles of Phaeodactylum tricornutum is also consistent with the occurrence of an Na+-ATPase. The available evidence on cytoplasmic pH is for pH 7.6 in a non-vacuolate marine pennate diatom Phaeodactylum tricornutum (Burns & Beardall, Reference Burns and Beardall1987), and pH 7.3 in the centric marine diatom Thalassiosira weissflogii at an external pH of 8 (Hervé et al., Reference Hervé, Derr, Douady, Quinet, Moisan and Lopez2002). The Ψ CO is −60 to −90 mV for the marine centric diatom Coscinodiscus wailesii (Gradmann & Boyd, Reference Gradmann and Boyd1995, Reference Gradmann and Boyd1999a, Reference Gradmann and Boyd1999b) and −84 mV for the marine centric diatom Odontella sinensis (Taylor et al., Reference Taylor, Brownlee and Wheeler2017). These values for a range of marine diatoms are consistent with the occurrence of active H+ efflux. For Na+ the only intracellular concentration values are for whole cells of the marine Coscinodiscus granii (46 mol m−3) and Coscinodiscus wailesii (125 mol m−3) (Kesseler, Reference Kesseler1974). Boyd & Gradmann (Reference Boyd and Gradmann1999) assume that these mean intracellular concentrations, dominated by the concentration in the largest cell compartment, the vacuole, also apply to the cytosol. With this assumption, and Ψ CO of −75 mV across the plasmalemma in the marine C. wailesii (Gradmann & Boyd, Reference Gradmann and Boyd1999a, Reference Gradmann and Boyd1999b), the Na+ free energy difference across the plasmalemma is 12 kJ mol−1, inside negative. Jones & Morel (Reference Jones and Morel1988) suggested that redox reactions in the plasmalemma of Thalassiosira weissflogii can act as an H+ efflux pump. Bertucci et al. (Reference Bertucci, Tambutté, Tambutté, Allemand and Zoccola2010) cite genomic evidence for a P-type H+-ATPase in the diatom Phaeodactylum tricornutum. Diatoms also have vacuolar H+ pyrophosphatases and V-type H+-ATPases (Bussard & Lopez, Reference Bussard and Lopez2014). While there is evidence from metazoans for the expression of V-type H+-ATPases in the plasmalemma (Beyenbach & Wieczorek, Reference Beyenbach and Wieczorek2006), there is no evidence as to whether this occurs in diatoms (Wieczorek et al., Reference Wieczorek, Brown, Grinstein, Ehrenfeld and Harvey1999; Bussard & Lopez, Reference Bussard and Lopez2014).

Almost all of the extant Ochrophyta: Phaeophyceae are marine. For the marine phaeophycean Ectocarpus siliculosus there is genomic evidence for a P Na+-ATPase (Uji et al., Reference Uji, Hirata, Mikami, Mizuto and Saga2012a, Reference Uji, Monma, Mizuta and Saga2012b). The only estimates of the electrochemical potential differences across the plasmalemma of the Phaeophyceae are for the development of the just-fertilized eggs of the fucoid Pelvetia fastigiata (Allen et al., Reference Allen, Jacobsen, Joaquen and Jaffe1972; Gibbon & Kropf, Reference Gibbon and Kropf1993) and the unfertilized eggs of the fucoid Fucus serratus (Taylor & Brownlee, Reference Taylor and Brownlee1993). Gibbon & Kropf (Reference Gibbon and Kropf1993) used microelectrodes to measure the Ψ CO of −60 mv, inside negative), and the pH difference (0.5 units, inside low) between the cytosol and seawater medium. Gibbon & Kropf (Reference Gibbon and Kropf1993) calculated the proton motive force across the plasmalemma, equivalent to an electrochemical potential difference, cytosol relative to medium for H+ of −2.5 kJ mol−1, i.e. close to electrochemical equilibrium. The data for ion content of Allen et al. (Reference Allen, Jacobsen, Joaquen and Jaffe1972) are for the whole zygote, rather than the cytosol, so calculations of electrochemical potential difference across the plasmalemma for K+, Na+ and Cl are less accurate than the value for H+ (Gibbon & Kropf, Reference Gibbon and Kropf1993). The calculations of Gibbon & Kropf (Reference Gibbon and Kropf1993) show that while K+ and Cl are near electrochemical equilibrium, the electrochemical potential difference, cytosol relative to medium, for Na+ is −14 kJ mol−1. Gibbon & Kropf (Reference Gibbon and Kropf1993) therefore suggest that the primary active transport at the plasmalemma is for Na+, with Na+:H+ antiport generating the much smaller H+ electrochemical potential difference. Taylor & Brownlee (Reference Taylor and Brownlee1993) examined the electrical properties of the plasmalemma, and the ion content, of unfertilized eggs of Fucus serratus in seawater. The Ψ CO using microelectrodes is −40 to −65 mV, with a mean of −50 mV. With the same cautions as for Pelvetia fastigiata zygotes, the driving force for H+ is 1 kJ mol−1 directed inwards, for Na+ 12 kJ mol−1, directed inwards, and for Cl, zero kJ mol−1, i.e. equilibrium. These values are consistent with the occurrence of a Na+ efflux, with very weak evidence of a H+ efflux pump, and for Cl at equilibrium. However, replacement of external Cl by isethionate makes the plasmalemma electrical potential difference 20 mV less negative, consistent with active electrogenic Cl influx, and a lower Cl concentration in the cytosol than in some other intracellular compartments. Further experiments are needed to examine this possibility.

Klenell et al. (Reference Klenell, Snoeijs and Pedersén2002, Reference Klenell, Snoeijs and Pedersén2004) suggested that Laminaria digitata and Laminaria saccharina (now Saccharina latissima) have a P-type H+-ATPase that is involved in acidifying part of the thallus surface; however, the evidence (inhibition by vanadate) does not distinguish a P-type H+ ATPase from a P-type Na+-ATPase in parallel with a H+:Na+ antiporter (see discussion above for Rhodophyta). Klenell et al. (Reference Klenell, Snoeijs and Pedersén2004) also used erythrosin B as an inhibitor of the plasmalemma P-type H+-ATPase; however, erythrosin B also inhibits a range of other processes (Gimmler, Reference Gimmler1988).

In the Ochrophyta: Raphidophyceae, inverted plasmalemma vesicles of the marine Heterosigma akashiwo show ATP-dependent Na+ accumulation. The Na+ accumulation was inhibited by vanadate, and was shown not to be an accumulation down an inside-negative electrical potential, or a result of Na+:H+ exchange following H+ accumulation by an H+-ATPase (Shono et al., Reference Shono, Wada and Fujii1995, Reference Shono, Hara, Wada and Fujii1996). These findings are supported by genomic evidence of a Na+-ATPase (Shono et al., Reference Shono, Wada, Hara and Fuji2001; Jo et al., Reference Jo, Shono, Wada, Ito, Nomoto and Hara2010; Uji et al., Reference Uji, Hirata, Mikami, Mizuto and Saga2012a).

For the Haptophyta the earliest known fossils of the calcified coccoliths of coccolithophores in the Class Prymnesiophyceae are from marine strata, and there is no evidence that the coccolithophores ever invaded fresh waters (Falkowski et al., Reference Falkowski, Katz, Knoll, Quigg, Raven, Schofield and Taylor2004). The coccolithophore Emiliania huxleyi (Haptophyta: Prymnesiophyceae) has a putative P-type H+-ATPase in the plasmalemma (Lohbeck et al., Reference Lohbeck, Riebesell and Reusch2014). The Ψ CO of a calcifying strain of Emiliania huxleyi (Sikes & Wilbur, Reference Sikes and Wilbur1982) is −81 mV using K+-valinomycin and −145 mV using a fluorescent dye (cf. Ritchie, Reference Ritchie1982, Reference Ritchie1984). However, using the preferable microelectrode technique, Taylor et al. (Reference Taylor, Chrachri, Wheeler, Goddard and Brownlee2011) found a Ψ CO of −46 mV. There are conflicting reports on the intracellular pH of Emiliania huxleyi (Dixon et al., Reference Dixon, Brownlee and Merrett1989; Nimer et al., Reference Nimer, Brownlee and Merrett1994; see also Suffrian et al., Reference Suffrian, Schulz, Gurowska, Riebesell and Bleich2011). Using the fluorescent dye method the internal pH was determined as 7.28 (Dixon et al., Reference Dixon, Brownlee and Merrett1989), 7.03 (Nimer et al., Reference Nimer, Brownlee and Merrett1994) and 7.0 (Gibbin et al., Reference Gibbin, Putnam, Davy and Gates2014), at an external pH of 8 in the presence of 2 mol m−3 inorganic carbon. Lower values for internal pH are found with the DMO method and/or in the absence of inorganic C (Dixon et al., Reference Dixon, Brownlee and Merrett1989; Nimer et al., Reference Nimer, Brownlee and Merrett1994). Another coccolithophore, Pleurochrysis sp., has a P-type Ca2+-ATPase (Araki & González, Reference Araki and González1998).

The earliest known fossil Dinophyta (Alveolata) are from marine sediments with subsequent invasion of fresh waters (Lenz et al., Reference Lenz, Wilde and Riegel2002; Falkowski et al., Reference Falkowski, Katz, Knoll, Quigg, Raven, Schofield and Taylor2004). Symbiotic dinoflagellates of the Symbiodiniaceae may have a Na+-ATPase in the plasmalemma (Goiran et al., Reference Goiran, Allemand and Galgani1997), although a H+-ATPase and H+-Na+ antiport has not been ruled out as the mechanism of Na+ efflux. Bertucci et al. (Reference Bertucci, Tambutté, Tambutté, Allemand and Zoccola2010) and Mies et al. (Reference Mies, Van Sluys, Metcalfe and Sumida2017a, Reference Mies, Voolstra, Castro, Pires, Calderon and Sumida2017b) showed that the gene for a plasmalemma-located P-type H+-ATPase in the Symbiodiniaceae shows increased expression during symbiosis with corals, possibly related to acidification of the perisymbiotic space (Bertucci et al., Reference Bertucci, Tambutté, Tambutté, Allemand and Zoccola2010). It is not clear which ion is pumped by the P-ATPase in the calcifying dinoflagellate Thoracosphaera helmii, but it is probably Ca2+ (Van de Waal et al., Reference Van de Waal, John, Ziveri, Hoins, Sluija and Röst2013).

Submerged marine flowering plants: seagrasses

Fernández et al. (Reference Fernández, Garcia-Sánchez and Felle1999), Garcia-Sánchez et al. (Reference Garcia-Sánchez, Jaime, Ramos, Sanders and Fernández2000) and Rubio et al. (Reference Rubio, Belver, Venene, Garcia-Sánchez and Fernández2011) provided electrophysiological evidence of a H+-ATPase of the leaf and root cell plasmalemmas of Zostera marina. The Ψ CO is −150 to −160 mV, inside negative, and is hyperpolarized by fusicoccin, a compound known to stimulate non-halophytic flowering plant plasmalemma H+-ATPases; the cytosolic pH is 7.3 (Fernández et al., Reference Fernández, Garcia-Sánchez and Felle1999; Garcia-Sánchez et al., Reference Garcia-Sánchez, Jaime, Ramos, Sanders and Fernández2000). With an external seawater pH of 8.0 and the cytosol −160 mV negative relative to the seawater, the H+ electrochemical potential difference, cytosol relative to medium, across the plasmalemma is thus −13.7 kJ mol−1 kJ per mol, favouring H+ entry. Rubio et al. (Reference Rubio, Linares-Rueda, Garcia-Sánchez and Fernández2005) showed that the cytosol Na+ concentration (measured with Na+-selective microelectrodes) of 10.7 ± 3.3 mol m−3 in cells of Zostera marina in seawater has a Ψ CO of −150 mV, so the electrochemical potential difference across the plasmalemma is 25 kJ mol−3, driving Na+ into the cytosol. Following Pak et al. (Reference Pak, Fukuhara and Nitta1995), Fukuhara et al. (Reference Fukuhara, Pak, Ohwaki, Tsujimura and Nitta1996) and Muramoto et al. (Reference Muramoto, Harada, Ohkaki, Takagi and Fukuhara2002) characterized a salt-tolerant P-type H+-ATPase in the plasmalemma of Zostera marina. Generation of the −25 kJ mol−1 gradient (cytosol relative to medium) for Na+ using a Na+:H+ antiporter and a −13.7 kJ mol−1 gradient for H+ requires a Na+:H+ ratio of 0.5 or lower. Rubio et al. (Reference Rubio, Garcia, Garcia-Sánchez, Niell, Felle and Fernández2017, Reference Rubio, Garcia-Pérez, Garcia-Sánchez and Fernández2018) showed that the cytosol pH and the electrical potential difference across the plasmalemma of the seagrass Posidonia oceanica are very similar to those of Zostera marina.

Emergent marine flowering plants: tidal (=salt) marsh plants (herbaceous) and mangroves (trees)

Brügemann & Janiesch (Reference Brügemann and Janiesch1989) compared the plasma membrane ATPase from control specimens and those grown with added NaCl of the tidal marsh plant Plantago maritima. Wu & Seliskar (Reference Wu and Seliskar1998) investigated salinity adaptations in the H+-ATPase of the tidal marsh plant Spartina patens. The H+-ATPase also energizes NaCl secretion by salt glands in recretohalophytic tidal marsh plants and mangroves (Yuan et al., Reference Yuan, Leng and Wang2016; Dassanayake & Larkin, Reference Dassanayake and Larkin2017). The suggestion that the Limonium salt gland has primary active transport involving a Cl-ATPase has not been substantiated (Raven, Reference Raven2017).

Evolutionary aspects of primary active ion pumps in marine photosynthetic organisms

The occurrence of primary active H+, Na+ and Cl at the plasmalemma of marine photosynthetic organisms is summarized in Table 3, encapsulating the outcomes of the analysis in the previous section. Among the Archaeplastida, the brackish Characeae and brackish and marine flowering plants only have P(II)-type H+-ATPases. However, the Streptophyta are not basal among the Archaeplastida, so it cannot be concluded that P(II)-type H+-ATPase is the ancestral energizer of the plasmalemma. No data seem to be available for Glaucophyta; among Rhodophyta, marine representatives have P(II)-type Na+-ATPases and acidophilic ‘freshwater’ Cyanidiophyceae have P(II)-type H+-ATPases. Marine Chlorophyta have both P(II)-type H+-ATPases and P(II)-type Na+-ATPases. In these two phyla of Archaeplastida it is likely that horizontal gene transfer has been involved, e.g. via virus-encoded P-ATPases: an example is a Ca2+-ATPase in a freshwater Chlorella virus (Bonza et al., Reference Bonza, Martin, Kang, Lewis, Greiner, Giacometti, Van Etten, De Michelis, Thiel and Moroni2010). A range of transporters are encoded by other viruses (Greiner et al., Reference Greiner, Moroni, Van Etten and Thiel2018) and horizontal gene transfer has been shown for ATPases in prokaryotes (Hilario & Gogarten, Reference Hilario and Gogarten1993) and other transporters in eukaryotic algae (Chan et al., Reference Chan, Reyes-Prieto and Bhattacharya2011, Reference Chan, Zäuner, Wheeler, Grossman, Prochnik, Blouin, Zhuang, Benning, Berg, Yarish, Ekiksen, Klein, Lin, Levine, Brawley and Bhattachaya2012). Similar horizontal gene transfer is required to account for the distribution of P(II)-type H+-ATPases and P(II)-type Na+-ATPases among marine algae with plastids derived from secondary or tertiary endosymbiosis (Chan et al., Reference Chan, Reyes-Prieto and Bhattacharya2011, Reference Chan, Zäuner, Wheeler, Grossman, Prochnik, Blouin, Zhuang, Benning, Berg, Yarish, Ekiksen, Klein, Lin, Levine, Brawley and Bhattachaya2012; Burki et al., Reference Burki, Roger, Brown and Simpson2020). The origin of the F-ATPase that pumps Cl in some ulvophycean Chlorophyta (Table 3) is unclear; metazoan Cl-ATPases seem to be P-ATPases (Gerencser & Zhang, Reference Gerencser and Zhang2003).

Table 3. Phylogenetic distribution of plasmalemma ion transport ATPases in marine photosynthetic eukaryotes

For references see text.

Conclusions

In Archaea there are ion-pumping rhodopsins that actively transport H+ out of, and Cl into, the cells; in one archaean there is an inward H+-pumping rhodopsin. In Bacteria all three of the ions H+, Na+ and Cl are moved (H+ and Na+ out, Cl in) by ion-pumping rhodopsins. There is an H+influx rhodopsin in Acetabularia, and (probably) H+ efflux rhodopsins in diatoms. Photochemistry based on bacteriochlorophyll exports H+ from the cytosol in some marine anoxygenic photosynthetic bacteria, but chlorophyll-based redox reactions do not export H+ from cells of marine cyanobacteria. Exergonic redox reactions export H+ and Na+ in photosynthetic bacteria, H+ in cyanobacteria and possibly H+ in eukaryotic algae. H+-and/or Na+-ATPases occur in the plasmalemma of all photosynthetic marine organisms tested. P-type H+ efflux ATPases occur in the marine Streptophyta, i.e. marine charophycean algae and seagrasses and emergent marine flowering plants. P-type Na+-ATPases are the main primary active ion pumps in the plasmalemma of other marine green algae and non-green algae. However, there may be P-type H+-ATPases in some cases, and a F-type Cl-ATPase occurs in the ulvophycean Acetabularia. Some assignments of P-type ATPases as H+ or as Na+ pumps using genomics are not conclusive. Despite the insights that genomics can provide, it seems that there are still large gaps in the availability of electrophysiological data from many of the ecologically (and economically) important marine (and freshwater) phototrophs. There is thus a need for more such data to improve our understanding of the functioning of primary active transport in these organisms.

Acknowledgements

The University of Dundee is a registered Scottish charity, No. 015096. Discussions over the years with Jim Barber, Mary Beilby, Mary Bisson, Jack Dainty, Geoff Findlay, Mario Giordano, Alex Hope, Tony Larkum, Enid MacRobbie, Ray Ritchie; Hugh Saddler, Dale Sanders, Vladimir Skulachev, F Andrew Smith, Roger Spanswick, N Alan Walker and Philip White have been very helpful.

References

Adams, PG, Cadby, AJ, Robinson, B, Tsukatani, Y, Tank, M, Wen, J, Blankenship, RE, Bryant, DA and Hunter, CN (2013) Comparison of physical characteristics of chlorosomes from three different phyla of green phototrophic bacteria. Biochimica Biophysica Acta 1827, 12351244.CrossRefGoogle ScholarPubMed
Allen, RD, Jacobsen, L, Joaquen, J and Jaffe, LF (1972) Ion concentrations in developing Pelvetia eggs. Developmental Biology 27, 538545.CrossRefGoogle ScholarPubMed
Amtmann, A and Gradmann, D (1994) Na+ transport in Acetabularia bypasses conductance of plasmalemma. Journal of Membrane Biology 139, 117125.CrossRefGoogle ScholarPubMed
Araki, Y and González, EL (1998) V- and P-type Ca2+-stimulated ATPase in a calcifying strain of Pleurochrysis sp. (Haptophyceae). Journal of Phycology 34, 7988.CrossRefGoogle Scholar
Badger, MR and Price, GD (2003) CO2 concentrating mechanisms in cyanobacteria: molecular components, their diversity and evolution. Journal of Experimental Botany 54, 609622.CrossRefGoogle ScholarPubMed
Balnokin, Y and Popova, L (1994) The ATP-driven Na+ pump in the plasma membrane of the marine unicellular alga, Platymonas viridis. FEBS Letters 343, 6164.CrossRefGoogle ScholarPubMed
Balnokin, Y, Popova, L and Gimmler, H (1997) Further evidence for an ATP-driven sodium pump in the marine alga Tetraselmis (Platymonas) viridis. Journal of Plant Physiology 150, 264270.CrossRefGoogle Scholar
Balnokin, YV, Popova, LG and Andreev, IM (1999) Electrogenicity of the Na+ -ATPase from the marine microalga Tetraselmis (Platymonas) viridis and associated H+ countertransport. FEBS Letters 462, 402406.CrossRefGoogle ScholarPubMed
Balnokin, YV, Popova, LG, Pagis, LY and Andreev, IM (2004) The Na+-translocating ATPase in the plasma membrane of the marine microalga Tetraselmis viridis catalyses Na+/H+ exchange. Planta 219, 332337.CrossRefGoogle Scholar
Barrero-Gil, J, Garciadeblás, B and Benito, B (2005) Sodium, potassium-ATPases in algae and oomycetes. Journal of Bioenergetics and Biomembranes 37, 269278.CrossRefGoogle ScholarPubMed
Beilby, MJ (2015) Salt tolerance at single cell level in giant-celled Characeae. Frontiers in Plant Science 6, art. 226.CrossRefGoogle ScholarPubMed
Bejá, O and Lanyi, JK (2014) Nature's toolkit for microbial rhodopsin. Proceedings of the National Academy of Sciences USA 111, 65386539.CrossRefGoogle ScholarPubMed
Bental, M, Degani, H and Avron, M (1986) 23Na-NMR studies of the intracellular sodium ion concentration in the halotolerant alga Dunaliella. Plant Physiology 87, 813817.CrossRefGoogle Scholar
Berg, IA, Kockelhorn, D, Ramos-Vera, WH, Say, RF, Zarzycki, J, Hügler, M, Alber, BE and Fuchs, G (2010) Autotrophic carbon fixation in Archaea. Nature Reviews Microbiology 8, 447460.CrossRefGoogle ScholarPubMed
Bergman, B, Siddiqui, PJA, Carpenter, EJ and Pescher, GA (1993) Cytochrome oxidase: subcellular distribution and relationship to nitrogenase expression in the nonheterocystous marine cyanobacterium Trichodesmium thiebaultii. Applied and Environmental Microbiology 59, 32393244.CrossRefGoogle Scholar
Bertucci, A, Tambutté, R, Tambutté, S, Allemand, D and Zoccola, D (2010) Symbiosis-dependent gene expression in coral dinoflagellate association: cloning and characterization of a P-type H+-ATPase gene. Proceedings of the Royal Society of London B 277, 8795.Google ScholarPubMed
Beyenbach, KW and Wieczorek, H (2006) The V-type H+-ATPase: molecular structure and function, physiological roles and regulation. Journal of Experimental Biology 209, 577589.CrossRefGoogle ScholarPubMed
Bhattacharya, P and Volcani, BE (1980) Sodium-dependent silicate transport in the apochlorotic marine diatom Nitzschia alba (Na+ gradient/Na+, K+-ATPase/membrane vesicles). Proceedings of the National Academy of Sciences USA 77, 63866390.CrossRefGoogle Scholar
Bisson, MA, Beilby, MJ and Shepherd, VA (2006) Electrophysiology of turgor regulation in marine siphonous green algae. Journal of Membrane Biology 211, 115.CrossRefGoogle ScholarPubMed
Blank, CE (2013 a) Origin and early evolution of photosynthetic eukaryotes; reinterpreting Proterozoic palaeobiology and biogeochemical processes in light of trait evolution. Journal of Phycology 49, 10401056.CrossRefGoogle Scholar
Blank, CE (2013 b) Phylogenetic distribution of compatible solute synthesis genes support a freshwater origin for cyanobacteria. Journal of Phycology 49, 880895.Google ScholarPubMed
Blank, CE and Sánchez-Baracaldo, P (2010) Timing of morphological and ecological innovations in the cyanobacteria – a key to understanding the rise in atmospheric oxygen. Geobiology 8, 123.CrossRefGoogle ScholarPubMed
Blinks, LR (1940) The relations of bioelectric phenomena to ionic permeability and to metabolism in large plant cells. Cold Spring Harbor Symposium 8, 204215.CrossRefGoogle Scholar
Blinks, LR (1949) The source of the bioelectric potentials in large plant cells. Proceedings of the National Academy of Sciences USA 35, 566575.CrossRefGoogle ScholarPubMed
Blount, RW and Levedahl, BH (1960) Active sodium and chloride transport in the single celled marine alga Halicystis ovalis. Acta Physiologia 49, 19.CrossRefGoogle ScholarPubMed
Bock, C, Jacob, A, Kirst, GO, Leibfritz, D and Mayer, A (1996) Metabolic changes of the Antarctic green alga Prasiola crispa subjected to water stress investigated by in vivo 31P NMR. Journal of Experimental Botany 47, 241249.CrossRefGoogle Scholar
Bonza, MC, Martin, H, Kang, M, Lewis, G, Greiner, T, Giacometti, S, Van Etten, JL, De Michelis, MI, Thiel, G and Moroni, A (2010) A functional calcium-transporting ATPase encoded by chlorella viruses. Journal of General Virology 91, 26202629.CrossRefGoogle ScholarPubMed
Boyd, CM and Gradmann, D (1999) Electrophysiology of the marine diatom Coscinodiscus wailesii. III. Uptake of nitrate and ammonium. Journal of Experimental Botany 50, 461467.Google Scholar
Brasier, MD (2013) Green algae (Chlorophyta) and the question of freshwater symbiogenesis in the early Proterozoic. Journal of Phycology 49, 10361039.CrossRefGoogle ScholarPubMed
Brewer, PG and Goldman, JC (1976) Alkalinity changes generated by phytoplankton growth. Limnology and Oceanography 21, 103117.CrossRefGoogle Scholar
Brügemann, W and Janiesch, P (1989) Comparison of plasma membrane ATPase from salt-treated and salt-free grown Plantago maritima. Journal of Plant Physiology 134, 2025.CrossRefGoogle Scholar
Bublitz, M, Morth, JP and Nissen, P (2011) P-type ATPases at a glance. Journal of Cell Science 124, 2525–2519.CrossRefGoogle ScholarPubMed
Burki, F, Roger, AJ, Brown, MW and Simpson, AGB (2020) The new tree of eukaryotes. Trends in Ecology and Evolution 35, 4355.CrossRefGoogle ScholarPubMed
Burns, BD and Beardall, J (1987) Utilization of inorganic carbon by marine microalgae. Journal of Marine Biology and Ecology 107, 7586.CrossRefGoogle Scholar
Bussard, A and Lopez, PJ (2014) Evolution of vacuolar pyrophosphatases and vacuolar H+ ATPases in diatoms. Journal of Marine Science and Technology 22, 5059.Google Scholar
Campbell, AM, Coble, AJ, Cohen, LD, Ch'ng, TH, Russo, KM, Long, EM and Armbrust, EV (2001) Identification and DNA sequence of a new H+-ATPase in the unicellular green alga Chlamydomonas reinhardtii (Chlorophyceae). Journal of Phycology 37, 536542.CrossRefGoogle Scholar
Chan, CX, Reyes-Prieto, A and Bhattacharya, D (2011) Red and green algal origin of diatom membrane transporters: insights into environmental adaptation and cell evolution. PLoS ONE 6, art.e29138.CrossRefGoogle ScholarPubMed
Chan, CX, Zäuner, A, Wheeler, G, Grossman, AR, Prochnik, SE, Blouin, NA, Zhuang, Y, Benning, C, Berg, GM, Yarish, C, Ekiksen, RL, Klein, AS, Lin, S, Levine, I, Brawley, SH and Bhattachaya, D (2012) Analysis of Porphyra membrane transporters demonstrates gene transfer among photosynthetic eukaryotes and numerous sodium-coupled transport systems. Plant Physiology 158, 20012012.CrossRefGoogle ScholarPubMed
Chen, T, Wang, W, Xu, K, Xu, Y, Ji, D, Chen, C and Xie, C (2019) K+ and Na+ transport contribute to K+/Na+ homeostasis in Pyropia haitanensis under hypersaline stress. Algal Research 40, art. 101526.CrossRefGoogle Scholar
Cohen, NR, Ellis, KA, Lampe, RH, McNair, H, Twining, BS, Maldonado, MT, Brzezinski, MA, Kizminov, FI, Thamatrakoln, K, Till, CP, Bruland, KW, Sunda, WG, Bangu, S and Marchetti, A (2017) Diatom transcriptional and physiological responses to changes in iron bioavailability across ocean provinces. Frontiers in Marine Science 4, art. 360.CrossRefGoogle Scholar
Dassanayake, M and Larkin, JC (2017) Making plants break a sweat: the structure, function, and evolution of plant salt glands. Frontiers in Plant Science 8, art. 406.Google ScholarPubMed
De Clerck, O and 35 co-authors (2018) Insights into the evolution of multicellularity from the sea lettuce genome. Current Biology 28, 29212933.CrossRefGoogle ScholarPubMed
Del Cortona, A, Jackson, CJ, Bucchini, F, Van Bel, M, D'hondt, S, Škaloud, P, Delwiche, CF, Knoll, AH, Raven, JA, Verbruggen, H, Vandepoel, K, De Clerck, O and Leliaert, F (2020) Neoproterozoic origin and multiple transitions to macroscopic growth in green seaweeds. Proceedings of the National Academy of Sciences USA 117, 25512559. doi: 10.1073/pnas.19.1006017.CrossRefGoogle ScholarPubMed
Dibrova, DV, Galperin, MY, Koonin, EV and Mulkidjanian, AY (2015) Ancient systems of sodium/potassium homeostasis as predecessors of membrane bioenergetics. Biochemistry (Moscow) 80, 495516.CrossRefGoogle ScholarPubMed
Dittami, SM, Heesch, S, Olsen, JL and Collén, J (2017) Transition between marine and freshwater environments provide new clues about the origins of multicellular plants and algae. Journal of Phycology 53, 731745.CrossRefGoogle ScholarPubMed
Dixon, GK, Brownlee, C and Merrett, MJ (1989) Measurement of internal pH in the coccolithophore Emiliania huxleyi using 1′,7′-bis’(2-carboxyethyl)-5(and −6)carboxy-fluorescein acetoxymethylester and digital imaging microscopy. Planta 178, 443449.CrossRefGoogle Scholar
Dodd, WA and Bidwell, RGS (1971) The effect of pH on the products of photosynthesis in 14CO2 by chloroplast preparations from Acetabularia mediterranea. Plant Physiology 47, 779783.CrossRefGoogle Scholar
Dodd, WA, Pitman, MG and West, KR (1966) Sodium and potassium transport in the marine alga Chaetomorpha darwinii. Australian Journal of Biological Sciences 19, 341354.CrossRefGoogle Scholar
Duchuzeau, A-L, Schoepp-Cothenet, B, Baymann, F, Russell, MJ and Nitscke, W (2014) Free energy conversion in the LUCA: quo vadis? Biochimica et Biophysica Acta 1837, 882888.Google Scholar
Ehrenfeld, J and Cousin, J-L (1984) Ionic regulation of the unicellular green alga Dunaiella tertiolecta: response to hypotonic shock. Journal of Membrane Biology 77, 4555.CrossRefGoogle Scholar
Emery, L, Whelan, S, Hirschi, KD and Pittman, JK (2012) Protein phylogenetic analysis of Ca2+/cation antiporters and insights into their evolution in plants. Frontiers in Plant Science 3, art. 1.CrossRefGoogle Scholar
Falkowski, PG, Katz, ME, Knoll, AH, Quigg, A, Raven, JA, Schofield, O and Taylor, FJ (2004) The evolution of modern eukaryotic phytoplankton. Science (New York, N.Y.) 305, 354360.CrossRefGoogle ScholarPubMed
Fernández, JA, Garcia-Sánchez, MJ and Felle, HH (1999) Physiological evidence for a proton pump and sodium exclusion mechanisms at the plasma membrane of the marine angiosperm Zostera marina L. Journal of Experimental Botany 50, 17631768.Google Scholar
Findlay, GP, Hope, AB, Pitman, MG, Smith, FA and Walker, NA (1971) Ionic relations of marine algae III Chaetomorpha membrane electrical properties and chloride fluxes. Australian Journal of Biological Sciences 24, 731746.CrossRefGoogle Scholar
Flynn, KJ, Blackford, JC, Baird, ME, Raven, JA, Clark, DR, Beardall, J, Brownlee, C, Fabian, H and Wheeler, GL (2012) Changes in pH at the exterior surface of plankton with ocean acidification. Nature Climate Change 2, 510513.CrossRefGoogle Scholar
Flynn, KJ, Öpik, H and Syrett, PJ (1987) The isolation of plasma membrane from the diatom Phaeodactylum tricornutum using an aqueous two-polymer phase system. Journal of General Microbiology 133, 93101.Google Scholar
Foflonker, F, Price, DC, Qiu, H, Palenik, B, Wang, S and Bhattacharya, D (2014) Genome of the halotolerant green alga Picochlorum sp. reveals strategies for thriving under fluctuating environmental conditions. Environmental Microbiology 17, 412426.CrossRefGoogle ScholarPubMed
Fukuhara, T, Pak, JY, Ohwaki, H, Tsujimura, H and Nitta, T (1996) Tissue-specific expression of the gene for a putative plasma membrane H+-ATPase in a seagrass. Plant Physiology 110, 3542.CrossRefGoogle Scholar
Fux, JE, Mehta, A, Moffat, J and Spafford, JD (2018) Eukaryotic voltage-gated sodium channels: on their origins, asymmetries, losses, diversification and adaptations. Frontiers in Physiology 9, art. 1406.CrossRefGoogle ScholarPubMed
Gabbay-Azaria, R, Schonfeld, M, Tel-Or, S, Messinger, R and Tel-Or, E (1992) Respiratory activity in the marine cyanobacterium Spirulina subsalsa and its role in salt tolerance. Archives of Microbiology 57, 183190.CrossRefGoogle Scholar
Garcia-Pichel, F, Ramirez-Reinat, E and Guo, Q (2010) Microbial excavation of solid carbonates powered by a P-type ATPase-mediated transcellular Ca2+ transport. Proceedings of the National Academy of Sciences USA 107, 2174921754.CrossRefGoogle ScholarPubMed
Garcia-Sánchez, MJ, Jaime, MP, Ramos, A, Sanders, D and Fernández, JA (2000) Sodium-dependent nitrate transport at the plasma membrane of leaf cells of the marine higher plant Zostera marina. Plant Physiology 122, 979–885.CrossRefGoogle ScholarPubMed
Gerencser, GA and Zhang, J (2003) Existence and nature of the chloride pump. Biochimica et Biophysica Acta 1618, 133139.CrossRefGoogle ScholarPubMed
Gibbin, EM, Putnam, HM, Davy, SK and Gates, RD (2014) Intracellular pH and its response to CO2-driven seawater acidification in symbiotic versus non-symbiotic coral cells. Journal of Experimental Biology 217, 19631969.CrossRefGoogle ScholarPubMed
Gibbon, BC and Kropf, DL (1993) Intracellular pH and its regulation in Pelvetia zygotes. Developmental Biology 157, 259268.CrossRefGoogle ScholarPubMed
Gimmler, H (1988) Erythrosine B – a specific inhibitor of plasmalemma ATPase in intact microalgae. Journal of Plant Physiology 132, 545551.CrossRefGoogle Scholar
Gimmler, H (2000) Primary sodium plasma membrane ATPases in salt-tolerant algae: facts and fictions. Journal of Experimental Botany 51, 11711178.CrossRefGoogle ScholarPubMed
Gimmler, H and Greenway, H (1983) Tetraphenylphosphonium (TPP+) is not suitable for assessment of electrical potentials in Chlorella emersonii. Plant Cell and Environment 6, 739744.Google Scholar
Gimmler, H, Kugel, H, Leibfritz, D and Mayer, A (1988) Cytoplasmic pH of Dunaliella parva and Dunaliella parva as monitored by in vivo 31P-NMR spectroscopy and the DMO method. Physiologia Plantarum 74, 521555.CrossRefGoogle Scholar
Ginzburg, M, Ratcliffe, RG and Southon, TE (1988) Phosphorus metabolism and intracellular pH in the halotolerant alga Dunaliella parva studied by 31P NMR. Biochimica et Biophysica Acta: Molecular Cell Research 969, 225235.CrossRefGoogle Scholar
Gleason, FH, Larkum, AWD, Raven, JA, Manohar, CS and Lilje, O (2019) Ecological implication of recently discovered and poorly studied sources of energy for true fungi, especially in extreme environments. Fungal Ecology 39, 380387.CrossRefGoogle Scholar
Goericke, R (2002) Bacteriochlorophyll a in the ocean: is anoxygenic bacterial photosynthesis important? Limnology and Oceanography 47, 290295.CrossRefGoogle Scholar
Goiran, C, Allemand, D and Galgani, I (1997) Transient Na+ stress in symbiotic dinoflagellates after isolation from coral-host cells and subsequent immersions in seawater. Marine Biology 129, 581589.CrossRefGoogle Scholar
Goldfarb, V and Gradmann, D (1983) ATPase activity in partially purified membranes of Acetabularia. Plant Cell Reports 2, 152155.CrossRefGoogle ScholarPubMed
Goldfarb, V, Sanders, D and Gradmann, D (1984) Reversal of electrogenic Cl pump in Acetabularia increases level and 32labelling of ATP. Journal of Experimental Botany 35, 645658.CrossRefGoogle Scholar
Gómez-Consarnau, L, Raven, JA, Levin, NM, Cutter, L, Wang, D, Seegers, B, Arístegui, J, Fuhrman J, A, Gasel, JM and Sañudo-Wilhelmy, SA (2019) Microbial rhodopsins are major contributors to the solar energy captured in the sea. Science Advances 5, eaaw8855.CrossRefGoogle Scholar
Gradmann, D and Boyd, CM (1995) Membrane voltage of marine phytoplankton, measured in the diatom Coscinodiscus radiatus. Marine Biology 124, 645650.CrossRefGoogle Scholar
Gradmann, D and Boyd, CM (1999a) Electrophysiology of the marine diatom Coscinodiscus wailesii I. Endogenous changes of membrane voltage and resistance. Journal of Experimental Botany 50, 445452.Google Scholar
Gradmann, D and Boyd, CM (1999b) Electrophysiology of the marine diatom Coscinodiscus wailesii II. Potassium currents. Journal of Experimental Botany 50, 453459.Google Scholar
Gradmann, D, Tittor, J and Goldfarb, V (1982) Electrogenic Cl pump in Acetabularia. Philosophical Transactions of the Royal Society B 299, 447457.Google ScholarPubMed
Graves, JS and Gutknecht, J (1976) Ion transport studies and determination of the cell wall elastic modulus in the marine alga Halicystis parvula. Journal of General Physiology 67, 579592.CrossRefGoogle ScholarPubMed
Graves, JS and Gutknecht, J (1977a) Chloride transport and membrane potential in the marine alga Halicystis parvula. Journal of Membrane Biology 36, 6581.CrossRefGoogle Scholar
Graves, JS and Gutknecht, J (1977b) Current-voltage relationship and voltage sensitivity in the Cl pump in Halicystis parvula. Journal of Membrane Biology 36, 8391.CrossRefGoogle Scholar
Greiner, T, Moroni, A, Van Etten, JI and Thiel, G (2018) Genes for membrane transport proteins: not so rare in viruses. Viruses 10, art. 456.CrossRefGoogle ScholarPubMed
Hayden, HS, Blomster, J, Maggs, CA, Silva, PC, Stanhope, MJ and Walland, JR (2003) Linnaeus was right all along: Ulva and Enteromorpha are not distinct genera. European Journal of Phycology 38, 277294.CrossRefGoogle Scholar
Helliwell, KE, Chrachri, A, Koester, JA, Wharam, S, Verret, F, Taylor, AR, Wheeler, GL and Brownlee, C (2019) Alternative mechanisms for fast Na+/Ca2+ signalling in eukaryotes via a novel class of single-domain voltage-gated channels. Current Biology 29, 15031511.CrossRefGoogle Scholar
Hervé, V, Derr, J, Douady, S, Quinet, M, Moisan, L and Lopez, JL (2002) Multiparametric analyses reveal the pH-dependence of silicon mineralisation in diatoms. PLoS ONE 7, art. e46722.Google Scholar
Hilario, E and Gogarten, JP (1993) Horizontal transfer of ATPase genes – the tree of life becomes a net of life. Biosystems 31, 111119.CrossRefGoogle ScholarPubMed
Hong, S and Pedersen, PL (2008) ATP synthase and the actions of inhibitors utilized to study its roles in human health, disease, and other scientific areas. Microbiology and Molecular Biology 72, 590641.CrossRefGoogle ScholarPubMed
Hope, AB and Walker, NA (1975) The Physiology of Giant Algal Cells. Cambridge: Cambridge University Press.Google Scholar
Ikeda, M, Schmid, R and Oesterhelt, D (1990 a) A chloride-translocating adenosine triphosphatase in Acetabularia acetabulum. 1. Purification and characterization of a novel type of adenosine triphosphatase that differs from chloroplast F1 adenosine triphosphatase. Biochemistry 29, 20572065.CrossRefGoogle Scholar
Ikeda, M, Schmid, R and Oesterhelt, D (1990 b) A chloride-translocating adenosine triphosphatase in Acetabularia acetabulum. 2. Reconstitution of the enzyme into liposomes and effect of net charge of liposomes on chloride permeability and reconstitution. Biochemistry 29, 20652070.CrossRefGoogle Scholar
Ikeda, M, Kadowaki, H, Ikeda, H, Moritani, C and Kanazawa, H (1997) Exchangeability of the b subunit of the Cl-translocating ATPase of Acetabularia acetabulum with the β-subunit of E. coli F1-ATPase: construction of the chimeric β-subunits and complementation studies. Biochimica Biophysica Acta 1322, 3340.CrossRefGoogle Scholar
Jackson, BZ (2016) Natural pH gradients in hydrothermal alkali vents were unlikely to have played a role in the origin of life. Journal of Molecular Evolution 83, 111.CrossRefGoogle ScholarPubMed
Jo, T, Shono, M, Wada, M, Ito, S, Nomoto, J and Hara, Y (2010) Homology modelling of an algal membrane protein, Heterosigma akashiwo, Na+-ATPase. Membrane 35, 8085.CrossRefGoogle Scholar
Johnson, DT, Wolfe-Simon, F, Pearson, A and Knoll, AH (2009) Anoxygenic photosynthesis modulated Proterozoic oxygen and sustained Earth's middle age. Proceedings of the National Academy of Sciences USA 106, 1692516929.CrossRefGoogle Scholar
Jones, GJ and Morel, FM (1988) Plasmalemma redox activity in the diatom Thalassiosira: a possible role for nitrate reductase. Plant Physiology 87, 143147.CrossRefGoogle ScholarPubMed
Kaplan, A, Scherer, S and Lerner, M (1989) Nature of the light-induced H+ efflux and Na+ uptake in Cyanobacteria. Plant Physiology 89, 12201225.CrossRefGoogle ScholarPubMed
Katz, A and Avron, M (1985) Determination of intracellular osmotic volume and sodium concentrations in Dunaliella. Plant Physiology 78, 817820.CrossRefGoogle ScholarPubMed
Katz, A and Pick, U (2001) Plasma membrane electron transport coupled to Na+ extrusion in the halotolerant alga Dunaliella. Biochimica et Biophysica Acta 1504, 423431.CrossRefGoogle ScholarPubMed
Katz, A, Bental, M, Degani, H and Avron, M (1991) In vivo pH regulation by a Na+/H+ antiporter in the halotolerant alga Dunaliella salina. Plant Physiology 96, 110115.CrossRefGoogle ScholarPubMed
Katz, A, Waridel, P, Shevchenko, A and Pick, U (2007) Salt-induced changes in the plasma membrane proteome of the halotolerant alga Dunaliella salina as revealed by blue native gel electrophoresis and nano-LC-MS/MS analysis. Molecular and Cellular Proteomics 6/9, 14591472.CrossRefGoogle Scholar
Kesseler, H (1974) Die anorganische-chemische Zussamensetzung dor Zellsaftes von Coscinodiscus granii (Bacillariophyceae, Centrales). Helgoländer Wissenschaftliche Meeresuntersuchungen 26, 481489.CrossRefGoogle Scholar
Khramov, DE, Matalin, DA, Karpichev, IV, Balnokin, YV and Popova, LG (2019) Expression of P-type ATPases of the marine green microalga Dunaliella maritima under hyperosmotic salt shock. Doklady Biochemistry and Biophysics 488, 327333.CrossRefGoogle ScholarPubMed
Kirchman, DL and Hanson, TE (2013) Bioenergetics of photo-heterotrophic bacteria in the oceans. Environmental Microbiology 5, 188199.Google Scholar
Kirst, GO and Bisson, MA (1982) Vacuolar and cytoplasmic pH, ion composition, and turgor pressure in Lamprothamnium as a function of external pH. Planta 155, 287295.CrossRefGoogle ScholarPubMed
Kishimoto, M, Shimajiri, Y, Oshima, A, Hase, A, Mikami, K and Akama, K (2013) Functional expression of an animal-type-Na+-ATPase gene from a marine red seaweed Porphyra yezoensis increases salinity tolerance in rice plants. Plant Biotechnology 30, 417422.CrossRefGoogle Scholar
Klenell, M, Snoeijs, P and Pedersén, M (2002) The involvement of a plasma membrane H+-ATPase in the blue-light enhancement of photosynthesis in Laminaria digitata (Phaeophyta). Journal of Phycology 38, 11431149.CrossRefGoogle Scholar
Klenell, M, Snoeijs, P and Pedersén, M (2004) Active carbon uptake in Laminaria digitata and L. saccharina (Phaeophyta) is driven by a proton pump in the plasma membrane. Hydrobiologia 514, 4153.CrossRefGoogle Scholar
Kolber, ZS, VanDover, CL, Niederman, RA and Falkowski, PG (2000) Bacterial photosynthesis in surface waters of the open ocean. Nature 407, 177179.CrossRefGoogle ScholarPubMed
Kolber, ZS, Gerald, F, Lang, AS, Beatty, JT, Blankenship, RE, VanDover, CL, Vetriani, M, Ratheber, C and Falkowski, PG (2001) Contribution of aerobic photoheterotrophic bacteria to the carbon cycle in the ocean. Science (New York, N.Y.) 292, 24922495.CrossRefGoogle ScholarPubMed
Komor, E, Cho, B-H, Schricker, S and Schobert, C (1989) Charge and acidity compensation during proton-sugar symport in Chlorella: the H+-ATPase does not fully compensate for the sugar-coupled proton influx. Planta 177, 917.CrossRefGoogle Scholar
Ksionzek, KB, Lechtenfeld, OJ, McAllister, SL, Schmitt-Kopplin, P, Geuer, JK, Geibert, W and Koch, BP (2016) Dissolved organic sulfur in the ocean: biogeochemistry of a petagram inventory. Science (New York, N.Y.) 354, 456459.CrossRefGoogle ScholarPubMed
Kuhlbrandt, W (2004) Biology, structure and function of P-type ATPases. Nature Reviews Microbiology Molecular Cell Biology 3, 282295.CrossRefGoogle Scholar
Larkum, AWD, Ritchie, RJ and Raven, JA (2018) Living of the sun: chlorophylls, bacteriochlorophylls and rhodopsins. Photosynthetica 56, 1143.CrossRefGoogle Scholar
Lea-Smith, DJ, Bombelli, P, Vasudevan, R and Howe, CJ (2016) Photosynthetic, respiratory and extracellular electron transport pathways in cyanobacteria. Biochimica et Biophysica Acta 1857, 247255.CrossRefGoogle ScholarPubMed
Lee, J, Ghosh, S and Saier, MH Jr (2017) Comparative genomic analysis of transport proteins encoded within the red algae Chondrus crispus, Galdieria sulphuraria and Cyanidioschyzon merolae. Journal of Phycology 53, 503521.CrossRefGoogle Scholar
Lenz, OK, Wilde, V and Riegel, W (2002) Distribution and palaeoecologic significance of the freshwater dinoflagellate cyst Messelodinium thielpfeifferae gen. et sp. nov. from the middle Eocene of Lake Messel, Germany. Palynology 31, 117134.Google Scholar
Lewis, LA (2017) Hold the salt: freshwater origin of primary plastids. Proceedings of the National Academy of Sciences USA 114, 97599760.CrossRefGoogle ScholarPubMed
Lohbeck, KT, Riebesell, U and Reusch, TBH (2014) Gene expression changes in the coccolithophore Emiliania huxleyi after 500 generations of selection to ocean acidification. Proceedings of the Royal Society of London B 281, 2014.0003.CrossRefGoogle ScholarPubMed
Lundberg, P, Welch, RG, Jensén, P and Vogel, HJ (1989) Phosphorus-31 and nitrogen-14 NMR studies of the uptake of phosphorus and nitrogen compounds in the marine macroalga Ulva lactuca. Plant Physiology 89, 13801387.CrossRefGoogle Scholar
Maeda, S-I and Omata, T (1997) Substrate-binding lipoprotein of the cyanobacterium Synechococcus sp. strain PCCC 7942 involved in the transport of nitrate and nitrite. Journal of Biological Chemistry 272, 30163041.CrossRefGoogle Scholar
Maeda, S-i, Murakami, A, Ito, H, Tanaka, A and Omata, T (2015) Functional characterization of the FNT family nitrite transporter of marine picocyanobacteria. Life (Chicago, ILL) 5, 432446.Google ScholarPubMed
Marchetti, A, Corlett, D, Hoplinson, BM, Ellis, K and Cassar, N (2015) Marine diatom proteorhodopsins and their role in coping with low iron availability. ISME Journal 9, 27452748.CrossRefGoogle ScholarPubMed
Martin, W and Russell, MJ (2007) The origin of biochemistry at an alkaline hydrothermal vent. Philosophical Transactions of the Royal Society of London B 367, 18871925.CrossRefGoogle Scholar
Mies, M, Van Sluys, MA, Metcalfe, CJ and Sumida, PYG (2017 a) Molecular evidence of symbiotic activity between Symbiodinium and Tridacna maxima larvae. Symbiosis 72, 1322.CrossRefGoogle Scholar
Mies, M, Voolstra, CR, Castro, TB, Pires, DO, Calderon, EN and Sumida, PYG (2017 b) Expression of a symbiosis-specific gene in Symbiodinium Type A1 associated with coral, nudibranch and giant clam larvae. Royal Society Open Science 4, 170253.CrossRefGoogle ScholarPubMed
Mitchell, P (1979) Compartmentation and communication in living systems. Ligand conduction: a general catalytic principle in chemical, osmotic and chemiosmotic reaction system (The Ninth Sir Hans Krebs Lecture). European Journal of Biochemistry 95, 120.CrossRefGoogle Scholar
Moritani, C, Ohnashi, T, Kadowaki, H, Tagaya, M, Lottspeich, F, Oesterhelt, D and Ikeda, M (1997) The primary structure of Cl-translocating ATPase, beta subunit of Acetabularia acetabulum, which belongs to the F-type ATPase family. Archives of Biochemistry and Biophysics 359, 115124.CrossRefGoogle Scholar
Mulkidjanian, AY, Dibrov, P and Galperin, MY (2008) The past and present of the sodium energetics: may the sodium-motive force be with you. Biochimica et Biophysica Acta 1777, 985992.CrossRefGoogle ScholarPubMed
Müller, ML, Irkens-Kiesecker, U, Rubinstein, B and Taiz, L (1996) On the mechanism of hyperacidification in lemon: comparison of the vacuolar H+-ATPase in the fruits and epicotyls. Journal of Biological Chemistry 271, 19181924.CrossRefGoogle ScholarPubMed
Müller, ML, Jensen, M and Taiz, L (1999) The vacuolar H+-ATPase of lemon fruits is regulated by variable H+/ATP coupling and slip. Journal of Biological Chemistry 274, 1070610716.CrossRefGoogle Scholar
Mullineaux, CW (2014) Co-existence of photosynthetic and respiratory activities in cyanobacterial thylakoid membranes. Biochimica et Biophysica Acta 1837, 503511.CrossRefGoogle ScholarPubMed
Mummert, H and Gradmann, D (1976) Voltage dependent potassium fluxes and the significance of action potentials in Acetabularia. Biochimica et Biophysica Acta 443, 44434450.Google ScholarPubMed
Mummert, H and Gradmann, D (1991 a) Ion fluxes in Acetabularia: vesicular shuttle. Journal of Membrane Biology 124, 255263.CrossRefGoogle ScholarPubMed
Mummert, H and Gradmann, D (1991 b) Action potentials in Acetabularia: measurements and simulation of voltage-gated fluxes. Journal of Membrane Biology 124, 265273.CrossRefGoogle Scholar
Muramoto, Y, Harada, A, Ohkaki, Y, Takagi, S and Fukuhara, T (2002) Salt-tolerant ATPase activity in the plasma membrane of the marine angiosperm Zostera marina L. Plant Cell Physiology 43, 11371143.Google Scholar
Nichols, DG and Ferguson, SJ (2013) Bioenergetics, 4th Edn. New York, NY: Academic Press/Elsevier.Google Scholar
Nimer, NA, Brownlee, C and Merrett, MJ (1994) Carbon dioxide availability, intracellular pH and growth rate of the coccolithophore Emiliania huxleyi. Marine Ecology Progress Series 100, 257262.CrossRefGoogle Scholar
Nishiyama, T et al. (2018) The Chara Genome: secondary complexity and implications for plant terrestrialisation. Cell 174, 448464.CrossRefGoogle Scholar
Nobel, PS (2009) Physicochemical and Environmental Plant Physiology, 4th Edn. Amsterdam: Elsevier.Google Scholar
Nowack, ECM (2014) Paulinella chromaphora – rethinking the transition from endosymbiont to organelles. Acta Societas Botanicorum Poloniae 83, 387397.CrossRefGoogle Scholar
Oesterhelt, D and Stoeckenius, W (1973) Functions of a new photoreceptor membrane. Proceedings of the National Academy of Sciences USA 70, 28532857.CrossRefGoogle ScholarPubMed
Ohhashi, T, Katsu, T and Ikeda, M (1992) Improvement of reconstitution of the Cl-translocating ATPase isolated from Acetabularia acetabulum into liposomes and several anion pump characteristics. Biochimica et Biophysica Acta 1106, 165170.CrossRefGoogle ScholarPubMed
Ohta, H, Shirakawa, H, Uchida, K, Yoshida, M, Matuo, Y and Ehami, I (1997) Cloning and sequencing of the gene encoding the plasma membrane H+-ATPase from an acidic red alga, Cyanidium caldarium. Biochimica et Biophysica Acta Bioenergetics 1319, 913.CrossRefGoogle ScholarPubMed
Omata, T, Price, GD, Badger, MR, Okamura, M, Gohta, S and Ogawa, T (1999) Identification of the ATP-binding cassette transporter involved in bicarbonate uptake in the cyanobacterium Synechococcus sp. strain PCC 7942. Proceedings of the National Academy of Sciences USA 96, 13571–13376.CrossRefGoogle ScholarPubMed
Oren-Shamir, M, Pick, U and Avron, M (1990) Plasma membrane potential of the alga Dunaliella, and its relation to osmoregulation. Plant Physiology 93, 403408.CrossRefGoogle ScholarPubMed
Pagis, LY, Popova, LG, Andreev, IM and Balnokin, YV (2001) Ion specificity of Na+-transporting systems in the plasma membrane of the halotolerant alga Tetraselmis (Platymonas) viridis. Russian Journal of Plant Physiology 48, 281286.CrossRefGoogle Scholar
Pagis, LY, Popova, LG, Andreev, IM and Balnokin, YV (2003) Comparative characterization of two primary pumps, H+-ATPase and Na+-ATPase, in the plasma membrane of the marine alga Tetraselmis viridis. Physiologia Plantarum 118, 514522.CrossRefGoogle Scholar
Pak, J-Y, Fukuhara, T and Nitta, T (1995) Discrete subcellular localization of membrane-bound ATPase activity in marine angiosperms and marine algae. Planta 196, 1522.CrossRefGoogle ScholarPubMed
Palmgren, M and Nissen, P (2011) P-type ATPases. Annual Review of Biophysics 40, 243266.CrossRefGoogle ScholarPubMed
Palmgren, MG, Sorenson, DM, Hallström, BM, Säll, T and Broberg, K (2020) Evolution of P2A and P5A ATPases: ancient gene duplications and the red algal connection to green plants revisited. Physiologia Plantarum 168, 630647. doi: 10.111/ppl.13008.CrossRefGoogle ScholarPubMed
Pedersen, CS, Axelsen, KB, Harper, JF and Palmgren, MG (2012) Evolution of plant P-type ATPases. Frontiers in Plant Science 3, art. 31.CrossRefGoogle ScholarPubMed
Pick, U, Kanni, L and Avron, M (1986) Determination of ion concentration and ion fluxes in the halotolerant alga Dunaliella salina. Plant Physiology 81, 9296.CrossRefGoogle ScholarPubMed
Ponce-Toledo, RI, Deschamps, P, Lôpez-García, P, Zivanovic, Y, Benzenara, K and Moreira, D (2017) An early-branching freshwater cyanobacterium at the origin of plastids. Current Biology 27, 386391.CrossRefGoogle ScholarPubMed
Popova, LG and Balnokin, YV (1992) H+-translocating ATPase and Na+/H+ antiport activities in the plasma membrane of the marine alga Platymonas viridis. FEBS Letters 3, 333336.CrossRefGoogle Scholar
Popova, L and Balnokin, YV (2013) Na+-ATPases of halotolerant algae. Russian Journal of Plant Physiology 68, 472482.CrossRefGoogle Scholar
Popova, LG, Shumkova, GA, Andreev, IM and Balnokin, YV (2005) Functional identification of electrogenic Na+-translocating ATPase in the plasma membrane of the halotolerant microalga Dunaliella maritima. FEBS Letters 579, 50025006.CrossRefGoogle ScholarPubMed
Popova, LG, Belyaev, DW, Shuvalov, AV, Yurchenko, AA, Matalin, DA, Khramov, DE, Orlova, YV and Balnokin, YV (2018) In silico analyses of transcriptomes of the marine green microalga Dunaliella tertiolecta: identification of sequences encoding P-type ATPases. Molecular Biology 52, 520531.CrossRefGoogle ScholarPubMed
Raven, JA (1984) Energetics and Transport in Aquatic Plants. New York, NY: AR Liss.Google Scholar
Raven, JA (2009 a) Functional evolution of photochemical energy transformations in oxygen-producing organisms. Functional Plant Biology 36, 505515.CrossRefGoogle ScholarPubMed
Raven, JA (2009 b) Contribution of anoxygenic and oxygenic phototrophy and chemolithotrophy to carbon and oxygen fluxes in aquatic environments. Aquatic Microbial Ecology 56, 3552.CrossRefGoogle Scholar
Raven, JA (2013) Half a century of pursuing the pervasive proton. Progress in Botany 74, 334.CrossRefGoogle Scholar
Raven, JA (2017) Chloride: essential micronutrient and multifunctional beneficial ion. Journal of Experimental Botany 68, 359367.Google Scholar
Raven, JA and Beardall, J (2016) The ins and outs of CO2. Journal of Experimental Botany 67, 113.CrossRefGoogle ScholarPubMed
Raven, JA and Hurd, CJ (2012) Ecophysiology of photosynthesis in macroalgae. Photosynthesis Research 113, 105125.CrossRefGoogle ScholarPubMed
Raven, JA and Smith, FA (1980) Intracellular pH regulation in the giant-celled marine alga Chaetomorpha darwinii. Journal of Experimental Botany 31, 13571369.CrossRefGoogle Scholar
Raven, JA and Smith, FA (1981) H+ transport in the evolution of photosynthesis. Biosystems 14, 95111.CrossRefGoogle ScholarPubMed
Raven, JA and Smith, FA (1982) Solute transport at the plasmalemma and the early evolution of cells. Biosystems 15, 1326.CrossRefGoogle ScholarPubMed
Reed, RD and Collins, JC (1981) Membrane potential measurements of marine macroalgae: Porphyra purpurea and Ulva lactuca. Plant Cell and Environment 4, 257260.Google Scholar
Reed, RD, Collins, JC and Russell, G (1981) The effects of salinity upon ion content and ion transport of the marine red alga Porphyra purpurea (Roth) C. Ag. Journal of Experimental Botany 32, 347367.CrossRefGoogle Scholar
Remis, D, Simonis, W and Gimmler, H (1992) Measurements of the transmembrane electrical potential of Dunaliella acidophila by microelectrodes. Archives of Microbiology 158, 350355.CrossRefGoogle Scholar
Ritchie, RJ (1982) Comparison of a lipophilic cation and microelectrodes to measure membrane potentials of the giant-celled algae Chara australis (Chlorophyta) and Griffithsia monilis (Rhodophyta). Journal of Membrane Biology 69, 5763.CrossRefGoogle Scholar
Ritchie, RJ (1984) A critical assessment of the use of lipophilic cations as membrane potential probes. Progress in Biophysics and Molecular Biology 43, 132.CrossRefGoogle ScholarPubMed
Ritchie, RJ (1985) Energetic considerations of ion transport in Enteromorpha intestinalis (L) Link. New Phytologist 100, 524.CrossRefGoogle Scholar
Ritchie, RJ (1988) The ionic relations of Ulva lactuca. Journal of Plant Physiology 133, 183192.CrossRefGoogle Scholar
Ritchie, RJ (1992) Sodium transport and the origin of the membrane potential in the cyanobacterium Synechococcus R-2 (Anacystis nidulans) PCC 7943. Journal of Plant Physiology 139, 320330.CrossRefGoogle Scholar
Roberts, DSE, Illot, I and Brownlee, C (1994) Cytoplasmic calcium and Fucus egg activation. Development (Cambridge, England) 120, 155163.Google Scholar
Rodríguez-Navarro, A and Benito, B (2010) Sodium or potassium efflux ATPase. A fungal, bryophyte and protozoal enzyme. Biochimica et Biophysica Acta 1799, 18411853.CrossRefGoogle Scholar
Rubio, L, Linares-Rueda, A, Garcia-Sánchez, MJ and Fernández, JA (2005) Physiological evidence for a sodium dependent high affinity phosphate and nitrate transport at the plasma membrane of leaf and root cells of Zostera marina L. Journal of Experimental Botany 56, 613622.CrossRefGoogle ScholarPubMed
Rubio, L, Belver, A, Venene, K, Garcia-Sánchez, MJ and Fernández, JA (2011) Evidence for a sodium efflux mechanism in the leaf cells of the seagrass Zostera marina L. Journal of Experimental Marine Biology and Ecology 402, 5664.CrossRefGoogle Scholar
Rubio, L, Garcia, D, Garcia-Sánchez, MJ, Niell, FX, Felle, HH and Fernández, JA (2017) Direct uptake of HCO3 in the marine angiosperm Posidonia oceanica (L.) Delile driven by a plasma membrane H+ economy. Plant Cell and Environment 40, 28202830.CrossRefGoogle ScholarPubMed
Rubio, L, Garcia-Pérez, D, Garcia-Sánchez, MJ and Fernández, JA (2018) Na+-dependent high-affinity nitrate, phosphate and amino acids transport in leaf cells of the seagrass Posidonia oceanica (L.) Delile. International Journal of Molecular Sciences 12, art. 1570.Google Scholar
Russell, MJ and Hall, AJ (1997) The emergence of life from iron monosulphide bubbles at a submarine hydrothermal vent. Journal of the Geological Society of London 154, 377402.CrossRefGoogle Scholar
Saddler, HD (1970a) The membrane potential of Acetabularia mediterranea. Journal of General Physiology 55, 802821.CrossRefGoogle Scholar
Saddler, HDW (1970b) The ionic relations of Acetabularia mediterranea. Journal of Experimental Botany 21, 345359.CrossRefGoogle Scholar
Saier, MH Jr (2000) Vectorial metabolism and evolution of transport systems. Journal of Bacteriology 182, 50295035.CrossRefGoogle ScholarPubMed
Saier, MH Jr, Reddy, VS, Tsu, BV, Ahmed, MS, Li, C and Moreno-Hagelsieb, G (2015) The Transporter Classification Database (TCDB): recent advances. Nucleic Acid Research 44, D372D379.CrossRefGoogle ScholarPubMed
Sánchez-Baracaldo, P, Ridgwell, A and Raven, JA (2014) A Neoproterozoic transition in the marine nitrogen cycle. Current Biology 24, 652657.CrossRefGoogle ScholarPubMed
Sánchez-Baracaldo, P, Raven, JA, Pisani, D and Knoll, AH (2017) Early photosynthetic eukaryotes inhabited low-salinity habitats. Proceedings of the National Academy of Sciences USA 114, E7737E7745.CrossRefGoogle ScholarPubMed
Scherer, S and Böger, P (1984) Vanadate-sensitive proton efflux by filamentous cysnobacteria. FEMS Microbiology Letters 22, 190192.CrossRefGoogle Scholar
Scherer, S, Stürzl, E and Böger, P (1984) Oxygen-dependent proton efflux in Cyanobacteria (blue-green algae). Journal of Bacteriology 158, 609614.CrossRefGoogle Scholar
Schultze, M, Forberich, B, Rexroth, S, Dyczmons, NG, Roegner, M and Appel, J (2009) Localization of cytochrome b6f complexes implies an incomplete respiratory chain in the cytoplasmic membranes of the cyanobacterium Synechocystis sp. PCC 6803. Biochimica et Biophysica Acta 1787, 14791485.CrossRefGoogle ScholarPubMed
Shevchenko, V, Mager, T, Kovalev, K, Polovinkin, V, Alekseev, A, Juettner, J, Chizhov, I, Bamann, C, Vavourakis, C, Ghai, R, Gushchin, I, Borshchevskiy, V, Rogachev, A, Melnikov, I, Popov, A, Balandin, T, Rodriguez-Valera, F, Manstein, DJ, Bueldt, G, Bamberg, E and Gordelly, V (2017) Inward H+ pump xenorhodopsin: mechanism and alternative optogenetic approach. Science Advances 3, e1603187.CrossRefGoogle ScholarPubMed
Shono, M, Wada, M and Fujii, T (1995) Partial purification of a Na+-ATPase from the plasma membrane of the marine alga Heterosigma akashiwo. Plant Physiology 108, 16151621.CrossRefGoogle ScholarPubMed
Shono, M, Hara, Y, Wada, M and Fujii, T (1996) A sodium pump in the plasma membrane of the marine alga Heterosigma akashiwo. Plant and Cell Physiology 37, 385388.CrossRefGoogle Scholar
Shono, M, Wada, M, Hara, Y and Fuji, T (2001) Molecular cloning of Na+-ATPase cDNA from a marine alga, Heterosigma akashiwo. Biochimica et Biophysica Acta 1511, 193199.CrossRefGoogle ScholarPubMed
Shumkova, GA, Popova, LG and Balnokin, YV (2000) Export of Na+ from cells of the halotolerant microalga Dunaliella maritima: Na+:H+ antiporter or primary Na+ pump? Biochemistry (Moscow) 65, 917923.Google ScholarPubMed
Sikes, CS and Wilbur, KH (1982) Functions of coccolithophore formation. Limnology and Oceanography 27, 1826.CrossRefGoogle Scholar
Siver, PA, Velez, M, Clivetti, M and Binda, P (2018) Early freshwater diatoms from the Cretaceous Battle formation in Western Canada. Palaios 33, 525534.CrossRefGoogle Scholar
Skulachev, VP (1984) Sodium bioenergetics. Trends in Biochemical Science 9, 483485.CrossRefGoogle Scholar
Skulachev, VP (1989) The sodium cycle: a novel type of bacterial energetics. Journal of Bioenegetics and Biomembranes 21, 635647.CrossRefGoogle ScholarPubMed
Slamovits, CH, Okamoto, N, Burri, L, James, ER and Keeling, PJ (2011) A bacterial proteorhodopsin proton pump in marine eukaryotes. Nature Communications 2, art. 183.CrossRefGoogle ScholarPubMed
Smahel, M, Hamann, A and Gradmann, D (1990) The prime plasmalemma ATPase of the halophilic alga Dunaliella bioculata: purification and characterisation. Planta 181, 496504.CrossRefGoogle Scholar
Smahel, H, Klieber, H-G and Gradmann, D (1992) Vanadate-sensitive ATPase in the plasmalemma of Acetabularia: biochemical and kinetic characteristics. Planta 188, 6264.CrossRefGoogle Scholar
Smith, FA and Raven, JA (1979) Intracellular pH and its regulation. Annual Review of Plant Physiology 30, 289311.CrossRefGoogle Scholar
Søndergaard, D and Pedersen, CNS (2015) PATBox: a toolbox for classification and analysis of P-type ATPases. PLoS ONE 10, art. e0139571.CrossRefGoogle ScholarPubMed
Soontharapirakkul, K and Incharoensakdi, A (2010) Na+-stimulated ATPase of alkalophilic halotolerant cyanobacterium Aphanothece halophytica translocates Na+ into proteoliposomes via Na+ uniport mechanism. BMC Biochemistry 11, art. 30.CrossRefGoogle Scholar
Soontharapirakkul, K, Promden, W, Yamada, N, Kageyama, H, Incharoensakdi, A, Iwamoto-Kihara, A and Takabe, T (2011) Halotolerant cyanobacterium Aphanothece halophytica contains an Na+-dependent F1FO-ATP synthase with a potential role in salt-stress tolerance. Journal of Biological Chemistry 286, 1016910176.CrossRefGoogle Scholar
Suffrian, K, Schulz, KG, Gurowska, MA, Riebesell, U and Bleich, M (2011) Cellular pH differences in Emiliania huxleyi reveal pronounced membrane proton permeability. New Phytologist 190, 595608.CrossRefGoogle Scholar
Tamogami, J, Kikuyawa, T, Wada, T, Demura, M, Kimura-Someya, T, Shirouzo, M, Yokayama, S, Miauchi, S, Simono, K and Kamo, N (2017) Evidence of two O-like intermediates in the photocycle of Acetabularia rhodopsin II, a light-driven proton pump from a marine alga. Biophysics Physicobiology 14, 4955.CrossRefGoogle Scholar
Taylor, AR (2009) A fast Na+/Ca2+-based action potential in a marine diatom. PLoS ONE 4, art. e4966.CrossRefGoogle Scholar
Taylor, AR and Brownlee, C (1993) Calcium and potassium currents in the Fucus egg. Planta 189, 109119.CrossRefGoogle Scholar
Taylor, AR, Chrachri, A, Wheeler, C, Goddard, H and Brownlee, C (2011) A voltage-gated H+ channel underlies pH homeostasis in calcifying coccolithophores. PLoS Biology 9, art. e1001085.CrossRefGoogle Scholar
Taylor, AR, Brownlee, C and Wheeler, G (2017) Coccolithophore cell biology: chalking up progress. Annual Review of Marine Science 9, 283310.CrossRefGoogle ScholarPubMed
Thever, MD and Saier, MH Jr (2009) Bioinformatic characterization of P-type ATPases encoded within the fully sequenced genomes of 26 eukaryotes. Journal of Membrane Biology 229, 115130.CrossRefGoogle ScholarPubMed
Thompson, SEM, Callow, JA, Callow, ME, Wheeler, GC, Taylor, AH and Brownlee, C (2007) Membrane recycling and calcium dynamics during settlement and adhesion of zoospores on the green alga Ulva linza. Plant Cell and Environment 30, 723744.CrossRefGoogle ScholarPubMed
Tragin, M and Vaulot, D (2018) Green microalgae in marine coastal waters: the Ocean Sampling Day (OSD) dataset. Scientific Reports 8, art. 14020.CrossRefGoogle ScholarPubMed
Uji, T, Hirata, R, Mikami, K, Mizuto, H and Saga, N (2012 a) Molecular characteristics and expression analysis of sodium pump genes in the marine red alga Porphyra yezoensis. Molecular Biology Reports 39, 79737980.CrossRefGoogle Scholar
Uji, T, Monma, R, Mizuta, H and Saga, N (2012 b) Molecular characterization and expression of two Na+/H+ antiporter genes in the marine red alga Porphyra yezoensis. Fisheries Science 78, 985991.CrossRefGoogle Scholar
Vader, A, Laughinghouse IV, HD, Griffiths, C, Jakobsen, KS and Gabrielsen, TM (2018) Proton-pumping rhodopsins are abundantly expressed by microbial eukaryotes in a high-Arctic fjord. Environmental Microbiology 20, 890902.CrossRefGoogle Scholar
Van de Waal, DB, John, U, Ziveri, G-J, Hoins, M, Sluija, A and Röst, B (2013) Ocean acidification in a marine dinoflagellate. PLoS ONE 8, e65987.CrossRefGoogle Scholar
Wada, T, Shimono, K, Kikukawa, T, Hato, M, Shinya, N, Ki, SY, Kimura-Somega, T, Shirouza, M, Taogami, J, Miyachi, S, Juing, KH, Kamo, N and Yokoyama, S (2011) Crystal structure of the eukaryotic light-driven proton-pumping rhodopsin, Acetabularia rhodopsin II, from marine algae. Journal of Molecular Biology 411, 986998.CrossRefGoogle Scholar
Walker, NA, Smith, FA and Cathers, IR (1980) Bicarbonate assimilation by freshwater charophytes and higher plants. I. Membrane transport of bicarbonate is not proven. Journal of Membrane Biology 12, 241256.Google Scholar
Wang, Q, Li, H and Post, AF (2000) Nitrate assimilation genes of the marine diazotrophic, filamentous cyanobacterium Trichodesmium sp. Strain WH9601. Journal of Bacteriology 182, 17641767.CrossRefGoogle ScholarPubMed
Wegmann, K (1986) Osmoregulation in eukaryotic algae. FEMS Microbiology Reviews 39, 3743.CrossRefGoogle Scholar
Weiss, M and Pick, U (1996) Primary structure and effect of pH on the expression of the plasma membrane H+-ATPase from Dunaliella acidophila and Dunaliella salina. Plant Physiology 112, 16931702.CrossRefGoogle ScholarPubMed
Wheeler, G, Helliwell, K and Brownlee, G (2019) Calcium signalling in algae. Perspectives in Phycology 6, 110.CrossRefGoogle Scholar
Wiangno, K, Raksajit, W and Incharoensakdi, A (2007) Presence of a Na+-type ATPase in the plasma membrane of the alkaliphilic halotolerant cyanobacterium Aphanothece halophytica. FEMS Microbiology Letters 270, 139145.CrossRefGoogle Scholar
Wieczorek, H, Brown, D, Grinstein, S, Ehrenfeld, J and Harvey, WR (1999) Animal plasma membrane energization by protein-motive V-ATPases. Bioessays 21, 637648.3.0.CO;2-W>CrossRefGoogle ScholarPubMed
Wolf, AH, Slayman, C and Gradmann, D (1995) Primary structure of the plasma membrane H+-ATPase from the halotolerant alga Dunaliella bioculata. Plant Molecular Biology 26, 657666.CrossRefGoogle Scholar
Wu, J and Seliskar, DM (1998) Salinity adaptation of plasma membrane H+-ATPase in the salt marsh plant Spartina patens: ATP hydrolysis and enzyme kinetics. Journal of Experimental Botany 49, 10051013.CrossRefGoogle Scholar
Yoshizawa, S, Kumuga, Y, Kim, H, Ogawa, Y, Hayashi, T, Iwasaki, W, DeLong, EF and Kogure, K (2014) Functional characterization of flavobacteria rhodopsin reveals a unique class of light-driven chloride pumps in bacteria. Proceedings of the National Academy of Sciences USA 111, 67326737.CrossRefGoogle Scholar
Yuan, F, Leng, B and Wang, B (2016) Progress in studying salt secretion from the salt glands in recretohalophytes: how do plants secrete salt? Frontiers in Plant Science 7, art. 977.CrossRefGoogle ScholarPubMed
Zhang, X, Ye, N, Liang, C, Mou, S, Fan, Z, Xu, J, Xu, D and Zhuang, Z (2012) De novo sequencing and analysis of the Ulva linza transcriptome to discover putative mechanisms associated with its successful colonization of coastal ecosystems. BMC Genomics 13, art. 565.CrossRefGoogle ScholarPubMed
Figure 0

Table 1. Energy sources for primary active transporters of H+, Na+ and Cl in the Archaea, Bacteria and Eukarya

Figure 1

Table 2. Intracellular H+ production and consumption in redox reactions, and calcification, in marine phytoplankton

Figure 2

Table 3. Phylogenetic distribution of plasmalemma ion transport ATPases in marine photosynthetic eukaryotes