Physiological roles of nonselective cation ... - Wiley Online Library

116 downloads 6905 Views 330KB Size Report
Feb 27, 2007 - Email: [email protected]. Received: 27 February ..... Na+ entry constitutes an efficient and relatively cheap form to restore the osmotic ... membrane voltage, suggesting these proteins lack any domains for sensing changes ...
Review

Blackwell Publishing Ltd

Tansley review Physiological roles of nonselective cation channels in plants: from salt stress to signalling and development

Author for correspondence: Vadim Demidchik Tel: +44 1206873322 Fax: +44 1206872592 Email: [email protected]

Vadim Demidchik1 and Frans J. M. Maathuis2 1

Department of Biological Sciences, University of Essex CO4 3SQ, Colchester, UK; 2Department of

Biology, Area 9, University of York, York YO10 5DD, UK

Received: 27 February 2007 Accepted: 18 April 2007

Contents Summary

387

VI.

Purine signalling

396

I.

Introduction

388

VII. Mechanosensitive ion channels

396

II.

Constitutive plasma membrane NSCCs

389

VIII. Vacuolar NSCCs

396

III.

NSCCs activated by reactive oxygen species (ROS-NSCCs)

392

IX.

Cation channels sensitive to elicitors

397

X.

NSCCs acting in concert

398

IV.

Cyclic nucleotide-gated channels (CNGCs)

393 XI.

Conclusions and perspectives

398

V.

Amino acid-gated NSCCs

395 References

400

Summary Key words: cytosolic free calcium, development, mineral nutrition, nonselective cation channels (NSCCs), plant neurotransmitter receptors, reactive oxygen species, salinity, signalling.

Nonselective cation channels (NSCCs) catalyse passive fluxes of cations through plant membranes. NSCCs do not, or only to a small extent, select between monovalent cations, and several are also permeable to divalent cations. Although a number of NSCC genes has been identified in plant genomes, a direct correlation between gene products and in vivo observed currents is still largely absent for most NSCCs. In this review, physiological functions and molecular properties of NSCCs are critically discussed. Recent studies have demonstrated that NSCCs are directly involved in a multitude of stress responses, growth and development, uptake of nutrients and calcium signalling. NSCCs can also function in the perception of external stimuli and as signal transducers for reactive oxygen species, pathogen elicitors, cyclic nucleotides, membrane stretch, amino acids and purines. New Phytologist (2007) 175: 387–404 © The Authors (2007). Journal compilation © New Phytologist (2007) doi: 10.1111/j.1469-8137.2007.02128.x

www.newphytologist.org

387

388 Review

Tansley review

Table 1 Glossary CaMBD CNBD CNGC Constitutive activity DA-NSCC Depolarization Elicitor FV channel GLRs HA-NSCC HACaC HR Hyperpolarization KOR Ligand MC MSLs NSCC Rectification ROS

SV channel VI-NSCC

Calmodulin-binding domain. A site found on cyclic nucleotide-gated channels that affects binding of ligand and therefore channel activity Cyclic nucleotide-binding domain. A site found on cyclic nucleotide-gated channels where the binding of ligands (cAMP or cGMP) occurs that leads to channel activation Cyclic nucleotide-gated channel. A type of nonselective ion channel that is activated by binding of cyclic nucleotides (cAMP or cGMP) Ion channel activity that can be recorded at all applied membrane voltages and does not require ligand for activation Depolarization-activated, nonselective cation channel. A nonselective cation channel that shows more activity at positive membrane voltages compared with negative membrane voltages Decrease in membrane voltage leading to less negative values Compound released during pathogen attack such as cell wall fragments. Perception of elicitors starts a signalling cascade involving ion channels ‘Fast vacuolar’ channel. A nonselective cation channel characterized in plant tonoplasts, possibly involved in K+ homeostasis Glutamate receptor-like genes. Gene family of putative glutamate receptors in Arabidopsis Hyperpolarization-activated, nonselective cation channel. A nonselective cation channel that shows more activity at negative membrane voltages compared with positive membrane voltages Hyperpolarization-activated, Ca2+ channel. Class of ion channel that has been characterized in planta as conducting Ca2+ influx. Further properties of this channel suggest HACaCs are a subclass of HA-NSCCs Hydroxyl radical Increase in membrane voltage leading to more negative values K+ outward rectifier. Class of K+-selective, voltage-dependent channels that is activated only at positive membrane voltages, leading to strong outwardly rectifying current Compound that binds to the protein. In the context of ion channels, many ligands are known that bind to the channel protein to instigate a transition from the closed to the open state Mechanosensitive channel. Ion channel whose gating is modulated by physical forces on the membrane Mechanosensitive-like channel genes. Gene family in Arabidopsis of putative mechanosensitive channels Nonselective cation channel. Ion channels that conduct cations but not anions and show no or little discrimination between different cations Phenomenon where transmembrane currents in one direction are different from those in the opposite direction Reactive oxygen species. Compounds that contain oxygen in a reactive state and cause oxidation of other compounds. Physiologically, most important ROS include singlet oxygen, hydrogen peroxide, superoxide anion radicals and hydroxyl radicals. Note that hydrogen peroxide and superoxide anion radical are not particularly reactive but are nevertheless considered as ROS ‘Slow vacuolar’ channel. A nonselective cation channel characterized in plant tonoplasts, possibly involved in Ca2+ signalling Voltage-independent, nonselective cation channel. A nonselective cation channel, whose activity is not modulated by the membrane voltage

I. Introduction The concept that plant membranes exhibit permeability for specific cations was developed at the beginning of the 20th century by Osterhout and others (reviewed by Osterhout, 1958). These ideas triggered intensive studies of cation conductances in plant membranes that culminated in the discovery and detailed characterization of many types of plant cation channels (reviewed by Hedrich & Schroeder, 1989; Tester, 1990; Yurin et al., 1991; White, 1998; Demidchik et al., 2002b; Véry & Sentenac, 2003). In plant plasma membranes, the predominant type of ion channel is K+-selective and voltage-dependent (see review by Véry & Sentenac, 2003). However, during many in planta studies, currents were observed in plasma membranes and vacuolar membranes, which lacked voltage dependence and often showed no or little selectivity amongst cations. Initially, these currents were often regarded as artefactual and merely constituting a ‘leak’ without much physiological relevance.

New Phytologist (2007) 175: 387–404

However, it became clear that discrete single-channel events underlie these currents, and more recently researchers have come to recognize the importance of nonselective cation channels (NSCCs; see Table 1 for glossary of terms used in this review). NSCCs typically are permeable to a wide range of monovalent cations. For example, most NSCCs demonstrate K+/Na+ selectivity ratios between 0.3 and 3, whereas this ratio is well over 10 in outward rectifying K+-selective channels, and 100 or more in inward rectifying K+-selective channels (Véry & Sentenac, 2003). Similarly, many NSCCs conduct Ca2+ with Ca2+:K+ selectivity ratios between 0.5 and 2, whereas these ratios in animal Ca2+-selective channels range from 10 to over 50. There are several areas of plant research where interest in NSCCs is substantial. These include Ca2+ nutrition, Ca2+ signalling and salinity stress. Although Ca2+ currents can be recorded from plant membranes, these invariably seem to be mediated by ion channels that are not, or are only weakly, selective for Ca2+. Such Ca2+-permeable NSCCs play a number

www.newphytologist.org © The Authors (2007). Journal compilation © New Phytologist (2007)

Tansley review

of physiological roles. In Arabidopsis guard cells, Ca2+-permeable NSCCs are involved in stomatal regulation. The same type of NSCC may contribute to relaying Ca2+ signals derived from plant pathogens and Ca2+ signalling across the vacuolar membrane (Johannes & Sanders, 1995; Peiter et al., 2005), whereas in root epidermis, NSCCs are thought to play a role in Ca2+ acquisition (Demidchik et al., 2002a). Another area in which NSCCs appear to play a critical role is plant salinity stress. A major question concerns the mechanism of Na+ entry into plant roots and, similar to the situation for Ca2+, no Na+-selective ion channels have been found in plants. Based on the similarity of the Ca2+-dependent block of NSCC current and of Na+ influx in intact tissue, NSCCs were proposed to form a major pathway for Na+ entry into plants (Tyerman et al., 1997) and many later studies appear to confirm this notion (Maathuis & Sanders, 2001; Demidchik & Tester, 2002; Essah et al., 2003). Here, we will update our current knowledge regarding plant NSCCs. Although NSCCs have been found in several types of plant membrane, this review will mainly focus on plasma mambrane and vacuolar membrane NSCCs and will discuss data that have recently appeared, in particular regarding the physiological roles of NSCCs in plants.

II. Constitutive plasma membrane NSCCs Gating of many ion channels is strictly dependent on membrane voltage, and unless the permissible voltage range is applied, no channel activity will be observed. Similarly, ligand-gated channels typically remain inactive without the presence of agonist. However, several types of ion channel have been observed that are active, irrespective of the membrane voltages and without the need for addition of any ligand. Such channels therefore appear to be permanently (constitutively) active at all applied membrane voltages. Nevertheless, the extent of channel activity may still be modulated by many cellular factors and by membrane voltage (Fig. 1). Constitutive NSCCs can therefore be subdivided according to their voltage dependence into voltage-insensitive, depolarization-activated and hyperpolarization-activated NSCCs. Constitutive NSCCs were among the first cation channels identified in plants. In addition, they were initially considered as constitutive plasma membrane leak conductances (Yurin et al., 1991) or as classes of K+ channel with low selectivity for K+ (Tester, 1990; White & Tester, 1992). 1. Depolarization-activated NSCCs (DA-NSCCs) The first patch-clamp study on constitutive NSCCs from higher plants was carried out by Stoeckel & Takeda (1989). In the plasma membrane of Haemanthus and Clivia endosperm cells, these authors found NSCCs that were activated by membrane depolarization and stimulated by cytosolic Ca2+. DA-NSCCs were also found in other preparations: Arabidopsis

Review

Fig. 1 Idealized current/voltage relationships showing the voltage range in which voltage-gated K+-selective channels activate (a) and the voltage range in which respective classes of nonselective cation channel (NSCC) activate (b). Inwardly rectifying K+ channels (KIR) open whenever membrane potentials are hyperpolarized (negative) and do not show activity at positive membrane potentials. The activation of outwardly rectifying K+ channels (KOR) shows the opposite behaviour. Note that for both types of channel no activity occurs across a large voltage range. Constitutive NSCCs show activity across the entire voltage range and this is not affected by membrane voltage in the case of voltage-independent NSCCs (VI-NSCC). Depolarization-activated NSCCs (DA-NSCCs) show more activity at positive membrane potentials, whereas hyperpolarization-activated NSCCs (HA-NSCCs) show the opposite behaviour.

cultured cells (Cerana & Colombo, 1992), Arabidopsis leaf mesophyll (Spalding et al., 1992; Shabala et al., 2006), Arabidopsis root epidermis (Shabala et al., 2006), Arabidopsis guard cells (Pei et al., 1998), Hordeum vulgare root xylem (Wegner & Raschke, 1994; de Boer & Wegner, 1997), Thlaspi spp. mesophyll cells (Piñeros & Kochian, 2003), Arabidopsis pollen tubes (Becker et al., 2004), Phaseolus vulgaris seed coats (Zhang et al., 2002) and Phaseolus vulgaris cotyledon dermal cells (Zhang et al., 2004). Some of the recorded channels showed slow activation kinetics (Cerana & Colombo, 1992; Wegner & Raschke, 1994; de Boer & Wegner, 1997) but in most cases a characteristically fast or ‘instantaneous’ change in current level occurred (Zhang et al., 2000, 2002, 2004; Piñeros & Kochian, 2003; Shabala et al., 2006). Activation kinetics and voltage dependence of some DANSCCs are reminiscent of those observed for outwardly rectifying K+-selective channels (K+ outward rectifiers, KORs). However, in contrast to KORs, DA-NSCCs are insensitive to blocking by extracellular or intracellular Na+ (Shabala et al., 2006). DA-NSCCs can be effectively blocked by extracellular Ca2+ (Shabala et al., 2006), but this has also been observed for many other types of cation channels. Other compounds that may affect DA-NSCCs are TEA+, nifedipine, diltiazem and verapamil (reviewed by Demidchik et al., 2002b; Shabala et al., 2006). The physiological functions of DA-NSCCs appear related to their ability to catalyse influx and efflux of monovalent cations and influx of divalent cations. The electrochemical gradient for most divalent cations is inwardly directed. This means that DA-NSCCs could form a conduit for the influx of divalent cations. Some indications that Ca2+ influx occurs through DANSCCs were found in maize root stelar cells (Roberts & Tester,

© The Authors (2007). Journal compilation © New Phytologist (2007) www.newphytologist.org

New Phytologist (2007) 175: 387–404

389

390 Review

Tansley review

1997). In addition, Piñeros & Kochian (2003) demonstrated that passive Zn2+ influx in the Zn-hyperaccumulating Thlaspi caerulescens can occur through DA-NSCC-type conductance. Involvement of DA-NSCCs in monovalent cation fluxes may be widespread. For example, several studies have shown that DA-NSCCs contribute to K+ loading into the xylem (de Boer & Wegner, 1997) and K+ redistribution in bean seeds (Zhang et al., 2002, 2004). The relatively high permeability for Na+ of many DA-NSCCs suggests they are also important contributors to the uptake and translocation of Na+, and thus their function impacts on salt tolerance. During salt stress, plant cells can rapidly accumulate Na+ to a very high concentration that interferes, in particular, with essential roles of K+ in the cytosol. Thus, maintaining high cytosolic K+/Na+ ratios is an important component of salinity tolerance (Maathuis & Amtmann, 1999; Shabala et al., 2006), and minimizing salt-induced K+ efflux and Na+ influx would contribute to increased tolerance. During salt stress, both intra- and extracellular [Na+] are likely to be high. In such conditions, KORs are substantially blocked by Na+, preventing loss of cellular K+ (Shabala et al., 2006). Elevated external Ca2+ will further block KORs (Sokolik & Yurin, 1986; Yurin et al., 1991; Marten et al., 1999; Shabala et al., 2006). In such conditions, DA-NSCCs become a more prominent component of the membrane permeability to monovalent cations, potentially allowing K+ loss and entry of Na+ (Shabala et al., 2006) (Fig. 2). Na+ entry constitutes an efficient and relatively cheap form to restore the osmotic balance and, as such, plays an important role in plant salinity tolerance. However, excessive Na+ influx causes Na+ toxicity, and relatively high DA-NSCC K+/ Na+ selectivities may therefore provide adaptive advantages in this respect. A comparative study on Arabidopsis thaliana and

Fig. 2 Inhibition of cation channels leads to the amelioration of Na+ toxicity in higher plants. External Ca2+, and intracellular and extracellular Na+ can all have inhibitory effects on cation channels. Na+ blocks the K+-selective outwardly rectifying channel (KOR), thus preventing cellular K+ loss. Ca2+ blocks KOR, voltage-independent nonselective cation channels (VI-NSCCs) and depolarizationactivated nonselective cation channels (DA-NSCCs), thus preventing excessive Na+ influx and some K+ efflux.

New Phytologist (2007) 175: 387–404

its halophytic relative Thellungiella halophila showed an approximately fourfold higher PK/PNa of DA-NSCC-like channels in roots of Thellungiella compared with Arabidopsis (Volkov et al., 2004; Volkov & Amtmann, 2006; Wang et al., 2006). The PK/PNa of other NSCCs that are involved in Na+ influx was also found to be lower in Thellungiella halophila. 2. Voltage-independent NSCCs (VI-NSCCs) The open probability of VI-NSCCs is not modulated by membrane voltage, suggesting these proteins lack any domains for sensing changes in membrane voltage. Thus, most VINSCCs equally conduct outward and inward current. VINSCCs are not inhibited by organic antagonists of Ca2+ channels such as nifedipine and verapamil, or by conventional blockers of K+ channels (TEA+, TTX, Cs+, Li+ and Na+). However, they are sensitive to Gd3+ and La3+. VI-NSCCs can be subdivided into two groups on the basis of blockage by cations such as Ca2+, Ba2+, Mg2+ and Zn2+. One class is partially blocked, whereas another class is both insensitive and permeable to these divalent cations. Inhibition by quinine is another aspect that pertains to some VI-NSCCs (Demidchik & Tester, 2002), but not others (White & Lemtiri-Chlieh, 1995). Most studies on VI-NSCCs have been carried out using root cells (White & Lemtiri-Chlieh, 1995; Roberts & Tester, 1997; Tyerman et al., 1997; White, 1997; Buschmann et al., 2000; Maathuis & Sanders, 2001; Demidchik & Tester, 2002; Demidchik et al., 2002a; Volkov et al., 2004; Murthy & Tester, 2006; Shabala et al., 2006; Volkov & Amtmann, 2006), which reflects the potential physiological significance of this transport system for the nonspecific uptake of cations. VI-NSCCs and VI-NSCC-like conductances have also been found in guard cells (Véry et al., 1998), Arabidopsis mesophyll cells (Shabala et al., 2006), cultured Arabidopsis cells (Amtmann et al., 1997), Arabidopsis pollen tubes (Becker et al., 2004), expanding Pisum sativum leaf epidermis (Elzenga & van Volkenburgh, 1994) and motor cells of Samanea saman (Yu et al., Moran, 2001), whereas some of the earlier recordings were obtained using green algae Nitella flexilis (Yurin et al., 1991; Sokolik, 1999). Although conclusive evidence is still lacking, it is widely considered that the main physiological function of VI-NSCCs is to catalyse uptake of cations. This may include monovalent cations like Na+ and NH4+ that do not permeate selective K+ channels, auxiliary uptake capacity for K+ when K+-selective channels are inhibited or not expressed, and the uptake of divalent cations such as Ca2+ and Mg2+ for which no selective pathways have been found in plant membranes at resting membrane potentials. This function can be highly beneficial where the acquisition of essential nutrients such as Ca2+, Mg2+, K+, NH4+, Mn2+ and Zn2+ is concerned, but also creates the potential for influx of harmful ions such as Na+, Cs+, Pb2+, Hg+ and Cd2+. Examples of both processes are discussed later.

www.newphytologist.org © The Authors (2007). Journal compilation © New Phytologist (2007)

Tansley review

All VI-NSCCs studied so far have been shown to be permeable to K+ (Demidchik et al., 2002b) and with the exception of VI-NSCCs in pea leaf epidermis (Elzenga & van Volkenburgh, 1994) and maize root cortex (Roberts & Tester, 1997), K+ is typically the most permeant ion. In most conditions, K+-selective inward rectifiers (KIRs) dominate passive K+ influx into root cells (Maathuis & Sanders, 1993), a contention that was given extra credence after Hirsch et al. (1998) showed that a loss of function in the main root KIR, AKT1, significantly reduced K+ uptake. However, a large component of the low-affinity K+ uptake remains in akt1-1 mutants, and although it is not clear what the exact mechanism is for this influx, VI-NSCCs in conjunction with KORs (Maathuis & Sanders, 1997) could mediate a significant part of this. Thus, VI-NSCCs may provide extra K+ uptake capacity, especially in conditions where the main low-affinity uptake pathway is inhibited. As was discussed for DA-NSCCs, VI-NSCCs are prime candidates for Na+ entry. Indeed, Na+ permeability of VINSCCs has been demonstrated in a range of tissues and species (Stoeckel & Takeda, 1989; Elzenga & van Volkenburgh, 1994; White & Lemtiri-Chlieh, 1995; Amtmann et al., 1997; Roberts & Tester, 1997; Tyerman et al., 1997; Véry et al., 1998; Maathuis & Sanders, 2001; Demidchik & Tester, 2002) and there is now substantial evidence that root Na+ influx is to a large extent catalysed by VI-NSCCs (Maathuis & Sanders, 2001; Demidchik & Tester, 2002; Demidchik et al., 2002b; White & Davenport, 2002; Tester & Davenport, 2003; Maathuis, 2006a; Shabala et al., 2006). This notion is further supported by the observation that Ca2+ has a direct blocking effect on Na+ currents through VI-NSCCs (see Fig. 2) and on Na+ influx into intact tissue (Tyerman et al., 1997; Essah et al., 2003). The latter phenomenon is thought to be a main part of the ameliorative action of Ca2+ on salinity. Na+ influx in intact tissue is also sensitive to VI-NSCC blockers such as quinine, lanthanides and histidine modifiers (Essah et al., 2003; Wang et al., 2006). Plants need large amounts of Ca2+, and understanding the mechanism of Ca2+ uptake from the soil is therefore of great value (Marschner, 1995; Welch, 1995). No Ca2+-selective channel has been found in plant membranes (Bothwell & Ng, 2005) and plant VI-NSCCs may be important in plant Ca2+ nutrition as a pathway for Ca2+ uptake, a notion that appears to contradict the blocking effect of Ca2+ on NSCCs (Fig. 2). The Ca2+ block derives from its high-affinity binding to the channel pore. However, the affinity of the pore for Ca2+ is low enough still to allow considerable Ca2+ permeation. A key role of VI-NSCCs in nutritional Ca2+ influx was demonstrated in Arabidopsis root epidermal cells (Demidchik et al., 2002a): 45Ca2+ flux measurements showed that Ca2+ uptake was blocked by Gd3+, a nonspecific blocker of VI-NSCCs. In addition, [Ca2+]cyt measured in Arabidopsis roots by Ca2+/ aequorin chemiluminometry revealed a linear voltage dependence, consistent with the voltage independence of VI-NSCCs.

Review

A further important feature of VI-NSCCs is their enhanced activity in cells of the elongation zone where Ca2+ influx is particularly high. A reason for this requirement may be the Ca2+-dependent exocytosis that sustains cell expansion. 3. Hyperpolarization-activated NSCCs (HA-NSCCs) Plants do not appear to have voltage-gated Ca2+-selective channels (Bothwell & Ng, 2005), but recordings on many plant membranes do show the presence of channels that conduct Ca2+ and these are believed to be important in the generation of Ca2+ signals. Although some of these activate via depolarization (Thion et al., 1998), most activate at very negative membrane voltages and are often described as hyperpolarization-activated Ca2+ channels (HACaCs) (Gelli & Blumwald, 1997; Gelli et al., 1997; Hamilton et al., 2000; Kiegle et al., 2000; Véry & Davies, 2000; Demidchik et al., 2002a, 2007). Unfortunately, a proper examination of monovalent ion selectivity in HACaCs is often lacking, although earlier reports (Schroeder & Hagiwara, 1990) and selectivity analyses of plant Ca2+-permeable cation channels in lipid bilayers (Aleksandrov et al., 1976; White & Tester, 1992; White & Davenport, 2002) and algal cells (Lunevsky et al., 1980, 1983) showed considerable permeability to monovalent ions. It is therefore likely that most HACaCs are in fact Ca2+permeable HA-NSCCs and we will regard both terms as interchangeable. Hyperpolarization-activated Ca2+ channels can catalyse the large Ca2+ influx that is required for plant elongation growth of root hairs and cells in the elongation zone (Kiegle et al., 2000; Véry & Davies, 2000; Demidchik et al., 2002a, 2007), and, similar to Ca2+-permeable VI-NSCCs, HACaCs are more active in growing tissues (Véry & Davies, 2000; Demidchik et al., 2007). In many cells, the resting membrane potential may be too positive for significant HACaC activity (Maathuis & Sanders, 1993; Demidchik et al., 2002a), but in such conditions Ca2+-permeable VI-NSCCs may elevate [Ca2+]cyt, which shifts the activation potential of HA-NSCCs to more positive voltages (Fig. 3). Alternatively, HACaC activity could be increased through the hyperpolarizing action of plasma membrane H+-ATPases (Miedema et al., 2001). The stimulation of HACaC activity by elicitors suggests that certain classes of HA-NSCCs are involved in Ca2+ signalling, which is part of the early response to pathogen attack (Gelli & Blumwald, 1997; Gelli et al., 1997). Interestingly, HACaC activity in root epidermal protoplasts appears approx. 40–60 min after formation of the whole-cell configuration (Demidchik et al., 2002a). This behaviour suggests that HACaCs either can be inserted into the plasma membrane or are activated by some unknown factor such as reactive oxygen species (ROS). Plant annexins have been shown to catalyse transmembrane Ca2+ fluxes (Clark & Roux, 1995) and, in the presence of Ca2+ and ROS, animal annexins move to the plasma membrane and form Ca2+-permeable

© The Authors (2007). Journal compilation © New Phytologist (2007) www.newphytologist.org

New Phytologist (2007) 175: 387–404

391

392 Review

Tansley review

Fig. 3 Constitutive Ca2+-permeable cation channels in the plasma membrane of higher plants. Left: the mechanism of possible involvement of voltage-independent nonselective cation channels (VI-NSCCs) in activation of hyperpolarization-activated Ca2+ channels (HACaCs) that results in a large elevation of [Ca2+]cyt and stimulation of exocytosis and cell elongation growth. Right: current–voltage relationships (I–V curves) of constitutive Ca2+-permeable cation channels. Notably, elevation of [Ca2+]cyt shifts the HACaC I-V curve to a more positive voltage range. DACaC, depolarization-activated Ca2+ channel.

channels (Gerke & Moss, 2002). Animal annexin-type Ca2+ conductances are similar to HACaCs: they are activated by intracellular Ca2+ and hyperpolarization. Thus, annexins could function as HACaC-like Ca2+ channels in plants (Clark & Roux, 1995; White et al., 2002). Apart from Ca2+ nutrition and signalling, HA-NSCCs may also have specialized tasks in specific membranes. A very special type of HA-NSCCs was found in the peribacteroid membrane that surrounds bacterial symbionts in nitrogen-fixing root nodules (Tyerman et al., 1995). These hyperpolarizationactivated channels showed permeability to monovalent and divalent cations, required cytosolic Mg2+ for their inward rectification and were inhibited by polyamines and Ca2+ in the symbiosome lumen (Whitehead et al., 1998, 2001; Roberts & Tyerman, 2002). The peribacteroid membrane controls fluxes of fixed nitrogen (NH3 or NH4+) between bacteria and the host cells (Obermeyer & Tyerman, 2005) and the HANSCCs found in this membrane mediate the NH4+ flux, thus playing an important role in the nitrogen metabolism of legumes (Tyerman et al., 1995). Interestingly, the HA-NSCCs present in the peribacteroid membrane have an extremely low unitary conductance of approx. 0.11 pS, which could only be determined using noise spectrum analysis. Such low unitary conductance is atypical for most ion channels but more often associated with carriers, or ion channels with complex gating (Tyerman et al., 1995; Obermeyer & Tyerman, 2005).

III. NSCCs activated by reactive oxygen species (ROS-NSCCs) Reactive oxygen species integrate signalling pathways involved in plant stress responses, growth and development, gravitropism, hormone action, and many other physiological phenomena (Apel & Hirt, 2004). However, the molecular mechanisms of ROS regulatory action are poorly understood. In animals,

New Phytologist (2007) 175: 387–404

ROS-activated cation channels mostly belong to NSCCs and K+ channel types, and have been shown to play a multitude of roles: from regulation of blood pressure to maintaining neuronal networks (reviewed by Lahiri et al., 2006). Plant ROS-NSCCs were initially detected in the green alga Nitella flexilis by Demidchik et al. (1996, 1997a,b, 2001). Extracellular application of redox-active transition metals, Cu2+ and Fe3+, which leads to the production of hydroxyl radicals (HRs), activated a voltage-independent, instantaneous, nonselective cationic conductance that was sensitive to divalent cations, protons and the organic Ca2+ channel blocker, nifedipine. Analysis of the temperature dependence of the Nitella conductance activated by Cu2+ suggested an ion channelbased mechanism with a typical Q10 of 1.2–1.6. The idea that HRs can activate cation channels was later examined in higher plants (Demidchik et al., 2003b; Foreman et al., 2003; Inoue et al., 2005). In Arabidopsis epidermal root cells, generation of extracellular HRs led to the activation of cation-selective channels with relative permeabilities of K+ (1.00) ≈ NH4+ (0.91) ≈ Na+ (0.71) ≈ Cs+ (0.67) > Ba2+ (0.32) ≈ Ca2+ (0.24) > TEA+ (0.09). A special aspect of ROS-NSCCs is their key role in elongation/expansion of plant cells (Demidchik et al., 2003b, 2007; Foreman et al., 2003) (Fig. 4). In elongating root hairs and in root elongation zone cells, HR-activated NSCCs showed significantly higher activity than in mature cells. The resulting increased Ca2+ influx and stimulated actin/myosin interaction lead to accelerated exocytosis, polar vesicle embedment and cell elongation (reviewed by Carol & Dolan, 2006). During this process, a plasma membrane localized NADPH oxidase was shown to produce the ROS necessary for the activation of Ca2+-permeable NSCCs (Foreman et al., 2003) (Fig. 4). Mutants lacking the oxidase (rhd2-1) produced far less extracellular ROS, did not form root hairs and exhibited stunted cell expansion in the elongation zone (Foreman et al., 2003).

www.newphytologist.org © The Authors (2007). Journal compilation © New Phytologist (2007)

Tansley review

Review

Fig. 4 Reactive oxygen species-activated nonselective cation channels (ROS-NSCCs) in the plant plasma membrane. ROS-NSCCs are activated by hydroxyl radicals and hydrogen peroxide (H2O2) and are involved in elongation growth and in stress responses. ROS necessary for activation of ROS-NSCCs can be produced by plasma membrane NADPH oxidases and cell wall-bound peroxidases.

Interestingly, NADPH oxidase itself is stimulated by cytosolic Ca2+, thus forming a positive feedback mechanism to amplify ROS signals (Fig. 4). Other oxidases, such as cell wall peroxidases, can also produce ROS for Ca2+-driven elongation growth (reviewed by Kawano, 2003), and activation of ROSNSCCs may also participate in pollen tube growth where a similar interaction between ROS and elevated [Ca2+]cyt has recently been found (Malho et al., 2006; McInnis et al., 2006). Hydrogen peroxide (H2O2) did not activate whole-cell currents in protoplasts isolated from the Arabidopsis mature root epidermis. However, H2O2 does induce inward Ca2+ currents in protoplasts from the epidermal elongation zone (Demidchik et al., 2007). Several mechanisms could be responsible for these contrasting results: (i) ROS-NSCCs in growing tissues may have a different structure and/or regulatory properties; (ii) young elongating cells may contain a higher density of H2O2-permeable aquaporins (Eisenbarth & Weig, 2005; Bienert et al., 2006), allowing H2O2 to access potential regulatory sites in the cytosol; (iii) elongating cells have a greater capacity to generate H2O2. Localization of H2O2 activation sites at the cytoplasmic side has recently been confirmed by Demidchik et al. (2007) (Fig. 4). These authors applied H2O2 extracellularly and intracellularly to mature epidermal protoplasts in different patch-clamp configurations. H2O2-induced activation was observed only when H2O2 was applied to the cytosolic side of the membrane. The same study showed that the recorded channels had no selectivity amongst divalent cations and a unitary Ca2+ conductance that was similar to

that of HACaCs in the same membranes. The latter finding suggests that ROS-NSCCs and HACaCs are the same channel. The data also suggest that there are at least two different types of ROS-NSCC: HR-activated, which were found in all tested root cell types; and H2O2-activated, in the elongation zone. Reactive oxygen species modulation of NSCCs tightly links activity of these channels to signalling processes that involve ROS (reviewed by Pitzschke et al., 2006). Of particular interest in this respect is abscisic acid (ABA) signalling (reviewed by Pei & Kuchitsu, 2005). This hormone is involved in the regulation of stomatal closure, seed dormancy, flowering, activation of antioxidants and defence reactions, stress responses and other phenomena, which are also known to be accompanied by ROS accumulation. ABA can stimulate NADPH oxidase-mediated generation of ROS, leading to the activation of Ca2+-permeable NSCCs, Ca2+ influx and stomatal closure (Pei et al., 2000; Kwak et al., 2003). Guard cell H2O2-activated NSCCs showed selectivity and voltage dependence similar to H2O2-activated channels in root epidermis. However, they were not studied at the single-channel level and further examination is required to reveal whether root and leaf ROS-NSCCs are the same.

IV. Cyclic nucleotide-gated channels (CNGCs) The second messengers 3′,5′-cyclic adenosine monophosphate (cAMP) and 3′,5′-cyclic guanosine monophosphate (cGMP) participate in many aspects of growth and development of

© The Authors (2007). Journal compilation © New Phytologist (2007) www.newphytologist.org

New Phytologist (2007) 175: 387–404

393

394 Review

Tansley review

Fig. 5 Generalized secondary structures of single subunits of plant cyclic nucleotide-gated channels (CNGCs; left) and glutamate receptors (GLRs; right), showing C and N termini, transmembrane domains and pore regions (P) which contain the selectivity filter. The CNGC N-terminus contains a domain that can bind cyclic nucleotides (CNBD) to activate channels, and a partially overlapping calmodulin (CaM)-binding domain that modulates the CNBD affinity for cyclic nucleotides. GLRs contain two putative extracellular substrate binding domains (S). Functional CNGCs are believed to contain four subunits, whereas GLRs may consist of four or five subunits.

higher plants. Examples include phytochrome signalling, gibberellic acid-induced signalling, pollen tube tip growth, plant cell cycle progression and salt stress tolerance (Maathuis & Sanders, 2001; Rubio et al., 2003; Talke et al., 2003; Newton & Smith, 2004). In animal cells, many cyclic nucleotide-based signalling cascades are relayed via the cyclic nucleotide-dependent kinases PKA and PKG. However, no orthologues of PKA and PKG have been found in plants. The number of plant gene products with putative cyclic nucleotide binding sites is limited to c. 40 in Arabidopsis (Maathuis, 2006b) and mainly consists of membrane transporters. Within this category, the cyclic nucleotide-gated ion channels (CNGCs) constitute the largest group. In animals, CNGCs function in the transduction of sensory input and in Ca2+ signalling (Biel et al., 1998) and are gated via binding of cAMP or cGMP to a cyclic nucleotidebinding domain (CNBD) near the C-terminus (Fig. 5). Additional control over gating is provided by a calmodulin-binding domain (CaMBD) at the N-terminus which interacts with the cyclic nucleotide-binding domain, thereby lowering the affinity for cAMP or cGMP. Similarly, plant CNGCs contain a C-terminal CNBD but differ from their animal counterparts regarding the CaMBD, with the plant CaMBD located at the C-terminus and partly overlapping the CNBD (Fig. 5). Nevertheless, CaM binding appears to have a comparable function in plant CNGCs, lowering their activity by preventing cyclic nucleotide binding. CNGCs have a generalized predicted structure of six transmembrane domains, S1–S6, with a pore domain (P loop) between S5 and S6, although the number of transmembrane spans varies considerably depending on the hydrophobicity algorithm that is used. The region that endows ion selectivity, the P loop selectivity sequence, is

New Phytologist (2007) 175: 387–404

significantly different in plant CNGCs compared with animal CNGCs. It has proved frustrating to achieve routine functional expression of plant CNGCs and this has greatly hampered the evaluation of their physiological roles. In a number of cases, heterologous expression in oocytes was achieved (Leng et al., 2002; Balague et al., 2003), in particular with Arabidopsis CNGC isoforms. The data showed that activation is cGMPand /or cAMP-dependent, that most CNGCs do not discriminate amongst monovalent cations, have a limited Ca2+ permeability and are blocked by Cs+ and external Mg2+. A notable exception is AtCNGC2 which exhibited a high degree of K+ selectivity with regard to Na+ (Hua et al., 2003), a feature that is unknown in animal CNGCs. Interestingly, AtCNGC1, AtCNGC2 and NtCBP4, but not AtCNGC4, all show considerable voltage dependence, resulting in a strong inward rectification. This phenomenon is unknown in animal CNGCs and it remains to be revealed where this rectification originates. Information on the physiological roles of plant CNGCs mainly derives from forward and reverse genetics and appears to cover two broad categories: plant pathogen interactions and plant cation nutrition. Null mutants in AtCNGC2 (Clough et al., 2000) and AtCNGC4 (Balague et al., 2003) showed that both these isoforms may participate in signalling events during pathogen attack. More recently, data from a deletion mutant leading to a CNGC11-CNGC12 chimera also suggested a role in the pathogen response for these CNGCs (Yoshioka et al., 2006). It is well documented that Ca2+ as well as K+ and Cl– fluxes occur during the early phase of plant defence responses and, in addition, cyclic nucleotides have been implicated in plant defence-related signalling. For example, ROS production requires cAMP and Ca2+ in French bean cells (Bindschedler et al., 2001), whereas exposure of tobacco suspension cells to nitric oxide causes a transient increase in cGMP concentrations (Durner et al., 1998). Thus, both Ca2+ and cyclic nucleotide-based signalling form part of defence responses where CNGCs may be involved. Several studies have reported on the role of CNGCs in plant nutritional aspects. Loss of function in CNGC1 led to mutants that showed lower shoot Ca2+ contents and altered gravitropic root response, and the authors concluded that CNGC1 may participate in plant Ca2+ uptake from the external medium (Ma et al., 2006). AtCNGC3 is predominantly expressed in epidermal and cortical root tissues and a null mutation in this gene altered both short-term Na+ influx and K+ uptake in high external K+ conditions, indicating it may play a role in nonselective monovalent cation uptake (Gobert et al., 2006). Apart from having a role in pathogen response which presumably is limited to shoot tissue, AtCNGC2 in roots influences the homeostasis of cations such as Ca2+. Mutant plants became hypersensitive to elevated Ca2+ without increasing tissue Ca2+ contents (Chan et al., 2003). Another isoform, CNGC10, that is also relatively highly expressed in

www.newphytologist.org © The Authors (2007). Journal compilation © New Phytologist (2007)

Tansley review

root tissue was able to complement the K+ uptake deficient phenotype of the akt1-1 loss of function mutant, showing it can form a root K+ uptake pathway and considerably augment K+ uptake when overexpressed in the akt1-1 genotype (Li et al., 2005). In addition, the application of membrane permeable analogues of cAMP and cGMP can affect Na+ influx, Na+ efflux and K+ influx (Maathuis & Sanders, 2001; Essah et al., 2003; Rubio et al., 2003; Maathuis, 2006b), suggesting that here, too, CNGCs may be involved. CNGCs may also impact on the uptake of toxic monovalent cations such as Cs+ (Hampton et al., 2005). Thus, it appears that plant CNGCs may function in roots as important contributors to nonselective uptake of cations, whereas in shoots their function may be mainly in early signalling events that form part of the pathogen response. In the context of signalling cascades, CNGCs may form important elements in crosstalk between cyclic nucleotides and Ca2+ (Talke et al., 2003), since CNGC activation is sensitive to both types of second messenger and CNGCs also conduct Ca2+. For example, the observation that externally applied membrane-permeable analogues of cyclic nucleotides causes transient elevations in [Ca2+]cyt (Volotovski et al., 1998) may indicate that Ca2+ signals are directly downstream of cyclic nucleotides, and are generated through the activity of CNGCs.

V. Amino acid-gated NSCCs Ionic conductances activated by amino acids, specifically by glutamate and glycine, are critical for synaptic transmission and other complex physiological phenomena in animals (Dingledine et al., 1999). Amino acid-activated conductances are mediated by ionotropic glutamate receptors. After binding glutamate or glycine, ionotropic glutamate receptors (iGluRs) form cation channels with variable selectivity, conductance, kinetics and pharmacology (Dingledine et al., 1999). Genes with similarities to those encoding animal ionotropic glutamate receptors have been found in plants (Lam et al., 1998). Twenty glutamate receptor-like genes (termed GLRs; Fig. 5) are found in the genome of Arabidopsis thaliana and are divided into three subgroups based on sequence similarity (Lacombe et al., 2001; Chiu et al., 2002). The family of GLR genes in plants shows a great divergence from its animal counterpart, particularly in the pore region (Davenport, 2002). Since functional analysis is lacking, it is still unknown what type of channel is encoded by GLRs, but on the basis of overall homology to their animal counterparts, GLRs are believed to be a subclass of ligand-gated NSCC (Davenport, 2002). In animals, many iGluRs are NSCCs. Some of them form Ca2+-permeable NSCCs with properties similar to constitutive VI-NSCCs of higher plants. Dennison & Spalding (2000) reported that extracellular application of glutamate generated transient increases in [Ca2+]cyt in Arabidopsis seedlings.

Review

Lanthanides inhibited this effect, as did removal of bath Ca2+, suggesting the Ca2+ entered the cytosol through glutamateactivated cation channels. Using the same experimental system, Dubos et al. (2003) have found that glycine addition could also elevate [Ca2+]cyt. They also showed that glycine has synergistic effects when added simultaneously with glutamate, probably because it increases the affinity of glutamate binding to the receptor. Demidchik et al. (2004) obtained data on protoplasts isolated from different root tissues and demonstrated that similar amounts of glutamate induced larger increases in [Ca2+]cyt in mature epidermal and cortical cells compared with cells from deeper tissues. These authors have also carried out the first electrophysiological characterization of glutamate-activated currents in plants and found that the probability of observing glutamate-activated conductances is low and increases with increasing glutamate concentrations. The recorded glutamate-activated conductances were nonselective for monovalent cations, Ca2+-permeable, voltageindependent, and revealed instantaneous activation kinetics. Channel activity was sensitive to quinine and lanthanides, resembling the pharmacology of constitutive VI-NSCCs. The functionality of plant ionotropic glutamate receptors has been questioned, particularly with respect to the availability of ligand in the extracellular compartment. However, several studies have shown that apoplastic concentrations of glutamate and glycine can range from 0.01 to 1 mM (Lohaus et al., 1995, 2001; Lohaus & Heldt, 1997). Such concentrations are high enough to activate ionotropic glutamate receptors, and apoplastic glutamate and glycine could therefore function in plant Ca2+ uptake and/or signalling. Despite the lack of electrophysiological characterization, effects of exogenous amino acids on plant cell physiology and phenotypic properties of glr knockout mutants have been a subject of intensive investigations during the last 5 years. A number of studies has shown that exposure to extracellular glutamate modifies Ca2+-dependent physiological processes, such as depolymerization of microtubules, cell elongation and responses to aluminium (Sivaguru et al., 2003), root branching (Walch-Liu et al., 2006) and sugar sensing (Dubos et al., 2005). Loss of function mutations in Oryza sativa GLR3.1 (Li et al., 2006) and AtGLR3.2 (Kim et al., 2001; Turano et al., 2002) showed their involvement in Ca2+ accumulation and Ca2+-mediated reactions, for example stress responses, programmed cell death, cell division and differentiation. AtGLR 1.1 may be involved in seed germination and ABA-mediated processes (Kang & Turano, 2003; Kang et al., 2004). Overexpression of a radish GLR (Kang et al., 2006) in Arabidopsis enhanced glutamate-activated transient [Ca2+]cyt elevation and altered Ca2+-mediated mechanisms such as necrosis, growth and development. Additionally, it resulted in enhanced resistance to a fungal pathogen, possibly because of the up-regulation of jasmonic acid-responsive defensin genes. Strong evidence that GLR3.3 forms Ca2+-permeable channels has recently been obtained by Qi et al. (2006). Depolarization and elevation

© The Authors (2007). Journal compilation © New Phytologist (2007) www.newphytologist.org

New Phytologist (2007) 175: 387–404

395

396 Review

Tansley review

of [Ca2+]cyt induced by glutamate were prevented in glr3.3 knockout mutants. Interestingly, in addition to glutamate, five other amino acids (glycine, alanine, serine, asparagine, and cysteine) and glutathione (g-glutamyl-cysteinyl-Gly) were demonstrated to be agonists of the GLR3.3-induced responses. Overall, these data strongly suggest that glutamate causes physiological effects through an increase in [Ca2+]cyt catalysed by iGluRs. Some of these effects, for example altered ABA signalling, can also be mediated by ROS, because an increase in [Ca2+]cyt activates membrane-bound NADPH oxidases (Sagi & Fluhr, 2001).

VI. Purine signalling In animals, extracellular purines such ATP and ADP function as signalling molecules, activating specific ionotropic (P2X) and G-protein-coupled (P2Y) receptors. Animal P2X receptors form NSCCs after interaction with ATP, ADP and sometimes AMP (Ralevic & Burnstock, 1998). Permeability to Ca2+ is an important characteristic of several P2X receptors (North, 2002), and P2X receptors are involved in physiological activities ranging from neurotransmission to cell death and cell proliferation (reviewed by Ralevic & Burnstock, 1998). Physiological effects of extracellular purines in plants have received little attention. In 1960s and 1970s, the effect of purines on plants was interpreted in terms of energy supplementation or Ca2+ chelation, but not in the context of signalling. The extracellular presence of purines was found to affect plant movement (Jaffe, 1973), K+ transport (Luttge et al., 1974), and activity of cell-degrading enzymes (Udvardy & Farkas, 1973). Although such effects were often observed in the presence of micromolar concentrations of ATP, for example when studying the Venus fly trap closure, they were not considered to be the result of receptor activity (Jaffe, 1973). Recent interest into the notion that extracellular purines can function as signalling agents in plants was triggered by two reports, showing: (i) depolarization of the plasma membrane by extracellular ATP and ADP in Arabidopsis root hairs (Lew & Dearnaley, 2000); and (ii) transient elevations in [Ca2+]cyt induced by extracellular purines (Demidchik et al., 2003a). In these studies, low micromolar concentrations of ADP and nonhydrolysable ATP analogues were found to be effective, indicating that signalling rather than energetics underlies purine activity. The effect of purines on [Ca2+]cyt was blocked by lanthanides and by nonspecific (suramin) and specific (pyridoxalphosphate-6-azophenyl-2′,4′-disulfonic acid) antagonists of animal P2X receptors, and absent when external Ca2+ was removed. Although conclusive evidence is still lacking, the accumulative data strongly suggest that purinergic receptors exist in plants and that these are involved in signalling events during stress and wounding.

New Phytologist (2007) 175: 387–404

VII. Mechanosensitive ion channels Many environmental cues lead to alterations in physical forces on membranes. In plants, such cues are in particular associated with changes in turgor but also mechanical perturbation for example by wind. Such external signals can be relayed into electrical and/or Ca2+ signals through the action of mechanosensitive channels (MCs) whose gating depends on changes in tension forces on the membrane. MC open probability is often positively correlated to membrane stretch, hence the terms MC and stretch activated channel are often used to denote the same type of channel. Several MCs have been characterized in plant plasma membranes using patch-clamp analyses showing both anion and cation permeability in MCs (Cosgrove & Hedrich, 1991; Pickard & Ding, 1993; Qi et al., 2004). In guard cells, MCs are believed to mediate anion and cation movement across the guard cell plasma membrane and thus contribute to volume and turgor regulation of stomata (Cosgrove & Hedrich, 1991). MCs are blocked by trivalent cations such as Gd3+ and Al3+, and although most MCs are nonselective, occasionally an MC conductance with considerable selectivity is described, such as the small-conductance (3 pS) Ca2+-selective MC in Vicia faba guard cells (Cosgrove & Hedrich, 1991). The nonselective nature of MCs causes a large Ca2+ influx upon activation and several authors have suggested that the resulting Ca2+ signal forms a crucial intermediate in the transduction of mechanical stimuli. For example, Sato et al. (2001) studied mechanical stimulation of the chloroplast avoidance response in ferns: mechanical pressure on protonemal cells resulted in chloroplasts moving away from the site within an hour. This response only occurred when extracellular Ca2+ was present and it was sensitive to Gd3+, suggestive of a Ca2+-permeable MC-based process. Pollen tubes were shown to contain several types of MC (Dutta & Robinson, 2004), one of which may be involved in maintaining the polar Ca2+ gradient necessary for directional growth of the tube. On the basis of homology to bacterial MC genes, a family of 10 MC-like genes was discovered in Arabidopsis (MSL1– 10, Haswell & Meyerowitz, 2006). MSL3 can rescue the osmotic shock sensitivity of bacterial mutants, whereas null mutations in MSL2 and MSL3 led to abnormalities in shape and size of Arabidopsis plastids. Both MSL2 and MSL3 localized to the plastid envelope and the authors hypothesize that these proteins mediate plastidic ion release in response to increased osmotic pressure within the plastid. Whether an intracellular differential in osmotic pressure that is large enough to activate MCs ever occurs remains a question.

VIII. Vacuolar NSCCs Plant vacuoles can constitute 90% of the cellular volume and play essential roles in turgor provision, chlorophyll breakdown, programmed cell death and development, mineral storage, and

www.newphytologist.org © The Authors (2007). Journal compilation © New Phytologist (2007)

Tansley review

as a depository for xenobiotics and toxic compounds. Vacuoles also constitute a major intracellular Ca2+ store and are therefore important in cellular signalling (Allen & Sanders, 1997). The amenability of vacuoles to patch-clamp methodology ensured that tonoplast channels were amongst the first and best characterized plant ion channels. The accumulative data show the presence of two NSCCs (Allen & Sanders, 1997): the ubiquitous slow vacuolar (SV) channel has K+/Na+ and K+/Ca2+selectivity ratios of around 1 and 4, respectively, is activated by tonoplast depolarization, has slow kinetics and requires elevated cytoplasmic Ca2+ concentrations. The fast vacuolar (FV) channel has a similar K+/Na+ selectivity ratio to the SV channel and is inhibited by elevated [Ca2+]cyt. The SV channel is now well established as a voltage-dependent NSCC which is sensitive to both cytoplasmic and luminal Ca2+ concentrations and further regulated by a host of mechanisms including phosphorylation (Bethke & Jones, 1997), 14-3-3 proteins (van den Wijngaard et al., 2001), organic cations and redox potential (Scholtz-Starke et al., 2005). The identification of the Arabidopsis SV channel as AtTPC1 (Peiter et al., 2005) led the way to delineating the role of the SV channel through overexpression and loss of function genetics. ABA-induced delay of seed germination was significantly less in a tpc1-1 knockout mutant, whereas it was augmented in overexpressing lines. In guard cells, ABAdependent closure was not affected by the expression of AtTPC1, but high external Ca2+, another well-documented closing stimulus, largely failed to evoke stomatal closure in the knockout mutant. The stomatal phenotype in tpc1-1 mutants is reminiscent of det1-3 and cas1-1 mutants, both of which are altered in their Ca2+ signature. Thus, the SV channel, at least in guard cells, might mediate Ca2+ release from the vacuole which contributes to cytoplasmic Ca2+ signatures. There are currently two sources of debate regarding the SV channel. Firstly, in species such as rice and tobacco the SV channel activity is also found in vacuolar membranes but orthologues of AtTPC1 were proposed to be expressed at the plasma membrane and not at the vacuolar membrane. This implies that orthologous proteins may be targeted to different membranes depending on plant species. It also suggests that rice and tobacco SV channels are not coded by the AtTPC1 homologues NtTPC1 and OsTPC1. Manipulation of NtTPC1 and OsTPC1 expression showed that these proteins may be involved in elicitor-induced Ca2+ signalling and in pathogen response (Kurusu et al., 2004). Secondly, a function of SV channels in Ca2+ signalling has been disputed. Although the SV channel can conduct Ca2+ in certain experimental conditions, it is questionable whether this occurs in vivo. SV channel activity requires high cytoplasmic Ca2+ concentrations and is severely blocked by luminal Ca2+. Thus, several studies have shown that, with physiological transtonoplast Ca2+ gradients and voltages, no SV channel activity is observed (Pottosin et al., 1997). Far less is known about the other tonoplast NSCC, the FV channel. Originally observed in red beet storage tissue (Hedrich

Review

& Neher, 1987), few further studies have been published. FV channels become increasingly inactive whenever the cytoplasmic Ca2+ concentration exceeds c. 200 nM, and FV open probability has been reported to be largely insensitive to tonoplast potential. Subsequent publications reported on the presence of FV channels in other tissues, such as barley mesophyll vacuoles (Tikhonova et al., 1997), showing moderate outward rectification and a biphasic voltage dependence. One of the physiological roles for the FV channel may be in maintaining cellular K+ homeostasis, since both luminal and cytoplasmic K+ concentrations affect FV open probability (Pottosin & Martínez-Estévez, 2003). Other putative roles for the FV channel include providing a shunt conductance for the VATPase (Davies & Sanders, 1995), osmoregulation and regulation of the tonoplast potential (Allen & Sanders, 1997).

IX. Cation channels sensitive to elicitors Elicitors are compounds released during pathogen attack. They can be produced by pathogens themselves or derive from host cell wall fragments and are important cues to plants in their capacity of evoking defence mechanisms, such as the hypersensitive response. Elicitors themselves may have poreforming ability (Klüsener & Weiler, 1999; Lee et al., 2001) and thus cause electrical signals directly when they are present. However, in most cases, it is assumed that their action is via host receptor proteins that transform perception of these chemical signals into electrical and Ca2+ signals. One of the earliest reports on this mechanism described the activation of Ca2+-permeable channels in tomato protoplasts by fungal elicitors through intermediate steps that included a putative G-protein and channel phosphorylation (Gelli et al., 1997). A similar process has been described in parsley protoplasts treated with Phytopthora-derived cell wall elicitor (Zimmermann et al., 1997). However, in tomato, hyperpolarization-activated channels carried the Ca2+ influx, whereas in parsley channel activation occurred when cells were far more depolarized. Addition of yeast elicitor and chitosan induced whole-cell hyperpolarization-activated currents in Arabidopsis guard cells (Klüesener et al., 2002). This Ca2+ current required the production of ROS through the action of NADPH oxidases suggestive of a linear pathway, elicitor > ROS > Ca2+, and pointing to a general mechanism where stress leads to the production of ROS and subsequent Ca2+ signals. Nevertheless, elicitor-induced Ca2+ influx has been reported to occur both before (Blume et al., 2000) and after (Kawano & Muto, 2000) ROS production and therefore different classes of Ca2+ channel and Ca2+ store are probably involved during the perception of elicitors. In tobacco suspension cells, cosuppression of NtTPC1 led to a significant decrease in the cryptogene-induced Ca2+ signal (Kadota et al., 2004). Similarly, overexpression of OsTPC1 generated an increase in Ca2+ signal in response to fungal elicitor and also a larger production of ROS. Both NtTPC1 and

© The Authors (2007). Journal compilation © New Phytologist (2007) www.newphytologist.org

New Phytologist (2007) 175: 387–404

397

398 Review

Tansley review

OsTPC1 are believed to be targeted to the plasma membrane, and the data suggest TPC1 may be one of the initial pathways that links elicitor perception to cytoplasmic Ca2+ signals and subsequent production of ROS. ROS, in turn, have been shown to directly activate HACaCs (Pei et al., 2000), thus creating a further amplification of the initial signal. Although an extremely high Ca2+ selectivity was reported for some elicitor-activated cation channels (Gelli et al., 1997), a proper evaluation of channel selectivity is absent for any of them. Such studies are required to show whether elicitor-activated cation channels form a separate group of highly Ca2+-selective cation channels or belong to NSCCs. The latter seems more likely since, as mentioned earlier, plants do not appear to have genes that encode Ca2+-selective channels.

X. NSCCs acting in concert In spite of belonging to a diverse range of gene families and having various predicted structures, NSCCs catalyse the same process – passive transport of cations. Influx and efflux of cations require careful regulation, since cations are involved in many crucial aspects of cellular physiology. A number of gene families and isoforms that encodes NSCCs may contribute to cell- and tissue-specific fine-tuning of signalling events and ion fluxes. There are many scenarios where the activity of various classes of NSCCs may converge to mediate highly relevant physiological processes. One example would be the short- and long-term response of plants to salt stress (Fig. 6). The early response to the sudden onset of salt stress probably reflects turgor changes and involves both cGMP-based and Ca2+-based signalling (Donaldson et al., 2004). The observed rapid rise in cellular cGMP may derive from a receptor kinasecyclase relay that senses turgor change and would lead to the activation of root cell CNGCs. As already described, CNGCs may form important nodes in signalling networks where cyclic nucleotide signals are converted into Ca2+ signals (Talke et al., 2003). The cytoplasmic Ca2+ signal could have many downstream targets and may include membrane oxidases leading to the production of ROS. Similar to other stresses, ROS have been recorded during salinity and drought stress, and via activation of ROS-NSCCs could sustain further Ca2+ influx. Sustained Ca2+cyt elevation often depends on Ca2+ release from internal stores, such as the vacuole, and this function could be mediated by SV-type channels, although no salinitydependent phenotype was observed in either tpc1-1 loss of function or TPC1-1 overexpressing plants (Peiter et al., 2005). Elevated cGMP can directly inactivate root VI-NSCCs and thus reduce the influx of harmful Na+ (Maathuis & Sanders, 2001; Rubio et al., 2003). In addition, both elevated Ca2+ and cGMP are also known to affect gene transcription, including those that encode NSCCs (Maathuis, 2006b). The latter may also affect monovalent cation homeostasis during salt stress and reduce Na+ uptake through NSCCs such as CNGC3 (Gobert et al., 2006).

New Phytologist (2007) 175: 387–404

Fig. 6 Nonselective cation channels working in concert. During the onset of NaCl stress, a rapid rise in cellular 3′,5′-cyclic guanosine monophosphate (cGMP) ensues, possibly via a receptor kinase that activates a guanyl cyclase. cGMP can deactivate voltageindependent nonselective cation channels (VI-NSCCs) and activate cyclic nucleotide-gated channels (CNGCs). The latter would allow a rise in cytoplasmic Ca2+, which in turn can directly activate further classes of NSCC, such as those that release Ca2+ from internal stores. Increased Ca2+ also leads to ROS production through activation of NADPH oxidases. Reactive oxygen species-activated NSCCs could also contribute to a sustained Ca2+ signal.

A second example where different types of NSCC interact is in the uptake of nutrients such as Ca2+. At resting membrane potentials, constitutive VI-NSCCs catalyse steady-state Ca2+ uptake for nutritional needs. HACaC activity will dominate at hyperpolarized voltages and/or when [Ca2+]cyt is elevated, and provides an additional Ca2+ loading that is required for the stimulation of exocytosis and growth (Fig. 6). A small increase in the activity of constitutive VI-NSCCs in growing tissues is probably sufficient to induce an initial increase in basal [Ca2+]cyt, leading to activation of HACaCs and further Ca2+ influx (Demidchik et al., 2002a). Similar amplification mechanisms could exist for the generation of Ca2+ signals in response to stress, signalling agents and other stimuli, such as gravity and hormones. Glutamate, glycine, purines, ROS, elicitors and membrane stretch all activate specific Ca2+-permeable NSCCs and NSCC-like conductances that elevate [Ca2+]cyt, which in turn stimulates HACaCs and NADPH oxidase (Fig. 6). As described for salt stress, the production of extracellular ROS may also form part of this positive feedback system.

XI. Conclusions and perspectives Many classes of NSCC can be distinguished on the basis of electrophysiological, biochemical and genomics data (see Table 2 for an inventory of various types of plant nonselective

www.newphytologist.org © The Authors (2007). Journal compilation © New Phytologist (2007)

Tansley review

Review

Table 2 An inventory of various types of plant nonselective ion channels and their physiological functions Channel type Vacuolar NSCCs SV (TPC1)

FV

Physiological functions

References

Tonoplast potential K+ homeostasis Ca2+ signalling Pathogen response Tonoplast potential Monovalent cation homeostasis

Hedrich & Neher (1987); Allen & Sanders (1997); Bethke & Jones (1997); Pottosin et al. (1997); van den Wijngaard et al. (2001); Kurusu et al. (2004); Scholz-Starke et al. (2004); Peiter et al. (2005) Hedrich & Neher (1987); Tikhonova et al. (1997)

Constitutive plasma membrane NSCCs DepolarizationMonovalent cation homeostasis activated NSCCs K+ loss during salinity Acquisition of divalent cations (Ca2+, Zn2+, etc.)

Stoeckel & Takeda (1989); Cerana & Colombo (1992); Spalding et al. (1992); Wegner & Raschke (1994); de Boer & Wegner (1997); White (1997); Pei et al. (1998); Zhang et al. (2000, 2002, 2004); Piñeros & Kochian (2003); Becker et al. (2004); Volkov et al. (2004); Shabala et al. (2006); Volkov & Amtmann (2006); Wang et al. (2006)

Voltage-independent NSCCs

Na+ uptake Ca2+ influx K+ uptake during salinity Ca2+-dependent elongation growth

Yurin et al. (1991); White & Tester (1992); Elzenga & van Volkenburgh (1994); White & Lemtiri-Chlieh (1995); Amtmann et al. (1997); Roberts & Tester (1997); Tyerman et al. (1997); White (1997); Véry et al. (1998); Sokolik (1999); Buschmann et al. (2000); Davenport & Tester (2000); Maathuis & Sanders (2001); Yu et al. (2001); Demidchik et al. (2002a); Demidchik & Tester (2002); White & Davenport (2002); Becker et al. (2004); Volkov et al. (2004); Murthy & Tester (2006); Shabala et al. (2006); Volkov & Amtmann (2006)

Hyperpolarizationactivated NSCCs and HACaCs

Polar elongation growth Bacterial NH4+ release in legumes

Schroeder & Hagiwara (1990); White & Tester (1992); Tyerman et al. (1995); Gelli & Blumwald (1997); Gelli et al. (1997); Whitehead et al. (1998); Davenport & Tester (2000); Hamilton et al. (2000); Kiegle et al. (2000); Véry & Davies (2000); Whitehead et al. (2001); Demidchik et al. (2002a); Roberts & Tyerman (2002); White & Davenport (2002); Obermeyer & Tyerman (2005); Demidchik et al. (2007)

ROS-activated NSCCs

Transition metal sensing Polar elongation growth Stomatal closure Stress signalling (ABA, salinity, pathogens, etc.)

Demidchik et al. (1996, 1997a,b, 2001, 2003b, 2007); Pei et al. (2000); Foreman et al. (2003); Kwak et al. (2003). Inoue et al. (2005)

Cyclic nucleotide-gated channels

Uptake of monovalents Na+ influx Ca2+ uptake Ca2+ signalling Pathogen response Hypersensitive response

Clough et al. (2000); Leng et al. (2002); Balague et al. (2003); Chan et al. (2003); Hua et al. (2003); Talke et al. (2003); Li et al. (2005); Gobert et al. (2006); Ma et al. (2006); Yoshioka et al. (2006)

Amino acid-gated NSCCs

Regulation of membrane potential Ca2+ transport Cytoskeleton function Sugar and light sensing Hypersensitive response ABA signalling

Lam et al. (1998); Dennison & Spalding (2000); Kim et al. (2001); Turano et al. (2002); Dubos et al. (2003); Kang & Turano (2003); Sivaguru et al. (2003); Demidchik et al. (2004); Kang et al. (2004); Dubos et al. (2005); Kang et al. (2006); Li et al. (2006); Qi et al. (2006); Walch-Liu et al. (2006)

Mechanosensitive ion channels

Osmotic adjustment Ca2+ signalling Stomatal function

Cosgrove & Hedrich (1991); Pickard & Ping-Ding (1993); Sato et al. (2001); Dutta & Robinson (2004); Qi et al. (2004); Haswell & Meyerowitz (2006)

Cation channels sensitive to elicitors

Ca2+ signalling during pathogen response

Gelli & Blumwald (1997); Gelli et al. (1997); Zimmermann et al. (1997); Klüsener & Weiler (1999); Kadota et al. (2004)

NSCC, nonselective cation channels; SV, slow vacuolar; FV, fast vacuolar; HACaCs, hyperpolarization-activated Ca2+ channels; ABA, abscisic acid. For an explanation of the category definitions, see the text. The physiological functions ascribed to the different channel types are based on published studies but specific evidence is lacking in many cases.

© The Authors (2007). Journal compilation © New Phytologist (2007) www.newphytologist.org

New Phytologist (2007) 175: 387–404

399

400 Review

Tansley review

ion channels). An increasingly recurrent theme is the important role this type of channel plays in signalling events. Many NSCCs show Ca2+ permeability and have been shown to mediate changes in [Ca2+]cyt that can be linked to important physiological processes, such as elicitor perception or ROS perception. Plant NSSCs therefore appear to substitute the Ca2+-selective channels in animal systems that carry out similar functions (Bothwell & Ng, 2005). The other main function of plant NSCCs is mainly confined to root tissues where they mediate uptake of important ions such as NH4+, Na+ and K+. The lack of selectivity implies less control over transmembrane fluxes and it remains unclear why these functions in plants are not mediated by selective ion channels, as is the case in animals. One main difference is the potential of large fluctuations in ionic conditions that plants may be exposed to, particularly root cells. Such fluctuations may require rapid adjustment of membrane voltage and osmotic potentials, functions that do not necessarily involve specific ions. The occurrence of salinity stress exemplifies this: a rapid reduction of the external water potential obliges cells to take up inorganic ions rapidly to increase cellular osmolarity. During K+ starvation, similar mechanisms may prevent excessive turgor loss. These processes are potentially beneficial but their nonselective nature would necessitate tight control over NSCC activity. In principle, plant Ca2+ signalling does not appear to be different from that observed in animals. Yet in animals, Ca2+ signalling proceeds through Ca2+-selective channels, whereas in plants NSCCs seem to provide the rise in [Ca2+]cyt. Opening of NSCCs inevitably leads to transmembrane fluxes of other cations, such as K+ and Na+, which may be more problematic in animal cells where extracellular concentrations of 100– 110 mM Na+ prevail. Most terrestrial plants are not exposed to large amounts of Na+ and hence such a disadvantage would not occur. However, it can be surmised that during salinity stress, opening of NSCCs to mediate Ca2+ signalling may well lead to toxic Na+ effects and it would be interesting to see whether halophytes have adaptations to avoid this. In many cases, different types of NSCC may be manifestations of the same protein. For example, ROS-, cyclic nucleotideand glutamate-activated NSCCs all show little or no voltage dependence and may very well constitute subclasses of VINSCCs. Additionally, ROS-NSCCs and constitutive HANSCCs demonstrate very similar unitary conductances (Véry & Davies, 2000; Demidchik et al., 2007). Both DA-NSCCs and HA-NSCCs are involved in monovalent cation uptake but in certain cell types could form elicitor-activated NSCClike conductances. A proper delineation of NSCC functions requires molecular and electrophysiological tools, but most importantly a direct link between specific NSCC conductances and NSCC-encoding genes. Only in one case has conclusive evidence been reported for such a link, for the vacuolar NSCC, TPC1. In other cases, heterologous expression of cyclic

New Phytologist (2007) 175: 387–404

nucleotide-gated channels led to nonselective currents (Leng et al., 2002; Balague et al., 2003), but the associated in vivo currents have not been reported. Several gene families have been identified, such as the CNGCs and GLRs that encode putative NSCCs, but in situ electrophysiological data are missing. Although it is imperative that the genomics and electrophysiological data for NSCCs can be combined, this is fraught with difficulties. Successful cloning and characterization of K+-selective channels (Sentenac et al., 1992; Hirsch et al., 1998) was based on yeast complementation assays and plant loss of function studies, both strategies that are not easily applied to NSCCs. The extensive gene families that encode putative NSCCs hamper loss of function studies, particularly because a large degree of functional redundancy may be present. Further transcriptomics and proteomics studies will aid in this respect, for example by revealing tissue- and membrane-specific expression of NSCC isoforms. The latter will greatly help in assigning potential functions but more importantly in targeting electrophysiological studies to the right cell type and cellular compartment. In vivo currents are often mediated by ion channels consisting of heteromers. Thus, to allow comparison between in vivo and heterologous systems, knowledge regarding channel subunit composition is urgently needed, for example through application of FRET-based studies. In combination, these approaches will provide vital steps in making further progress regarding the physiological functions of NSCCs.

References Aleksandrov AA, Berestovsky GN, Volkova SP, Vostrikov IY, Zherelova OM, Kravchik S, Lunevsky VZ. 1976. Reconstitution of single calcium-sodium channels of the cells in lipid bilayer. Doklady Akademii Nauk SSSR 227: 723–726. Allen GJ, Sanders D. 1997. Vacuolar ion channels of higher plants. Advances in Botanical Research 25: 217–252. Amtmann A, Laurie S, Leigh RA, Sanders D. 1997. Multiple inward channels provide flexibility on Na+/K+ discrimination at the plasma membrane of barley suspension culture cells. Journal of Experimental Botany 48: 481–497. Apel K, Hirt H. 2004. Reactive oxygen species: Metabolism, oxidative stress, and signal transduction. Annual Reviews of Plant Biology 55: 373–399. Balague C, Lin BQ, Alcon C, Flottes G, Malmstrom S, Kohler C, Neuhaus G, Pelletier G, Gaymard F, Roby D. 2003. HLM1, an essential signaling component in the hypersensitive response, is a member of the cyclic nucleotide-gated channel ion channel family. The Plant Cell 15: 365–379. Becker D, Geiger D, Dunkel M, Roller A, Bertl A, Latz A, Carpaneto A, Dietrich P, Roelfsema MRG, Voelker C, Schmidt D, Mueller-Roeber B, Czempinski K, Hedrich R. 2004. AtTPK4, an Arabidopsis tandem-pore K+ channel, poised to control the pollen membrane voltage in a pH- and Ca2+-dependent manner. Proceedings of the National Academy of Sciences, USA 101: 15621–15626. Bethke PC, Jones RL. 1997. Reversible protein phosphorylation regulates the activity of the slow-vacuolar ion channel. Plant Journal 11: 1227–1235. Biel M, Sautter A, Ludwig A, Hofmann F, Zong XG. 1998. Cyclic nucleotide-gated channels – mediators of NO: cGMP- regulated processes. Naunyn-Schmiedebergs Archives Pharmacology 358: 140–144.

www.newphytologist.org © The Authors (2007). Journal compilation © New Phytologist (2007)

Tansley review Bienert GP, Schjoerring JK, Jahn TP. 2006. Membrane transport of hydrogen peroxide. Biochimica Biophysica Acta 1758: 994 –1003. Bindschedler LV, Minibayeva F, Gardner SL, Gerrish C, Davies DR, Bolwell GP. 2001. Early signalling events in the apoplastic oxidative burst in suspension cultured French bean cells involve cAMP and Ca2+. New Phytologist 151: 185 –194. Blume B, Nurnberger T, Nass N, Scheel D. 2000. Receptor-mediated increase in cytoplasmic free calcium required for activation of pathogen defense in parsley. Plant Cell 12: 1425–1440. de Boer AH, Wegner LH. 1997. Regulatory mechanisms of ion channels in xylem parenchyma cells. Journal of Experimental Botany 48: 441–449. Bothwell JHF, Ng CKY. 2005. The evolution of Ca2+ signalling in photosynthetic eukaryotes. New Phytologist 166: 21–38. Buschmann PH, Vaidynathan R, Gassmann W, Schroeder JI. 2000. Enhancement of Na+ uptake currents, time dependent inward-rectifying K+ channel currents, and K+ channel transcripts by K+ starvation in wheat root cells. Plant Physiology 122: 1387–1397. Carol RJ, Dolan L. 2006. The role of reactive oxygen species in cell growth: lessons from root hairs. Journal of Experimental Botany 57: 1829–1834. Cerana R, Colombo R. 1992. K+ and Cl– conductance of Arabidopsis thaliana plasma membrane at depolarised voltages. Botanica Acta 105: 273–277. Chan CWM, Schorrak LM, Smith RK, Bent AF, Sussman MR. 2003. A cyclic nucleotide-gated ion channel, CNGC2, is crucial for plant development and adaptation to calcium stress. Plant Physiology 132: 728–731. Chiu JC, Brenner ED, DeSalle R, Nitabach MN, Holmes TC, Coruzzi GM. 2002. Phylogenetic and expression analysis of the glutamate-receptor-like gene family in Arabidopsis thaliana. Molecular Biology and Evolution 19: 1066 –1082. Clark GB, Roux SJ. 1995. Annexins of plant cells. Plant Physiology 109: 1133–1139. Clough SJ, Fengler KA, Yu I, Lippok B, Smith RK, Bent A. 2000. The Arabidopsis dnd1 ‘defense, no death’ gene encodes a mutated cyclic nucleotide-gated ion channel. Proceedings of the National Academy of Sciences, USA 97: 9323 –9328. Cosgrove DJ, Hedrich R. 1991. Stretch-activated chloride, potassium, and calcium channels coexisting in plasma membranes of guard cells of Vicia faba L. Planta 186: 143 –153. Davenport R. 2002. Glutamate receptors in plants. Annals of Botany 90: 549–557. Davenport RJ, Tester M. 2000. A weakly voltage-dependent, nonselective cation channel mediates toxic sodium influx in wheat. Plant Physiology 122: 823–834. Davies JM, Sanders D. 1995. ATP, pH and Mg2+ modulate a cation current in Beta vulgaris vacuoles: a possible shunt conductance for the vacuolar H+-ATPase. Journal of Membrane Biology 145: 75 – 86. Demidchik V, Adobea P, Tester MA. 2004. Glutamate activates sodium and calcium currents in the plasma membrane of Arabidopsis root cells. Planta 219: 167–175. Demidchik V, Bowen HC, Maathuis FJM, Shabala SN, Tester MA, White PJ, Davies JM. 2002a. Arabidopsis thaliana root nonselective cation channels mediate calcium uptake and are involved in growth. Plant Journal 32: 799–808. Demidchik V, Davenport RJ, Tester MA. 2002b. Nonselective cation channels in plants. Annual Reviews of Plant Biology 53: 67–107. Demidchik V, Nichols C, Oliynyk M, Glover B, Davies JM. 2003a. Is extracellular ATP a signalling agent in plants? Plant Physiology 133: 456–461. Demidchik V, Shabala SN, Coutts KB, Tester MA, Davies JM. 2003b. Free oxygen radicals regulate plasma membrane Ca2+- and K+-permeable channels in plant root cells. Journal of Cell Sciences 116: 81–88. Demidchik V, Shabala S, Davies J. 2007. Spatial variation in H2O2 response of Arabidopsis thaliana root epidermal Ca2+ flux and plasma membrane Ca2+ channels. Plant Journal 49: 377–386.

Review

Demidchik VV, Sokolik AI, Yurin VM. 1996. The copper ion influence on functioning of plant cell plasmalemma H+-ATPase. Doklady Akademii Nauk Belarusi 40: 84–87. Demidchik VV, Sokolik AI, Yurin VM. 1997a. Mechanisms of conductance modification in plant cell membranes under the action of trivalent iron ions. Doklady Akademii Nauk Belarusi 41: 83–87. Demidchik V, Sokolik A, Yurin V. 1997b. The effect of Cu2+ on ion transport systems of the plant cell plasmalemma. Plant Physiology 114: 1313–1325. Demidchik V, Sokolik A, Yurin V. 2001. Characteristics of non-specific permeability and H+-ATPase inhibition induced in the plasma membrane of Nitella flexilis by excessive Cu2+. Planta 212: 583–590. Demidchik V, Tester MA. 2002. Sodium fluxes through nonselective cation channels in the plant plasma membrane of protoplasts from Arabidopsis roots. Plant Physiology 128: 379–387. Dennison KL, Spalding EP. 2000. Glutamate-gated calcium fluxes in Arabidopsis. Plant Physiology 124: 1511–1514. Dingledine R, Borges K, Bowie D, Traynelis SF. 1999. The glutamate receptor ion channels. Pharmacological Reviews 51: 7–61. Donaldson L, Ludidi N, Knight MR, Gehring C, Denby K. 2004. Salt and osmotic stress cause rapid increases in Arabidopsis thaliana cGMP levels. FEBS Letters 569: 317–320. Dubos C, Huggins D, Grant GH, Knight MR, Campbell MM. 2003. A role for glycine in the gating of plant NMDA-like receptors. Plant Journal 35: 800–810. Dubos C, Willment J, Huggins D, Grant GH, Campbell MM. 2005. Kanamycin reveals the role played by glutamate receptors in shaping plant resource allocation. Plant Journal 43: 348–355. Durner J, Wendehenne D, Klessig DF. 1998. Defense gene induction in tobacco by nitric oxide, cyclic GMP, and cyclic ADP-ribose. Proceedings of the National Academy of Sciences, USA 95: 10328–10333. Dutta R, Robinson KR. 2004. Identification and characterization of stretch-activated ionchannels in pollen protoplasts. Plant Physiology 135: 1398–1406. Eisenbarth DA, Weig AR. 2005. Dynamics of aquaporins and water relations during hypocotyl elongation in Ricinus communis L seedlings. Journal of Experimental Botany 56: 1831–1842. Elzenga JTM, van Volkenburgh E. 1994. Characterization of ion channels in the plasma membrane of epidermal cells of expanding pea (Pisum sativum arg) leaves. Journal of Membrane Biology 137: 227–235. Essah PA, Davenport R, Tester M. 2003. Sodium influx and accumulation in Arabidopsis. Plant Physiology 133: 307–318. Foreman J, Demidchik V, Bothwell JHF, Mylona P, Miedema H, Torres MA, Linstead P, Costa S, Brownlee C, Jones JDG, Davies JM, Dolan L. 2003. Reactive oxygen species produced by NADPH oxidase regulate plant cell growth. Nature 422: 442–446. Gelli A, Blumwald E. 1997. Hyperpolarisation-activated Ca2+-permeable channels in the plasma membrane of tomato cells. Journal of Membrane Biology 155: 35–45. Gelli A, Higgins VJ, Blumwald E. 1997. Activation of plant plasma membrane Ca2+-permeable channels by race-specific fungal elicitors. Plant Physiology 113: 269–279. Gerke V, Moss S. 2002. Annexins: from structure to function. Physiological Reviews 82: 331–371. Gobert A, Park G, Amtmann A, Sanders D, Maathuis FJM. 2006. Arabidopsis thaliana cyclic nucleotide gated channel 3 forms a nonselective ion transporter involved in germination and cation transport. Journal of Experimental Botany 57: 791–800. Hamilton DWA, Hills A, Kohler B, Blatt MR. 2000. Ca2+ channels at the plasma membrane of stomatal guard cells are activated by hyperpolarization and abscisic acid. Proceedings of the National Academy of Sciences, USA 97: 4967–4972. Hampton CR, Broadley MR, White PJ. 2005. Short review: The mechanisms of radiocaesium uptake by arabidopsis roots. Nukleonika 50: S3–S8.

© The Authors (2007). Journal compilation © New Phytologist (2007) www.newphytologist.org

New Phytologist (2007) 175: 387–404

401

402 Review

Tansley review

Haswell ES, Meyerowitz EM. 2006. MscS-like proteins control plastid size and shape in Arabidopsis thaliana. Current Biology 16: 1–11. Hedrich R, Neher E. 1987. Cytoplasmic calcium regulates voltage-dependent ion channels in plant vacuoles. Nature 329: 833–835. Hedrich R, Schroeder JI. 1989. The physiology of ion channels and electrogenic pumps in higher plants. Annual Reviews of Plant Physiology and Plant Molecular Biology 40: 539–569. Hirsch RE, Lewis BD, Spalding EP, Sussman MR. 1998. A role for the AKT1 potassium channel in plant nutrition. Science 280: 918 –921. Hua BG, Mercier RW, Leng Q, Berkowitz GA. 2003. Plants do it differently. A new basis for potassium/sodium selectivity in the pore of an ion channel. Plant Physiology 132: 1353 –1361. Inoue H, Kudo T, Kamada H, Kimura M, Yamaguchi I, Hamamoto H. 2005. Copper elicits an increase in cytosolic free calcium in cultured tobacco cells. Plant Physiology and Biochemistry 43: 1089–1094. Jaffe MJ. 1973. The role of ATP in mechanically stimulated rapid closure of the Venus’s-Flytrap. Plant Physiology 51: 17–18. Johannes E, Sanders D. 1995. The voltage-gated Ca2+ release channel in the vacuolar membrane of sugar beet resides in two activity states. FEBS Letters 365: 1–6. Kadota Y, Furuichi T, Ogasawara Y, Goh T, Higashi K, Muto S, Kuchitsu K. 2004. Identification of putative voltage-dependent Ca2+-permeable channels involved in cryptogein-induced Ca2+ transients and defense responses in tobacco BY-2 cells. Biochemical and Biophysical Research Communications 317: 823 –830. Kang S, Kim HB, Lee H, Choi JY, Heu S, Oh CJ, Kwon SI, An CS. 2006. Overexpression in Arabidopsis of a plasma membrane-targeting glutamate receptor from small radish increases glutamate-mediated Ca2+ influx and delays fungal infection. Molelules and Cells 21: 418– 427. Kang J, Mehta S, Turano FJ. 2004. The putative glutamate receptor 1.1 (AtGLR1.1) in Arabidopsis thaliana regulates abscisic acid biosynthesis and signaling to control development and water loss. Plant Cell Physiology 45: 1380–1389. Kang J, Turano FJ. 2003. The putative glutamate receptor 1.1 (AtGLR1.1) functions as a regulator of carbon and nitrogen metabolism in Arabidopsis thaliana. Proceedings of the National Academy of Sciences, USA 100: 6872–6877. Kawano T. 2003. Roles of the reactive oxygen species-generating peroxidase reactions in plant defense and growth induction. Plant Cell Reports 21: 829–837. Kawano T, Muto S. 2000. Mechanism of peroxidase actions for salicylic acid-induced generation of active oxygen species and an increase in cytosolic calcium in tobacco cell suspension culture. Journal of Experimental Botany 51: 685 – 693. Kiegle E, Gilliham M, Haseloff J, Tester M. 2000. Hyperpolarisationactivated calcium currents found only in cells from the elongation zone of Arabidopsis thaliana roots. Plant Journal 21: 225–229. Kim SA, Kwak JM, Jae SK, Wang MH, Nam HG. 2001. Overexpression of the AtGluR2 gene encoding an Arabidopsis homolog of mammalian glutamate receptors impairs calcium utilization and sensitivity to ionic stress in transgenic plants. Plant Cell Physiology 42: 74–84. Klüsener B, Weiler EW. 1999. Pore-forming properties of elicitors of plant defense reactions and cellulolytic enzymes. FEBS Letters 459: 263–266. Klüsener B, Young JJ, Murata Y, Allen GJ, Mori IC, Hugouvieux V, Schroeder JI. 2002. Convergence of calcium signaling pathways of pathogenic elicitors and abscisic acid in Arabidopsis guard cells. Plant Physiology 130: 2152–2163. Kurusu T, Sakurai Y, Miyao A, Hirochika H, Kuchitsu K. 2004. Identification of a putative voltage-gated Ca2+-permeable channel (OsTPC1) involved in Ca2+ influx and regulation of growth and development in rice. Plant Cell Physiology 45: 693 –702. Kwak JM, Mori IC, Pei ZM, Leonhardt N, Torres MA, Dangl JL, Bloom RE, Bodde S, Jones JDG, Schroeder JI. 2003. NADPH oxidase AtrbohD and AtrbohF genes function in ROS-dependent ABA signaling in Arabidopsis. EMBO Journal 22: 2623–2633.

New Phytologist (2007) 175: 387–404

Lacombe B, Becker D, Hedrich R, Chiu J, DeSalle R, Heinemann S, Hollmann M, Kwak J, Le Novere N, Nam HG, Sakmann B, Schroeder JI, Spalding EP, Tester M, Turano FJ, Coruzzi G. 2001. On the identity of plant glutamate receptors. Science 292: 1486–1487. Lahiri S, Roy A, Baby SM, Hoshi T, Semenza GL, Prabhakar NR. 2006. Oxygen sensing in the body. Progress in Biophysics and Molecular Biology 91: 249–286. Lam H-M, Chiu J, Hsieh M-H, Meisel L, Oliviera IC, Shin M, Coruzzi G. 1998. Glutamate receptor genes in plants. Nature 396: 125–126. Lee J, Klusener B, Tsiamis G, Stevens C, Neyt C, Tampakaki AP, Panopoulos NJ, Noller J, Weiler EW, Cornelis GR, Mansfield JW, Nurnberger T. 2001. HrpZ(Psph) from the plant pathogen Pseudomonas syringae pv. phaseolicola binds to lipid bilayers and forms an ion-conducting pore in vitro. Proceedings of the National Academy of Sciences USA 98: 289–294. Leng Q, Mercier RW, Hua BG, Fromm H, Berkowitz GA. 2002. Electrophysiological analysis of cloned cyclic nucleotide-gated ion channels. Plant Physiology 128: 400–410. Lew RR, Dearnaley JDW. 2000. Extracellular nucleotide effects on electrical properties of growing Arabidopsis thaliana root hairs. Plant Science 153: 1–6. Li XL, Borsics T, Harrington HM, Christopher DA. 2005. Arabidopsis AtCNGC10 rescues potassium channel mutants of E-coli, yeast and Arabidopsis and is regulated by calcium/calmodulin and cyclic GMP in E-coli. Functional Plant Biology 32: 643–653. Li J, Zhu SH, Song XW, Shen Y, Chen HMYuJ, Yi KK, Liu YF, Karplus VJ, Wu P, Deng XW. 2006. A rice glutamate receptor-like gene is critical for the division and survival of individual cells in the root apical meristem. The Plant Cell 18: 340–349. Lohaus G, Heldt H-W. 1997. Assimilation of gaseous ammonia and the transport of its products in barley and spinach leaves. Journal of Experimental Botany 48: 1779–1786. Lohaus G, Pennewiss K, Sattelmacher B, Hussmann M, Muehling KH. 2001. Is the infiltration-centrifugation technique appropriate for the isolation of apoplastic fluid? A critical evaluation with different plant species. Physiologia Plantarum 111: 457–465. Lohaus G, Winter H, Riens B, Heldt H-W. 1995. Further studies of the phloem loading process in leaves of barley and spinach – the comparison of metabolite concentrations in the apoplastic compartment with those in the cytosolic compartment and in the sieve tubes. Botanica Acta 108: 270–275. Lunevsky VZ, Zherelova OM, Aleksandrov AA, Vinokurov MG, Berestovsky GN. 1980. Model of selective filter of a calcium channel in Characeae algal cell. Biofizika 25: 685–691. Lunevsky VZ, Zherelova OM, Vostrikov IY, Berestobsky GN. 1983. Excitation of characeae cell membranes as a result of activation of calcium and chloride channels. Journal of Membrane Biology 72: 43–58. Luttge U, Schoch EV, Ball E. 1974. Can externally applied ATP supply energy to active ion uptake mechanisms of intact plant cells? Australian Journal of Plant Physiology 1: 211–220. Maathuis FJM. 2006a. The role of monovalent cation transporters in plant responses to salinity. Journal of Experimental Botany 57: 1137–1147. Maathuis FJM. 2006b. cGMP modulates gene transcription and cation transport in Arabidopsis roots. Plant Journal 45: 700–711. Maathuis FJM, Amtmann A. 1999. K+ nutrition and Na+ toxicity: the basis of cellular K+/Na+ ratios. Annals of Botany 84: 123–133. Maathuis FJM, Sanders D. 1993. Energization of potassium uptake in Arabidopsis thaliana. Planta 191: 302–307. Maathuis FJM, Sanders D. 1997. Regulation of K+ uptake by external K+ levels. Journal of Experimental Botany 48: 451–458. Maathuis FJM, Sanders D. 2001. Sodium uptake in Arabidopsis thaliana roots is regulated by cyclic nucleotides. Plant Physiology 127: 1617–1625. Malho R, Liu Q, Monteiro D, Rato C, Camacho L, Dinis A. 2006. Signalling pathways in pollen germination and tube growth. Protoplasma 228: 21–30.

www.newphytologist.org © The Authors (2007). Journal compilation © New Phytologist (2007)

Tansley review Marschner H. 1995. Mineral nutrition of higher plants, 2nd edn. London, UK: Academic Press. Marten I, Hoth S, Deeken R, Ache P, Ketchum KA, Hoshi T, Hedrich R. 1999. AKT3, a phloem-localized K+ channel, is blocked by protons. Proceedings of the National Academy of Sciences, USA 96: 7581–7586. McInnis SM, Desikan R, Hancock JT, Hiscock SJ. 2006. Production of reactive oxygen species and reactive nitrogen species by angiosperm stigmas and pollen: potential signalling crosstalk? New Phytologist 172: 221–228. Miedema H, Bothwell JHF, Brownlee C, Davies JM. 2001. Calcium uptake by plant cells – channels and pumps acting in concert. Trends in Plant Science 6: 514–519. Murthy M, Tester M. 2006. Cation currents in protoplasts from the roots of a Na+ hyperaccumulating mutant of Capsicum annuum. Journal of Experimental Botany 57: 1171–1180. Newton RP, Smith CJ. 2004. Cyclic nucleotides. Phytochemistry 65: 2423–2437. North RA. 2002. Molecular physiology of P2X receptors. Physiological Reviews 82: 1013–1067. Obermeyer G, Tyerman SD. 2005. NH4+ currents across the peribacteroid membrane of soybean. Macroscopic and microscopic properties, inhibition by Mg2+, and temperature dependence indicate a subpicoSiemens channel finely regulated by divalent cations. Plant Physiology 139: 1015 –1029. Osterhout WJV. 1958. The use of aquatic plants in the study of some fundamental problems. Ann Rev Plant Physiology 8: 1–11. Pei ZM, Kuchitsu K. 2005. Early ABA signaling events in guard cells. Journal of Plant Growth Regulation 24: 296 –307. Pei ZM, Murata Y, Benning G, Thomine S, Klusener B, Allen GJ, Grill E, Schroeder JI. 2000. Calcium channels activated by hydrogen peroxide mediate abscisic acid signalling in guard cells. Nature 406: 731–734. Pei ZM, Schroeder JI, Schwarz M. 1998. Background ion channel activities in Arabidopsis guard cells and review of ion channel regulation by protein phosphorylation events. Journal of Experimental Botany 49: 319–328. Peiter E, Maathuis FJM, Mills LN, Knight H, Pelloux M, Hetherington AM, Sanders D. 2005. The vacuolar Ca2+-activated channel TPC1 regulates germination and stomatal movement. Nature 434: 404–408. Pickard BG, Ping-Ding J. 1993. The mechanosensory calcium-selective ion channels: Key component of a plasmalemmal control centre? Australian Journal of Plant Physiology 20: 439– 459. Piñeros MA, Kochian LV. 2003. Differences in whole-cell and singlechannel ion currents across the plasma membrane of mesophyll cells from two closely related Thlaspi species. Plant Physiology 131: 583 –594. Pitzschke A, Forzani C, Hirt H. 2006. Reactive oxygen species signaling in plants. Antioxidant and Redox Signalling 8: 1757–1764. Pottosin II, Martínez-Estévez M. 2003. Regulation of the fast vacuolar channel by cytosolic and vacuolar potassium. Biophysical Journal 84: 977–986. Pottosin II, Tikhonova LI, Hedrich R, Schoenknecht G. 1997. Slowly activating vacuolar channels cannot mediate Ca2+-induced Ca2+ release. Plant Journal 12: 1387–1398. Qi Z, Kishigami A, Nakagawa Y, Iida H, Sokabe M. 2004. A mechanosensitive anion channel in Arabidopsis thaliana mesophyll cells. Plant and Cell Physiology 45: 1704–1708. Qi Z, Stephens NR, Spalding EP. 2006. Calcium entry mediated by GLR3.3, an Arabidopsis glutamate receptor with a broad agonist profile. Plant Physiology 142: 963 –971. Ralevic V, Burnstock G. 1998. Receptors for purines and pyrimidines. Pharmacological Reviews 50: 413 – 492. Roberts SK, Tester M. 1997. A patch clamp study of Na+ transport in maize roots. Journal of Experimental Botany 48: 431– 440. Roberts DM, Tyerman SD. 2002. Voltage-dependent cation channels permeable to NH4+, K+ and Ca2+ in the symbiosome membrane of the model legume Lotus japonicus. Plant Physiology 128: 370 –378.

Review

Rubio F, Flores P, Navarro JM, Martinez V. 2003. Effects of Ca2+, K+ and cGMP on Na+ uptake in pepper plants. Plant Science 165: 1043 –1049. Sagi M, Fluhr R. 2001. Superoxide production by plant homologues of the gp91phox NADPH oxidase. Modulation of activity by calcium and by tobacco mosaic virus infection. Plant Physiology 126: 1281–1290. Sato Y, Wada M, Kadota A. 2001. External Ca2+ is essential for chloroplast movement induced by mechanical stimulation but not by light stimulation. Plant Physiology 127: 497–504. Scholz-Starke J, Gambale F, Carpaneto A. 2005. Modulation of plant ion channels by oxidizing and reducing agents. Archives of Biochemistry and Biophysics 434: 43–50. Schroeder JI, Hagiwara S. 1990. Repetitive increases in cytosolic Ca2+ of guard cells by abscisic acid activation of nonselective Ca2+ permeable channels. Proceedings of the National Academy of Sciences, USA 87: 9305–9309. Sentenac H, Bonneaud N, Minet M, Lacroute F, Salmon JM, Gaymard F, Grignon C. 1992. Cloning and expression in yeast of a plant potassium ion transport system. Science 256: 663–665. Shabala S, Demidchik V, Shabala L, Cuin TA, Smith SJ, Miller AJ, Davies JM, Newman IA. 2006. Extracellular Ca2+ ameliorates NaCl-induced K+ loss from Arabidopsis root and leaf cells by controlling plasma membrane K+-permeable channels. Plant Physiology 141: 1653–1665. Sivaguru M, Pike S, Gassmann W, Baskin TI. 2003. Aluminum rapidly depolymerizes cortical microtubules and depolarizes the plasma membrane: evidence that these responses are mediated by a glutamate receptor. Plant and Cell Physiology 44: 667–675. Sokolik AI. 1999. Nonselective ion conductivity of plasmalemma as an important component of the system of membrane transport of ions in plants. Doklady Akademii Nauk Belarusi 43: 77–80. Sokolik AI, Yurin VM. 1986. Potassium channels in plasmalemma of Nitella cells at rest. Journal of Membrane Biology 89: 9–22. Spalding EP, Slayman CL, Goldsmith MHM, Gradmann D, Bertl A. 1992. Ion channels in Arabidopsis plasma membrane – transport characteristics and involvement in light-induced voltage changes. Plant Physiology 99: 96–102. Stoeckel H, Takeda K. 1989. Calcium-activated, voltage-dependent, nonselective cation currents in endosperm plasma membrane from higher plants. Proceedings of Royal Society, London Series B 237: 213–231. Talke IN, Blaudez D, Maathuis FJM, Sanders D. 2003. CNGCs: prime targets of plant cyclic nucleotide signalling? Trends of Plant Science 8: 286–293. Tester M. 1990. Plant ion channels: whole-cell and single channel studies. New Phytologist. 114: 305–340. Tester M, Davenport R. 2003. Na+ tolerance and Na+ transport in higher plants. Annals of Botany 91: 503–527. Thion L, Mazars C, Nacry P, Bouchez D, Moreau M, Ranjeva R, Thuleau P. 1998. Plasma membrane depolarization-activated calcium channels, stimulated by microtubule-depolymerizing drugs in wild-type Arabidopsis thaliana protoplasts, display constitutively large activities and a longer half-life in ton 2 mutant cells affected in the organization of cortical microtubules. Plant Journal 13: 603–610. Tikhonova LI, Pottosin II, Dietz KJ, Schonknecht G. 1997. Fast-activating cation channel in barley mesophyll vacuoles. Inhibition by calcium. Plant Journal 11: 1059–1070. Turano FJ, Muhitch MJ, Felker FC, McMahon MB. 2002. The putative glutamate receptor 3.2 from Arabidopsis thaliana (AtGLR3.2) is an integral membrane peptide that accumulates in rapidly growing tissues and persists in vascular-associated tissues. Plant Science 163: 43–51. Tyerman SD, Skerrett M, Garrill A, Findlay GP, Leigh RA. 1997. Pathways for the permeation of Na+ and Cl– into protoplasts derived from the cortex of wheat roots. Journal of Experimental Botany 48: 459–480. Tyerman SD, Whitehead LF, Day DA. 1995. A channel-like transporter for NH4+ on the symbiotic interface of N2-fixing plants. Nature 378: 629–632.

© The Authors (2007). Journal compilation © New Phytologist (2007) www.newphytologist.org

New Phytologist (2007) 175: 387–404

403

404 Review

Tansley review

Udvardy J, Farkas GL. 1973. ATP stimulates the formation of nucleases in excised Avena leaves. Zurnal Pflanzenphysiologie 69: 394– 401. Véry A-A, Davies JM. 2000. Hyperpolarization-activated calcium channels at the tip of Arabidopsis root hairs. Proceedings of the National Academy of Sciences, USA 97: 9801–9806. Véry A-A, Robinson MF, Mansfield TA, Sanders D. 1998. Guard cell cation channels are involved in NaCl-induced stomatal closure in a halophyte. Plant Journal 14: 509–521. Véry A-A, Sentenac H. 2003. Molecular mechanisms and regulation of K+ transport in higher plants. Annual Reviews of Plant Biology 54: 575–603. Volkov V, Amtmann A. 2006. Thellungiella halophila, a salt-tolerant relative of Arabidopsis thaliana, has specific root ion-channel features supporting K+/Na+ homeostasis under salinity stress. Plant Journal 48: 342–353. Volkov V, Wang B, Dominy P, Fricke W, Amtmann A. 2004. Thellungiella halophila, a salt-tolerant relative of Arabidopsis thaliana, possesses effective mechanisms to discriminate between potassium and sodium. Plant, Cell & Environment 27: 1–14. Volotovski ID, Sokolovsky SG, Molchan OV, Knight MR. 1998. Second messengers mediate increases in cytosolic calcium in tobacco protoplasts. Plant Physiology 117: 1023 –1030. Walch-Liu P, Liu LH, Remans T, Tester M, Forde BG. 2006. Evidence that 1-glutamate can act as an exogenous signal to modulate root growth and branching in Arabidopsis thaliana. Plant and Cell Physiology 47: 1045–1057. Wang B, Davenport RJ, Volkov V, Amtmann A. 2006. Low unidirectional sodium influx into root cells restricts net sodium accumulation in Thellungiella halophila, a salt-tolerant relative of Arabidopsis thaliana. Journal of Experimental Botany 57: 1161–1170. Wegner LH, Raschke K. 1994. Ion channels in the xylem parenchyma of barley roots. A procedure to isolate protoplasts from this tissue and patch-clamp exploration of salt passageways into xylem vessels. Plant Physiology 105: 799–813. Welch RM. 1995. Micronutrient nutrition of higher plants. Critical Reviews in Plant Sciences 14: 49–82. White PJ. 1997. Cation channels in the plasma membrane of rye roots. Journal of Experimental Botany 48: 499–514. White PJ. 1998. Calcium channels in the plasma membrane of root cells. Annals of Botany 81: 173 –183. White PJ, Bowen HC, Demidchik V, Nichols C, Davies JM. 2002. Genes for calcium-permeable channels in the plasma membrane of plant root cells. Biochimica Biophysica Acta 1564: 299–309.

White PJ, Davenport RJ. 2002. The voltage-independent cation channel in the plasma membrane of wheat roots is permeable to divalent cations and may be involved in cytosolic Ca2+ homeostasis. Plant Physiology 130: 1386–1395. White PJ, Lemtiri-Chlieh F. 1995. Potassium currents across the plasma membrane of protoplasts derived from rye roots: a patch-clamp study. Journal of Experimental Botany 46: 497–511. White PJ, Tester MA. 1992. Potassium channels from the plasma membrane of rye roots characterized following incorporation into planar lipid bilayers. Planta 186: 188–202. Whitehead LF, Day DA, Tyerman SD. 1998. Divalent cation gating of an ammonium permeable channel in the symbiotic membrane of soybean nodules. Plant Journal 16: 313–324. Whitehead LF, Tyerman SD, Day DA. 2001. Polyamines as potential regulators of nutrient exchange across the peribacteroid membrane in soybean root nodules. Australian Journal of Plant Physiology 28: 675– 681. van den Wijngaard PWJ, Bunney TD, Roobeek I, Schonknecht G, de Boer AH. 2001. Slow vacuolar channels from barley mesophyll cells are regulated by 14-3-3 proteins. FEBS Letters 488: 100–104. Yoshioka K, Moeder W, Kang HG, Kachroo P, Masmoudi K, Berkowitz G, Klessig DF. 2006. The chimeric Arabidopsis cyclic nucleotide-gated ion channel11/12 activates multiple pathogen resistance responses. The Plant Cell 18: 747–763. Yu L, Moshelion M, Moran N. 2001. Extracellular protons inhibit the activity of inward-rectifying potassium channels in the motor cells of Samanea saman Pulvini. Plant Physiology 127: 1310–1322. Yurin VM, Sokolik AI, Kudryashev AP. 1991. Regulation of ion transport through plant cell membranes. Minsk, Belarus: Science and Engineering. Zhang WH, Skerrett M, Walker NA, Patrick JW, Tyerman SD. 2002. Nonselective currents and channels in plasma membrane of coat cells in developing Phaseolus vulgaris L. Seeds. Plant Physiology 128: 388–399. Zhang WH, Walker NA, Patrick JW, Tyerman SD. 2004. Calcium-dependent K+ current in plasma membranes of dermal cells of developing bean cotyledons. Plant, Cell & Environment 27: 251–262. Zhang WH, Walker NA, Tyerman SD, Patrick JW. 2000. Fast activation of a time-dependent outward current in protoplasts derived from coats of developing Phaseolus vulgaris seeds. Planta 211: 894–898. Zimmermann S, Nurnberger T, Frachisse JM, Wirtz W, Guern J, Hedrich R, Scheel D. 1997. Receptor-mediated activation of a plant Ca2+-permeable ion channel involved in pathogen defense. Proceedings of the National Academy of Sciences, USA 94: 2751–2755.

About New Phytologist • New Phytologist is owned by a non-profit-making charitable trust dedicated to the promotion of plant science, facilitating projects from symposia to open access for our Tansley reviews. Complete information is available at www.newphytologist.org. • Regular papers, Letters, Research reviews, Rapid reports and both Modelling/Theory and Methods papers are encouraged. We are committed to rapid processing, from online submission through to publication ‘as-ready’ via OnlineEarly – our average submission to decision time is just 30 days. Online-only colour is free, and essential print colour costs will be met if necessary. We also provide 25 offprints as well as a PDF for each article. • For online summaries and ToC alerts, go to the website and click on ‘Journal online’. You can take out a personal subscription to the journal for a fraction of the institutional price. Rates start at £131 in Europe/$244 in the USA & Canada for the online edition (click on ‘Subscribe’ at the website). • If you have any questions, do get in touch with Central Office ([email protected]; tel +44 1524 594691) or, for a local contact in North America, the US Office ([email protected]; tel +1 865 576 5261).

New Phytologist (2007) 175: 387–404

www.newphytologist.org © The Authors (2007). Journal compilation © New Phytologist (2007)